Sample records for o3 hono hcho

  1. Nocturnal Vertical Gradients of O3, NO2, NO3, HONO, HCHO, and SO2 in Los Angeles, CA, during CalNex 2010

    NASA Astrophysics Data System (ADS)

    Tsai, J.; Pikelnaya, O.; Hurlock, S. C.; Wong, K.; Cheung, R.; Haman, C. L.; Lefer, B. L.; Stutz, J.

    2010-12-01

    Nocturnal chemistry, through the conversion and removal of air pollutants, plays an important role in determining the initial condition for photochemistry during the following day. In the stable nocturnal boundary layer (NBL) the interplay between suppressed vertical mixing and surface emissions of NOx and VOCs can result in pronounced vertical trace gas profiles. The resulting altitude dependence of nocturnal chemistry makes the interpretation of ground observations challenging. In particular, the quantification of the nocturnal loss of NOx, due to NO3 and N2O5 chemistry, requires observations throughout the entire vertical extent of the NBL. The formation of daytime radical precursors, such as HONO, is also altitude dependent. An accurate assessment of their impact on daytime chemistry requires measurements of their profiles during the night and morning. Here we present observations from the CalNex-LA experiment, which took place from May 15 to June 15, 2010 on the east side of the Los Angeles Basin, CA. A Long-Path Differential Optical Absorption Spectrometer (LP-DOAS) was set up on the roof of the Millikan library (265 m asl, 35m agl) on the campus of the California Institute of Technology. Four retroreflector arrays were mounted about 5 -7 km North-East of the instrument at 310m, 353m, 487m and 788 m asl. The vertical profiles of NO3, HONO, NO2, O3, HCHO, and SO2 were retrieved at altitude intervals of 35-78m, 78-121m, 121-255m and 255-556m above the ground. During many nights vertical gradients were observed, with elevated NO2 and HONO concentrations near the surface and larger ozone and NO3 concentrations aloft. Simultaneous ceilometer observations of the NBL structure show the impact of meteorology on the vertical trace gas distributions. We will discuss the consequences of trace gases gradients on the nocturnal NOx budget.

  2. Evaluating emissions of HCHO, HONO, NO2, and SO2 from point sources using portable Imaging DOAS

    NASA Astrophysics Data System (ADS)

    Pikelnaya, O.; Tsai, C.; Herndon, S. C.; Wood, E. C.; Fu, D.; Lefer, B. L.; Flynn, J. H.; Stutz, J.

    2011-12-01

    Our ability to quantitatively describe urban air pollution to a large extent depends on an accurate understanding of anthropogenic emissions. In areas with a high density of individual point sources of pollution, such as petrochemical facilities with multiple flares or regions with active commercial ship traffic, this is particularly challenging as access to facilities and ships is often restricted. Direct formaldehyde emissions from flares may play an important role for ozone chemistry, acting as an initial radical precursor and enhancing the degradation of co-emitted hydrocarbons. HONO is also recognized as an important OH source throughout the day. However, very little is known about direct HCHO and HONO emissions. Imaging Differential Optical Absorption Spectroscopy (I-DOAS), a relatively new remote sensing technique, provides an opportunity to investigate emissions from these sources from a distance, making this technique attractive for fence-line monitoring. In this presentation, we will describe I-DOAS measurements during the FLAIR campaign in the spring/summer of 2009. We performed measurements outside of various industrial facilities in the larger Houston area as well as in the Houston Ship Channel to visualize and quantify the emissions of HCHO, NO2, HONO, and SO2 from flares of petrochemical facilities and ship smoke stacks. We will present the column density images of pollutant plumes as well as fluxes from individual flares calculated from I-DOAS observations. Fluxes from individual flares and smoke stacks determined from the I-DOAS measurements vary widely in time and by the emission sources. We will also present HONO/NOx ratios in ship smoke stacks derived from the combination of I-DOAS and in-situ measurements, and discuss other trace gas ratios in plumes derived from the I-DOAS observations. Finally, we will show images of HCHO, NO2 and SO2 plumes from control burn forest fires observed in November of 2009 at Vandenberg Air Force Base, Santa Maria

  3. Vertical columns of NO2, HONO, HCHO, CHOCHO and aerosol extinction: diurnal and seasonal variations in context of CalNex and CARES

    NASA Astrophysics Data System (ADS)

    Ortega, I.; Coburn, S.; Oetjen, H.; Sinreich, R.; Thalman, R. M.; Waxman, E.; Volkamer, R.

    2011-12-01

    We present results from two ground-based University of Colorado Multi Axis Differential Optical Absorption Spectroscopy (CU-MAX-DOAS) instruments that were deployed during the CALNEX and CARES 2010 field campaigns. Ground based CU-MAX-DOAS measurements were carried out through Dec 2010, and measured vertical column abundances of nitrogen dioxide (NO2), nitrous acid (HONO), formaldehyde (HCHO), glyoxal (CHOCHO), and aerosol extinction, which is determined indirectly from observing the oxygen dimers (O4). The measurements were acquired on the top of Millikan library at Caltech, Pasadena, CA, at the Fontana Arrows site located 60 Km east of Caltech, and for a limited period also downwind of Sacramento at T1 site during CARES. In the South Coast Air Basin, the MAX-DOAS instruments at both sites collected an extended time series of use to test satellites, and atmospheric chemistry models. We determine the state of the planetary boundary layer by comparing the columns observations with in-situ sensors, and place the CALNEX and CARES measurements intensive into seasonal context.

  4. DFT study on the interaction of TiO2 (001) surface with HCHO molecules

    NASA Astrophysics Data System (ADS)

    Wu, Guofei; Zhao, Cuihua; Guo, Changqing; Chen, Jianhua; Zhang, Yibing; Li, Yuqiong

    2018-01-01

    The interactions of formaldehyde (HCHO) molecule with TiO2 (001) surface were studied using density functional theory calculations. HCHO molecules are dissociated by the cleavage of Csbnd H bonds after adsorption on TiO2 surface. The strong interactions between HCHO melecules and TiO2 surface are largely attributed to the bonding of hydrogen of HCHO and oxygen of TiO2 surface, which is mainly from the hybridization of the H 1s, O 2p and O 2s. The newly formed Hsbnd O bonds cause the structure changes of TiO2 surface, and lead to the cleavage of Osbnd Ti bond of TiO2 surface. The Csbnd O bond that the dissociated remains of HCHO and newly formed Hsbnd O bond can be oxidized to form carbon dioxide and water in subsequent action by oxygen from the atomosphere. The charges transfer from HCHO to TiO2 surface, and the sum amount of the charges transferred from four HCHO molecules to TiO2 surface is bigger than that from one HCHO molecule to TiO2 surface due to the combined interaction of four HCHO molecules with TiO2 surface.

  5. Template-free fabrication of hierarchical In2O3 hollow microspheres with superior HCHO-sensing properties

    NASA Astrophysics Data System (ADS)

    Zhang, Su; Song, Peng; Tian, Zhebin; Wang, Qi

    2018-05-01

    Hierarchical In2O3 hollow microspheres were successfully prepared via a facile and low-cost hydrothermal method. Their morphology and structure were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), and the Brunauer-Emmett-Teller (BET) approach. The SEM and TEM results revealed that the as-obtained hollow In2O3 microspheres is composed of In2O3 nanospheres with 200-400 nm in diameter, and the size of In2O3 microspheres is about 2-4 μm. The specific surface area of the as-prepared In2O3 is about 40.94 m2/g. The sensor based on hierarchical In2O3 hollow microspheres displays excellent sensing properties to 10 ppm HCHO, and the optimum operating temperature is relatively low (200 °C). The response value of the as-fabricated sensor to 10 ppm HCHO is about 20. Due to the sensor based on hierarchical In2O3 hollow microspheres has many advantages, such as facile preparation and excellent gas-sensing properties, it has a wide range of prospects in practical applications.

  6. Sources of Nitrous Acid, Formaldehyde, and Hydroxyl Radical in Doha, Qatar.

    NASA Astrophysics Data System (ADS)

    Ackermann, Luis; Rappenglueck, Bernhard; Ayoub, Mohammed

    2017-04-01

    One of the most important species in the atmosphere is the hydroxyl radical (OH), due to its role controlling the oxidizing capacity of an air shed. The main formation processes of OH include the photolysis of ozone (O3), nitrous acid (HONO), formaldehyde (HCHO), and the ozonolysis of alkenes. Still, the sources of HONO in the atmosphere are not sufficiently well known, with indications that heterogeneous reactions on surfaces may contribute to the observed concentrations. The city of Doha in Qatar presents a unique opportunity to explore photochemical processes including the effects of high particulates concentrations under extreme weather conditions (high temperatures and humidity) and complex emission sources. Two Intensive Observational Periods (IOP) were conducted in Doha in 2016, one during the winter and the other during the summer. These consisted of meteorological measurements, ozone (O3), nitrous acid (HONO), formaldehyde (HCHO), nitrogen monoxide (NO), direct nitrogen dioxide (NO2), sulfur dioxide (SO2), carbon monoxide (CO), as well as particulate matter with an aerodynamic diameter ≤ 10 μm and 2.5 μm (PM10 and PM2.5). In addition photolysis rates of HONO, HCHO, NO2, and singlet oxygen (O1D) were measured. The photostationary state concentration of OH was calculated from its known sources and sinks. The maximum hourly average concentration of OH was determined to be around 1.1 ppt for summer and 0.5 ppt for winter IOP. For the 24-hr average, the photolysis of HONO was the main precursor for OH production with 54.3 % and 72.7 % (summer and winter IOP), while the photolysis of O3 was responsible for 23.8 % and 19.7 % and the photolysis of HCHO accounted for 21.9 % and 7.6 % (summer and winter IOP, respectively). In this study we present source apportionment analysis for the radical precursors HONO and HCHO during the winter and summer IOP and its diurnal variation and elucidate their impact on OH production. We also infer NOx vs VOC limitation of O3

  7. Co3O4 nanorod-supported Pt with enhanced performance for catalytic HCHO oxidation at room temperature

    NASA Astrophysics Data System (ADS)

    Yan, Zhaoxiong; Xu, Zhihua; Cheng, Bei; Jiang, Chuanjia

    2017-05-01

    Formaldehyde (HCHO) removal from air at room (ambient) temperature by effective catalysts is of significance for improving indoor air quality, and catalysts with high efficiency and good recyclability are highly desirable. In this study, platinum (Pt) supported on nanorod-shaped Co3O4 (Pt/Co3O4) was prepared by calcination of microwave-assisted synthesized Co3O4 precursor followed by NaBH4-reduction of Pt precursor. The as-prepared Co3O4 exhibited a morphology of nanorods with lengths of 400-700 nm and diameters of approximately 40-50 nm, which were self-assembled by nanoparticles. The Pt/Co3O4 catalyst exhibited a superior catalytic performance for HCHO oxidation at room temperature compared to Pt supported on commercial Co3O4 (Pt/Co3O4-c) and Pt supported on commercial TiO2 (Pt/TiO2), which is mainly due to the high oxygen mobility resulting from its distinct nanorod morphology, strong metal-support interaction between Pt and Co3O4, and the intrinsic redox nature of the Co3O4 support. This study provides new insights into the fabrication of high-performance catalysts for indoor air purification.

  8. Vibrational spectroscopy of NO + (H2O)n: Evidence for the intracluster reaction NO + (H2O)n --> H3O + (H2O)n - 2 (HONO) at n => 4

    NASA Astrophysics Data System (ADS)

    Choi, Jong-Ho; Kuwata, Keith T.; Haas, Bernd-Michael; Cao, Yibin; Johnson, Matthew S.; Okumura, Mitchio

    1994-05-01

    Infrared spectra of mass-selected clusters NO+(H2O)n for n=1 to 5 were recorded from 2700 to 3800 cm-1 by vibrational predissociation spectroscopy. Vibrational frequencies and intensities were also calculated for n=1 and 2 at the second-order Møller-Plesset (MP2) level, to aid in the interpretation of the spectra, and at the singles and doubles coupled cluster (CCSD) level energies of n=1 isomers were computed at the MP2 geometries. The smaller clusters (n=1 to 3) were complexes of H2O ligands bound to a nitrosonium ion NO+ core. They possessed perturbed H2O stretch bands and dissociated by loss of H2O. The H2O antisymmetric stretch was absent in n=1 and gradually increased in intensity with n. In the n=4 clusters, we found evidence for the beginning of a second solvation shell as well as the onset of an intracluster reaction that formed HONO. These clusters exhibited additional weak, broad bands between 3200 and 3400 cm-1 and two new minor photodissociation channels, loss of HONO and loss of two H2O molecules. The reaction appeared to go to completion within the n=5 clusters. The primary dissociation channel was loss of HONO, and seven vibrational bands were observed. From an analysis of the spectrum, we concluded that the n=5 cluster rearranged to form H3O+(H2O)3(HONO), i.e., an adduct of the reaction products.

  9. SO2 Initiates the Efficient Conversion of NO2 to HONO on MgO Surface.

    PubMed

    Ma, Qingxin; Wang, Tao; Liu, Chang; He, Hong; Wang, Zhe; Wang, Weihao; Liang, Yutong

    2017-04-04

    Nitrous acid (HONO) is an important source of hydroxyl radical (OH) that determines the fate of many chemically active and climate relevant trace gases. However, the sources and the formation mechanisms of HONO remain poorly understood. In this study, the effect of SO 2 on the heterogeneous reactions of NO 2 on MgO as a mineral dust surrogate was investigated. The reactivity of MgO to NO 2 is weak, while coexisting SO 2 can increase the uptake coefficients of NO 2 on MgO by 2-3 orders of magnitude. The uptake coefficients of NO 2 on SO 2 -aged MgO are independent of NO 2 concentrations in the range of 20-160 ppbv and relative humidity (0-70%RH). The reaction mechanism was demonstrated to be a redox reaction between NO 2 and surface sulfite. In the presence of SO 2 , NO 2 was reduced to nitrite under dry conditions, which could be further converted to gas-phase HONO in humid conditions. These results suggest that the reductive effect of SO 2 on the heterogeneous conversion of NO 2 to HONO may have a significant contribution to the unknown sources of HONO observed in polluted areas (for example, in China).

  10. LIF instrument development, in situ measurement at South Pole and one-dimensional air-snowpack modeling of atmospheric nitrous acid (HONO)

    NASA Astrophysics Data System (ADS)

    Liao, Wei

    Atmospheric nitrous acid (HONO) is a significant and sometimes dominant OH source in Polar Regions. In the polar atmosphere, measurements of HONO are an important part of understanding the dynamics of snow-air chemistry and atmospheric photochemistry. The low levels of HONO present in such regions necessitate the development of instrumentation with low detection limits. An improved method of detecting HONO is developed using photo-fragmentation and laser-induced fluorescence. The detection limit of this method is 2-3 pptv for ten-minute integration time with 35% uncertainty. The ANTCI 2003 measurements confirm the high N oxides observed previously in ISCAT 1998 and 2000. The median LIF observed mixing ratio of HONO 10m above the snow was 5.8 pptv (mean value 6.3 pptv) with a maximum of 18.2pptv on Nov 30th, Dec 1st, 3rd, 15th, 17th, 21st, 22nd, 25th, 27th and 28th. The LIF HONO observations are compared to concurrent HONO observations performed by mist chamber/ion chromatography (MC/IC). Both the LIF and MC/IC techniques observed enhanced HONO; however, the MC/IC observations were higher than the LIF observations by a factor of 7.2+/-2.3 in the median. It is suggested that the MC/IC technique might suffer from interference from HNO4. As in ISCAT 2000, the abundance of both HONO measurements exceeds the pure gas phase model predictions, with LIF higher than the pure gas phase model by a factor of 1.92+/-0.67, which implies snow emission of HONO must occur. The LIF measured HONO concentrations are not high enough to significantly influence the NOx budget during ANTCI 2003, but will increase the modeled HOx over-prediction by 28%+/-15% and lead to a dramatic over-prediction of measured OH by 157%+/-35%. Given the short lifetime of HONO, these differences are hard to reconcile with observed low OH levels unless there is a missing HO x sink. It appears, however, that HONO competes with O3 and HCHO as the dominant source of OH at South Pole during ANTCI 2003. Since pure

  11. Investigating isotopic signatures of atmospheric nitrous acid (HONO)

    NASA Astrophysics Data System (ADS)

    Chai, J.; Miller, D. J.; Hastings, M. G.

    2016-12-01

    Nitrous acid (HONO) is an important reactive nitrogen species that can be easily photolyzed to nitrogen oxide and hydroxyl radical in the troposphere. HONO greatly influences atmospheric oxidation capacity, affecting the formation of tropospheric ozone (O3) and secondary aerosol. Recent studies have indicated that in addition to heterogeneous NOx reactions, biomass burning, soil emission and photolysis of nitric acid (HNO3) on surfaces (e.g. aerosol particles and soot) are also important sources of HONO. However, these sources have not yet been well constrained. The stable isotope ratios in nitrate have been successfully used to trace NOx sources and oxidation chemistry in the atmosphere. Can the isotopic signatures of HONO be used to trace NOx oxidation and renoxification pathways? For this purpose, we have built an annular denuder HONO collection system for the stable isotope study of HONO. Preliminary tests show successful collection and recovery of HONO synthesized in our lab. Nitrogen and oxygen isotopic analysis of the recovered HONO also shows consistent isotopic signatures. Results from field applications of this method in near road and on road environments, agricultural settings, and laboratory based biomass burns will be presented.

  12. HONO (nitrous acid) emissions from acidic northern soils

    NASA Astrophysics Data System (ADS)

    Maljanen, Marja; Yli-Pirilä, Pasi; Joutsensaari, Jorma; Martikainen, Pertti J.

    2015-04-01

    The photolysis of HONO (nitrous acid) is an important source of OH radical, the key oxidizing agent in the atmosphere, contributing also to removal of atmospheric methane (CH4), the second most important greenhouse gas after carbon dioxide (CO2). The emissions of HONO from soils have been recently reported in few studies. Soil HONO emissions are regarded as missing sources of HONO when considering the chemical reactions in the atmosphere. The soil-derived HONO has been connected to soil nitrite (NO2-) and also directly to the activity of ammonia oxidizing bacteria, which has been studied with one pure culture. Our hypothesis was that boreal acidic soils with high nitrification activity could be also sources of HONO and the emissions of HONO are connected with nitrification. We selected a range of dominant northern acidic soils and showed in microcosm experiments that soils which have the highest nitrous oxide (N2O) and nitric oxide (NO) emissions (drained peatlands) also have the highest HONO production rates. The emissions of HONO are thus linked to nitrogen cycle and also NO and N2O emissions. Natural peatlands and boreal coniferous forests on mineral soils had the lowest HONO emissions. It is known that in natural peatlands with high water table and in boreal coniferous forest soils, low nitrification activity (microbial production of nitrite and nitrate) limits their N2O production. Low availability of nitrite in these soils is the likely reason also for their low HONO production rates. We also studied the origin of HONO in one peat soil with acetylene and other nitrification inhibitors and we found that HONO production is not closely connected to ammonium oxidation (nitrification). Acetylene blocked NO emissions but did not affect HONO or N2O emissions, thus there is another source behind HONO emission from these soils than ammonium oxidation. It is still an open question if this process is microbial or chemical origin.

  13. Radical Sources in the Uintah Basin during 2013 Winter Ozone Episodes

    NASA Astrophysics Data System (ADS)

    Roberts, J. M.; Yuan, B.; Veres, P. R.; Warneke, C.; De Gouw, J. A.; Geiger, F.; Brown, S. S.; Edwards, P. M.; Wild, R.; Min, K.; Bates, T. S.; Quinn, P.; Banta, R. M.; Zamora, R. J.; McLaren, R.; Young, C.; Kercher, J. P.; Thornton, J. A.; Williams, E. J.

    2013-12-01

    Winter time O3 in excess of the NAAQS, 75 ppbv, has been observed in several geographic basins in Wyoming and Utah that are heavily impacted by emissions from oil and gas operations. The timing and circumstances of these high O3 events imply that radical sources such as HONO, HCHO, and perhaps ClNO2 are significant relative to the traditional O3-photolysis channel. Here we present data from the 2013 Uintah Basin Winter Ozone Study (UBWOS) that show that HONO and HCHO were the major sources of radicals during O3 episodes. This result stands in contrast to the results obtained in more typical urban atmospheres, such as the CalNEx 2010 measurements in Pasadena, where O3 photolysis was found to be the major radical source. The precise contribution of each radical source during UBWOS 2013 awaits further work on the fluxes to and from snow surfaces, and verification of HONO measurement techniques. Such a coupling of radical and NOx sources complicates the traditional NOx vs.VOC paradigm in which one or the other quantity determines the best O3 control strategy. This amplifies the need for a quantitative understanding of NOx to HONO conversion mechanisms.

  14. HONO (nitrous acid) emissions from acidic northern soils

    NASA Astrophysics Data System (ADS)

    Maljanen, Marja; Yli-Pirilä, Pasi; Joutsensaari, Jorma; Sulassaari, Sirkka; Martikainen, Pertti J.

    2014-05-01

    The photolysis of HONO (nitrous acid) is an important source of OH radical, the key oxidizing agent in the atmosphere, contributing also to removal of atmospheric methane (CH4), the second most important greenhouse gas after carbon dioxide (CO2). There are missing sources of HONO when considering the chemical reactions in the atmosphere. Soil could be such a missing source. Emissions of HONO from soils studied in laboratory incubations have been recently reported. The soil-derived HONO has been connected to soil nitrite (NO2-) and a study with an ammonium oxidizing bacterium has shown that HONO could be produced in ammonium oxidation. Our hypothesis was that boreal acidic soils with high nitrification activity could be important sources of HONO. We selected a range of dominant northern acidic soils and showed in microcosm experiments that soils which have the highest nitrous oxide (N2O) and nitric oxide (NO) emissions (drained peatlands) also have the highest HONO production rates. The emissions of HONO are thus linked to nitrogen cycle processes. In contrast to drained peatlands, natural peatlands with high water table and boreal coniferous forests on mineral soils with low nitrification capacity had low HONO emissions. It is known that in natural peatlands with high water table and in boreal coniferous forest soils, low nitrification activity (microbial production of nitrite and nitrate) limits their N2O production. Low nitrification rate and low availability of nitrite in these soils are the likely reasons for their low HONO production rates. We studied the origin of HONO in one drained peat soil by inhibiting nitrification with acetylene. Acetylene blocked NO emissions but did not affect HONO or N2O emissions, thus ammonium oxidation is not the direct mechanism for the HONO emission in this soil. It is still an open question if HONO originates directly from some microbial process like ammonium oxidation or chemically from nitrite produced in microbial processes.

  15. Emissions of nitrous acid (HONO), nitric oxide (NO) and nitrous oxide (N2O) from boreal agricultural soil - Effect of N fertilization

    NASA Astrophysics Data System (ADS)

    Bhattarai, Hem Raj; Virkajärvi, Perttu; -Yli Pirilä, Pasi; Maljanen, Marja

    2017-04-01

    There is no doubt that nitrogen (N) fertilization has crucial role in increasing food production. However, in parallel it can cause severe impact in environment such as eutrophication, surface/groundwater pollution via nitrate (NO3-) leaching and emissions of N trace gases. Fertilization increases the emissions of nitrous oxide (N2O) which is 260 stronger greenhouse gas than carbon dioxide (CO2). It also enhances the emissions of nitric oxide (NO); an oxidized and very reactive form of nitrogen which can fluctuate the ozone (O3) concentration in atmosphere and cause acidification. The effects of N- fertilization on the emission of N2O and NO from agricultural soil are well known. However, the effects of N fertilization on nitrous acid (HONO) emissions are unknown. Few studies have shown that HONO is emitted from soil but they lack to interlink fertilization and HONO emission. HONO accounts for 17-34 % of hydroxyl (OH-) radical production? in the atmosphere, OH- radicals have vital role in atmospheric chemistry; they can cause photochemical smog, form O3, oxidize volatile organic compounds and also atmospheric methane (CH4). We formulated hypothesis that N fertilization will increase the HONO emissions as it does for N2O and NO. To study this, we took soil samples from agricultural soil receiving different amount of N-fertilizer (0, 250 and 450 kg ha-1) in eastern Finland. HONO emissions were measured by dynamic chamber technique connected with LOPAP (Quma Elektronik & Analytik GmbH), NO by NOx analyzer (Thermo scientific) and static chamber technique and gas chromatograph was used for N2O gas sampling and analysis. Several soil parameters were also measured to establish the relationship between the soil properties, fertilization rate and HONO emission. This study is important because eventually it will open up more questions regarding the forms of N loss from soils and impact of fertilization on atmospheric chemistry.

  16. A method for the measurement of atmospheric HONO based on DNPH derivatization and HPLC analysis

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Zhou, X.; Qiao, H.; Deng, G.

    1999-10-15

    A simple measurement technique was developed for atmospheric HONO based on aqueous scrubbing using a coil sampler followed by 2,4-dinitrophenylhydrazine (DNPH) derivatization and high-performance liquid chromatographic (HPLC) analysis. Quantitative sampling efficiency was obtained using a 1 mM phosphate buffer, pH 7.0, as the scrubbing solution at a gas sampling flow rate of 2 L min{sup {minus}1} and a liquid flow rate of 0.24 mL min{sup {minus}1}. Derivation of the scrubbed nitrous acid by DNPH was fast and was completed within 5 min in a derivatization medium containing 300 {micro}M DNPH and 8 mM HCI at 45 C. The azide derivativemore » was separated from DNPH reagent and carbonyl derivatives by reverse-phase HPLC and was detected with an UV detector at 309 nm. The detection limit is {le}5 pptv and may be lowered to 1 pptv with further DNPH purification. Interferences from NO, NO{sub 2} PAN, O{sub 3}, HNO{sub 3}, and HCHO were studied and found to be negligible. Ambient HONO concentration was measured simultaneously in downtown Albany, NY, by this method and by an ion chromatographic technique after sampling using a fritted bubbler. The results, from 70 pptv during the day to 1.7 ppbv in the early morning, were in very good agreement from the two techniques, within {+-} 20%.« less

  17. Global CO emission estimates inferred from assimilation of MOPITT and IASI CO data, together with observations of O3, NO2, HNO3, and HCHO.

    NASA Astrophysics Data System (ADS)

    Zhang, X.; Jones, D. B. A.; Keller, M.; Jiang, Z.; Bourassa, A. E.; Degenstein, D. A.; Clerbaux, C.; Pierre-Francois, C.

    2017-12-01

    Atmospheric carbon monoxide (CO) emissions estimated from inverse modeling analyses exhibit large uncertainties, due, in part, to discrepancies in the tropospheric chemistry in atmospheric models. We attempt to reduce the uncertainties in CO emission estimates by constraining the modeled abundance of ozone (O3), nitrogen dioxide (NO2), nitric acid (HNO3), and formaldehyde (HCHO), which are constituents that play a key role in tropospheric chemistry. Using the GEOS-Chem four-dimensional variational (4D-Var) data assimilation system, we estimate CO emissions by assimilating observations of CO from the Measurement of Pollution In the Troposphere (MOPITT) and the Infrared Atmospheric Sounding Interferometer (IASI), together with observations of O3 from the Optical Spectrograph and InfraRed Imager System (OSIRIS) and IASI, NO2 and HCHO from the Ozone Monitoring Instrument (OMI), and HNO3 from the Microwave Limb Sounder (MLS). Our experiments evaluate the inferred CO emission estimates from major anthropogenic, biomass burning and biogenic sources. Moreover, we also infer surface emissions of nitrogen oxides (NOx = NO + NO2) and isoprene. Our results reveal that this multiple species chemical data assimilation produces a chemical consistent state that effectively adjusts the CO-O3-OH coupling in the model. The O3-induced changes in OH are particularly large in the tropics. Overall, our analysis results in a better constrained tropospheric chemical state.

  18. Ab initio electron correlated studies on the intracluster reaction of NO+ (H2O)(n) → H3O+ (H2O)(n-2) (HONO) (n = 4 and 5).

    PubMed

    Asada, Toshio; Nagaoka, Masataka; Koseki, Shiro

    2011-01-28

    Hydrated nitrosonium ion clusters NO(+)(H(2)O)(n) (n = 4 and 5) were investigated by using MP2/aug-cc-pVTZ level of theory to clarify isomeric reaction pathways for formation of HONO and fully hydrated hydride ions. We found some new isomers and transition state structures in each hydration number, whose lowest activation energies of the intracluster reactions were found to be 4.1 and 3.4 kcal mol(-1) for n = 4 and n = 5, respectively. These thermodynamic properties and full quantum mechanical molecular dynamics simulation suggest that product isomers with HONO and fully hydrated hydride ions can be obtained at n = 4 and n = 5 in terms of excess hydration binding energies which can overcome these activation barriers.

  19. Fluxes of Nitrous Acid (HONO) above an Agricultural Field Side near Paris

    NASA Astrophysics Data System (ADS)

    Laufs, S.; Cazaunau, M.; Stella, P.; Loubet, B.; Kurtenbach, R.; Cellier, P.; Mellouki, W.; Kleffmann, J.

    2012-04-01

    HONO is an important precursor of the OH radical, the detergent of the atmosphere. Field measurements show high diurnal HONO mixing ratios that cannot be explained by chemical models with known gas phase chemistry. Therefore, daytime sources of HONO are still under discussion. During the last decade many experimental investigation were performed to study heterogeneous production of HONO like the photo enhanced reduction of NO2 on humic acids or photolysis of HNO3 on surfaces. Recently, nitrite produced by bacteria, present in soil, was discussed as a source of HONO as well. In addition gas phase sources like the photolysis of nitrophenols, or the reaction of excited NO2 are discussed. Gradient measurements show high mixing ratios of HONO even above the boundary layer. However, beside intensive investigations on the sources of HONO, it is still an open question whether heterogeneous or gas phase sources are more important in the atmosphere. Flux measurements could represent a method to find the origin of missing sources of HONO. Until now instruments are not sensitive and fast enough to do Eddy correlation measurements for HONO. Alternatively, HONO fluxes are estimated by the Aerodynamic Gradient (AGM), or Relaxed Eddy Accumulation (REA) methods. Here we present HONO fluxes estimated by AGM and the LOPAP technique (Long Path Absorption Photometer) above an agricultural field in Grignon, Paris (48°51'N, 1°58'E). Fluxes during different seasons and different types of vegetations including bare soil will be presented and compared with chemical corrected fluxes of NO, NO2 and O3, or other parameters.

  20. Seasonal behavior of carbonyls and source characterization of formaldehyde (HCHO) in ambient air

    NASA Astrophysics Data System (ADS)

    Lui, K. H.; Ho, Steven Sai Hang; Louie, Peter K. K.; Chan, C. S.; Lee, S. C.; Hu, Di; Chan, P. W.; Lee, Jeffrey Chi Wai; Ho, K. F.

    2017-03-01

    Gas-phase formaldehyde (HCHO) is an intermediate and a sensitive indicator for volatile organic compounds (VOCs) oxidation, which drives tropospheric ozone production. Effective photochemical pollution control strategies demand a thorough understanding of photochemical oxidation precursors, making differentiation between sources of primary and secondary generated HCHO inevitable. Spatial and seasonal variations of airborne carbonyls based on two years of measurements (2012-2013), coupled with a correlation-based HCHO source apportionment analysis, were determined for three sampling locations in Hong Kong (denoted HT, TC, and YL). Formaldehyde and acetaldehyde were the two most abundant compounds of the total quantified carbonyls. Pearson's correlation analysis (r > 0.7) implies that formaldehyde and acetaldehyde possibly share similar sources. The total carbonyl concentration trends (HT < TC < YL) reflect location characteristics (urban > rural). A regression analysis further quantifies the relative primary HCHO source contributions at HT (∼13%), TC (∼21%), and YL (∼40%), showing more direct vehicular emissions in urban than rural areas. Relative secondary source contributions at YL (∼36%) and TC (∼31%) resemble each other, implying similar urban source contributions. Relative background source contributions at TC could be due to a closed structure microenvironment that favors the trapping of HCHO. Comparable seasonal differences are observed at all stations. The results of this study will aid in the development of a new regional ozone (O3) control policy, as ambient HCHO can enhance O3 production and also be produced from atmospheric VOCs oxidation (secondary HCHO).

  1. Deciphering the role of radical precursors during the Second Texas Air Quality Study.

    PubMed

    Olaguer, Eduardo P; Rappenglück, Bernhard; Lefer, Barry; Stutz, Jochen; Dibb, Jack; Griffin, Robert; Brune, William H; Shauck, Maxwell; Buhr, Martin; Jeffries, Harvey; Vizuete, William; Pinto, Joseph P

    2009-11-01

    The Texas Environmental Research Consortium (TERC) funded significant components of the Second Texas Air Quality Study (TexAQS II), including the TexAQS II Radical and Aerosol Measurement Project (TRAMP) and instrumented flights by a Piper Aztec aircraft. These experiments called attention to the role of short-lived radical sources such as formaldehyde (HCHO) and nitrous acid (HONO) in increasing ozone productivity. TRAMP instruments recorded daytime HCHO pulses as large as 32 parts per billion (ppb) originating from upwind industrial activities in the Houston Ship Channel, where in situ surface monitors detected HCHO peaks as large as 52 ppb. Moreover, Ship Channel petrochemical flares were observed to produce plumes of apparent primary HCHO. In one such combustion plume that was depleted of ozone by large emissions of oxides of nitrogen (NOx), the Piper Aztec measured a ratio of HCHO to carbon monoxide (CO) 3 times that of mobile sources. HCHO from uncounted primary sources or ozonolysis of underestimated olefin emissions could significantly increase ozone productivity in Houston beyond previous expectations. Simulations with the CAMx model show that additional emissions of HCHO from industrial flares or mobile sources can increase peak ozone in Houston by up to 30 ppb. Other findings from TexAQS II include significant concentrations of HONO throughout the day, well in excess of current air quality model predictions, with large nocturnal vertical gradients indicating a surface or near-surface source of HONO, and large concentrations of nighttime radicals (approximately30 parts per trillion [ppt] HO2). HONO may be formed heterogeneously on urban canopy or particulate matter surfaces and may be enhanced by organic aerosol of industrial or motor vehicular origin, such as through conversion of nitric acid (HNO3). Additional HONO sources may increase daytime ozone by more than 10 ppb. Improving the representation of primary and secondary HCHO and HONO in air quality

  2. Investigation of a potential HCHO measurement artifact from ISOPOOH

    PubMed Central

    St. Clair, Jason M.; Rivera-Rios, Jean C.; Crounse, John D.; Praske, Eric; Kim, Michelle J.; Wolfe, Glenn M.; Keutsch, Frank N.; Wennberg, Paul O.; Hanisco, Thomas F.

    2018-01-01

    Recent laboratory experiments have shown that a first generation isoprene oxidation product, ISOPOOH, can decompose to methyl vinyl ketone (MVK) and methacrolein (MACR) on instrument surfaces, leading to overestimates of MVK and MACR concentrations. Formaldehyde (HCHO) was suggested as a decomposition co-product, raising concern that in situ HCHO measurements may also be affected by an ISOPOOH interference. The HCHO measurement artifact from ISOPOOH for the NASA In Situ Airborne Formaldehyde instrument (ISAF) was investigated for the two major ISOPOOH isomers, (1,2)-ISOPOOH and (4,3)-ISOPOOH, under dry and humid conditions. The dry conversion of ISOPOOH to HCHO was 3±2% and 6±4% for (1,2)-ISOPOOH and (4,3)-ISOPOOH, respectively. Under humid (RH= 40-60%) conditions, conversion to HCHO was 6±4% for (1,2)-ISOPOOH and 10±5% for (4,3)-ISOPOOH. The measurement artifact caused by conversion of ISOPOOH to HCHO in the ISAF instrument was estimated for data obtained on the 2013 September 6 flight of the Studies of Emissions and Atmospheric Composition, Clouds and Climate Coupling by Regional Surveys (SEAC4RS) campaign. Prompt ISOPOOH conversion to HCHO was the source for <4% of the observed HCHO, including in the high-isoprene boundary layer. Time-delayed conversion, where previous exposure to ISOPOOH affects measured HCHO later in flight, was conservatively estimated to be < 10% of observed HCHO and is significant only when high ISOPOOH sampling periods immediately precede periods of low HCHO. PMID:29636831

  3. Investigation of a Potential HCHO Measurement Artifact from ISOPOOH

    NASA Technical Reports Server (NTRS)

    St Clair, Jason M.; Rivera-Rios, Jean C.; Crounse, John D.; Praske, Eric; Kim, Michelle J.; Wolfe, Glenn M.; Keutche, Frank N.; Wennberg, Paul O.; Hanisco, Thomas F.

    2016-01-01

    Recent laboratory experiments have shown that a first generation isoprene oxidation product, ISOPOOH, can decompose to methyl vinyl ketone (MVK) and methacrolein (MACR) on instrument surfaces, leading to overestimates of MVK and MACR concentrations. Formaldehyde (HCHO) was suggested as a decomposition co-product, raising concern that in situ HCHO measurements may also be affected by an ISOPOOH interference. The HCHO measurement artifact from ISOPOOH for the NASA In Situ Airborne Formaldehyde instrument (ISAF) was investigated for the two major ISOPOOH isomers, (1,2)-ISOPOOH and (4,3)-ISOPOOH, under dry and humid conditions. The dry conversion of ISOPOOH to HCHO was 3+/-2% and 6+/-4% for (1,2)-ISOPOOH and (4,3)-ISOPOOH, respectively. Under humid (RH= 40-60%) conditions, conversion to HCHO was 6+/-4% for (1,2)-ISOPOOH and 10+/-5% for (4,3)-ISOPOOH. The measurement artifact caused by conversion of ISOPOOH to HCHO in the ISAF instrument was estimated for data obtained on the 2013 September 6 flight of the Studies of Emissions and Atmospheric Composition, Clouds and Climate Coupling by Regional Surveys (SEAC4RS) campaign. Prompt ISOPOOH conversion to HCHO was the source for <4% of the observed HCHO, including in the high-isoprene boundary layer. Time-delayed conversion, where previous exposure to ISOPOOH affects measured HCHO later in flight, was conservatively estimated to be < 10% of observed HCHO and is significant only when high ISOPOOH sampling periods immediately precede periods of low HCHO.

  4. Ambient formaldehyde and its contributing factor to ozone and OH radical in a rural area

    NASA Astrophysics Data System (ADS)

    Xiaoyan, Wang; Huixiang, Wang; Shaoli, Wang

    2010-06-01

    Formaldehyde (HCHO), as well as correlative pollutants was measured from 1 to 31 July in 2007 at Mazhuang, a rural site located in the east of China. Gaseous HCHO was scrubbed from the air with an acidic 2,4-dinitrophenylhydrazine (DNPH) solution, which leaded to the reaction of HCHO with DNPH and produced a stable product, 2,4-dinitrophenylhydrazone, followed by online analysis by high-performance liquid chromatography (HPLC) coupled with Ultraviolet detector. During the observation period, mixing ratios of HCHO ranged from 0.2 ppbv to 6.2 ppbv, with an average of 1.5 ± 0.67 ppbv. HCHO shows an evident diurnal variation, the maximum appeared during 12:00-14:00. The average concentration diurnal variations of measured HCHO, ozone (O 3), Methylhydroperoxides (MHP, CH 3OOH), hydrogen peroxide (H 2O 2), nitrogen oxides (NO x) and meteorological parameters were compared. The similar variations of HCHO, O 3 and radiation imply that photo-oxidation of hydrocarbons might be the major source for HCHO. Based on the maximum incremental reactivity (MIR) coefficient of HCHO, the calculation shows that HCHO contributes about 20% to total observed O 3 during the study period. In order to compare the contributions of O 3, HCHO and HONO to OH radical, photolysis rate parameters ( J-values) of the three compounds were calculated by the Tropospheric Ultraviolet and Visible (TUV) Radiation Model (4.4 version). Based on the comparison, this study reaches the conclusion that O 3 is the dominant source of OH radical at Mazhuang. This study also uses P(HCHO)/P(O 3) which represents the ratio of contrbutions of HCHO and O 3 to OH radical, to discuss the action of HCHO in OH radical soucers. The result shows that P(HCHO)/P(O 3) is 12.5% on average, with the maximum of 21.0% at 13:00 P.M. and minimum of 7.5% before 9:00 A.M. and after 17:00 P.M..Therefore HCHO is also an important source of OH radical and cannot be ignored.

  5. Equilibrium of particle nitrite with gas phase HONO: Tropospheric measurements in the high Arctic during polar sunrise

    NASA Astrophysics Data System (ADS)

    Li, Shao-Meng

    1994-12-01

    Gas phase HONO(g) and nitrite in particles of <5-μm size were measured in the troposphere during the Polar Sunrise Experiment at Alert, Northwest Territories, Canada, during January 19 to April 20, 1992, using denuder-filter pack sampling and IC-UV detection. The measurements indicated that HONO(g) existed at concentrations of up to 70 ppt before polar sunrise but gradually decreased to 5-10 ppt after sunrise. The calculated OH formation rate from HONO(g) photolysis was greater than from the photolysis of both O3 and CH2O by more than one order of magnitude during the sunlit period and led to moderately high levels of OH, e.g., 3×105 molecules cm-3 OH at noontime on April 5. Particle nitrite measurements showed a gradual increase in concentrations with increasing solar insolation, but the concentrations were generally less than 10 ppt. The pH and the sulfate molar concentrations of the particles and the water vapor mixing ratio indicate that the particles were highly acidic being approximately 70% (W/W) H2SO4 solution. In such highly concentrated H2SO4 solution, most particle nitrite should exist as hydrated nitrosonium ion H2ONO+. Taking this into consideration, the particle nitrite was in an approximate equilibrium with the measured HONO(g). This equilibrium, with HONO(g) rapidly photolyzed, was a good indication that the particles were effective sources of HONO(g) and implied rapid production of particle N(+III) during this period. Two possible pathways leading to the formation of particle N(+III) species are suggested, i.e., reduction of HNO3(aq) by SO2(g) and reduction of NO3-; (aq) by Br- (aq). However, N2O5 reaction with NaBr cannot be ruled out as the alternative HONO(g) formation mechanism which bypasses the equilibrium.

  6. Evidence of Aerosols as a Media for Rapid Daytime HONO Production over China

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Liu, Zhen; Wang, Yuhang; Costabile, Francesa

    Current knowledge of daytime HONO sources remains incomplete. A large missing daytime HONO source has been found in many places around the world, including polluted regions in China. Conventional understanding and recent studies attributed this missing source mainly to ground surface processes or gas-phase chemistry, while assuming aerosols to be an insignificant media for HONO production. We analyze in situ observations of HONO and its precursors at an urban site in Beijing, China, and report an apparent dependence of the missing HONO source strength on aerosol surface area and solar ultraviolet radiation. Based on extensive correlation analysis and process-modeling, wemore » propose that the rapid daytime HONO production in Beijing can be explained by enhanced hydrolytic disproportionation of NO2 on aqueous aerosol surfaces due to catalysis by dicarboxylic acid anions. The combination of high abundance of NO2, aromatic hydrocarbons, and aerosols over broad regions in China likely leads to elevated HONO levels, rapid OH production, and enhanced oxidizing capacity on a regional basis. Our findings call for attention to aerosols as a media for daytime heterogeneous HONO production in polluted regions like Beijing. This study also highlights the complex and uncertain heterogeneous chemistry in China, which merits future efforts of reconciling regional modeling and laboratory experiments, in order to understand and mitigate the regional particulate and O3 pollutions over China.« less

  7. The relationship between nighttime formation of gaseous HONO and nocturnal stability in an urban environment

    NASA Astrophysics Data System (ADS)

    McLaren, Robert; Wojtal, Patryk; Taylor, Peter

    2014-05-01

    Nitrous acid (HONO) is an important radical precursor in the troposphere that accumulates overnight giving rise to a significant photolytic production of the hydroxyl radical, OH, in the boundary layer during early morning hours the next day. It is understood that HONO is formed in the dark through the heterogeneous hydrolysis of NO2 on surfaces (2 NO2+ H2O -> HONO + HNO3) in a first order process, largely dominated by hydrolysis on ground surfaces and a smaller contribution from aerosol surfaces. Despite progress, the dark heterogeneous mechanism of HONO formation is still not well understood, mirroring our lack of consensus on the daytime production of HONO. We have measured HONO at night in an urban area (York University, Toronto, Canada) by DOAS for over one year. This rich dataset was analyzed with a view to understanding the nocturnal formation mechanism, and possible links to the daytime HONO formation mechanism. Frequently, "steady-states" of HONO are observed at night; d[HONO]/dt ~ 0, which follow after a rapid buildup of HONO during sunset at rates of several ppb hr-1. These steady-state levels of HONO are found to be independent of the mixing ratio of NO2 throughout the night. On other occasions, steady-states are not observed and HONO continues to increase throughout the night, highly correlated with the levels of NO2 d([HONO]/[NO2])/dt ~ 0). We have found that a very significant predictor of the type of behavior is the nocturnal stability of the atmosphere, measured by the thermal gradient, ΔT=T9.5m-T1.0m, and wind speed. The steady-state behavior is found to occur almost exclusively on unstable nights with higher wind speeds and ΔT ~ 0, when mixing of air in the lower atmosphere is more efficient. The non steady-state behavior of HONO is observed on stable nights with low wind speeds and large thermal gradients, ΔT > 2oC indicating limited vertical mixing. The observation of NO2 independent steady-states of HONO under conditions of efficient

  8. The Reduction of HNO3 to HONO by Volatile Organic Compounds Associated with Rush Hour Traffic

    NASA Astrophysics Data System (ADS)

    Rutter, A. P.; Malloy, Q.; Scheuer, E.; Gutierrez, C.; Calzada, M.; Dibb, J. E.; Griffin, R. J.

    2012-12-01

    Nitrous acid (HONO) is an important source of OH radicals in urban environments. However, the sources of HONO are not completely understood, which makes modeling urban atmospheric chemistry difficult. During a previous field study in Houston, TX a correlation was observed between increases in HONO and organic aerosol freshly emitted by motor vehicle traffic during morning rush hours (Ziemba et al., 2010). This source of HONO could not be explained by primary HONO emissions, and the hypothesis was drawn that the HONO was being formed from the reduction of HNO3 by the organic aerosols emitted by motor vehicles. To test this hypothesis, nitric acid (HNO3) was combined in a flow tube with aerosols made from automobile engine oil, which were used as a model for the organic aerosols emitted by rush hour traffic. Reduction of the HNO3 to HONO was observed, although the reaction was found to occur with the volatile organic carbon compounds (VOCs) found in the aerosol vapor, and not the particle surfaces. To explore this further Teflon raschig rings were added to the flow tube to increase surface area but the reaction was not enhanced, confirming the reaction to be homogeneous. The HONO formation observed ranged between 0.1 and 0.5 ppb hr-1 with a mean of 0.3±0.1 ppb hr-1, for typical nitric acid concentrations of 4-5 ppb and estimated concentrations of the reactive components in the engine oil vapor of between 200 and 300 ppt. The observations in this study compared well to the cited field study which observed formation rates between 0.05 and 0.5 ppb hr-1 with an average of 0.3±0.15 ppb hr-1. Water vapor was found to decrease the HONO formation rate by 0.1 ppb hr-1 for every1% of increase in the water mixing ratio. Reference Ziemba L.D., Dibb J.E., Griffin R.J., Anderson C.H., Whitlow S.I., Lefer B.L., Rappenglueck B., and Flynn J. (2010) Heterogeneous conversion of nitric acid to nitrous acid on the surface of primary organic aerosol in an urban atmosphere. Atmospheric

  9. Global and Regional Impacts of HONO on the Chemical Composition of Clouds and Aerosols

    NASA Technical Reports Server (NTRS)

    Elshorbany, Y. F.; Crutzen, P. J.; Steil, B.; Pozzer, A.; Tost, H.; Lelieveld, J.

    2014-01-01

    Recently, realistic simulation of nitrous acid (HONO) based on the HONO / NOx ratio of 0.02 was found to have a significant impact on the global budgets of HOx (OH + HO2) and gas phase oxidation products in polluted regions, especially in winter when other photolytic sources are of minor importance. It has been reported that chemistry-transport models underestimate sulphate concentrations, mostly during winter. Here we show that simulating realistic HONO levels can significantly enhance aerosol sulphate (S(VI)) due to the increased formation of H2SO4. Even though in-cloud aqueous phase oxidation of dissolved SO2 (S(IV)) is the main source of S(VI), it appears that HONO related enhancement of H2O2 does not significantly affect sulphate because of the predominantly S(IV) limited conditions, except over eastern Asia. Nitrate is also increased via enhanced gaseous HNO3 formation and N2O5 hydrolysis on aerosol particles. Ammonium nitrate is enhanced in ammonia-rich regions but not under ammonia-limited conditions. Furthermore, particle number concentrations are also higher, accompanied by the transfer from hydrophobic to hydrophilic aerosol modes. This implies a significant impact on the particle lifetime and cloud nucleating properties. The HONO induced enhancements of all species studied are relatively strong in winter though negligible in summer. Simulating realistic HONO levels is found to improve the model measurement agreement of sulphate aerosols, most apparent over the US. Our results underscore the importance of HONO for the atmospheric oxidizing capacity and corroborate the central role of cloud chemical processing in S(IV) formation

  10. Tropospheric HONO Distribution and Chemistry in the Southeastern U.S.

    NASA Astrophysics Data System (ADS)

    Zhou, X.; Ye, C.; Pu, D.; Stutz, J.; Festa, J.; Spolaor, M.; Weinheimer, A. J.; Campos, T. L.; Haggerty, J. A.; Cantrell, C. A.; Mauldin, L.; Guenther, A. B.; Hornbrook, R. S.; Apel, E. C.; Jensen, J. B.

    2014-12-01

    During the NOMADSS field campaign, nitrous acid (HONO) and particulate nitrate (pNO3) was measured on NCAR C-130 research aircraft during five research flights over the Southeast U.S. Aerosol samples were also collected on Teflon filters for the determination of pNO3 photolysis rate constants in the laboratory. Daytime HONO concentrations range from low ppt in free troposphere to 10-20 ppt in the boundary layer in the background air masses, to up to 40 ppt in the industrial and urban plumes. While daytime HONO sink is well defined, dominated by its photolysis, daytime sources vary in different types of air masses: pNO3 photolysis appears to be the major HONO source in the background terrestrial air masses in both the boundary layer and the free troposphere. With an average pNO3 photolysis rate constant of (2.8±1.7)×10-4 s-1, p-NO3 photolysis becomes to be an effective pathway to recycle HNO3 to NOx in the troposphere, with HONO as a dominant intermediate product. Within the high-NOx industrial plumes encountered, HONO is predominantly produced by secondary formation processes involving NOx as the precursor. Away from ground surface, no significant nighttime HONO accumulation exists in the background terrestrial air mass.

  11. Gradient Measurements of Nitrous Acid (hono)

    NASA Astrophysics Data System (ADS)

    Kleffmann, J.; Kurtenbach, R.; Lörzer, J.; Wiesen, P.; Kalthoff, N.; Vogel, B.; Vogel, H.

    Nitrous acid (HONO) plays an important role in photochemical air pollution due to its photodissociation by solar UV radiation into hydroxyl radicals and thus significantly enhances photooxidation processes. Furthermore, HONO is an important indoor pol- lutant, which can react with amines leading to nitrosamines, which are known to be carcinogenic. Despite its importance in atmospheric chemistry the mechanisms lead- ing to HONO formation are still not completely understood at present. Although it is commonly proposed that HONO is formed by heterogeneous processes, i.e. by the conversion of NO2 on wet surfaces, it is still under discussion whether HONO produc- tion is dominated by the surface of particles or by the ground surface. Simultaneous vertical profile measurements of HONO, the precursor NO2 and the aerosol surface area, which could answer this question are not available at present. Accordingly, in the present study night-time HONO, NO2 and particle surface area gradients in the altitude range 10-190 m were measured on the meteorological tower at the Forschungszentrum Karlsruhe/Germany using a new, very sensitive HONO in- strument (LOPAP), a commercial NOx monitor and a SMPS system. For all gradient measurements during the campaign it was observed that the [HONO]/[NO2] ratio decreased with increasing altitude. In contrast, the particle sur- face area was found to be more or less constant. Accordingly, no correlation between the [HONO]/[NO2] ratio and the particle surface area was observed showing that HONO formation was dominated by processes on ground surfaces and that signifi- cant HONO formation on particle surfaces could be excluded for the measurement site.

  12. Global and Regional Impacts of HONO on the Chemical Composition of Clouds and Aerosols

    NASA Technical Reports Server (NTRS)

    Elshorbany, Y. F.; Crutzen, P. J.; Steil, B.; Pozzer, A.; Tost, H.; Lelieveld, J.

    2014-01-01

    Recently, realistic simulation of nitrous acid (HONO) based on the HONO/NO(sub x) ratio of 0.02 was found to have a significant impact on the global budgets of HO(sub x) (OH + HO2) and gas phase oxidation products in polluted regions, especially in winter when other photolytic sources are of minor importance. It has been reported that chemistry-transport models underestimate sulphate concentrations, mostly during winter. Here we show that simulating realistic HONO levels can significantly enhance aerosol sulphate (S(VI)) due to the increased formation of H2SO4. Even though in-cloud aqueous phase oxidation of dissolved SO2 (S(IV)) is the main source of S(VI), it appears that HONO related enhancement of H2O2 does not significantly affect sulphate because of the predominantly S(IV) limited conditions, except over eastern Asia. Nitrate is also increased via enhanced gaseous HNO3 formation and N2O5 hydrolysis on aerosol particles. Ammonium nitrate is enhanced in ammonia-rich regions but not under ammonia-limited conditions. Furthermore, particle number concentrations are also higher, accompanied by the transfer from hydrophobic to hydrophilic aerosol modes. This implies a significant impact on the particle lifetime and cloud nucleating properties. The HONO induced enhancements of all species studied are relatively strong in winter though negligible in summer. Simulating realistic HONO levels is found to improve the model measurement agreement of sulphate aerosols, most apparent over the US. Our results underscore the importance of HONO for the atmospheric oxidizing capacity and corroborate the central role of cloud chemical processing in S(IV) formation.

  13. HONO fluxes from soil surfaces: an overview

    NASA Astrophysics Data System (ADS)

    Wu, Dianming; Sörgel, Matthias; Tamm, Alexandra; Ruckteschler, Nina; Rodriguez-Caballero, Emilio; Cheng, Yafang; Pöschl, Ulrich; Weber, Bettina

    2016-04-01

    Gaseous nitrous acid (HONO) contributes up to 80% of atmospheric hydroxyl (OH) radicals and is also linked to health risks through reactions with tobacco smoke forming carcinogens. Field and modeling results suggested a large unknown HONO source in the troposphere during daytime. By measuring near ground HONO mixing ratio, up to 30% of HONO can be released from forest, rural and urban ground as well as snow surfaces. This source has been proposed to heterogeneous reactions of nitrogen dioxide (NO2) on humic acid surfaces or nitric acid photolysis. Laboratory studies showed that HONO emissions from bulk soil samples can reach 258 ng m-2 s-1 (in term of nitrogen), which corresponding to 1.1 × 1012 molecules cm-2 s-1and ˜ 100 times higher than most of the field studies, as measured by a dynamic chamber system. The potential mechanisms for soil HONO emissions include chemical equilibrium of acid-base reaction and gas-liquid partitioning between soil nitrite and HONO, but the positive correlation of HONO fluxes with pH (largest at neutral and slightly alkaline) points to the dominance of the formation process by ammonia-oxidizing bacteria (AOB). In general soil surface acidity, nitrite concentration and abundance of ammonia-oxidizing bacteria mainly regulate the HONO release from soil. A recent study showed that biological soil crusts in drylands can also emit large quantities of HONO and NO, corresponding to ˜20% of global nitrogen oxide emissions from soils under natural vegetation. Due to large concentrations of microorganisms in biological soil crusts, particularly high HONO and NO emissions were measured after wetting events. Considering large areas of arid and arable lands as well as peatlands, up to 70% of global soils are able to emitting HONO. However, the discrepancy between large soil HONO emissions measured in lab and low contributions of HONO flux from ground surfaces in field as well as the role of microorganisms should be further investigated.

  14. Is Forest Ground and Soil a Net Source or Sink for HONO?

    NASA Astrophysics Data System (ADS)

    Kim, T.; Kim, K.; Zhou, X.

    2017-12-01

    Ambient measurements and chamber experiments were conducted at the PROPHET site during the PROPHET-AMOS 2016 field campaign, to investigate the exchange of nitrous acid (HONO) between the forest ground and the atmosphere. HONO concentrations measured at 1.3 m and 10 cm above the ground surface consistently showed positive gradients with height, suggesting that the ground surface was a net sink for HONO. The HONO concentration gradients were significantly more pronounced during rainy and foggy periods than during dry periods, indicating an enhancement of HONO deposition onto the wet ground surface. Significant loss of HONO from the gas phase to the ground surface in an open-bottom chamber supports the argument that forest ground is a net HONO sink via deposition. Despite the ground surface was not a net HONO source, HONO was found to accumulate in the atmosphere within the forest canopy during the first half of the night. Heterogeneous reactions of NO2 on the surfaces of tree trunks and branches is proposed to be responsible for the observed nighttime HONO production.

  15. Detailed budget analysis of HONO in central London reveals a missing daytime source

    NASA Astrophysics Data System (ADS)

    Lee, J. D.; Whalley, L. K.; Heard, D. E.; Stone, D.; Dunmore, R. E.; Hamilton, J. F.; Young, D. E.; Allan, J. D.; Laufs, S.; Kleffmann, J.

    2016-03-01

    Measurements of HONO were carried out at an urban background site near central London as part of the Clean air for London (ClearfLo) project in summer 2012. Data were collected from 22 July to 18 August 2014, with peak values of up to 1.8 ppbV at night and non-zero values of between 0.2 and 0.6 ppbV seen during the day. A wide range of other gas phase, aerosol, radiation, and meteorological measurements were made concurrently at the same site, allowing a detailed analysis of the chemistry to be carried out. The peak HONO/NOx ratio of 0.04 is seen at ˜ 02:00 UTC, with the presence of a second, daytime, peak in HONO/NOx of similar magnitude to the night-time peak, suggesting a significant secondary daytime HONO source. A photostationary state calculation of HONO involving formation from the reaction of OH and NO and loss from photolysis, reaction with OH, and dry deposition shows a significant underestimation during the day, with calculated values being close to 0, compared to the measurement average of 0.4 ppbV at midday. The addition of further HONO sources from the literature, including dark conversion of NO2 on surfaces, direct emission, photolysis of ortho-substituted nitrophenols, the postulated formation from the reaction of HO2 × H2O with NO2, photolysis of adsorbed HNO3 on ground and aerosols, and HONO produced by photosensitized conversion of NO2 on the surface increases the daytime modelled HONO to 0.1 ppbV, still leaving a significant missing daytime source. The missing HONO is plotted against a series of parameters including NO2 and OH reactivity (used as a proxy for organic material), with little correlation seen. Much better correlation is observed with the product of these species with j(NO2), in particular NO2 and the product of NO2 with OH reactivity. This suggests the missing HONO source is in some way related to NO2 and also requires sunlight. Increasing the photosensitized surface conversion rate of NO2 by a factor of 10 to a mean daytime first

  16. Resolving the HONO formation mechanism in the ionosphere via ab initio molecular dynamic simulations

    PubMed Central

    He, Rongxing; Li, Lei; Zhong, Jie; Zhu, Chongqin; Francisco, Joseph S.; Zeng, Xiao Cheng

    2016-01-01

    Solar emission produces copious nitrosonium ions (NO+) in the D layer of the ionosphere, 60 to 90 km above the Earth’s surface. NO+ is believed to transfer its charge to water clusters in that region, leading to the formation of gaseous nitrous acid (HONO) and protonated water cluster. The dynamics of this reaction at the ionospheric temperature (200–220 K) and the associated mechanistic details are largely unknown. Using ab initio molecular dynamics (AIMD) simulations and transition-state search, key structures of the water hydrates—tetrahydrate NO+(H2O)4 and pentahydrate NO+(H2O)5—are identified and shown to be responsible for HONO formation in the ionosphere. The critical tetrahydrate NO+(H2O)4 exhibits a chain-like structure through which all of the lowest-energy isomers must go. However, most lowest-energy isomers of pentahydrate NO+(H2O)5 can be converted to the HONO-containing product, encountering very low barriers, via a chain-like or a three-armed, star-like structure. Although these structures are not the global minima, at 220 K, most lowest-energy NO+(H2O)4 and NO+(H2O)5 isomers tend to channel through these highly populated isomers toward HONO formation. PMID:27071120

  17. Resolving the HONO formation mechanism in the ionosphere via ab initio molecular dynamic simulations.

    PubMed

    He, Rongxing; Li, Lei; Zhong, Jie; Zhu, Chongqin; Francisco, Joseph S; Zeng, Xiao Cheng

    2016-04-26

    Solar emission produces copious nitrosonium ions (NO(+)) in the D layer of the ionosphere, 60 to 90 km above the Earth's surface. NO(+) is believed to transfer its charge to water clusters in that region, leading to the formation of gaseous nitrous acid (HONO) and protonated water cluster. The dynamics of this reaction at the ionospheric temperature (200-220 K) and the associated mechanistic details are largely unknown. Using ab initio molecular dynamics (AIMD) simulations and transition-state search, key structures of the water hydrates-tetrahydrate NO(+)(H2O)4 and pentahydrate NO(+)(H2O)5-are identified and shown to be responsible for HONO formation in the ionosphere. The critical tetrahydrate NO(+)(H2O)4 exhibits a chain-like structure through which all of the lowest-energy isomers must go. However, most lowest-energy isomers of pentahydrate NO(+)(H2O)5 can be converted to the HONO-containing product, encountering very low barriers, via a chain-like or a three-armed, star-like structure. Although these structures are not the global minima, at 220 K, most lowest-energy NO(+)(H2O)4 and NO(+)(H2O)5 isomers tend to channel through these highly populated isomers toward HONO formation.

  18. Infrared Spectrum of N-Oxidohydroxylamine [ONH(OH)] Produced in Reaction H + Hono in Solid Para-Hydrogen

    NASA Astrophysics Data System (ADS)

    Haupa, Karolina Anna; Lee, Yuan-Pern

    2017-06-01

    Hydrogenation reactions in the N/O chemical network are important for an understanding of the mechanism of formation of organic molecules in dark interstellar clouds, but many reactions remain unknown. We present the results of the reaction H + HONO in solid {para}-hydrogen ({p}-H_{2}) at 3.3 K investigated with infrared spectra. Two methods that produced hydrogen atoms were the irradiation of HONO molecules in {p}-H_{2} at 365 nm to produce OH radicals that reacted readily with nearby H_{2} to produce mobile H atoms, and irradiation of Cl_{2} molecules (co-deposited with HONO) in {p}-H_{2} at 405 nm to produce Cl atoms that reacted readily with nearby H_{2} to produce mobile H atoms. In both experiments, we assigned IR lines at 3549.6 (νb{1}), 1465.0 (νb{3}), 1372.2 (νb{4}), 895.6/898.5 (νb{6}), and 630.9 (νb{7}) \\wn to N-oxidohydroxylamine [ONH(OH)], the primary product of HONO hydrogenation. The assignments were derived according to the consideration of possible reactions and comparison of observed vibrational wavenumbers and their IR intensities with values predicted with the B3LYP/aug-cc-pVTZ method of quantum-chemical calculations. The agreement between observed and calculated D/H- and ^{15}N/^{14}N-isotopic ratios further supports these assignments. The role of this reaction in the N/O chemical network in dark interstellar clouds is discussed.

  19. Nitrous acid in a street canyon environment: Sources and contributions to local oxidation capacity

    NASA Astrophysics Data System (ADS)

    Yun, Hui; Wang, Zhe; Zha, Qiaozhi; Wang, Weihao; Xue, Likun; Zhang, Li; Li, Qinyi; Cui, Long; Lee, Shuncheng; Poon, Steven C. N.; Wang, Tao

    2017-10-01

    Nitrous acid (HONO) plays an important role in radical formation and photochemical oxidation processes in the boundary layer. However, its impact on the chemistry in a street canyon microenvironment has not been thoroughly investigated. In this study, we measured HONO in a street canyon in urban Hong Kong and used an observation-based box model (OBM) with the Master Chemical Mechanism (MCM v3.3.1) to investigate the contribution of HONO to local oxidation chemistry. The observed HONO mixing ratios were in the range of 0.4-13.9 ppbv, with an average of 3.91 ppbv in the daytime and 2.86 ppbv at night. A mean HONO/NOx emission ratio of 1.0% (±0.5%) from vehicle traffic was derived. OBM simulations constrained by the observed HONO showed that the maximum concentrations of OH, HO2, and RO2 reached 4.65 × 106, 4.40 × 106, and 1.83 × 106 molecules cm-3, which were 7.9, 5.0, and 7.5 times, respectively, the results in the case without HONO constrained. Photolysis of HONO contributed to 86.5% of the total primary radical production rates and led to efficient NO2 and O3 production under the condition of weak regional transport of O3. The formation of HNO3 contributed to 98.4% of the total radical termination rates. Our results suggest that HONO could significantly increase the atmospheric oxidation capacity in a street canyon and enhance the secondary formation of HNO3 and HCHO, which can damage outdoor building materials and pose health risks to pedestrians.

  20. Development of an activated carbon filter to remove NO2 and HONO in indoor air.

    PubMed

    Yoo, Jun Young; Park, Chan Jung; Kim, Ki Yeong; Son, Youn-Suk; Kang, Choong-Min; Wolfson, Jack M; Jung, In-Ha; Lee, Sung-Joo; Koutrakis, Petros

    2015-05-30

    To obtain the optimum removal efficiency of NO2 and HONO by coated activated carbon (ACs), the influencing factors, including the loading rate, metal and non-metal precursors, and mixture ratios, were investigated. The NOx removal efficiency (RE) for K, with the same loading (1.0 wt.%), was generally higher than for those loaded with Cu or Mn. The RE of NO2 was also higher when KOH was used as the K precursor, compared to other K precursors (KI, KNO3, and KMnO4). In addition, the REs by the ACs loaded with K were approximately 38-55% higher than those by uncoated ACs. Overall, the REs (above 95%) of HONO and NOx with 3% KOH were the highest of the coated AC filters that were tested. Additionally, the REs of NOx and HONO using a mixing ratio of 6 (2.5% PABA (p-aminobenzoic acid)+6% H3PO4):4 (3% KOH) were the highest of all the coatings tested (both metal and non-metal). The results of this study show that AC loaded with various coatings has the potential to effectively reduce NO2 and HONO levels in indoor air. Copyright © 2015 Elsevier B.V. All rights reserved.

  1. Reaction of H + HONO in solid para-hydrogen: infrared spectrum of ˙ONH(OH).

    PubMed

    Haupa, Karolina Anna; Tielens, Alexander Godfried Gerardus Maria; Lee, Yuan-Pern

    2017-06-21

    Hydrogenation reactions in the N/O chemical network are important for an understanding of the mechanism of formation of organic molecules in dark interstellar clouds, but many reactions remain unknown. We present the results of the reaction H + HONO in solid para-hydrogen (p-H 2 ) at 3.3 K investigated with infrared spectra. Two methods that produced hydrogen atoms were the irradiation of HONO molecules in p-H 2 at 365 nm to produce OH radicals that reacted readily with nearby H 2 to produce mobile H atoms, and irradiation of Cl 2 molecules (co-deposited with HONO) in p-H 2 at 405 nm to produce Cl atoms that reacted, upon IR irradiation of the p-H 2 matrix, readily with nearby H 2 to produce mobile H atoms. In both experiments, we assigned IR lines at 3549.6 (ν 1 ), 1465.0 (ν 3 ), 1372.2 (ν 4 ), 898.5/895.6 (ν 6 ), and 630.9 (ν 7 ) cm -1 to hydroxy(oxido)-λ 5 -azanyl radical [˙ONH(OH)], the primary product of HONO hydrogenation. Two weak lines at 3603.4 and 991.0 cm -1 are tentatively assigned to the dihydroxy-λ 5 -azanyl radical, ˙N(OH) 2 . The assignments were derived according to the consideration of possible reactions and comparison of observed vibrational wavenumbers and their IR intensities with values predicted quantum-chemically with the B3LYP/aug-cc-pVTZ method. The agreement between observed and calculated D/H- and 15 N/ 14 N-isotopic ratios further supports these assignments. The role of this reaction in the N/O chemical network in dark interstellar clouds is discussed.

  2. Significant HONO concentration at a semi-rural site in the Pearl River Delta during a severe pollution period and its impact on atmospheric oxidation capacity

    NASA Astrophysics Data System (ADS)

    Yun, H.; Wang, T.; Wang, W.; Yu, C.; Xia, M.; Xue, L.; Wang, Z.; Zhang, N.; Poon, S.; Zhou, Y.; Yue, D.; Zhai, Y.

    2017-12-01

    Nitrous acid (HONO) is an important source of hydroxyl radical (OH) in the boundary layer, and has considerable impact on atmospheric oxidation capacity and ozone formation. However, the abundance of HONO and subsequent effects under severe pollution conditions, especially in winter, has not been thoroughly investigated. We conducted an intensive observation at a semi-rural site (Heshan) in the center of the Pearl River Delta (PRD) in January 2017. Extremely high HONO concentrations (up to 9.0 ppbv) were observed with a LOng-Path Absorption Photometer (LOPAP) in a severe pollution episode with especially high PM2.5 ( 400 μg m-3) and O3 ( 160 ppbv). HONO sustained at a relatively high level in the morning and had peaks even in the afternoon. An observation-based box model (OBM) built on Master Chemical Mechanism (MCM v3.3.1) was used to simulate the formation of HONO and its contribution to the radical concentrations. The results showed that HONO was the dominant source of primary radicals (= OH+HO2+RO2) and governed the in-situ production of ozone. Currently-identified HONO sources were added into the model to reveal the formation process of HONO during both the nighttime and daytime, and the relative importance of these sources will be discussed.

  3. Wavelength-Resolved Photon Fluxes of Indoor Light Sources: Implications for HOx Production

    NASA Astrophysics Data System (ADS)

    Kowal, S.; Kahan, T.

    2017-12-01

    Only a handful of studies have considered photolytic reactions indoors because photon fluxes at short wavelengths are generally considered to be negligible. We have measured wavelength resolved photon fluxes from indoor light sources including incandescent, halogen, compact fluorescent (CFL), and light emitting diodes (LED). In addition, fluorescent tubes, used in many offices and industrial buildings, and sunlight through windows were measured. The measured photon fluxes were used to calculate photolysis rate constants for potential indoor hydroxyl and peroxy radical (OH and HO2, "HOx") precursors: acetaldehyde (CH3CHO), formaldehyde (HCHO), hydrogen peroxide (H2O2), nitrous acid (HONO) and ozone (O3). Rate constants in conjunction with typical indoor concentrations were used to predict HOx production rates under various lighting conditions. Our results illustrate that all light sources except LEDs emit light at high enough energy to photolyze HOx precursors. Under typical lighting conditions only fluorescent tubes and sunlight will initiate significant photochemical HOx formation, and HONO and HCHO will be the only molecules that will have a strong influence on HOx levels indoors. Data from our experiments can be used in indoor air models to better predict HOx levels indoors.

  4. Investigation of effective line intensities of trans-HONO near 1255 cm-1 using continuous-wave quantum cascade laser spectrometers

    NASA Astrophysics Data System (ADS)

    Cui, Xiaojuan; Dong, Fengzhong; Sigrist, Markus W.; Zhang, Zhirong; Wu, Bian; Xia, Hua; Pang, Tao; Sun, Pengshuai; Fertein, Eric; Chen, Weidong

    2016-10-01

    Effective line intensities of P branch transitions of trans-nitrous acid (HONO) in the ν3 H-O-N bending mode near 1255 cm-1 have been determined by scaling measured HONO absorption intensities by continuous-wave quantum cascade laser absorption spectroscopy to reference values. Gaseous HONO samples were synthetized in the laboratory using the reaction of H2SO4 and NaNO2 solutions and the heterogeneous formation on surfaces in the presence of ambient water vapor and NO2 gas in a sealed gas sampling bag. The quantification of HONO was performed using a denuder associated with a NOx analyzer. Observed absorption line strengths for the trans conformer are found to be by a factor of approximately 1.17 higher than previously reported line strengths.

  5. Observations of Radical Precursors during TexAQS II: Findings and Implications

    NASA Astrophysics Data System (ADS)

    Olaguer, E. P.; Lefer, B. L.; Rappenglueck, B.; Pinto, J. P.

    2009-12-01

    The Texas Environmental Research Consortium (TERC) sponsored and helped organize significant components of the Second Texas Air Quality Study (TexAQS II). Some of the TERC-sponsored experiments, most notably those associated with the TexAQS II Radical and Aerosol Measurement Project (TRAMP) sited on top of the Moody Tower at the University of Houston, found evidence for the importance of short-lived radical sources such as formaldehyde (HCHO) and nitrous acid (HONO) in increasing ozone productivity. During TRAMP, daytime HCHO pulses as large as 32 ppb were observed and attributed to industrial activities upwind in the Houston Ship Channel (HSC), and HCHO peaks as large as 52 ppb were detected by in-situ surface monitors in the HSC. In addition, an instrumented Piper Aztec aircraft observed plumes of apparent primary formaldehyde in flares from petrochemical facilities in the HSC. In one such combustion plume, depleted of ozone by large NOx emissions, the Piper Aztec measured an HCHO-to-CO ratio three times that of mobile sources. HCHO from uncounted primary sources or ozonolysis of underestimated olefin emissions could significantly increase ozone productivity in Houston beyond previous expectations. Simulations with the CAMx model show that additional emissions of HCHO from industrial flares can increase peak ozone in Houston by up to 30 ppb, depending on conditions in the planetary boundary layer. Other findings from TexAQS II include significant concentrations of HONO throughout the day, well in excess of current air quality model predictions, with large nocturnal vertical gradients indicating a surface or near-surface source of HONO, and large concentrations of night-time radicals (~30 ppt HO2). Additional HONO sources could increase daytime ozone by more than 10 ppb. Improving the representation of primary and secondary HCHO and HONO in air quality models could enhance the effectiveness of simulated control strategies, and thus make ozone attainment

  6. Inter-comparison of HONO field measurements and its summertime variation in Seoul, Korea

    NASA Astrophysics Data System (ADS)

    Kim, J.; Lee, G.; Lee, D.; Cho, S.

    2017-12-01

    HONO is a key source of OH radical responsible for atmospheric oxidative capacity and plays an important role in heterogeneous oxidation of some species. To understand the oxidative mechanisms that lead to urban ozone and aerosol formation we need to know the sources and behavior of this trace gas. Despite its importance, HONO budgets, especially in the urban conditions in Korea are not well understood. In this study, HONO measurement was conducted in Olympic Park located in Seoul, Korea from May 19 to June 15 of 2016 using Quantum Cascade-Tunable Infrared Laser Differential Absorption Spectrometer (QC-TILDAS), High Efficiency Diffusion Scrubber-Ion Chromatography (HEDS-IC) and Monitor for AeRosols & Gases in ambient Air (MARGA). Overall, the measurements obtained with the three instruments agreed within analysis uncertainty. The resulting detection limits of all instruments were close to 0.10 ppbv for HONO. HONO concentrations over the measurement period varied from the detection limit to 3.46 ppbv (QC-TILDAS), 3.03 ppbv (HEDS-IC) and 4.81 ppbv (MARGA), respectively. Using a chemical box model including varying PBL heights and emissions, major paths of HONO production and its contributions to ozone was identified. The model showed significant underestimation compared to observations, which suggests additional unknown HONO production or direct HONO emission.

  7. Oxidative capacity of the Mexico City atmosphere - Part 1: A radical source perspective

    NASA Astrophysics Data System (ADS)

    Volkamer, R.; Sheehy, P. M.; Molina, L. T.; Molina, M. J.

    2007-04-01

    A detailed analysis of OH, HO2 and RO2 radical sources is presented for the near field photochemical regime inside the Mexico City Metropolitan Area (MCMA). During spring of 2003 (MCMA-2003 field campaign) an extensive set of measurements was collected to quantify time resolved ROx (sum of OH, HO2, RO2) radical production rates from day- and nighttime radical sources. The Master Chemical Mechanism (MCMv3.1) was constrained by measurements of (1) concentration time-profiles of photosensitive radical precursors, i.e., nitrous acid (HONO), formaldehyde (HCHO), ozone (O3), glyoxal (CHOCHO), and other oxygenated volatile organic compounds (OVOCs); (2) respective photolysis-frequencies (J-values); (3) concentration time-profiles of alkanes, alkenes, and aromatic VOCs (103 compound are treated) and oxidants, i.e., OH- and NO3 radicals, O3; and (4) NO, NO2, meteorological and other parameters. The ROx production rate was calculated directly from these observations; MCM was used to estimate further ROx production from unconstrained sources, and express overall ROx production as OH-equivalents (i.e., taking into account the propagation efficiencies of RO2 and HO2 radicals into OH radicals). Daytime radical production is found to be about 10-25 times higher than at night; it does not track the abundance of sunlight. 12-h average daytime contributions of individual sources are: HCHO and O3 photolysis, each about 20%; O3/alkene reactions and HONO photolysis, each about 15%; unmeasured sources about 30%. While the direct contribution of O3/alkene reactions appears to be moderately small, source-apportionment of ambient HCHO and HONO identifies O3/alkene reactions as being largely responsible for jump-starting photochemistry about one hour after sunrise. The peak radical production is found to be higher than in any other urban influenced environment studied to date; further, differences exist in the timing of radical production. Our measurements and analysis comprise a database

  8. BrO, OClO and HCHO Observations from the EOS-Aura Ozone Monitoring Instrument

    NASA Astrophysics Data System (ADS)

    Kurosu, T. P.; Chance, K.; Sioris, C. E.

    2004-12-01

    The Ozone Monitoring Instrument (OMI) was launched on 15 July 2004 on the EOS-Aura platform into a sun-synchronous, polar orbit with an equator crossing time of 13:45h (ascending node). OMI is a nadir-viewing near-UV/Visible spectrometer, covering the spectral region of 270 nm to 500 nm with a resolution between 0.45 nm and 1.0 nm and a nominal ground footprint of 13 km×24 km. Global coverage is achieved in one day. The very high spatial resolution of OMI measurements sets a new standard for trace gas and air quality monitoring from space. Combined with daily global coverage, this significantly advances our ability to answer outstanding questions on air pollution, including the determination of BrO sources in mid and low latitudes, BrO--O3 anti-correlations as a function of latitude, and the production of formaldehyde in cities of the developing world. We introduce the design of the OMI operational retrieval algorithm for BrO, OClO and HCHO. Based on a direct (non-DOAS) non-linear fitting approach, it includes wavelength calibration for radiances and irradiances, an undersampling correction, and the characterization of the instrument slit function. We will present results of BrO (global distribution, and tropospheric contributions from the break-up ice shelves and volcanic emissions), formaldehyde (over regions of isoprene emissions, forest fires, and heavy urban pollution), and, contingent upon the availability of suitable OMI observations, OClO (under ozone hole conditions). Where available, trace gas retrievals from OMI will be compared to results from the SCIAMACHY and GOME instruments.

  9. The Oxidant Production over Antarctic Land and its Export (OPALE) project: An overview of data collected in summer 2011-2012 at Concordia

    NASA Astrophysics Data System (ADS)

    Kukui, Alexandre; Legrand, Michel; Frey, Markus; Preunkert, Susanne; Savarino, Joel; Gallée, Hubert; Vicars, William; Gil Roca, Jaime; Jourdain, Bruno

    2015-04-01

    The need to characterize the oxidative capacity of the atmosphere of East Antarctica motivated the OPALE investigations at the top of the high plateau (Concordia) where processes are suspected to differ from those already identified at South Pole. For instance, in contrast to South Pole experiencing 24-hour sunlight, the solar irradiance at Concordia has a strong diurnal cycle. This has consequences on intensity of snow emissions as well as on the dynamic of the boundary layer. Concordia is also the inland site where the longest chemical ice core records have been extracted. Investigations made at Concordia in December 2011-January 2012 included OH and RO2 together with concurrent measurements of NO, NO2, HONO, O3, H2O2, HCHO, photolysis rates as well as meteorological parameters and physics of the boundary layer. HONO was investigated by deploying for the first time in Antarctica an absorption photometer (LOPAP), an analyser supposed to be free of interferences with numerous chemical species. Also investigated for the first time is the excess of 17O of ozone with a newly developed fast sampling method. The diurnal cycle of snow emissions was also documented for NOx and HCHO that strongly influence the level of radicals. The concentrations of OH and RO2 radicals (median values of 3x106 and 1x108 in molecule cm-3, respectively) were found to be comparable to those observed at South Pole confirming that the elevated oxidative capacity is a common characteristic of near-surface atmospheric layer for most of the Antarctic plateau. Similar to the SP findings the major factor explaining to high radical levels at Concordia was the high levels of NO leading to fast recycling of RO2 to OH. At the same time, in contrast to the SP where the radical levels are controlled by NO levels mostly via changing boundary layer properties, OH and RO2 at Concordia show strong diurnal variability. The variability of NOx at Concordia is also determined by the solar diurnal cycle via an

  10. Retrieval of tropospheric HCHO in El Salvador using ground based DOAS

    NASA Astrophysics Data System (ADS)

    Abarca, W.; Gamez, K.; Rudamas, C.

    2017-12-01

    Formaldehyde (HCHO) is the most abundant carbonyl in the atmosphere, being an intermediate product in the oxidation of most volatile organic compounds (VOCs). HCHO is carcinogenic, and highly water soluble [1]. HCHO can originate from biomass burning and fossil fuel combustion and has been observed from satellite and ground-based sensors by using the Differential Optical Absorption Spectroscopy (DOAS) technique [2].DOAS products can be used for air quality monitoring, validation of chemical transport models, validation of satellite tropospheric column density retrievals, among others [3]. In this study, we report on column density levels of HCHO measured by ground based Multi-Axis -DOAS in different locations of El Salvador in March, 2015. We have not observed large differences of the HCHO column density values at different viewing directions. This result points out a reasonably polluted and hazy atmosphere in the measuring sites, as reported by other authors [4]. Average values ranging from 1016 to 1017 molecules / cm2 has been obtained. The contribution of vehicular traffic and biomass burning to the column density levels in these sites of El Salvador will be discussed. [1] A. R. Garcia et al., Atmos. Chem. Phys. 6, 4545 (2006) [2] E. Peters et al., Atmos. Chem. Phys. 12, 11179 (2012) [3] T. Vlemmix, et al. Atmos. Meas. Tech., 8, 941-963, 2015 [4] A. Heckel et al., Atmos. Chem. Phys. 5, (2005)

  11. Nitrous acid (HONO) measurements during winter haze events in Beijing

    NASA Astrophysics Data System (ADS)

    Bloss, W.; Kramer, L. J.; Crilley, L.; Lee, J. D.; Squires, F. A.; Tong, S.

    2017-12-01

    Daytime HONO levels can reach several parts per billion in megacities during winter haze events and hence act as the dominant (primary) precursor to OH radicals in the urban boundary layer, and affect NOx abundance. Understanding the sources of HONO is therefore important to quantify atmospheric oxidative capacity and secondary pollutant formation during such haze events. Despite decades of research, there are still large uncertainties in HONO formation mechanisms, and as a result models often substantially underestimate peak HONO levels. In this study, measurements of HONO were performed at the Institute of Atmospheric Physics (IAP) site located in central Beijing during Nov/Dec 2016, across both haze and non-haze events. Using a commercial long-path absorption photometer (LOPAP), vertical profiles of HONO concentrations up to a height of 260 m on the IAP Meteorological Tower were performed, as well as continuous near-surface measurements. Preliminary results showed that HONO levels near the ground were very high during the winter haze events with concentrations over 10 ppbV observed. Typically, during the vertical profiles a negative gradient was observed, indicating a large HONO source close to the surface. However, during some of the profiles elevated HONO concentrations were also observed at higher altitudes pointing to a strong source within the boundary layer. Co-located NOx and SO2 measurements are used to elucidate potential HONO sources from direct emissions, homogeneous gas phase reactions and heterogeneous conversion of NO2 on surfaces. Results from ground level HONO/NOx ratios show a midday peak during clean periods indicating a photo-enhanced process, which was not apparent during hazy days. The potential impact of these findings on the OH radical budget in wintertime Beijing will be discussed.

  12. Sensitivities of winter ozone pollution events in oil and gas producing regions to VOCs, NOx and radicals (Invited)

    NASA Astrophysics Data System (ADS)

    Edwards, P. M.; Aikin, K.; De Gouw, J. A.; Dube, W. P.; Geiger, F.; Gilman, J.; Helmig, D.; Holloway, J.; Kercher, J. P.; Koss, A.; Lerner, B. M.; Martin, R. S.; McLaren, R.; Min, K.; Parrish, D. D.; Peischl, J.; Roberts, J. M.; Ryerson, T. B.; Thornton, J. A.; Veres, P. R.; Warneke, C.; Wild, R. J.; Williams, E. J.; Young, C.; Yuan, B.; Brown, S. S.

    2013-12-01

    The Uintah Basin in northeastern Utah, a region of intense oil and gas extraction, experienced ozone (O3) mixing ratios well above limits set by air quality standards for multiple days during three of the last four winters. The Uintah Basin Winter Ozone Study (UBWOS) consisted of two field intensives, in early 2012 and 2013, with the goal of addressing current uncertainties in the chemical and physical processes that drive wintertime O3 production in regions of oil and gas development. The data from these two study periods provide an excellent comparison of high and low O3 production years, as meteorological conditions during the winter of 2011-2012 resulted in no elevated O3 mixing ratios, in contrast to the winter of 2012-2013 when observed O3 mixing ratios were the highest yet recorded in the Uintah Basin. Box modeling studies, using the Master Chemical Mechanism (MCM v3.2) chemistry scheme, have been used to investigate our understanding of O3 photochemistry in this unusual emissions environment. Simulations identify O3 production in 2012 to be highly radical limited, with less conventional radical sources, such as HCHO, HONO, and ClNO2 photolysis, playing a central role. Consequently, O3 production during 2012 was highly VOC sensitive, despite the much larger mixing ratio of total non-methane hydrocarbons relative to NO¬x. Conditions during UBWOS 2013 resulted in significantly higher O3 precursor species concentrations than during 2012, including the concentrations of the radical precursors HCHO and HONO. Simulations constrained to the 2013 data show the effects of these changes in pre-cursor concentrations on the radical budget, and thus on local O3 photochemistry and its sensitivities during a wintertime O3 pollution episode.

  13. Photoenhanced uptakes of NO2 by indoor surfaces: A new HONO source

    NASA Astrophysics Data System (ADS)

    Gligorovski, S.; Bartolomei, V.; Soergel, M.; Gomez Alvarez, E.; Zetzsch, C.; Wortham, H.

    2012-12-01

    Nitrous acid (HONO) is a known household pollutant that can lead to human respiratory tract irritation. HONO acts as the nitrosating agent, e.g. by the formation of the so-called third-hand smoke after wall reactions of HONO with nicotine (1). HONO can be generated indoors directly during combustion processes or indirectly via heterogeneous NO2 reactions with adsorbed water on diverse surfaces (2). Recently a new source was identified as another path of HONO formation in the troposphere (3). Namely, the light-induced heterogeneous reaction of NO2 with adsorbed organics (known as photosensitizers) on various surfaces such as roads, buildings, rocks or plants leads to enhanced HONO production. The detected values of HONO indoors vary in the range between 2 and 25 parts per billion (ppb). However, like outdoors, the processes leading to HONO formation indoors are not completely understood (4). Indoor photolysis radiation sources include exterior sunlight (λ>350 nm) that enters typically through the windows and indoor illumination sources, i.e., rare gas/mercury fluorescent light bulbs and tungsten and tungsten/halogen light bulbs among others. The present work is showing the importance of indoor sources of HONO recently identified or postulated. We have tested a number of common household chemical agents commonly used for cleaning purposes or coatings of domestic surfaces to better identify different indoor HONO sources. We used a heterogeneous flow tube technique to test the HONO production potentials of these household chemical agents under different experimental conditions, namely with and without light and at different relative humidity levels and different NO2 concentrations. We report uptake kinetics measurements of the heterogeneous reaction of gas phase NO2 with lacquer and paint coated on the walls of the reactor. The flow tube was irradiated with four near-ultraviolet (UV) emitting lamps (range of wavelengths 300-420nm). We observed that the heterogeneous

  14. Characterizing agricultural soil nitrous acid (HONO) and nitric oxide (NO) emissions with their nitrogen isotopic composition

    NASA Astrophysics Data System (ADS)

    Chai, J.; Miller, D. J.; Guo, F.; Dell, C. J.; Karsten, H.; Hastings, M. G.

    2017-12-01

    Nitrous acid (HONO) is a major source of atmospheric hydroxyl radical (OH), which greatly impacts air quality and climate. Fertilized soils may be important sources of HONO in addition to nitric oxide (NO). However, soil HONO emissions are especially challenging to quantify due to huge spatial and temporal variation as well as unknown HONO chemistry. With no in-situ measurements available, soil HONO emissions are highly uncertain. Isotopic analysis of HONO may provide a tool for tracking these sources. We characterize in situ soil HONO and NO fluxes and their nitrogen isotopic composition (δ15N) across manure management and meteorological conditions during a sustainable dairy cropping study in State College, Pennsylvania. HONO and NO were simultaneously collected at hourly resolution from a custom-coated dynamic soil flux chamber ( 3 LPM) using annular denuder system (ADS) coupled with an alkaline-permanganate NOx collection system for offline isotopic analysis of δ15N with ±0.6 ‰ (HONO) and ±1.5 ‰ (NO) precision. The ADS method was tested using laboratory generated HONO flowing through the chamber to verify near 100% collection (with no isotopic fractionation) and suitability for soil HONO collection. Corn-soybean rotation plots (rain-fed) were sampled following dairy manure application with no-till shallow-disk injection (112 kg N ha-1) and broadcast with tillage incorporation (129 kg N ha-1) during spring 2017. Soil HONO fluxes (n=10) ranged from 0.1-0.6 ng N-HONO m-2 s-1, 4-28% of total HONO+NO mass fluxes. HONO and NO fluxes were correlated, with both declining during the measurement period. The soil δ15N-HONO flux weighted mean ±1σ of -15 ± 6‰ was less negative than δ15N of simultaneously collected NO (-29 ± 8‰). This can potentially be explained by fractionations associated with microbial conversion of nitrite, abiotic production of HONO from soil nitrite, and uptake and release with changing soil moisture. Our results have implications for

  15. Detailed budget analysis of HONO in central London reveals a missing daytime source

    NASA Astrophysics Data System (ADS)

    Lee, J. D.; Whalley, L. K.; Heard, D. E.; Stone, D.; Dunmore, R. E.; Hamilton, J. F.; Young, D. E.; Allan, J. D.; Laufs, S.; Kleffmann, J.

    2015-08-01

    Measurements of HONO were carried out at an urban background site near central London as part of the Clean air for London (ClearfLo) project in summer 2012. Data was collected from 22 July-18 August 2014, with peak values of up to 1.8 ppbV at night and non-zero values of between 0.2 and 0.6 ppbV seen during the day. A wide range of other gas phase, aerosol, radiation and meteorological measurements were made concurrently at the same site, allowing a detailed analysis of the chemistry to be carried out. The peak HONO/NOx ratio of 0.04 is seen at ~ 02:00 UTC, with the presence of a second, daytime peak in HONO/NOx of similar magnitude to the night-time peak suggesting a significant secondary daytime HONO source. A photostationary state calculation of HONO involving formation from the reaction of OH and NO and loss from photolysis, reaction with OH and dry deposition shows a significant underestimation during the day, with calculated values being close to zero, compared to the measurement average of 0.4 ppbV at midday. The addition of further HONO sources, including postulated formation from the reaction of HO2 with NO2 and photolysis of HNO3, increases the daytime modelled HONO to 0.1 ppbV, still leaving a significant extra daytime source. The missing HONO is plotted against a series of parameters including NO2 and OH reactivity, with little correlation seen. Much better correlation is observed with the product of these species with j(NO2), in particular NO2 and the product of NO2 with OH reactivity. This suggests the missing HONO source is in some way related to NO2 and also requires sunlight. The effect of the missing HONO to OH radical production is also investigated and it is shown that the model needs to be constrained to measured HONO in order to accurately reproduce the OH radical measurements.

  16. NO2 and HCHO variability in Mexico City from MAX-DOAS measurements

    NASA Astrophysics Data System (ADS)

    Grutter, M.; Friedrich, M. M.; Rivera, C. I.; Arellano, E. J.; Stremme, W.

    2015-12-01

    Atmospheric studies in large cities are of great relevance since pollution affects air quality and human health. A network of Multi Axis Differential Optical Absorption Spectrometers (MAX-DOAS) has been established in strategic sites within the Mexico City metropolitan area. Four instruments are now in operation with the aim to study the variability and spatial distribution of key pollutants, providing results of O4, NO2 and HCHO slant column densities (SCD). A numerical code has been written to retrieve gas profiles of NO2 and HCHO using radiative transfer simulations. We present the first results of the variability of these trace gases which will bring new insight in the current knowledge of transport patterns, emissions as well as frequency and origin of extraordinary events. Results of the vertical column densities (VCD) valiability of NO2 and HCHO in Mexico City are presented. These studies are useful to validate current and future satellite observatopns such as OMI, TROPOMI and TEMPO.

  17. Investigation of the 3D distribution of tropospheric formaldehyde (HCHO) at the city of Mainz (Germany) using measurements of a 4 azimuth MAX-DOAS instrument

    NASA Astrophysics Data System (ADS)

    Donner, Sebastian; Gu, Myojeong; Remmers, Julia; Wang, Yang; Wagner, Thomas

    2017-04-01

    The Differential Optical Absorption Spectroscopy (DOAS)-method allows to investigate the distribution of different atmospheric trace gases (e.g. NO2, SO2, HCHO...) simultaneously. This is done by analysing the absorptions of these species in spectra of scattered sunlight. Multi-AXis (MAX)-DOAS measurements observe scattered sun light under different elevation angles. From such measurements tropospheric vertical column densities (VCDs) and vertical profiles of the measured trace gases and aerosols can be determined. We performed measurements using a 4 azimuth MAX-DOAS system on the roof of the Max Planck Institute for Chemistry in Mainz/Germany since 2013. This instrument observes scattered sunlight in 4 separate orthogonal azimuth directions. We derive vertical profiles of trace gases in these 4 different azimuth directions. From these results we can investigate the 3D distribution of the trace gases. Mainz is located at the edge of the Rhine-Main area which is one of the densest populated areas in Germany. Therefore it experiences episodes of high and low pollution depending on the meteorological conditions. In this study we focus on formaldehyde (HCHO). It is either emitted directly by industries and other anthropogenic and biogenic activities. Usually higher amounts are produced by photochemical reactions from precursor substances (secondary production), where it plays an important role in photochemical smog chemistry and O3 chemistry. As it is an intermediate product of basic oxidation cycles of other hydrocarbons (also referred to as volatile organic compounds (VOCs)) especially in summer its concentrations are determined by the abundances of VOCs. Therefore HCHO observations can be used as an indicator for VOCs. Up to now we have nearly 4 years (starting from May 2013) of almost continuous data which provides already a quite large dataset. In this work we present a first overview of our HCHO results including time series of HCHO columns, a first comparison of

  18. Measurement of HONO Production From Traffic in a UK Road Tunnel

    NASA Astrophysics Data System (ADS)

    Kramer, L. J.; Crilley, L.; Adams, T. J.; Ball, S. M.; Pope, F.; Bloss, W.

    2016-12-01

    Nitrous Acid (HONO) has an important role in the boundary layer as a source of hydroxyl radicals (OH) which can oxidize VOCs and, in the presence of NOx, lead to the formation of ozone. In urban areas with high traffic density, vehicular emissions can be an important source of HONO, however, there are limited real-world studies on HONO emissions from vehicles and large uncertainties on emission values from different traffic fleets (e.g. diesel, gasoline cars, and light- and heavy-duty vehicles). Here, we will present preliminary results from measurements of HONO, nitrogen oxides, CO2 and particulate matter performed over the summer in a road tunnel in Birmingham, UK. A broadband cavity enhanced absorption spectroscopy system (BBCEAS) was deployed to perform high temporal resolution measurements (20 s) of HONO and NO2, alongside commercial analysers for NO, NOy, CO2 and PM. Using information on vehicle density and traffic fleet, emissions ratios of HONO/NOx and estimates of direct HONO emissions will be presented.

  19. Laser-Based and Ultra-Portable Gas Sensor for Indoor and Outdoor Formaldehyde (HCHO) Monitoring

    NASA Astrophysics Data System (ADS)

    Shutter, J. D.; Allen, N.; Paul, J.; Thiebaud, J.; So, S.; Scherer, J. J.; Keutsch, F. N.

    2017-12-01

    While used as a key tracer of oxidative chemistry in the atmosphere, formaldehyde (HCHO) is also a known human carcinogen and is listed and regulated by the United States EPA as a hazardous air pollutant. Combustion processes and photochemical oxidation of volatile organic compounds (VOCs) are the major outdoor sources of HCHO, and building materials and household products are ubiquitous sources of indoor HCHO. Due to the ease with which humans can be exposed to HCHO, it is imperative to monitor levels of both indoor and outdoor HCHO exposure in both short and long-term studies.High-quality direct and indirect methods of quantifying HCHO mixing ratios exist, but instrument size and user-friendliness can make them cumbersome or impractical for certain types of indoor and long-term outdoor measurements. In this study, we present urban HCHO measurements by using a new, commercially-available, ppbv-level accurate HCHO gas sensor (Aeris Technologies' MIRA Pico VOC Laser-Based Gas Analyzer) that is highly portable (29 cm x 20 cm x 10 cm), lightweight (3 kg), easy-to-use, and has low power (15 W) consumption. Using an ultra-compact multipass cell, an absorption path length of 13 m is achieved, resulting in a sensor capable of achieving ppbv/s sensitivity levels with no significant spectral interferences.To demonstrate the utility of the gas sensor for emissions measurements, a GPS was attached to the sensor's housing in order to map mobile HCHO measurements in real-time around the Boston, Massachusetts, metro area. Furthermore, the sensor was placed in residential and industrial environments to show its usefulness for indoor and outdoor pollution measurements. Lastly, we show the feasibility of using the HCHO sensor (or a network of them) in long-term monitoring stations for hazardous air pollutants.

  20. CARIBIC DOAS observations of nitrous acid and formaldehyde in a large convective cloud

    NASA Astrophysics Data System (ADS)

    Heue, K.-P.; Riede, H.; Walter, D.; Brenninkmeijer, C. A. M.; Wagner, T.; Frieß, U.; Platt, U.; Zahn, A.; Stratmann, G.; Ziereis, H.

    2014-07-01

    The chemistry in large thunderstorm clouds is influenced by local lightning-NOx production and uplift of boundary layer air. Under these circumstances trace gases like nitrous acid (HONO) or formaldehyde (HCHO) are expected to be formed or to reach the tropopause region. However, up to now only few observations of HONO at this altitude have been reported. Here we report on a case study where enhancements in HONO, HCHO and nitrogen oxides (NOx) were observed by the CARIBIC flying laboratory (Civil Aircraft for the Regular Investigation of the atmosphere Based on an Instrument Container). The event took place in a convective system over the Caribbean Sea in August 2011. Inside the cloud the light path reaches up to 100 km. Therefore the DOAS instrument on CARIBIC was very sensitive to the tracers inside the cloud. Based on the enhanced slant column densities of HONO, HCHO and NO2, average mixing ratios of 37, 468 and 210 ppt, respectively, were calculated. These data represent averages for constant mixing ratios inside the cloud. However, a large dependency on the assumed profile is found; for HONO a mixing ratio of 160 ppt is retrieved if the total amount is assumed to be situated in the uppermost 2 km of the cloud. The NO in situ instrument measured peaks up to 5 ppb NO inside the cloud; the background in the cloud was about 1.3 ppb, and hence clearly above the average outside the cloud (≈ 150 ppt). The high variability and the fact that the enhancements were observed over a pristine marine area led to the conclusion that, in all likelihood, the high NO concentrations were caused by lighting. This assumption is supported by the number of flashes that the World Wide Lightning Location Network (WWLLN) counted in this area before and during the overpass. The chemical box model CAABA is used to estimate the NO and HCHO source strengths which are necessary to explain our measurements. For NO a source strength of 10 × 109 molec cm-2 s-1 km-1 is found, which

  1. Light-induced nitrous acid (HONO) production from NO2 heterogeneous reactions on household chemicals

    NASA Astrophysics Data System (ADS)

    Gómez Alvarez, Elena; Sörgel, Matthias; Gligorovski, Sasho; Bassil, Sabina; Bartolomei, Vincent; Coulomb, Bruno; Zetzsch, Cornelius; Wortham, Henri

    2014-10-01

    Nitrous acid (HONO) can be generated in various indoor environments directly during combustion processes or indirectly via heterogeneous NO2 reactions with water adsorbed layers on diverse surfaces. Indoors not only the concentrations of NO2 are higher but the surface to volume (S/V) ratios are larger and therefore the potential of HONO production is significantly elevated compared to outdoors. It has been claimed that the UV solar light is largely attenuated indoors. Here, we show that solar light (λ > 340 nm) penetrates indoors and can influence the heterogeneous reactions of gas-phase NO2 with various household surfaces. The NO2 to HONO conversion mediated by light on surfaces covered with domestic chemicals has been determined at atmospherically relevant conditions i.e. 50 ppb NO2 and 50% RH. The formation rates of HONO were enhanced in presence of light for all the studied surfaces and are determined in the following order: 1.3·109 molecules cm-2 s-1 for borosilicate glass, 1.7·109 molecules cm-2 s-1 for bathroom cleaner, 1.0·1010 molecules cm-2 s-1 on alkaline detergent (floor cleaner), 1.3·1010 molecules cm-2 s-1 for white wall paint and 2.7·1010 molecules cm-2 s-1 for lacquer. These results highlight the potential of household chemicals, used for cleaning purposes to generate HONO indoors through light-enhanced NO2 heterogeneous reactions. The results obtained have been applied to predict the timely evolution of HONO in a real indoor environment using a dynamic mass balance model. A steady state mixing ratio of HONO has been estimated at 1.6 ppb assuming a contribution from glass, paint and lacquer and considering the photolysis of HONO as the most important loss process.

  2. A comparison of measured HONO uptake and release with calculated source strengths in a heterogeneous forest environment

    NASA Astrophysics Data System (ADS)

    Sörgel, Matthias; Trebs, Ivonne; Wu, Dianming; Held, Andreas

    2015-04-01

    Vertical mixing ratio profiles of nitrous acid (HONO) were measured in a clearing and on the forest floor in a rural forest environment (in the south-east of Germany) by applying a lift system to move the sampling unit of the LOng Path Absorption Photometer (LOPAP) up and down. For the forest floor, HONO was found to be predominantly deposited, whereas net deposition was dominating in the clearing only during nighttime and net emissions were observed during daytime. For selected days, net fluxes of HONO were calculated from the measured profiles using the aerodynamic gradient method. The emission fluxes were in the range of 0.02 to 0.07 nmol m-2 s-1, and, thus were in the lower range of previous observations. These fluxes were compared to the strengths of postulated HONO sources and to the amount of HONO needed to sustain photolysis in the boundary layer. Laboratory measurements of different soil samples from both sites revealed an upper limit for soil biogenic HONO emission fluxes of 0.025 nmol m-2 s-1. HONO formation by light induced NO2 conversion was calculated to be below 0.03 nmol m-2 s-1 for the investigated days, which is comparable to the potential soil fluxes. Due to light saturation at low irradiance, this reaction pathway was largely found to be independent of light intensity, i.e. it was only dependent on ambient NO2. We used three different approaches based on measured leaf nitrate loadings for calculating HONO formation from HNO3 photolysis. While the first two approaches based on empirical HONO formation rates yielded values in the same order of magnitude as the estimated fluxes, the third approach based on available kinetic data of the postulated pathway failed to produce noticeable amounts of HONO. Estimates based on reported cross sections of adsorbed HNO3 indicate that the lifetime of adsorbed HNO3 was only about 15 min, which would imply a substantial renoxification. Although the photolysis of HNO3 was significantly enhanced at the surface, the

  3. Which processes drive observed variations of HCHO columns over India?

    NASA Astrophysics Data System (ADS)

    Surl, Luke; Palmer, Paul I.; González Abad, Gonzalo

    2018-04-01

    We interpret HCHO column variations observed by the Ozone Monitoring Instrument (OMI), aboard the NASA Aura satellite, over India during 2014 using the GEOS-Chem atmospheric chemistry and transport model. We use a nested version of the model with a horizontal resolution of approximately 25 km. HCHO columns are related to local emissions of volatile organic compounds (VOCs) with a spatial smearing that increases with the VOC lifetime. Over India, HCHO has biogenic, pyrogenic, and anthropogenic VOC sources. Using a 0-D photochemistry model, we find that isoprene has the largest molar yield of HCHO which is typically realized within a few hours. We also find that forested regions that neighbour major urban conurbations are exposed to high levels of nitrogen oxides. This results in depleted hydroxyl radical concentrations and a delay in the production of HCHO from isoprene oxidation. We find that propene is the only anthropogenic VOC emitted in major Indian cities that produces HCHO at a comparable (but slower) rate to isoprene. The GEOS-Chem model reproduces the broad-scale annual mean HCHO column distribution observed by OMI (r = 0.6), which is dominated by a distinctive meridional gradient in the northern half of the country, and by localized regions of high columns that coincide with forests. Major discrepancies are noted over the Indo-Gangetic Plain (IGP) and Delhi. We find that the model has more skill at reproducing observations during winter (JF) and pre-monsoon (MAM) months with Pearson correlations r > 0.5 but with a positive model bias of x≃1×1015 molec cm-2. During the monsoon season (JJAS) we reproduce only a diffuse version of the observed meridional gradient (r = 0.4). We find that on a continental scale most of the HCHO column seasonal cycle is explained by monthly variations in surface temperature (r = 0.9), suggesting a role for biogenic VOCs, in agreement with the 0-D and GEOS-Chem model calculations. We also find that the seasonal cycle during

  4. Photo-induced formation of nitrous acid (HONO) on protein surfaces

    NASA Astrophysics Data System (ADS)

    Meusel, Hannah; Elshorbany, Yasin; Bartels-Rausch, Thorsten; Selzle, Kathrin; Lelieveld, Jos; Ammann, Markus; Pöschl, Ulrich; Su, Hang; Cheng, Yafang

    2014-05-01

    The study of nitrous acid (HONO) is of great interest, as the photolysis of HONO leads to the OH radical, which is the most important oxidant in the troposphere. HONO is directly emitted by combustion of fossil fuel and from soil biogenic nitrite (Su et al., 2011), and can also be formed by gas phase reactions of NO and OH and heterogeneous reactions of NO2. Previous atmospheric measurements have shown unexpectedly high HONO concentrations during daytime. Measured mixing ratios were about one order of magnitude higher than model simulations (Kleffmann et al. 2005, Vogel et al. 2003). The additional daytime source of HONO might be attributed to the photolysis of adsorbed nitric acid or heterogeneous photochemistry of NO2 on organic substrates, such as humic acids or polyphenolic compounds (Stemmler et al., 2006), or indirectly through nitration of phenols and subsequent photolysis of nitrophenols (Sosedova et al., 2011, Bejan et al., 2006). An important reactive surface for the heterogeneous formation of HONO could involve proteins, which are ubiquitous in the environment. They are part of coarse biological aerosol particles like pollen grains, fine particles (fragments of pollen, microorganism, plant debris) and dissolved in rainwater, soil and road dust (Miguel et al. 1999). In this project a thin film of bovine serum albumin (BSA), a model protein with 67 kDa and 21 tyrosine residues per molecule, is irradiated and exposed to nitrogen dioxide in humidified nitrogen. The formation of HONO is measured with long path absorption photometry (LOPAP). The generated HONO is in the range of 100 to 1100 ppt depending on light intensity, NO2 concentration and film thickness. Light induced HONO formation on protein surfaces is stable over the 20-hours experiment of irradiation and exposure. On the other hand, light activated proteins reacting with NO2 form nitrated proteins, as detected by liquid chromatography (LC-DAD). Our experiments on tetranitromethane (TNM) nitrated

  5. Light-induced protein nitration and degradation with HONO emission

    NASA Astrophysics Data System (ADS)

    Meusel, Hannah; Elshorbany, Yasin; Kuhn, Uwe; Bartels-Rausch, Thorsten; Reinmuth-Selzle, Kathrin; Kampf, Christopher J.; Li, Guo; Wang, Xiaoxiang; Lelieveld, Jos; Pöschl, Ulrich; Hoffmann, Thorsten; Su, Hang; Ammann, Markus; Cheng, Yafang

    2017-10-01

    Proteins can be nitrated by air pollutants (NO2), enhancing their allergenic potential. This work provides insight into protein nitration and subsequent decomposition in the presence of solar radiation. We also investigated light-induced formation of nitrous acid (HONO) from protein surfaces that were nitrated either online with instantaneous gas-phase exposure to NO2 or offline by an efficient nitration agent (tetranitromethane, TNM). Bovine serum albumin (BSA) and ovalbumin (OVA) were used as model substances for proteins. Nitration degrees of about 1 % were derived applying NO2 concentrations of 100 ppb under VIS/UV illuminated conditions, while simultaneous decomposition of (nitrated) proteins was also found during long-term (20 h) irradiation exposure. Measurements of gas exchange on TNM-nitrated proteins revealed that HONO can be formed and released even without contribution of instantaneous heterogeneous NO2 conversion. NO2 exposure was found to increase HONO emissions substantially. In particular, a strong dependence of HONO emissions on light intensity, relative humidity, NO2 concentrations and the applied coating thickness was found. The 20 h long-term studies revealed sustained HONO formation, even when concentrations of the intact (nitrated) proteins were too low to be detected after the gas exchange measurements. A reaction mechanism for the NO2 conversion based on the Langmuir-Hinshelwood kinetics is proposed.

  6. Potential sources of nitrous acid (HONO) and their impacts on ozone: A WRF-Chem study in a polluted subtropical region

    NASA Astrophysics Data System (ADS)

    Zhang, Li; Wang, Tao; Zhang, Qiang; Zheng, Junyu; Xu, Zheng; Lv, Mengyao

    2016-04-01

    Current chemical transport models commonly undersimulate the atmospheric concentration of nitrous acid (HONO), which plays an important role in atmospheric chemistry, due to the lack or inappropriate representations of some sources in the models. In the present study, we parameterized up-to-date HONO sources into a state-of-the-art three-dimensional chemical transport model (Weather Research and Forecasting model coupled with Chemistry: WRF-Chem). These sources included (1) heterogeneous reactions on ground surfaces with the photoenhanced effect on HONO production, (2) photoenhanced reactions on aerosol surfaces, (3) direct vehicle and vessel emissions, (4) potential conversion of NO2 at the ocean surface, and (5) emissions from soil bacteria. The revised WRF-Chem was applied to explore the sources of the high HONO concentrations (0.45-2.71 ppb) observed at a suburban site located within complex land types (with artificial land covers, ocean, and forests) in Hong Kong. With the addition of these sources, the revised model substantially reproduced the observed HONO levels. The heterogeneous conversions of NO2 on ground surfaces dominated HONO sources contributing about 42% to the observed HONO mixing ratios, with emissions from soil bacterial contributing around 29%, followed by the oceanic source (~9%), photochemical formation via NO and OH (~6%), conversion on aerosol surfaces (~3%), and traffic emissions (~2%). The results suggest that HONO sources in suburban areas could be more complex and diverse than those in urban or rural areas and that the bacterial and/or ocean processes need to be considered in HONO production in forested and/or coastal areas. Sensitivity tests showed that the simulated HONO was sensitive to the uptake coefficient of NO2 on the surfaces. Incorporation of the aforementioned HONO sources significantly improved the simulations of ozone, resulting in increases of ground-level ozone concentrations by 6-12% over urban areas in Hong Kong and

  7. Soil HONO Emissions and Its Potential Impact on the Atmospheric Chemistry and Nitrogen Cycle

    NASA Astrophysics Data System (ADS)

    Su, H.; Chen, C.; Zhang, Q.; Poeschl, U.; Cheng, Y.

    2014-12-01

    Hydroxyl radicals (OH) are a key species in atmospheric photochemistry. In the lower atmosphere, up to ~30% of the primary OH radical production is attributed to the photolysis of nitrous acid (HONO), and field observations suggest a large missing source of HONO. The dominant sources of N(III) in soil, however, are biological nitrification and denitrification processes, which produce nitrite ions from ammonium (by nitrifying microbes) as well as from nitrate (by denitrifying microbes). We show that soil nitrite can release HONO and explain the reported strength and diurnal variation of the missing source. The HONO emissions rates are estimated to be comparable to that of nitric oxide (NO) and could be an important source of atmospheric reactive nitrogen. Fertilized soils appear to be particularly strong sources of HONO. Thus, agricultural activities and land-use changes may strongly influence the oxidizing capacity of the atmosphere. A new HONO-DNDC model was developed to simulate the evolution of HONO emissions in agriculture ecosystems. Because of the widespread occurrence of nitrite-producing microbes and increasing N and acid deposition, the release of HONO from soil may also be important in natural environments, including forests and boreal regions. Reference: Su, H. et al., Soil Nitrite as a Source of Atmospheric HONO and OH Radicals, Science, 333, 1616-1618, 10.1126/science.1207687, 2011.

  8. 3 dimensional distributions of NO2, CHOCHO, and HCHO measured by the University of Colorado 2D-MAX-DOAS during MAD-CAT

    NASA Astrophysics Data System (ADS)

    Ortega, Ivan; Sinreich, Roman; Volkamer, Rainer

    2014-05-01

    We present results of 2 dimensional Multi Axis-DOAS (2D-MAX-DOAS) measurements to infer 3-dimensional measurements of trace gases by characterizing boundary layer vertical profiles and near surface azimuth horizontal distribution of NO2 (14 angles covering 360°). We combine the established optimal estimation inversion with a new parameterization approach; the first method to derive NO2 tropospheric vertical profiles and boundary layer height and the second one to retrieve the azimuth horizontal distribution of near surface NO2 mixing ratios, both at multiple wavelengths (350 nm, 450 nm, and 560 nm). This was conducted for three cloud-free days in the framework of the intensive Multi Axis DOAS Comparison campaign for Aerosols and Trace gases (MAD-CAT) in Mainz, Germany 2013. By retrieving NO2 at multiple wavelengths range-resolved distributions of NO2 are derived using an 'Onion-peeling' approach, i.e., exploiting the fact that the optical path lengths at different wavelengths probe different horizontal air masses. We also measure glyoxal (CHOCHO) and formaldehyde (HCHO) distributions, and present to our knowledge the first 3-dimesional trace-gas distribution measurements of CHOCHO by a ground-based instrument. We expand the 2D-MAX-DOAS capabilities to calculate azimuth ratios of HCHO-to-NO2 (RFN) and CHOCHO-to-NO2 (RGN) to pinpoint volatile organic compound (VOC) oxidation chemistry and CHOCHO-to-HCHO (RGF) ratios as an indicator of biogenic and/or anthropogenic VOC emissions. The results of RFN correlate well with RGN and we identify azimuth variations that indicate gradients in the VOC/NOx chemistry that leads to O3 and secondary aerosol production. While there is a clear diurnal pattern in the RFN and RGN, no such variations are observed in the RGF, which shows rather constant values below 0.04 throughout the day, consistent with previous measurements, and indicative of urban air masses.

  9. Ag-Modified In2O3/ZnO Nanobundles with High Formaldehyde Gas-Sensing Performance

    PubMed Central

    Fang, Fang; Bai, Lu; Song, Dongsheng; Yang, Hongping; Sun, Xiaoming; Sun, Hongyu; Zhu, Jing

    2015-01-01

    Ag-modified In2O3/ZnO bundles with micro/nano porous structures have been designed and synthesized with by hydrothermal method continuing with dehydration process. Each bundle consists of nanoparticles, where nanogaps of 10–30 nm are present between the nanoparticles, leading to a porous structure. This porous structure brings high surface area and fast gas diffusion, enhancing the gas sensitivity. Consequently, the HCHO gas-sensing performance of the Ag-modified In2O3/ZnO bundles have been tested, with the formaldehyde-detection limit of 100 ppb (parts per billion) and the response and recover times as short as 6 s and 3 s, respectively, at 300 °C and the detection limit of 100 ppb, response time of 12 s and recover times of 6 s at 100 °C. The HCHO sensing detect limitation matches the health standard limitation on the concentration of formaldehyde for indoor air. Moreover, the strategy to synthesize the nanobundles is just two-step heating and easy to scale up. Therefore, the Ag-modified In2O3/ZnO bundles are ready for industrialization and practical applications. PMID:26287205

  10. Summertime photochemistry during CAREBeijing-2007: ROx budgets and O3 formation

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Liu, Zhen; Wang, Y.; Gu, Dasa

    2012-08-28

    We analyze summertime photochemistry near the surface in Beijing, China, using a 1-D photochemical model (Regional chEmical and trAnsport Model, REAM-1D) constrained by in situ observations, focusing on the budgets of ROx (OH + HO2 + RO2) radicals and O3 formation. While the modeling analysis focuses on near-surface photochemical budgets, the implications for the budget of O3 in the planetary boundary layer are also discussed. In terms of daytime average, the total ROx primary production rate near the surface in Beijing is 6.6 ppbv per hour (ppbv h{sup 1}, among the highest found in urban atmospheres. The largest primary ROxmore » source in Beijing is photolysis of oxygenated volatile organic compounds (OVOCs), which produces HO2 and RO2 at 2.5 ppbv h{sup 1}1 and 1.7 ppbv h{sup 1}, respectively. Photolysis of excess HONO from an unknown heterogeneous source is the predominant primary OH source at 2.2 ppbv h{sup 1}, much larger than that of O1D+H2O (0.4 ppbv h{sup 1}). The largest ROx sink is via OH + NO2 reaction (1.6 ppbv h{sup 1}), followed by formation of RO2NO2 (1.0 ppbv h{sup 1}) and RONO2 (0.7 ppbv h{sup 1}). Due to the large aerosol surface area, aerosol uptake of HO2 appears to be another important radical sink, although the estimate of its magnitude is highly variable depending on the uptake coefficient value used. The daytime average O3 production and loss rates near the surface are 32 ppbv h{sup 1} and 6.2 ppbv h{sup 1}, respectively. Assuming NO2 to be the source of excess HONO, the NO2 to HONO transformation leads to considerable O3 loss and reduction of its lifetime. Our observation-constrained modeling analysis suggests that oxidation of VOCs (especially aromatics) and heterogeneous reactions (e.g. HONO formation and aerosol uptake HO2) play potentially critical roles in the primary radical budget and O3 formation in Beijing. One important ramification is that O3 production is neither NOx nor VOC limited, but in a transition regime where reduction

  11. The Unique OMI HCHO/NO2 Feature During the 2008 Beijing Summer Olympics: Implications for Ozone Production Sensitivity

    NASA Technical Reports Server (NTRS)

    Witte, J. C.; Duncan, B. N.; Douglass, A. R.; Kurosu, T. P.; Chance, K.; Retscher, C.

    2010-01-01

    In preparation of the Beijing Summer Olympic and Paralympics Games, strict controls were imposed between July and September 2008 on motor vehicle traffic and industrial emissions to improve air quality for the competitors. We assessed chemical sensitivity of ozone production to these controls using Ozone Monitoring Instrument (OMI) column measurements of formaldehyde (HCHO) and nitrogen dioxide (NO2), where their ratio serves as a proxy for the sensitivity. During the emission controls, HCHO/NO2 increased and indicated a NOx-limited regime, in contrast to the same period in the preceding three years when the ratio indicates volatile organic carbon (VOC)-limited and mixed NOx-VOC-limited regimes. After the emission controls were lifted, observed NO2 and HCHO/NO2 returned to their previous values. The 2005-2008 OMI record shows that this transition in regimes was unique as ozone production in Beijing was rarely NOx-limited. OMI measured summertime increases in HCHO of around 13% in 2008 compared to prior years, the same time period during which MODIS vegetation indices increased. The OMI HCHO increase may be due to higher biogenic emissions of HCHO precursors, associated with Beijing's greening initiative for the Olympics. However, NO2 and HCHO were also found to be well-correlated during the summer months. This indicates an anthropogenic VOC contribution from vehicle emissions to OMI HCHO and is a plausible explanation for the relative HCHO minimum observed in August 2008, concurrent with a minimum in traffic emissions. We calculated positive trends in 2005-2008 OMI HCHO and NO2 of about +1 x 10(exp 14) Molec/ square M-2 and +3 x 10(exp 13) molec CM-2 per month, respectively. The positive trend in NO2 may be an indicator of increasing vehicular traffic since 2005, while the positive trend in HCHO may be due to a combined increase in anthropogenic and biogenic emissions since 2005.

  12. Measurements of HONO, NO, NOy and SO2 in aircraft exhaust plumes at cruise

    NASA Astrophysics Data System (ADS)

    Jurkat, T.; Voigt, C.; Arnold, F.; Schlager, H.; Kleffmann, J.; Aufmhoff, H.; Schäuble, D.; Schaefer, M.; Schumann, U.

    2011-05-01

    Measurements of gaseous nitrogen and sulfur oxide emissions in young aircraft exhaust plumes give insight into chemical oxidation processes inside aircraft engines. Particularly, the OH-induced formation of nitrous acid (HONO) from nitrogen oxide (NO) and sulfuric acid (H2SO4) from sulfur dioxide (SO2) inside the turbine which is highly uncertain, need detailed analysis to address the climate impact of aviation. We report on airborne in situ measurements at cruise altitudes of HONO, NO, NOy, and SO2 in 9 wakes of 8 different types of modern jet airliners, including for the first time also an A380. Measurements of HONO and SO2 were made with an ITCIMS (Ion Trap Chemical Ionization Mass Spectrometer) using a new ion-reaction scheme involving SF5- reagent ions. The measured molar ratios HONO/NO and HONO/NOy with averages of 0.038 ± 0.010 and 0.027 ± 0.005 were found to decrease systematically with increasing NOx emission-index (EI NOx). We calculate an average EI HONO of 0.31 ± 0.12 g NO2 kg-1. Using reliable measurements of HONO and NOy, which are less adhesive than H2SO4 to the inlet walls, we derive the OH-induced conversion fraction of fuel sulfur to sulfuric acid $\\varepsilon$ with an average of 2.2 ± 0.5 %. $\\varepsilon$ also tends to decrease with increasing EI NOx, consistent with earlier model simulations. The lowest HONO/NO, HONO/NOy and $\\varepsilon$ was observed for the largest passenger aircraft A380.

  13. Potential role of the nitroacidium ion on HONO emissions from the snowpack.

    PubMed

    Hellebust, Stig; Roddis, Tristan; Sodeau, John R

    2007-02-22

    The effects of photolysis on frozen, thin films of water-ice containing nitrogen dioxide (as its dimer dinitrogen tetroxide) have been investigated using a combination of Fourier transform reflection-absorption infrared (FT-RAIR) spectroscopy and mass spectrometry. The release of HONO is ascribed to a mechanism in which nitrosonium nitrate (NO+NO3-) is formed. Subsequent solvation of the cation leads to the nitroacidium ion, H2ONO+, i.e., protonated nitrous acid. The pathway proposed explains why the field measurement of HONO at different polar sites is often contradictory.

  14. Investigation of the role of the calvin cycle and C1 metabolism during HCHO metabolism in gaseous HCHO-treated petunia under light and dark conditions using 13C-NMR.

    PubMed

    Sun, Huiqun; Zhang, Wei; Tang, Lijuan; Han, Shuang; Wang, Xinjia; Zhou, Shengen; Li, Kunzhi; Chen, Limei

    2015-01-01

    It has been shown that formaldehyde (HCHO) absorbed by plants can be assimilated through the Calvin cycle or C1 metabolism. Our previous study indicated that Petunia hybrida could effectively eliminate HCHO from HCHO-polluted air. To understand the roles of C1 metabolism and the Calvin cycle during HCHO metabolism and detoxification in petunia plants treated with gaseous H(13)CHO under light and dark conditions. Aseptically grown petunia plants were treated with gaseous H(13)CHO under dark and light conditions. The metabolites generated from HCHO detoxification in petunia were investigated using (13)C-NMR. [2-(13)C]glycine (Gly) was generated via C1 metabolism and [U-(13)C]glucose (Gluc) was produced through the Calvin cycle simultaneously in petunia treated with low-level gaseous H(13)CHO under light conditions. Generation of [2-(13)C]Gly decreased whereas [U-(13) C]Gluc and [U-(13)C]fructose (Fruc) production increased greatly under high-level gaseous H(13)CHO stress in the light. In contrast, [U-(13)C]Gluc and [U-(13)C] Fruc production decreased greatly and [2-(13)C]Gly generation increased significantly under low-level and high-level gaseous H(13)CHO stress in the dark. C1 metabolism and the Calvin cycle contributed differently to HCHO metabolism and detoxification in gaseous H(13CHO-treated petunia plants. As the level of gaseous HCHO increased, the role of C1 metabolism decreased and the role of the Calvin cycle increased under light conditions. However, opposite changes were observed in petunia plants under dark conditions. Copyright © 2015 John Wiley & Sons, Ltd.

  15. The Role of Iron-Bearing Minerals in NO2 to HONO Conversion on Soil Surfaces.

    PubMed

    Kebede, Mulu A; Bish, David L; Losovyj, Yaroslav; Engelhard, Mark H; Raff, Jonathan D

    2016-08-16

    Nitrous acid (HONO) accumulates in the nocturnal boundary layer where it is an important source of daytime hydroxyl radicals. Although there is clear evidence for the involvement of heterogeneous reactions of NO2 on surfaces as a source of HONO, mechanisms remain poorly understood. We used coated-wall flow tube measurements of NO2 reactivity on environmentally relevant surfaces (Fe (hydr)oxides, clay minerals, and soil from Arizona and the Saharan Desert) and detailed mineralogical characterization of substrates to show that reduction of NO2 by Fe-bearing minerals in soil can be a more important source of HONO than the putative NO2 hydrolysis mechanism. The magnitude of NO2-to-HONO conversion depends on the amount of Fe(2+) present in substrates and soil surface acidity. Studies examining the dependence of HONO flux on substrate pH revealed that HONO is formed at soil pH < 5 from the reaction between NO2 and Fe(2+)(aq) present in thin films of water coating the surface, whereas in the range of pH 5-8 HONO stems from reaction of NO2 with structural iron or surface complexed Fe(2+) followed by protonation of nitrite via surface Fe-OH2(+) groups. Reduction of NO2 on ubiquitous Fe-bearing minerals in soil may explain HONO accumulation in the nocturnal boundary layer and the enhanced [HONO]/[NO2] ratios observed during dust storms in urban areas.

  16. An Atmospheric Constraint on the NO2 Dependence of Daytime Near-Surface Nitrous Acid (HONO).

    PubMed

    Pusede, Sally E; VandenBoer, Trevor C; Murphy, Jennifer G; Markovic, Milos Z; Young, Cora J; Veres, Patrick R; Roberts, James M; Washenfelder, Rebecca A; Brown, Steven S; Ren, Xinrong; Tsai, Catalina; Stutz, Jochen; Brune, William H; Browne, Eleanor C; Wooldridge, Paul J; Graham, Ashley R; Weber, Robin; Goldstein, Allen H; Dusanter, Sebastien; Griffith, Stephen M; Stevens, Philip S; Lefer, Barry L; Cohen, Ronald C

    2015-11-03

    Recent observations suggest a large and unknown daytime source of nitrous acid (HONO) to the atmosphere. Multiple mechanisms have been proposed, many of which involve chemistry that reduces nitrogen dioxide (NO2) on some time scale. To examine the NO2 dependence of the daytime HONO source, we compare weekday and weekend measurements of NO2 and HONO in two U.S. cities. We find that daytime HONO does not increase proportionally to increases in same-day NO2, i.e., the local NO2 concentration at that time and several hours earlier. We discuss various published HONO formation pathways in the context of this constraint.

  17. Consistency of the Health of the Nation Outcome Scales (HoNOS) at inpatient-to-community transition.

    PubMed

    Luo, Wei; Harvey, Richard; Tran, Truyen; Phung, Dinh; Venkatesh, Svetha; Connor, Jason P

    2016-04-27

    The Health of the Nation Outcome Scales (HoNOS) are mandated outcome-measures in many mental-health jurisdictions. When HoNOS are used in different care settings, it is important to assess if setting specific bias exists. This article examines the consistency of HoNOS in a sample of psychiatric patients transitioned from acute inpatient care and community centres. A regional mental health service with both acute and community facilities. 111 psychiatric patients were transferred from inpatient care to community care from 2012 to 2014. Their HoNOS scores were extracted from a clinical database; Each inpatient-discharge assessment was followed by a community-intake assessment, with the median period between assessments being 4 days (range 0-14). Assessor experience and professional background were recorded. The difference of HoNOS at inpatient-discharge and community-intake were assessed with Pearson correlation, Cohen's κ and effect size. Inpatient-discharge HoNOS was on average lower than community-intake HoNOS. The average HoNOS was 8.05 at discharge (median 7, range 1-22), and 12.16 at intake (median 12, range 1-25), an average increase of 4.11 (SD 6.97). Pearson correlation between two total scores was 0.073 (95% CI -0.095 to 0.238) and Cohen's κ was 0.02 (95% CI -0.02 to 0.06). Differences did not appear to depend on assessor experience or professional background. Systematic change in the HoNOS occurs at inpatient-to-community transition. Some caution should be exercised in making direct comparisons between inpatient HoNOS and community HoNOS scores. Published by the BMJ Publishing Group Limited. For permission to use (where not already granted under a licence) please go to http://www.bmj.com/company/products-services/rights-and-licensing/

  18. Measurement of nitrous acid (HONO) by external-cavity quantum cascade laser based quartz-enhanced photoacoustic absorption spectroscopy

    NASA Astrophysics Data System (ADS)

    Yi, Hongming; Maamary, Rabih; Gao, Xiaoming; Sigrist, Markus W.; Fertein, Eric; Chen, Weidong

    2016-04-01

    Spectroscopic detection of short-lived gaseous nitrous acid (HONO) at 1254.85 cm-1 was realized by off-beam coupled quartz-enhanced photoacoustic spectroscopy (QEPAS) in conjunction with an external cavity quantum cascade lasers (EC-QCL). High sensitivity monitoring of HONO was performed within a very small gas-sample volume (of ~40 mm3) allowing a significant reduction (of about 4 orders of magnitude) of air sampling residence time which is highly desired for accurate quantification of chemically reactive short-lived species. Calibration of the developed QEPAS-based HONO sensor was carried out by means of lab-generated HONO samples whose concentrations were determined by simultaneous measurements of direct HONO absorption spectra in a 109.5 m multipass cell using a distributed feedback (DBF) QCL. A minimum detection limit (MDL @ SNR=1) of 66 ppbv HONO was achieved at 70 mbar using a laser output power of 50 mW and 1 s integration time, which corresponded to a normalized noise equivalent absorption coefficient of 3.6×10-8 cm-1.W/Hz1/2. This MDL was down to 7 ppbv at the optimal integration time of 150 s. The corresponding minimum detected absorption coefficient (SNR=1) is ~1.1×10-7 cm-1 (MDL: ~3 ppbv) in 1 s and ~1.1×10-8 cm-1 (MDL~330 pptv) in 150 s, respectively, with 1 W laser power. Acknowledgements The authors acknowledge financial supports from the CaPPA project (ANR-10-LABX-005) and the CPER CLIMIBIO program. References H. Yi, R. Maamary, X. Gao, M. W. Sigrist, E. Fertein, W. Chen, "Short-lived species detection of nitrous acid by external-cavity quantum cascade laser based quartz-enhanced photoacoustic absorption spectroscopy", Appl. Phys. Lett. 106 (2015) 101109

  19. Enhanced photochemical conversion of NO2 to HONO on humic acids in the presence of benzophenone.

    PubMed

    Han, Chong; Yang, Wangjin; Yang, He; Xue, Xiangxin

    2017-12-01

    The photochemical conversion of NO 2 to HONO on humic acids (HA) in the presence of benzophenone (BP) was investigated using a flow tube reactor coupled to a NO x analyzer at ambient pressure. BP significantly enhanced the reduction of NO 2 to HONO on HA under simulated sunlight, as shown by the increase of NO 2 uptake coefficient (γ) and HONO yield with the mass ratio of BP to HA. The γ and HONO yield on the mixtures of HA and BP obviously depended on the environmental conditions. Both γ and HONO yield increased with the increase of irradiation intensity and temperature, whereas they decreased with pH. The γ exhibited a negative dependence on the NO 2 concentration, which had slight influences on the HONO yield. There were maximum values for the γ and HONO yield at relative humidity (RH) of 22%. Finally, atmospheric implications about the photochemical reaction of NO 2 and HA in the presence of photosensitive species were discussed. Copyright © 2017 Elsevier Ltd. All rights reserved.

  20. Intercomparison of field measurements of nitrous acid (HONO) during the SHARP Campaign

    EPA Science Inventory

    Because of the importance of HONO as a radical reservoir, consistent and accurate measurements of its concentration are needed. As part of the SHARP (Study of Houston Atmospheric Radical Precursors), time series of HONO were obtained by five different measurement techniques on th...

  1. Gaseous nitrous acid (HONO) and nitrogen oxides (NOx) emission from gasoline and diesel vehicles under real-world driving test cycles.

    PubMed

    Trinh, Ha T; Imanishi, Katsuma; Morikawa, Tazuko; Hagino, Hiroyuki; Takenaka, Norimichi

    2017-04-01

    Reactive nitrogen species emission from the exhausts of gasoline and diesel vehicles, including nitrogen oxides (NO x ) and nitrous acid (HONO), contributes as a significant source of photochemical oxidant precursors in the ambient air. Multiple laboratory and on-road exhaust measurements have been performed to estimate the NO x emission factors from various vehicles and their contribution to atmospheric pollution. Meanwhile, HONO emission from vehicle exhaust has been under-measured despite the fact that HONO can contribute up to 60% of the total hydroxyl budget during daytime and its formation pathway is not fully understood. A profound traffic-induced HONO to NO x ratio of 0.8%, established by Kurtenbach et al. since 2001, has been widely applied in various simulation studies and possibly linked to under-estimation of HONO mixing ratios and OH radical budget in the morning. The HONO/NO x ratios from direct traffic emission have become debatable when it lacks measurements for direct HONO emission from vehicles upon the fast-changing emission reduction technology. Several recent studies have reported updated values for this ratio. This study has reported the measurement of HONO and NO x emission as well as the estimation of exhaust-induced HONO/NO x ratios from gasoline and diesel vehicles using different chassis dynamometer tests under various real-world driving cycles. For the tested gasoline vehicle, which was equipped with three-way catalyst after-treatment device, HONO/NO x ratios ranged from 0 to 0.95 % with very low average HONO concentrations. For the tested diesel vehicle equipped with diesel particulate active reduction device, HONO/NO x ratios varied from 0.16 to 1.00 %. The HONO/NO x ratios in diesel exhaust were inversely proportional to the average speeds of the tested vehicles. Photolysis of HONO is a dominant source of morning OH radicals. Conventional traffic-induced HONO/NO x ratio of 0.8% has possibly linked to underestimation of the total HONO

  2. A review and update of the Health of the Nation Outcome Scales (HoNOS).

    PubMed

    James, Mick; Painter, Jon; Buckingham, Bill; Stewart, Malcolm W

    2018-04-01

    Aims and method The Health of the Nation Outcome Scales (HoNOS) and its older adults' version (HoNOS 65+) have been used widely for 20 years, but their glossaries have not been revised to reflect clinicians' experiences or changes in service delivery. The Royal College of Psychiatrists convened an international advisory board, with UK, Australian and New Zealand expertise, to identify desirable amendments. The aim was to improve rater experience by removing ambiguity and inconsistency in the glossary rather than more radical revision. Changes proposed to the HoNOS are reported. HoNOS 65+ changes will be reported separately. Based on the views and experience of the countries involved, a series of amendments were identified. Clinical implications While effective clinician training remains critically important, these revisions aim to improve intra- and interrater reliability and improve validity. Next steps will depend on feedback from HoNOS users. Reliability and validity testing will depend on funding. Declaration of interest None.

  3. Key role of pH in the photochemical conversion of NO2 to HONO on humic acid

    NASA Astrophysics Data System (ADS)

    Han, Chong; Yang, Wangjin; Wu, Qianqian; Yang, He; Xue, Xiangxin

    2016-10-01

    The heterogeneous photochemical reactions of NO2 with humic acid (HA) were performed using a flow tube reactor coupled to a NOx analyzer. The effects of the pH on the uptake coefficient (γ) of NO2 and HONO and NO yields were investigated in detail. With increasing the pH in the range of 2-12, γ was almost constant with an average value of (4.21 ± 0.46) × 10-6, whereas the HONO yield and NO yield linearly decreased from (81.07 ± 4.07)% and (10.35 ± 3.86)% to (13.87 ± 9.15)% and (1.51 ± 0.94)%, respectively. According to the characterization of HA compositions and possible reaction paths, it can be concluded that the pH may influence the transfer of protons and the equilibrium of HONO with NO2- by varying the contents of carboxyl and phenol groups in HA, which should primarily contribute to the change in the HONO yield with the pH.

  4. Anthropogenic Emissions of Highly Reactive Volatile Organic Compounds (HRVOCs) Inferred from Oversampling of OMI HCHO Columns

    NASA Technical Reports Server (NTRS)

    Zhu, Lei; Jacob, Daniel; Mickley, Loretta; Marais, Eloise; Zhang, Aoxing; Cohan, Daniel; Yoshida, Yasuko; Duncan, Bryan; Abad, Gonzalo Gonzalez; Chance, Kelly; hide

    2014-01-01

    Satellite observations of formaldehyde (HCHO) columns provide top-down constraints on emissions of highly reactive volatile organic compounds (HRVOCs). This approach has been used previously to constrain emissions of isoprene from vegetation, but application to US anthropogenic emissions has been stymied by lack of a discernable HCHO signal. Here we show that oversampling of HCHO data from the Ozone Monitoring Instrument (OMI) for 2005 - 2008 enables quantitative detection of urban and industrial plumes in eastern Texas including Houston, Port Arthur, and Dallas-Fort Worth. By spatially integrating the individual urban-industrial HCHO plumes observed by OMI we can constrain the corresponding HCHO-weighted HRVOC emissions. Application to the Houston plume indicates a HCHO source of 260 plus or minus 110 kmol h-1 and implies a factor of 5.5 plus or minus 2.4 underestimate of anthropogenic HRVOC emissions in the US Environmental Protection Agency inventory. With this approach we are able to monitor the trend in HRVOC emissions over the US, in particular from the oil-gas industry, over the past decade.

  5. Car MAX-DOAS measurements of the tropospheric Formaldehyde (HCHO) column around Bucharest (Romania) and in the Rhein-Main area (Germany)

    NASA Astrophysics Data System (ADS)

    Donner, Sebastian; Shaiganfar, Reza; Riffel, Katharina; Dörner, Steffen; Lampel, Johannes; Remmers, Julia; Wagner, Thomas

    2016-04-01

    The DOAS (differential optical absorption spectroscopy)-method analyses the absorptions of atmospheric trace gases in spectra of scattered sun light. It is an excellent way to determine the concentrations of different trace gases (e.g. NO2, SO2, HCHO…) simultaneously. MAX (Multi-AXis)-DOAS measurements observe scattered sun light under different elevation angles. From such measurements tropospheric vertical column densities (VCDs) or even vertical profiles of the measured trace gases and aerosols can be determined. We performed mobile MAX-DOAS measurements using two instruments on the roof of a car in summer 2015 in Romania during the AROMAT2 campaign and in the Winter/Spring 2016 in the Rhein-Main area (Germany). The latter is one of the densest populated areas in Germany. One instrument is a commercial Mini-MAX-DOAS instrument from the Hoffmann company, the other a self-built instrument using an AVANTES spectrometer with better optical characteristics. The instruments were looking in two different directions (one forward and one backward). Mobile MAX-DOAS measurements cover a quite large area in a short period of time. This enables to map existing gradients of concentrations of tropospheric trace gases, e.g. NO2 and HCHO. The results of those measurements then can be used to validate satellite measurements or can be compared to model results. In this study we focus on formaldehyde (HCHO). In small amounts it is emitted directly by industries and other anthropogenic and biogenic activities. Large amounts are mostly secondary produced. As it is an intermediate product of basic oxidation cycles of other hydrocarbons its concentrations are determined by the abundances of other hydrocarbons. Therefore it can be used as an indicator for volatile organic compounds (VOCs). Furthermore HCHO plays an important role in photochemical smog chemistry and tropospheric O3 chemistry. In this work we present the measurement setup and preliminary HCHO results of the AROMAT2

  6. Intercomparison of HONO Measurements Made Using Wet-Chemical (NITROMAC) and Spectroscopic (IBBCEAS & LP/FAGE) Techniques

    NASA Astrophysics Data System (ADS)

    Dusanter, S.; Lew, M.; Bottorff, B.; Bechara, J.; Mielke, L. H.; Berke, A.; Raff, J. D.; Stevens, P. S.; Afif, C.

    2013-12-01

    A good understanding of the oxidative capacity of the atmosphere is important to tackle fundamental issues related to climate change and air quality. The hydroxyl radical (OH) is the dominant oxidant in the daytime troposphere and an accurate description of its sources in atmospheric models is of utmost importance. Recent field studies indicate higher-than-expected concentrations of HONO during the daytime, suggesting that the photolysis of HONO may be an important underestimated source of OH. Understanding the tropospheric HONO budget requires confidence in analytical instrumentation capable of selectively measuring HONO. In this presentation, we discuss an intercomparison study of HONO measurements performed during summer 2013 at the edge of a hardwood forest in Southern Indiana. This exercise involved a wet chemical technique (NITROMAC), an Incoherent Broad-Band Cavity Enhanced Absorption Spectroscopy instrument (IBBCEAS), and a Laser-Photofragmentation/Fluorescence Assay by Gas Expansion instrument (LP/FAGE). The agreement observed between the three techniques will be discussed for both ambient measurements and cross calibration experiments.

  7. Direct emission of nitrous acid (HONO) from gasoline cars in China determined by vehicle chassis dynamometer experiments

    NASA Astrophysics Data System (ADS)

    Liu, Yuhan; Lu, Keding; Ma, Yufang; Yang, Xinping; Zhang, Wenbin; Wu, Yusheng; Peng, Jianfei; Shuai, Shijin; Hu, Min; Zhang, Yuanhang

    2017-11-01

    HONO plays a key role in atmospheric chemistry, and while its importance is well-known, the sources of HONO are still not completely understood. As a component of ambient HONO sources, direct emission from vehicles is an area that should be extensively studied. In this study, we determined the HONO emission index for typical gasoline vehicles in the car population of China through a chassis dynamometer with different types of engines (PFI/GDI), starting conditions (cold/warm) and running styles (Beijing cycle). Emission ratios of HONO to nitrogen oxide (NOX) for the Chinese gasoline cars are determined to be in the range of (0.03-0.42) % and an averaged value is about 0.18%, which are comparable to those reported in the few studies available in Europe, the United States and Japan for gasoline cars while smaller for those of the diesel cars. The atmospheric impact of the direct HONO emission from gasoline cars was analyzed for a typical urban site in Beijing, significant contributions of the direct emission toward the HONO budget were found during morning rush hours or twilight conditions to be 8-12%.

  8. Analysis of Strong Wintertime Ozone Events in an Area of Extensive Oil and Gas Extraction

    NASA Astrophysics Data System (ADS)

    Rappenglück, Bernhard; Ackermann, Luis; Alvarez, Sergio; Golovko, Julia; Buhr, Martin; Field, Robert; Soltis, Jeff; Montague, Derek C.; Hauze, Bill; Scott, Adamson; Risch, Dan; Wilkerson, George; Bush, David; Stoeckenius, Till; Keslar, Cara

    2015-04-01

    During recent years, elevated ozone (O3) values have been observed repeatedly in the Upper Green River Basin (UGRB), Wyoming during wintertime. This paper presents an analysis of high ozone days in late winter 2011 (1-hour average up to 166 ppbv). Intensive Observational Periods (IOPs) were performed which included comprehensive surface and boundary layer measurements. Low windspeeds in combination with low mixing layer heights (~50 m agl) are essential for accumulation of pollutants. Air masses contain substantial amounts of reactive nitrogen (NOx) and non-methane hydrocarbons (NMHC) emitted from fossil fuel exploration activities in the Pinedale Anticline. On IOP days in the morning hours reactive nitrogen (up to 69%), then aromatics and alkanes (each ~10-15%; mostly ethane and propane) are major contributors to the hydroxyl (OH) reactivity. This time frame largely coincides with lowest NMHC/NOx ratios (~50), reflecting a relatively low NMHC mixture, and a change from a NOx-limited regime towards a NMHC limited regime. OH production on IOP days is mainly due to nitrous acid (HONO). On a 24-hr basis and as determined for a measurement height of 1.80 m above the surface HONO photolysis on IOP days can contribute ~83% to OH production on average, followed by alkene ozonolysis (~9%). Photolysis by ozone and HCHO photolysis contributes about 4% each to hydroxyl formation. High HONO levels (maximum hourly median on IOP days: 1,096 pptv) are favored by a combination of shallow boundary layer conditions and enhanced photolysis rates due to the high albedo of the snow surface. HONO is most likely formed through (i) abundant nitric acid (HNO3) produced in atmospheric oxidation of NOx, deposited onto the snow surface and undergoing photo-enhanced heterogeneous conversion to HONO and (ii) combustion related emission of HONO. HONO production is confined to the lowermost 10 m of the boundary layer. HONO, serves as the most important precursor for OH, strongly enhanced due to

  9. Nitrogen Oxides in the Nocturnal Boundary Layer: Chemistry of Nitrous Acid (HONO) and the Nitrate Radical (N03)

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Jochen Stutz

    altitude dependent. Measurements at one altitude, for example at the ground, where most air quality monitoring stations are located, are not representative for the rest of the NBL. Our model also revealed that radical chemistry is, in general, altitude dependent at night. We distinguish three regions: an unreactive, NO rich, ground layer; an upper, O3 and NO3 dominated layer, and a reactive mixing layer, where RO2 radicals are mixed from aloft with NO from the ground. In this reactive layer an active radical chemistry and elevated OH radical levels can be found. The downward transport of N2O5 and HO2NO2, followed by their thermal decay, was also identified as a radical source in this layer. Our observations also gave insight into the formation of HONO in the NBL. Based on our field experiments we were able to show that the NO2 to HONO conversion was relative humidity dependent. While this fact was well known, we found that it is most likely the uptake of HONO onto surfaces which is R.H. dependent, rather than the NO2 to HONO conversion. This finding led to the proposal of a new NO2 to HONO conversion mechanism, which is based on solid physical chemical principles. Noteworthy is also the observation of enhanced NO2 to HONO conversion during a dust storm event in Phoenix. The final activity in our project investigated the influence of the urban canopy, i.e. building walls and surfaces, on nocturnal chemistry. For the first time the surface area of a city was determined based on a Geographical Information System database of the city of Santa Monica. The surface to volume areas found in this study showed that, in the 2 lower part of the NBL, buildings provide a much larger surface area than the aerosol. In addition, buildings take up a considerable amount of the volume near the ground. The expansion of our model and sensitivity studies based on the Santa Monica data revealed that the surface area of buildings considerably influences HONO levels in urban areas. The volume

  10. The Enhanced Formaldehyde-Sensing Properties of P3HT-ZnO Hybrid Thin Film OTFT Sensor and Further Insight into Its Stability

    PubMed Central

    Tai, Huiling; Li, Xian; Jiang, Yadong; Xie, Guangzhong; Du, Xiaosong

    2015-01-01

    A thin-film transistor (TFT) having an organic–inorganic hybrid thin film combines the advantage of TFT sensors and the enhanced sensing performance of hybrid materials. In this work, poly(3-hexylthiophene) (P3HT)-zinc oxide (ZnO) nanoparticles' hybrid thin film was fabricated by a spraying process as the active layer of TFT for the employment of a room temperature operated formaldehyde (HCHO) gas sensor. The effects of ZnO nanoparticles on morphological and compositional features, electronic and HCHO-sensing properties of P3HT-ZnO thin film were systematically investigated. The results showed that P3HT-ZnO hybrid thin film sensor exhibited considerable improvement of sensing response (more than two times) and reversibility compared to the pristine P3HT film sensor. An accumulation p-n heterojunction mechanism model was developed to understand the mechanism of enhanced sensing properties by incorporation of ZnO nanoparticles. X-ray photoelectron spectroscope (XPS) and atomic force microscopy (AFM) characterizations were used to investigate the stability of the sensor in-depth, which reveals the performance deterioration was due to the changes of element composition and the chemical state of hybrid thin film surface induced by light and oxygen. Our study demonstrated that P3HT-ZnO hybrid thin film TFT sensor is beneficial in the advancement of novel room temperature HCHO sensing technology. PMID:25608214

  11. The enhanced formaldehyde-sensing properties of P3HT-ZnO hybrid thin film OTFT sensor and further insight into its stability.

    PubMed

    Tai, Huiling; Li, Xian; Jiang, Yadong; Xie, Guangzhong; Du, Xiaosong

    2015-01-19

    A thin-film transistor (TFT) having an organic-inorganic hybrid thin film combines the advantage of TFT sensors and the enhanced sensing performance of hybrid materials. In this work, poly(3-hexylthiophene) (P3HT)-zinc oxide (ZnO) nanoparticles' hybrid thin film was fabricated by a spraying process as the active layer of TFT for the employment of a room temperature operated formaldehyde (HCHO) gas sensor. The effects of ZnO nanoparticles on morphological and compositional features, electronic and HCHO-sensing properties of P3HT-ZnO thin film were systematically investigated. The results showed that P3HT-ZnO hybrid thin film sensor exhibited considerable improvement of sensing response (more than two times) and reversibility compared to the pristine P3HT film sensor. An accumulation p-n heterojunction mechanism model was developed to understand the mechanism of enhanced sensing properties by incorporation of ZnO nanoparticles. X-ray photoelectron spectroscope (XPS) and atomic force microscopy (AFM) characterizations were used to investigate the stability of the sensor in-depth, which reveals the performance deterioration was due to the changes of element composition and the chemical state of hybrid thin film surface induced by light and oxygen. Our study demonstrated that P3HT-ZnO hybrid thin film TFT sensor is beneficial in the advancement of novel room temperature HCHO sensing technology.

  12. Observation and modeling of ambient nitrous acid (HONO) at a rural site (Wangdu) in the North China Plain in summer 2014

    NASA Astrophysics Data System (ADS)

    Liu, Y.; Lu, K.; Li, X.; Dong, H.; Tan, Z.; Wu, Y.; Zeng, L.; Bohn, B.; Broch, S.; Zou, Q.; Fuchs, H.; Hofzumahaus, A.; Holland, F.; Rohrer, F.; Min, K. E.; Brown, S. S.; Wahner, A.; Zhang, Y.

    2017-12-01

    The photolysis of HONO, was frequently determined to be the major OH primary production channels in the city clusters of China. However, the source of ambient HONO is still a mystery needs to be elucidated. In the framework of an integrated field experiment performed at a rural site in the North China Plain (NCP), HONO was measured by two LOPAP instruments and one CEAS instrument in order to obtain the best quality of the observations and comprehensive parameters that support the exploration of the HONO budget were completely available. In general, the observed HONO concentrations by these three different instruments showed reasonable good agreement as indicated by the high correlation efficiencies while the difference of the regression slopes showed the HONO measured uncertainty is about 15% which is twice about the known uncertainties from each instrument. The diurnal variation of the observed HONO concentrations showed high value (1 -5 ppb) at night and small values (100 - 500 ppt) around noon time which is similar to other reported campaigns. With the assumption of photo stationary state, the missing HONO source is determined experimentally constrained to the observed HONO, jHONO, OH, and NO. Possible HONO formation rates from the chemical reactions as proposed from literatures were calculated in the framework of an observational based box model. It is found that heteorogeneous uptake of NO2 on the ground and the photolysis of nitrate were the major HONO sources for this site when constrained with the recommended kinetc parameters. Uncertainty analysis showed that the uptake coefficient of NO2 and the photolysis frequency of nitrate are the important kinetc factors to be determined in the future studies. In addition, the direct emission of HONO from soil is found to be very important for a few days after fertilization.

  13. MAX-DOAS measurements of HONO slant column densities during the MAD-CAT campaign: inter-comparison, sensitivity studies on spectral analysis settings, and error budget

    NASA Astrophysics Data System (ADS)

    Wang, Yang; Beirle, Steffen; Hendrick, Francois; Hilboll, Andreas; Jin, Junli; Kyuberis, Aleksandra A.; Lampel, Johannes; Li, Ang; Luo, Yuhan; Lodi, Lorenzo; Ma, Jianzhong; Navarro, Monica; Ortega, Ivan; Peters, Enno; Polyansky, Oleg L.; Remmers, Julia; Richter, Andreas; Puentedura, Olga; Van Roozendael, Michel; Seyler, André; Tennyson, Jonathan; Volkamer, Rainer; Xie, Pinhua; Zobov, Nikolai F.; Wagner, Thomas

    2017-10-01

    In order to promote the development of the passive DOAS technique the Multi Axis DOAS - Comparison campaign for Aerosols and Trace gases (MAD-CAT) was held at the Max Planck Institute for Chemistry in Mainz, Germany, from June to October 2013. Here, we systematically compare the differential slant column densities (dSCDs) of nitrous acid (HONO) derived from measurements of seven different instruments. We also compare the tropospheric difference of SCDs (delta SCD) of HONO, namely the difference of the SCDs for the non-zenith observations and the zenith observation of the same elevation sequence. Different research groups analysed the spectra from their own instruments using their individual fit software. All the fit errors of HONO dSCDs from the instruments with cooled large-size detectors are mostly in the range of 0.1 to 0.3 × 1015 molecules cm-2 for an integration time of 1 min. The fit error for the mini MAX-DOAS is around 0.7 × 1015 molecules cm-2. Although the HONO delta SCDs are normally smaller than 6 × 1015 molecules cm-2, consistent time series of HONO delta SCDs are retrieved from the measurements of different instruments. Both fits with a sequential Fraunhofer reference spectrum (FRS) and a daily noon FRS lead to similar consistency. Apart from the mini-MAX-DOAS, the systematic absolute differences of HONO delta SCDs between the instruments are smaller than 0.63 × 1015 molecules cm-2. The correlation coefficients are higher than 0.7 and the slopes of linear regressions deviate from unity by less than 16 % for the elevation angle of 1°. The correlations decrease with an increase in elevation angle. All the participants also analysed synthetic spectra using the same baseline DOAS settings to evaluate the systematic errors of HONO results from their respective fit programs. In general the errors are smaller than 0.3 × 1015 molecules cm-2, which is about half of the systematic difference between the real measurements.The differences of HONO delta SCDs

  14. Measurement of HONO, HNCO, and other inorganic acids by negative-ion proton-transfer chemical-ionization mass spectrometry (NI-PT-CIMS): application to biomass burning emissions

    NASA Astrophysics Data System (ADS)

    Roberts, J. M.; Veres, P.; Warneke, C.; Neuman, J. A.; Washenfelder, R. A.; Brown, S. S.; Baasandorj, M.; Burkholder, J. B.; Burling, I. R.; Johnson, T. J.; Yokelson, R. J.; de Gouw, J.

    2010-07-01

    A negative-ion proton-transfer chemical ionization mass spectrometric technique (NI-PT-CIMS), using acetate as the reagent ion, was applied to the measurement of volatile inorganic acids of atmospheric interest: hydrochloric (HCl), nitrous (HONO), nitric (HNO3), and isocyanic (HNCO) acids. Gas phase calibrations through the sampling inlet showed the method to be intrinsically sensitive (6-16 cts/pptv), but prone to inlet effects for HNO3 and HCl. The ion chemistry was found to be insensitive to water vapor concentrations, in agreement with previous studies of carboxylic acids. The inlet equilibration times for HNCO and HONO were 2 to 4 s, allowing for measurement in biomass burning studies. Several potential interferences in HONO measurements were examined: decomposition of HNO3·NO3- clusters within the CIMS, and NO2-water production on inlet surfaces, and were quite minor (≤1%, 3.3%, respectively). The detection limits of the method were limited by the instrument backgrounds in the ion source and flow tube, and were estimated to range between 16 and 50 pptv (parts per trillion by volume) for a 1 min average. The comparison of HONO measured by CIMS and by in situ FTIR showed good correlation and agreement to within 17%. The method provided rapid and accurate measurements of HNCO and HONO in controlled biomass burning studies, in which both acids were seen to be important products.

  15. Measurement of HONO, HNCO, and other inorganic acids by negative-ion proton-transfer chemical-ionization mass spectrometry (NI-PT-CIMS): application to biomass burning emissions

    NASA Astrophysics Data System (ADS)

    Roberts, J. M.; Veres, P.; Warneke, C.; Neuman, J. A.; Washenfelder, R. A.; Brown, S. S.; Baasandorj, M.; Burkholder, J. B.; Burling, I. R.; Johnson, T. J.; Yokelson, R. J.; de Gouw, J.

    2010-01-01

    A negative-ion proton transfer chemical ionization mass spectrometric technique (NI-PT-CIMS), using acetate as the reagent ion, was applied to the measurement of volatile inorganic acids of atmospheric interest: hydrochloric (HCl), nitrous (HONO), nitric (HNO3), and isocyanic (HNCO) acids. Gas phase calibrations through the sampling inlet showed the method to be intrinsically sensitive (6-16 cts/pptv), but prone to inlet effects for HNO3 and HCl. The ion chemistry was found to be insensitive to water vapor concentrations, in agreement with previous studies of carboxylic acids. The inlet equilibration times for HNCO and HONO were 2 to 4 s, allowing for measurement in biomass burning studies. Several potential interferences in HONO measurements were examined: decomposition of HNO3·NO3- clusters within the CIMS, and NO2-water production on inlet surfaces, and were quite minor (≤1%, 3.3%, respectively). The detection limits of the method were limited by the instrument backgrounds in the ion source and flow tube, and were estimated to range between 16 and 50 pptv (parts per trillion by volume) for a 1 min average. The comparison of HONO measured by CIMS and by in situ FTIR showed good correlation and agreement to within 17%. The method provided rapid and accurate measurements of HNCO and HONO in controlled biomass burning studies, in which both acids were seen to be important products.

  16. Measurement of HONO, HNCO, and Other Inorganic Acids by Negative-ion Proton-Transfer Chemical-Ionization Mass Spectrometry (NI-PT-CIMS):Application to Biomass Burning Emissions.

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Roberts, James M.; Veres, Patrick; Warneke, Carsten

    2010-07-23

    A negative-ion proton transfer chemical ionization mass spectrometric technique (NI-PT-CIMS), using acetate as the reagent ion, was applied to the measurement of volatile inorganic acids of atmospheric interest: hydrochloric (HCl), nitrous (HONO) nitric (HNO3), and isocyanic (HNCO) acids. Gas phase calibrations through the sampling inlet showed the method to be intrinsically sensitive (6-16 cts/pptv), but prone to inlet effects for HNO3 and HCl. The ion chemistry was found to be insensitive to water vapor concentrations, in agreement with previous studies of carboxylic acids. The inlet equilibration times for HNCO and HONO were 2 to 4 seconds, allowing for measurement inmore » biomass burning studies. Several potential interferences in HONO measurements were examined: decomposition of HNO3•NO3- clusters within the CIMS, and NO2-water production on inlet surfaces, and were quite minor (>_1%, 3.3%, respectively). The detection limits of the method were limited by the instrument backgrounds in the ion source and flow tube, and were estimated to range between 16 and 50 pptv (parts per trillion by volume). The comparison of HONO measured by CIMS and by in situ FTIR showed good correlation and agreement to within 17%. The method provided rapid and accurate measurements of HNCO and HONO in controlled biomass burning studies, and suggest both as products of biomass burning.« less

  17. Heterogeneous Reactivity of NO2 with Photocatalytic Paints: A Possible Source of Nitrous Acid (HONO) in the Indoor Environment

    NASA Astrophysics Data System (ADS)

    Gligorovski, S.; Bartolomei, V.; Gandolfo, A.; Gomez Alvarez, E.; Kleffmann, J.; Wortham, H.

    2014-12-01

    There is an increasing concern about the indoor air environment, where we spend most of our time. Common methods of improving indoor air quality include controlling pollution sources, increasing ventilation rates or using air purifiers. Photocatalytic remediation technology was suggested as a new possibility to eliminate indoor air pollutants instead of just diluting or disposing them. In the present study, heterogeneous reactions of NO2 were studied on photocatalytic paints containing different size and quantity of TiO2. The heterogeneous reactions were conducted in a photo reactor under simulated atmospheric conditions. The flat pyrex rectangular plates covered with the paint were inserted into the reactor. These plates have been sprayed with the photocatalytic paints at our industrial partner's (ALLIOS) facilities using a high precision procedure that allowed the application of a thin layer of a given thickness of the paint. This allows a homogeneous coverage of the surface with the paint and an accurate determination of the exact amount of paint exposed to gaseous NO2. We demonstrate that the indoor photocatalytic paints which contain TiO2 can substantially reduce the concentrations of nitrogen dioxide (NO2). We show that the efficiency of nitrogen dioxide (NO2) removal increase with the quantity of TiO2 in the range 0 - 7 %. The geometric uptake coefficients increase from 5 · 10-6 to 1.6 · 10-5 under light irradiation of the paints. On the other hand, during the reactions of NO2 with this paint (7 % of TiO2) nitric oxide (NO) and nitrous acid (HONO) are formed. Nitrous acid (HONO) is an important harmful indoor pollutant and its photolysis leads to the formation of highly reactive OH radicals (Gomez Alvarez et al., 2013). Maximum conversion efficiencies of NO2to HONO and NO of 15 % and 33 % were observed at 30 % RH, respectively. Thus, the quantity of TiO2 embedded in the paint is an important parameter regarding the nitrogen oxides (NOx = NO + NO2

  18. The Primary and Recycling Sources of OH During the NACHTT-2011 Campaign: HONO as an Important OH Primary Source in the Wintertime

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Kim, S.; VandenBoer, Trevor; Young, Cora

    2014-06-10

    We present OH observation results during the NACHTT-11 field campaign at the Boulder Atmospheric Observatory in Weld County, Colorado. The observed OH levels during the daytime (at noon) were ~ 2.7 × 106 molecules cm-3 at the ground level (2 m above ground level, AGL). HONO and ozone photolysis were the two dominant photochemical OH production pathways during the field campaign. However, alkene ozonolysis, found an important source for OH by two previous winter season OH observations, was a minor contribution to OH primary production (~5 %). To evaluate recycling sources of OH from HO2 and RO2, an observation constrainedmore » University of Washington Chemical Mechanism (UWCM) box model is employed to simulated ambient OH levels with different model scenarios. For the base run without constraining observed HONO, the model simulated OH significantly underestimates the observed OH level (20.8 times in the morning and 7.2 times in the daytime). This indicates that the known HONO sources incorporated in the UWCM model cannot explain the observed HONO level. Once HONO is constrained by the observation, the discrepancy between observation and model simulation improves (5.1 times in the morning and 2.1 times in the daytime) but still out of the measurement uncertainty range (35 %). We explore two possible reasons for the observed unexplainably high wintertime OH levels. First, potential roles of Cl atoms produce organic peroxy radicals from the reactions between Cl atmos and alkane compounds. However, the Cl levels during the observation period are estimated very low (~ 103 atoms cm-3) to explain the enhanced OH levels. Second, Impacts of higher HONO levels on the ground was evaluated. Strong HONO gradient towards ground was observed especially during the early morning (6 am to 8 am) was observed and the lowest level available for the HONO observation during the campaign is 5 m AGL. Once we assume the twice of the observed HONO levels averaged between 5 m to 15 m at 2

  19. Heterogeneous Chemistry of HONO on Liquid Sulfuric Acid: A New Mechanism of Chlorine Activation on Stratospheric Sulfate Aerosols

    NASA Technical Reports Server (NTRS)

    Zhang, Renyi; Leu, Ming-Taun; Keyser, Leon F.

    1996-01-01

    Heterogeneous chemistry of nitrous acid (HONO) on liquid sulfuric acid (H2SO4) Was investigated at conditions that prevail in the stratosphere. The measured uptake coefficient (gamma) of HONO on H2SO4 increased with increasing acid content, ranging from 0.03 for 65 wt % to about 0.1 for 74 wt %. In the aqueous phase, HONO underwent irreversible reaction with H2SO4 to form nitrosylsulfuric acid (NO(+)HSO4(-). At temperatures below 230 K, NO(+)HSO4(-) was observed to be stable and accumulated in concentrated solutions (less than 70 wt % H2SO4) but was unstable and quickly regenerated HONO in dilute solutions (less than 70 wt %). HCl reacted with HONO dissolved in sulfuric acid, releasing gaseous nitrosyl chloride (ClNO). The reaction probability between HCl and HONO varied from 0.01 to 0.02 for 60-72 wt % H2SO4. In the stratosphere, ClNO photodissociates rapidly to yield atomic chlorine, which catalytically destroys ozone. Analysis of the laboratory data reveals that the reaction of HCl with HONO on sulfate aerosols can affect stratospheric ozone balance during elevated sulfuric acid loadings after volcanic eruptions or due to emissions from the projected high-speed civil transport (HSCT). The present results may have important implications on the assessment of environmental acceptability of HSCT.

  20. Long-term (2005-2014) trends in formaldehyde (HCHO) columns across North America as seen by the OMI satellite instrument: Evidence of changing emissions of volatile organic compounds

    NASA Astrophysics Data System (ADS)

    Zhu, Lei; Mickley, Loretta J.; Jacob, Daniel J.; Marais, Eloïse A.; Sheng, Jianxiong; Hu, Lu; Abad, Gonzalo González; Chance, Kelly

    2017-07-01

    Satellite observations of formaldehyde (HCHO) columns provide top-down information on emissions of highly reactive volatile organic compounds (VOCs). We examine the long-term trends in HCHO columns observed by the Ozone Monitoring Instrument from 2005 to 2014 across North America. Biogenic isoprene is the dominant source of HCHO, and its emission has a large temperature dependence. After correcting for this dependence, we find a general pattern of increases in much of North America but decreases in the southeastern U.S. Over the Houston-Galveston-Brazoria industrial area, HCHO columns decreased by 2.2% a-1 from 2005 to 2014, consistent with trends in emissions of anthropogenic VOCs. Over the Cold Lake Oil Sands in the southern Alberta in Canada, HCHO columns increased by 3.8% a-1, consistent with the increase in crude oil production there. HCHO variability in the northwestern U.S. and Midwest could be related to afforestation and corn silage production. Although NOx levels can affect the HCHO yield from isoprene oxidation, we find that decreases in anthropogenic NOx emissions made only a small contribution to the observed HCHO trends.Plain Language SummaryWe use satellite observations to diagnose long-term trends in <span class="hlt">HCHO</span> columns across North America from 2005 to 2014. <span class="hlt">HCHO</span> generally increased from 2005-2009 to 2010-2014 but decreased in the southeastern U.S. We find significant regional trends in excess of 20% related to decreases in urban anthropogenic VOC emissions (Houston metropolitan area) and increases in oil/gas production (oil sands in western Canada). Significant regional trends in the northwestern U.S. and in the Midwest may be driven by afforestation and agricultural activity. The impact of declining NO<fi>x</fi> emission over the U.S. on <span class="hlt">HCHO</span> columns is likely small over this time frame.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_5 --> <div id="page_6" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="101"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.A13C0235C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.A13C0235C"><span>Effect of Coatings on the Uptake Rate and <span class="hlt">HONO</span> Yield in Heterogeneous Reaction of Soot with NO2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cruz-Quiñones, M.; Khalizov, A. F.; Zhang, R.</p> <p>2009-12-01</p> <p>Heterogeneous reaction of nitrogen dioxide on carbon soot aerosols has been suggested as a possible source of nighttime nitrous acid (<span class="hlt">HONO</span>) in atmosphere boundary layer. Available laboratory data show significant variability in the measured reaction probabilities and <span class="hlt">HONO</span> yields, making it difficult to asses the atmospheric significance of this process. Moreover, little is known of how aging of soot aerosol through internal mixing with other atmospheric trace constituents will affect the heterogeneous reactivity and <span class="hlt">HONO</span> production. In this work, the heterogeneous reaction of NO2 on fresh and aged soot films leading to <span class="hlt">HONO</span> formation was studied through a series of kinetic uptake experiments and <span class="hlt">HONO</span> yield measurements. Soot samples were prepared by incomplete combustion of propane and kerosene fuels under lean and rich flame conditions. Experiments were performed in a low-pressure, fast-flow reactor coupled to a chemical ionization mass spectrometer (CIMS), using atmospheric-level NO2 concentrations. Heterogeneous uptake coefficients, γ(geom) and γ(BET), were calculated using geometric and internal BET soot surface areas, respectively. The uptake coefficient and the <span class="hlt">HONO</span> yield depend on the type of fuel and combustion regime and are the highest for soot samples prepared using rich kerosene flame. Although, the internal surface area of soot measured by BET method is a factor of 50 to 500 larger than the geometric surface area, only the top soot layers are involved in heterogeneous reaction with NO2 as follows from the observed weak dependence of γ(geom) and decrease in γ(BET) with increasing sample mass. Heating the soot samples before exposure to NO2 increases the BET surface area, the <span class="hlt">HONO</span> yield, and the NO2 uptake coefficient due to the removal of the organic fraction from the soot backbone that unblocks active sites and makes them accessible for physical adsorption and chemical reactions. Our results support the oxidation-reduction mechanism involving</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AGUFM.A21B0851M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AGUFM.A21B0851M"><span>Identification of tropospheric emissions sources from satellite observations: Synergistic use of <span class="hlt">HCHO</span>, NO2, and SO2 trace gas measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marbach, T.; Beirle, S.; Khokhar, F.; Platt, U.</p> <p>2005-12-01</p> <p>We present case studies for combined <span class="hlt">HCHO</span>, NO2, and SO2 satellite observations, derived from GOME measurements. Launched on the ERS-2 satellite in April 1995, GOME has already performed continuous operations over 8 years providing global observations of the different trace gases. In this way, satellite observations provide unique opportunities for the identifications of trace gas sources. The satellite <span class="hlt">HCHO</span> observations provide information concerning the localization of biomass burning (intense source of <span class="hlt">HCHO</span>). The principal biomass burning areas can be observed in the Amazon basin region and in central Africa Weaker <span class="hlt">HCHO</span> sources (south east of the United States, northern part of the Amazon basin, and over the African tropical forest), not correlated with biomass burning, could be due to biogenic isoprene emissions. The <span class="hlt">HCHO</span> data can be compared with NO2 and SO2 results to identify more precisely the tropospheric sources (biomass burning events, human activities, additional sources like volcanic emissions). Biomass burning are important tropospheric sources for both <span class="hlt">HCHO</span> and NO2. Nevertheless <span class="hlt">HCHO</span> reflects more precisely the biomass burning as it appears in all biomass burning events. NO2 correlate with <span class="hlt">HCHO</span> over Africa (grassland fires) but not over Indonesia (forest fires). In south America, an augmentation of the NO2 concentrations can be observed with the fire shift from the forest to grassland vegetation. So there seems to be a dependence between the NO2 emissions during biomass burning and the vegetation type. Other high <span class="hlt">HCHO</span>, SO2, and NO2 emissions can be correlated with climatic events like the El Nino in 1997, which induced dry conditions in Indonesia causing many forest fires.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19261298','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19261298"><span>Assessing depression outcome in patients with moderate dementia: sensitivity of the <span class="hlt">HoNOS</span>65+ scale.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Canuto, Alessandra; Rudhard-Thomazic, Valérie; Herrmann, François R; Delaloye, Christophe; Giannakopoulos, Panteleimon; Weber, Kerstin</p> <p>2009-08-15</p> <p>To date, there is no widely accepted clinical scale to monitor the evolution of depressive symptoms in demented patients. We assessed the sensitivity to treatment of a validated French version of the Health of the Nation Outcome Scale (<span class="hlt">HoNOS</span>) 65+ compared to five routinely used scales. Thirty elderly inpatients with ICD-10 diagnosis of dementia and depression were evaluated at admission and discharge using paired t-test. Using the Brief Psychiatric Rating Scale (BPRS) "depressive mood" item as gold standard, a receiver operating characteristic curve (ROC) analysis assessed the validity of <span class="hlt">HoNOS</span>65+F "depressive symptoms" item score changes. Unlike Geriatric Depression Scale, Mini Mental State Examination and Activities of Daily Living scores, BPRS scores decreased and Global Assessment Functioning Scale score increased significantly from admission to discharge. Amongst <span class="hlt">HoNOS</span>65+F items, "behavioural disturbance", "depressive symptoms", "activities of daily life" and "drug management" items showed highly significant changes between the first and last day of hospitalization. The ROC analysis revealed that changes in the <span class="hlt">HoNOS</span>65+F "depressive symptoms" item correctly classified 93% of the cases with good sensitivity (0.95) and specificity (0.88) values. These data suggest that the <span class="hlt">HoNOS</span>65+F "depressive symptoms" item may provide a valid assessment of the evolution of depressive symptoms in demented patients.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26629972','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26629972"><span>Three-Dimensional Ordered Mesoporous Mn<span class="hlt">O</span>2-Supported Ag Nanoparticles for Catalytic Removal of Formaldehyde.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bai, Bingyang; Qiao, Qi; Arandiyan, Hamidreza; Li, Junhua; Hao, Jiming</p> <p>2016-03-01</p> <p>Three-dimensional (<span class="hlt">3</span>D) ordered mesoporous Ag/Mn<span class="hlt">O</span>2 catalyst was prepared by impregnation method based on <span class="hlt">3</span>D-Mn<span class="hlt">O</span>2 and used for catalytic oxidation of <span class="hlt">HCHO</span>. Ag nanoparticles are uniformly distributed on the polycrystalline wall of <span class="hlt">3</span>D-Mn<span class="hlt">O</span>2. The addition of Ag does not change the <span class="hlt">3</span>D ordered mesoporous structure of the Ag/Mn<span class="hlt">O</span>2, but does reduce the pore size and surface area. Ag nanoparticles provide sufficient active site for the oxidation reaction of <span class="hlt">HCHO</span>, and Ag (111) crystal facets in the Ag/Mn<span class="hlt">O</span>2 are active faces. The 8.9% Ag/Mn<span class="hlt">O</span>2 catalyst shows a higher normalized rate (10.1 nmol·s(-1)·m(-2) at 110 °C) and TOF (0.007 s(-1) at 110 °C) under 1300 ppm of <span class="hlt">HCHO</span> and 150 000 h(-1) of GHSV, and its apparent activation energy of the reaction is the lowest (39.1 kJ/mol). More Ag active sites, higher low-temperature reducibility, more abundant surface lattice oxygen species, oxygen vacancies, and lattice defects generated from interaction Ag with Mn<span class="hlt">O</span>2 are responsible for the excellent catalytic performance of <span class="hlt">HCHO</span> oxidation on the 8.9% Ag/Mn<span class="hlt">O</span>2 catalyst. The 8.9% Ag/Mn<span class="hlt">O</span>2 catalyst remained highly active and stable under space velocity increasing from 60 000 to 150 000 h(-1), under initial <span class="hlt">HCHO</span> concentration increasing from 500 to 1300 ppm, and under the presence of humidity, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ACP....18..799M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ACP....18..799M"><span>Emission of nitrous acid from soil and biological soil crusts represents an important source of <span class="hlt">HONO</span> in the remote atmosphere in Cyprus</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Meusel, Hannah; Tamm, Alexandra; Kuhn, Uwe; Wu, Dianming; Lena Leifke, Anna; Fiedler, Sabine; Ruckteschler, Nina; Yordanova, Petya; Lang-Yona, Naama; Pöhlker, Mira; Lelieveld, Jos; Hoffmann, Thorsten; Pöschl, Ulrich; Su, Hang; Weber, Bettina; Cheng, Yafang</p> <p>2018-01-01</p> <p>Soil and biological soil crusts can emit nitrous acid (<span class="hlt">HONO</span>) and nitric oxide (NO). The terrestrial ground surface in arid and semiarid regions is anticipated to play an important role in the local atmospheric <span class="hlt">HONO</span> budget, deemed to represent one of the unaccounted-for <span class="hlt">HONO</span> sources frequently observed in field studies. In this study <span class="hlt">HONO</span> and NO emissions from a representative variety of soil and biological soil crust samples from the Mediterranean island Cyprus were investigated under controlled laboratory conditions. A wide range of fluxes was observed, ranging from 0.6 to 264 ng m-2 s-1 <span class="hlt">HONO</span>-N at optimal soil water content (20-30 % of water holding capacity, WHC). Maximum NO-N fluxes at this WHC were lower (0.8-121 ng m-2 s-1). The highest emissions of both reactive nitrogen species were found from bare soil, followed by light and dark cyanobacteria-dominated biological soil crusts (biocrusts), correlating well with the sample nutrient levels (nitrite and nitrate). Extrapolations of lab-based <span class="hlt">HONO</span> emission studies agree well with the unaccounted-for <span class="hlt">HONO</span> source derived previously for the extensive CYPHEX field campaign, i.e., emissions from soil and biocrusts may essentially close the Cyprus <span class="hlt">HONO</span> budget.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25221921','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25221921"><span>C1 metabolism plays an important role during formaldehyde metabolism and detoxification in petunia under liquid <span class="hlt">HCHO</span> stress.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Wei; Tang, Lijuan; Sun, Huiqun; Han, Shuang; Wang, Xinjia; Zhou, Shengen; Li, Kunzhi; Chen, Limei</p> <p>2014-10-01</p> <p>Petunia hybrida is a model ornamental plant grown worldwide. To understand the <span class="hlt">HCHO</span>-uptake efficiency and metabolic mechanism of petunia, the aseptic petunia plants were treated in <span class="hlt">HCHO</span> solutions. An analysis of <span class="hlt">HCHO</span>-uptake showed that petunia plants effectively removed <span class="hlt">HCHO</span> from 2, 4 and 6 mM <span class="hlt">HCHO</span> solutions. The (13)C NMR analyses indicated that H(13)CHO was primarily used to synthesize [5-(13)C]methionine (Met) via C1 metabolism in petunia plants treated with 2 mM H(13)CHO. Pretreatment with cyclosporin A (CSA) or l-carnitine (LC), the inhibitors of mitochondrial permeability transition pores, did not affect the synthesis of [5-(13)C]Met in petunia plants under 2 mM H(13)CHO stress, indicating that the Met-generated pathway may function in the cytoplasm. Under 4 or 6 mM liquid H(13)CHO stress, H(13)CHO metabolism in petunia plants produced considerable amount of H(13)COOH and [2-(13)C]glycine (Gly) through C1 metabolism and a small amount of [U-(13)C]Gluc via the Calvin Cycle. Pretreatment with CSA or LC significantly inhibited the production of [2-(13)C]Gly in 6 mM H(13)CHO-treated petunia plants, which suggests that chloroplasts and peroxisomes might be involved in the generation of [2-(13)C]Gly. These results revealed that the C1 metabolism played an important role, whereas the Calvin Cycle had only a small contribution during <span class="hlt">HCHO</span> metabolism and detoxification in petunia under liquid <span class="hlt">HCHO</span> stress. Copyright © 2014 Elsevier Masson SAS. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JGRD..119.5583P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JGRD..119.5583P"><span>Intercomparison of field measurements of nitrous acid (<span class="hlt">HONO</span>) during the SHARP campaign</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pinto, J. P.; Dibb, J.; Lee, B. H.; Rappenglück, B.; Wood, E. C.; Levy, M.; Zhang, R.-Y.; Lefer, B.; Ren, X.-R.; Stutz, J.; Tsai, C.; Ackermann, L.; Golovko, J.; Herndon, S. C.; Oakes, M.; Meng, Q.-Y.; Munger, J. W.; Zahniser, M.; Zheng, J.</p> <p>2014-05-01</p> <p>Because of the importance of <span class="hlt">HONO</span> as a radical reservoir, consistent and accurate measurements of its concentration are needed. As part of SHARP (Study of Houston Atmospheric Radical Precursors), time series of <span class="hlt">HONO</span> were obtained by six different measurement techniques on the roof of the Moody Tower at the University of Houston. Techniques used were long path differential optical absorption spectroscopy (DOAS), stripping coil-visible absorption photometry (SC-AP), long path absorption photometry (LOPAP®), mist chamber/ion chromatography (MC-IC), quantum cascade-tunable infrared laser differential absorption spectroscopy (QC-TILDAS), and ion drift-chemical ionization mass spectrometry (ID-CIMS). Various combinations of techniques were in operation from 15 April through 31 May 2009. All instruments recorded a similar diurnal pattern of <span class="hlt">HONO</span> concentrations with higher median and mean values during the night than during the day. Highest values were observed in the final 2 weeks of the campaign. Inlets for the MC-IC, SC-AP, and QC-TILDAS were collocated and agreed most closely with each other based on several measures. Largest differences between pairs of measurements were evident during the day for concentrations < 100 parts per trillion (ppt). Above 200 ppt, concentrations from the SC-AP, MC-IC, and QC-TILDAS converged to within about 20%, with slightly larger discrepancies when DOAS was considered. During the first 2 weeks, <span class="hlt">HONO</span> measured by ID-CIMS agreed with these techniques, but ID-CIMS reported higher values during the afternoon and evening of the final 4 weeks, possibly from interference from unknown sources. A number of factors, including building related sources, likely affected measured concentrations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JChPh.148f4301W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JChPh.148f4301W"><span>Detection of transient infrared absorption of SO<span class="hlt">3</span> and 1,<span class="hlt">3</span>,2-dioxathietane-2,2-dioxide [cyc-(CH2)<span class="hlt">O</span>(SO2)<span class="hlt">O</span>] in the reaction CH2OO+SO2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Yi-Ying; Dash, Manas Ranjan; Chung, Chao-Yu; Lee, Yuan-Pern</p> <p>2018-02-01</p> <p>We recorded time-resolved infrared absorption spectra of transient species produced on irradiation at 308 nm of a flowing mixture of CH2I2/<span class="hlt">O</span>2/N2/SO2 at 298 K. Bands of CH2OO were observed initially upon irradiation; their decrease in intensity was accompanied by the appearance of an intense band at 1391.5 cm-1 that is associated with the degenerate SO-stretching mode of SO<span class="hlt">3</span>, two major bands of <span class="hlt">HCHO</span> at 1502 and 1745 cm-1, and five new bands near >1340, 1225, 1100, 940, and 880 cm-1. The band near 1340 cm-1 was interfered by absorption of SO2 and SO<span class="hlt">3</span>, so its band maximum might be greater than 1340 cm-1. SO<span class="hlt">3</span> in its internally excited states was produced initially and became thermalized at a later period. The rotational contour of the band of thermalized SO<span class="hlt">3</span> agrees satisfactorily with the reported spectrum of SO<span class="hlt">3</span>. These five new bands are tentatively assigned to an intermediate 1,<span class="hlt">3</span>,2-dioxathietane-2,2-dioxide [cyc-(CH2)<span class="hlt">O</span>(SO2)<span class="hlt">O</span>] according to comparison with anharmonic vibrational wavenumbers and relative IR intensities predicted for this intermediate. Observation of a small amount of cyc-(CH2)<span class="hlt">O</span>(SO2)<span class="hlt">O</span> is consistent with the expected reaction according to the potential energy scheme predicted previously. SO<span class="hlt">3</span>+<span class="hlt">HCHO</span> are the major products of the title reaction. The other predicted product channel HCOOH+SO2 was unobserved and its branching ratio was estimated to be <5%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29448796','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29448796"><span>Detection of transient infrared absorption of SO<span class="hlt">3</span> and 1,<span class="hlt">3</span>,2-dioxathietane-2,2-dioxide [cyc-(CH2)<span class="hlt">O</span>(SO2)<span class="hlt">O</span>] in the reaction CH2OO+SO2.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Yi-Ying; Dash, Manas Ranjan; Chung, Chao-Yu; Lee, Yuan-Pern</p> <p>2018-02-14</p> <p>We recorded time-resolved infrared absorption spectra of transient species produced on irradiation at 308 nm of a flowing mixture of CH 2 I 2 /<span class="hlt">O</span> 2 /N 2 /SO 2 at 298 K. Bands of CH 2 OO were observed initially upon irradiation; their decrease in intensity was accompanied by the appearance of an intense band at 1391.5 cm -1 that is associated with the degenerate SO-stretching mode of SO <span class="hlt">3</span> , two major bands of <span class="hlt">HCHO</span> at 1502 and 1745 cm -1 , and five new bands near >1340, 1225, 1100, 940, and 880 cm -1 . The band near 1340 cm -1 was interfered by absorption of SO 2 and SO <span class="hlt">3</span> , so its band maximum might be greater than 1340 cm -1 . SO <span class="hlt">3</span> in its internally excited states was produced initially and became thermalized at a later period. The rotational contour of the band of thermalized SO <span class="hlt">3</span> agrees satisfactorily with the reported spectrum of SO <span class="hlt">3</span> . These five new bands are tentatively assigned to an intermediate 1,<span class="hlt">3</span>,2-dioxathietane-2,2-dioxide [cyc-(CH 2 )<span class="hlt">O</span>(SO 2 )<span class="hlt">O</span>] according to comparison with anharmonic vibrational wavenumbers and relative IR intensities predicted for this intermediate. Observation of a small amount of cyc-(CH 2 )<span class="hlt">O</span>(SO 2 )<span class="hlt">O</span> is consistent with the expected reaction according to the potential energy scheme predicted previously. SO <span class="hlt">3</span> +<span class="hlt">HCHO</span> are the major products of the title reaction. The other predicted product channel HCOOH+SO 2 was unobserved and its branching ratio was estimated to be <5%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.A11D0125B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.A11D0125B"><span>Laboratory Investigation of Trace Gas Emissions from Biomass Burning on DoD Bases</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burling, I. R.; Yokelson, R. J.; Griffith, D. W.; Roberts, J. M.; Veres, P. R.; Warneke, C.; Johnson, T. J.</p> <p>2009-12-01</p> <p>Vegetation representing fuels commonly managed with prescribed fires was collected from five DoD bases and burned under controlled conditions at the USFS Firelab in Missoula, MT. The smoke emissions were measured with a large suite of state-of-the-art instrumentation. Seventy-seven fires were conducted and the smoke composition data will improve DoD land managers’ ability to assess the impact of prescribed fires on local air quality. A key instrument used in the measurement of the gas phase species in smoke was an open-path FTIR (OP-FTIR) spectrometer, built and operated by the Universities of Montana and Wollongong. The OP-FTIR has to date detected and quantified 20 gas phase species - CO2, CO, H2<span class="hlt">O</span>, N2<span class="hlt">O</span>, NO2, NO, <span class="hlt">HONO</span>, NH<span class="hlt">3</span>, HCl, SO2, CH4, CH<span class="hlt">3</span>OH, <span class="hlt">HCHO</span>, HCOOH, C2H2, C2H4, CH<span class="hlt">3</span>COOH, HCN, propylene and furan. The spectra were analyzed using a non-linear least squares fitting routine that included reference spectra recently acquired at the Pacific Northwest National Laboratories. Preliminary results from the OP-FTIR analysis are reported here. Of particular interest, gas-phase nitrous acid (<span class="hlt">HONO</span>) was detected simultaneously by the OP-FTIR and negative-ion proton-transfer chemical ionization spectrometer (NI-PT-CIMS), with preliminary fire-integrated molar emission ratios (relative to NOx) ranging from approximately 0.03 to 0.20, depending on the vegetation type. <span class="hlt">HONO</span> is an important precursor in the production of OH, the primary oxidizing species in the atmosphere. There existed little previous data documenting <span class="hlt">HONO</span> emissions from either wild or prescribed fires. The non-methane organic emissions were dominated by oxygenated species, which can be further oxidized and thus involved in secondary aerosol formation. Elevated amounts of gas-phase HCl were also detected in the smoke, with the amounts varying depending on location and vegetation type.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1569386','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1569386"><span>Formation and fate of gaseous and particulate mutagens and carcinogens in real and simulated atmospheres.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Pitts, J N</p> <p>1983-01-01</p> <p>The growing use of coal for heating and electric power generation and diesel engines in light duty motor vehicles will increase not only the existing atmospheric concentrations of criteria pollutants such as NO2, SO2, <span class="hlt">O</span><span class="hlt">3</span> and fine particulates, but also the concentrations of a number of highly reactive gaseous copollutants such as <span class="hlt">HONO</span>, <span class="hlt">HONO</span>2, PAN and the nitrate radical, NO<span class="hlt">3</span>. These gaseous noncriteria pollutants are of interest not only because of their roles in the chemistry of the "clean" and polluted troposphere, including "acid rain," but also because they may pose health risks disproportionate to their relatively low ambient concentrations, and through complex heterogeneous reactions, they may serve as precursors or catalysts in the formation of "nonclassical" particulate mutagens and carcinogens such as certain nitroarenes associated with combustion generated particulate polycyclic organic matter (POM). Results of research efforts to establish current ambient levels of these noncriteria pollutants and to develop an understanding of their sources, formation and sinks are reported here. First, long pathlength (greater than or equal to 1 km) infrared and UV-visible spectroscopic studies of ambient levels of gaseous <span class="hlt">HONO</span>, NO<span class="hlt">3</span>, <span class="hlt">HONO</span>2, PAN, <span class="hlt">HCHO</span> and HCOOH in southern California atmospheres are described, and data given on their ambient concentrations. Second, an integrated chemical/microbiological investigation is described. It is directed toward identifying the nature of direct-acting mutagens found in extracts of diesel and ambient POM, as well as those formed upon exposure of environmentally relevant PAH to simulated natural and polluted atmospheres. The identification of certain of these mutagens, including a newly identified class of mutagenic PAH-lactones is discussed, along with the mechanisms of their formation and fate in the natural and polluted troposphere. PMID:6337822</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20100031236&hterms=air+quality&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dair%2Bquality','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20100031236&hterms=air+quality&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dair%2Bquality"><span>The Utility of the OMI <span class="hlt">HCHO</span> and NO2 Data Products in Air Quality Decision- Making Activities</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Duncan, Bryan N.</p> <p>2010-01-01</p> <p>We will present three related air quality applications of the OMI <span class="hlt">HCHO</span> (formaldehyde) and NO2 (nitrogen dioxide) data products, which we us to support mission planning of an OMI-like instrument for the proposed GEO-CAPE satellite that has as one of its objectives to study air quality from space. First, we will discuss a novel and practical application of the data products to the "weight of evidence" in the air quality decision-making process (e.g., State Implementation Plan (SIP)) for a city, region, or state to demonstrate that it is making progress toward attainment of the National Ambient Air Quality Standard (NAAQS) for ozone. Any trend, or lack thereof, in the observed OMI <span class="hlt">HCHO</span>/NO2, which we use as an air quality indicator, may support that an emission control strategy implemented to reduce ozone is or is not occurring for a metropolitan area. Second, we will discuss how we use variations in the OMI <span class="hlt">HCHO</span> product as a proxy for variability in the biogenic hydrocarbon, isoprene, which is an important player for the formation of high levels of ozone and the dominant source of <span class="hlt">HCHO</span> in the eastern U.S. Third, we will discuss the variability of NO2 in the U.S. as indicated by the OMI NO2 product. In addition, we will show the impact of the 2005 hurricanes on pollutant emissions, including those associated with the intensive oil extraction and refining activities, in the Gulf of Mexico region using the OMI NO2 product. The variability of <span class="hlt">HCHO</span> and NO2 as indicated by OMI helps us to understand changes in the OMI <span class="hlt">HCHO</span>/NO2 and the implications for ozone formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.7603P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.7603P"><span>A new DOAS instrument on long-distance IAGOS-CARIBIC flights and airborne DOAS applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Penth, Lara; Frieß, Udo; Pöhler, Denis; Platt, Ulrich; Zahn, Andreas</p> <p>2017-04-01</p> <p>Within the IAGOS-CARIBIC project airborne DOAS (Differential Optical Absorption Spectroscopy) measurements of atmospheric trace gases are performed aboard a commercial long range passenger aircraft from Lufthansa since 2005. They provide a unique dataset for episodic, long-term and seasonal observations. The DOAS instrument is the only remote sensing technique aboard. DOAS is a well-established remote sensing technique to retrieve trace gas columns in the atmosphere from scattered light spectra of the sun. A series of trace gas species can be observed simultaneously, including nitrogen dioxide (NO2), sulphur dioxide (SO2), bromine oxide (Br<span class="hlt">O</span>), nitrous acid (<span class="hlt">HONO</span>), formaldehyde (<span class="hlt">HCHO</span>) and ozone (<span class="hlt">O</span><span class="hlt">3</span>). Since DOAS is a contact-free measurement technique, it is specially well suited for measuring highly reactive trace gases. It is widely used on different platforms and the airborne DOAS measurements are filling the gap between ground-based measurements and satellite data. The CARIBIC DOAS instrument is divided into an instrument unit within the CARIBIC container in the cargo hold of the aircraft, a telescope unit, which is specially designed for the permanently mounted pylon underneath the aircraft, and fiber optics in between. The instrument unit consists of three temperature stabilized spectrometers and the readout and control electronics. The telescope unit contains three telescopes, which observe scattered sunlight to the right under the elevation angles of +10˚ , -10˚ and -82˚ (nadir) relative to the horizon. This measurement geometry allows the separation of boundary layer, free tropospheric and stratospheric trace gas columns along the flight track. A new DOAS instrument was designed and installed in 2016 (first flights expected from March 2017) to improve the detection limits of NO2, SO2, Br<span class="hlt">O</span>, <span class="hlt">HCHO</span>, <span class="hlt">HONO</span>, <span class="hlt">O</span><span class="hlt">3</span> and <span class="hlt">O</span>4. Furthermore, an extended wavelength range allows to measure in addition iodine monoxide (a potentially important oxidant in the free troposphere</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23108328','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23108328"><span>Kinetics of the benzyl + <span class="hlt">O</span>(<span class="hlt">3</span>P) reaction: a quantum chemical/statistical reaction rate theory study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>da Silva, Gabriel; Bozzelli, Joseph W</p> <p>2012-12-14</p> <p>The resonance stabilized benzyl radical is an important intermediate in the combustion of aromatic hydrocarbons and in polycyclic aromatic hydrocarbon (PAH) formation in flames. Despite being a free radical, benzyl is relatively stable in thermal, oxidizing environments, and is predominantly removed through bimolecular reactions with open-shell species other than <span class="hlt">O</span>(2). In this study the reaction of benzyl with ground-state atomic oxygen, <span class="hlt">O</span>((<span class="hlt">3</span>)P), is examined using quantum chemistry and statistical reaction rate theory. C(7)H(7)<span class="hlt">O</span> energy surfaces are generated at the G<span class="hlt">3</span>SX level, and include several novel pathways. Transition state theory is used to describe elementary reaction kinetics, with canonical variational transition state theory applied for barrierless <span class="hlt">O</span> atom association with benzyl. Apparent rate constants and branching ratios to different product sets are obtained as a function of temperature and pressure from solving the time-dependent master equation, with RRKM theory for microcanonical k(E). These simulations indicate that the benzyl + <span class="hlt">O</span> reaction predominantly forms the phenyl radical (C(6)H(5)) plus formaldehyde (<span class="hlt">HCHO</span>), with lesser quantities of the C(7)H(6)<span class="hlt">O</span> products benzaldehyde, ortho-quinone methide, and para-quinone methide (+H), along with minor amounts of the formyl radical (HCO) + benzene. Addition of <span class="hlt">O</span>((<span class="hlt">3</span>)P) to the methylene site in benzyl produces a highly vibrationally excited C(7)H(7)<span class="hlt">O</span>* adduct, the benzoxyl radical, which can β-scission to benzaldehyde + H and phenyl + <span class="hlt">HCHO</span>. In order to account for the experimental observation of benzene as the major reaction product, a roaming radical mechanism is proposed that converts the nascent products phenyl and <span class="hlt">HCHO</span> to benzene + HCO. Oxygen atom addition at the ortho and para ring sites in benzyl, which has not been previously considered, is shown to lead to the quinone methides + H; these species are less-stable isomers of benzaldehyde that are proposed as important combustion intermediates, but</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ACP....10.6969V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ACP....10.6969V"><span>Oxidative capacity of the Mexico City atmosphere - Part 1: A radical source perspective</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Volkamer, R.; Sheehy, P.; Molina, L. T.; Molina, M. J.</p> <p>2010-07-01</p> <p>A detailed analysis of OH, HO2 and RO2 radical sources is presented for the near field photochemical regime inside the Mexico City Metropolitan Area (MCMA). During spring of 2003 (MCMA-2003 field campaign) an extensive set of measurements was collected to quantify time-resolved ROx (sum of OH, HO2, RO2) radical production rates from day- and nighttime radical sources. The Master Chemical Mechanism (MCMv<span class="hlt">3</span>.1) was constrained by measurements of (1) concentration time-profiles of photosensitive radical precursors, i.e., nitrous acid (<span class="hlt">HONO</span>), formaldehyde (<span class="hlt">HCHO</span>), ozone (<span class="hlt">O</span><span class="hlt">3</span>), glyoxal (CHOCHO), and other oxygenated volatile organic compounds (OVOCs); (2) respective photolysis-frequencies (J-values); (<span class="hlt">3</span>) concentration time-profiles of alkanes, alkenes, and aromatic VOCs (103 compound are treated) and oxidants, i.e., OH- and NO<span class="hlt">3</span> radicals, <span class="hlt">O</span><span class="hlt">3</span>; and (4) NO, NO2, meteorological and other parameters. The ROx production rate was calculated directly from these observations; the MCM was used to estimate further ROx production from unconstrained sources, and express overall ROx production as OH-equivalents (i.e., taking into account the propagation efficiencies of RO2 and HO2 radicals into OH radicals). Daytime radical production is found to be about 10-25 times higher than at night; it does not track the abundance of sunlight. 12-h average daytime contributions of individual sources are: Oxygenated VOC other than <span class="hlt">HCHO</span> about 33%; <span class="hlt">HCHO</span> and <span class="hlt">O</span><span class="hlt">3</span> photolysis each about 20%; <span class="hlt">O</span><span class="hlt">3</span>/alkene reactions and <span class="hlt">HONO</span> photolysis each about 12%, other sources <<span class="hlt">3</span>%. Nitryl chloride photolysis could potentially contribute ~15% additional radicals, while NO2* + water makes - if any - a very small contribution (~2%). The peak radical production of ~7.5 107 molec cm-<span class="hlt">3</span> s-1 is found already at 10:00 a.m., i.e., more than 2.5 h before solar noon. <span class="hlt">O</span><span class="hlt">3</span>/alkene reactions are indirectly responsible for ~33% of these radicals. Our measurements and analysis comprise a database that enables testing of the representation of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.A51C0032B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.A51C0032B"><span>Combustion Processes Indoors: a Source of High OH Radical Concentrations Through the Photolysis of <span class="hlt">Hono</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bartolomei, V.; Gomez Alvarez, E.; Glor, M.; Gligorovski, S.; Temime-Roussel, B.; Quivet, E.; Strekowski, R.; Zetzsch, C.; Held, A. B.; Wortham, H.</p> <p>2013-12-01</p> <p>Hydroxyl radical (OH) is one of the most important oxidant species in the atmosphere controlling its self-oxidizing capacity. The main sources of OH radicals are photolysis of ozone and photolysis of nitrous acid (<span class="hlt">HONO</span>), among the others. In the indoor air, the ozonolysis of alkenes has been suggested as the main OH formation pathway. The possibility for OH formation through photolytic pathways in the indoor environment has been, up to now, ignored (Gómez Alvarez et al., 2012). Models and indirect measurements to the present time predicted concentrations of OH radicals in the order of 104 -105 cm-<span class="hlt">3</span>. Recently, by direct measurements we have detected high OH radical concentrations of 1.8 106 cm-<span class="hlt">3</span> in a classroom in Marseille and we demonstrated that its main source is the photolysis of <span class="hlt">HONO</span> (Gómez Alvarez et al., 2013). The concentrations of <span class="hlt">HONO</span> are quite high indoors, reaching levels in the order of a few tens of ppbV (Gómez Alvarez et al., 2013). This is mainly due to 1) direct combustion sources and 2) heterogeneous reactions of NO2 on the numerous surfaces present in the indoor environment. <span class="hlt">HONO</span> levels of 30 ppb were measured in a previous campaign carried out in Bayreuth in July 2012 as direct emissions from the combustion of a candle. The combination between so high concentrations of <span class="hlt">HONO</span> and higher than expected light transmissions indoors (or indoor artificial lighting) could have a significant impact on the OH concentrations indoors which could feasibly become considerably higher than we measured in our school campaign (Gomez Alvarez et al., 2013). In order to evaluate these upper limits under combustion conditions in the indoor environment, we have carried out a campaign in the LOTASC chamber (Bayreuth, Germany). For this aim, the exhaust fumes from the burning of a commonly used domestic candle have been introduced in the chamber. The chamber was irradiated under well research indoor lighting conditions. A thorough characterization of light intensities</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24738832','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24738832"><span>Sodium-promoted Pd/Ti<span class="hlt">O</span>2 for catalytic oxidation of formaldehyde at ambient temperature.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Changbin; Li, Yaobin; Wang, Yafei; He, Hong</p> <p>2014-05-20</p> <p>Catalytic oxidation of formaldehyde (<span class="hlt">HCHO</span>) to CO2 at ambient conditions is of great interest for indoor <span class="hlt">HCHO</span> purification. Here, we report that sodium-doped Pd/Ti<span class="hlt">O</span>2 is a highly effective catalyst for the catalytic oxidation of <span class="hlt">HCHO</span> at room temperature. It was observed that Na doping has a dramatic promotion effect on the Pd/Ti<span class="hlt">O</span>2 catalyst and that nearly 100% <span class="hlt">HCHO</span> conversion could be achieved over the 2Na-Pd/Ti<span class="hlt">O</span>2 catalyst at a GHSV of 95000 h(-1) and <span class="hlt">HCHO</span> inlet concentration of 140 ppm at 25 °C. The mechanism of the Na-promotion effect was investigated by using Brunauer-Emmett-Teller (BET), X-ray diffraction (XRD), CO chemisorption, Temperature-programmed reduction by H2 (H2-TPR), X-ray photoelectron spectroscopy (XPS) and temperature-programmed desorption of <span class="hlt">O</span>2 (<span class="hlt">O</span>2-TPD) methods. The results showed that Na species addition can induce and further stabilize a negatively charged and well-dispersed Pd species, which then facilitates the activation of H2<span class="hlt">O</span> and chemisorbed oxygen, therefore resulting in the high performance of the 2Na-Pd/Ti<span class="hlt">O</span>2 catalyst for the ambient <span class="hlt">HCHO</span> destruction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ACP....18.5931H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ACP....18.5931H"><span>Ship-based MAX-DOAS measurements of tropospheric NO2, SO2, and <span class="hlt">HCHO</span> distribution along the Yangtze River</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hong, Qianqian; Liu, Cheng; Chan, Ka Lok; Hu, Qihou; Xie, Zhouqing; Liu, Haoran; Si, Fuqi; Liu, Jianguo</p> <p>2018-04-01</p> <p>In this paper, we present ship-based Multi-Axis Differential Optical Absorption Spectroscopy (MAX-DOAS) measurements of tropospheric trace gases' distribution along the Yangtze River during winter 2015. The measurements were performed along the Yangtze River between Shanghai and Wuhan, covering major industrial areas in eastern China. Tropospheric vertical column densities (VCDs) of nitrogen dioxide (NO2), sulfur dioxide (SO2), and formaldehyde (<span class="hlt">HCHO</span>) were retrieved using the air mass factor calculated by the radiative transfer model. Enhanced tropospheric NO2 and SO2 VCDs were detected over downwind areas of industrial zones over the Yangtze River. In addition, spatial distributions of atmospheric pollutants are strongly affected by meteorological conditions; i.e., positive correlations were found between concentration of pollutants and wind speed over these areas, indicating strong influence of transportation of pollutants from high-emission upwind areas along the Yangtze River. Comparison of tropospheric NO2 VCDs between ship-based MAX-DOAS and Ozone Monitoring Instrument (OMI) satellite observations shows good agreement with each other, with a Pearson correlation coefficient (R) of 0.82. In this study, the NO2 / SO2 ratio was used to estimate the relative contributions of industrial sources and vehicle emissions to ambient NO2 levels. Analysis results of the NO2 / SO2 ratio show a higher contribution of industrial NO2 emissions in Jiangsu Province, while NO2 levels in Jiangxi and Hubei provinces are mainly related to vehicle emissions. These results indicate that different pollution control strategies should be applied in different provinces. In addition, multiple linear regression analysis of ambient carbon monoxide (CO) and odd oxygen (Ox) indicated that the primary emission and secondary formation of <span class="hlt">HCHO</span> contribute 54.4 ± <span class="hlt">3</span>.7 % and 39.<span class="hlt">3</span> ± 4.<span class="hlt">3</span> % to the ambient <span class="hlt">HCHO</span>, respectively. The largest contribution from primary emissions in winter suggested that</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.6314D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.6314D"><span>Development of a harmonised multi sensor retrieval scheme for <span class="hlt">HCHO</span> within the Quality Assurance For Essential Climate Variables (QA4ECV) project</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>De Smedt, Isabelle; Richter, Andreas; Beirle, Steffen; Danckaert, Thomas; Van Roozendael, Michel; Yu, Huan; Bösch, Tim; Hilboll, Andreas; Peters, Enno; Doerner, Steffen; Wagner, Thomas; Wang, Yang; Lorente, Alba; Eskes, Henk; Van Geffen, Jos; Boersma, Folkert</p> <p>2016-04-01</p> <p>One of the main goals of the QA4ECV project is to define community best-practices for the generation of multi-decadal ECV data records from satellite instruments. QA4ECV will develop retrieval algorithms for the Land ECVs surface albedo, leaf area index (LAI), and fraction of active photosynthetic radiation (fAPAR), as well as for the Atmosphere ECV ozone and aerosol precursors nitrogen dioxide (NO2), formaldehyde (<span class="hlt">HCHO</span>), and carbon monoxide (CO). Here we assess best practices and provide recommendations for the retrieval of <span class="hlt">HCHO</span>. Best practices are established based on (1) a detailed intercomparison exercise between the QA4ECV partner's for each specific algorithm processing steps, (2) the feasibility of implementation, and (<span class="hlt">3</span>) the requirement to generate consistent multi-sensor multi-decadal data records. We propose a fitting window covering the 328.5-346 nm spectral interval for the morning sensors (GOME, SCIAMACHY and GOME-2) and an extension to 328.5-359 nm for OMI and GOME-2, allowed by improved quality of the recorded spectra. A high level of consistency between group algorithms is found when the retrieval settings are carefully aligned. However, the retrieval of slant columns is highly sensitive to any change in the selected settings. The use of a mean background radiance as DOAS reference spectrum allows for a stabilization of the retrievals. A background correction based on the reference sector method is recommended for implementation in the QA4ECV <span class="hlt">HCHO</span> algorithm as it further reduces retrieval uncertainties. <span class="hlt">HCHO</span> AMFs using different radiative transfer codes show a good overall consistency when harmonized settings are used. As for NO2, it is proposed to use a priori <span class="hlt">HCHO</span> profiles from the TM5 model. These are provided on a 1°x1° latitude-longitude grid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5130232','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5130232"><span>Utility of the Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) in Predicting Mental Health Service Costs for Patients with Common Mental Health Problems: Historical Cohort Study</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Twomey, Conal; Prina, A. Matthew; Baldwin, David S.; Das-Munshi, Jayati; Kingdon, David; Koeser, Leonardo; Prince, Martin J.; Stewart, Robert; Tulloch, Alex D.; Cieza, Alarcos</p> <p>2016-01-01</p> <p>Background Few countries have made much progress in implementing transparent and efficient systems for the allocation of mental health care resources. In England there are ongoing efforts by the National Health Service (NHS) to develop mental health ‘payment by results’ (PbR). The system depends on the ability of patient ‘clusters’ derived from the Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) to predict costs. We therefore investigated the associations of individual <span class="hlt">HoNOS</span> items and the Total <span class="hlt">HoNOS</span> score at baseline with mental health service costs at one year follow-up. Methods An historical cohort study using secondary care patient records from the UK financial year 2012–2013. Included were 1,343 patients with ‘common mental health problems’, represented by ICD-10 disorders between F32-48. Costs were based on patient contacts with community-based and hospital-based mental health services. The costs outcome was transformed into ‘high costs’ vs ‘regular costs’ in main analyses. Results After adjustment for covariates, 11 <span class="hlt">HoNOS</span> items were not associated with costs. The exception was ‘self-injury’ with an odds ratio of 1.41 (95% CI 1.10–2.99). Population attributable fractions (PAFs) for the contribution of <span class="hlt">HoNOS</span> items to high costs ranged from 0.6% (physical illness) to 22.4% (self-injury). After adjustment, the Total <span class="hlt">HoNOS</span> score was not associated with costs (OR 1.03, 95% CI 0.99–1.07). However, the PAF (33.<span class="hlt">3</span>%) demonstrated that it might account for a modest proportion of the incidence of high costs. Conclusions Our findings provide limited support for the utility of the self-injury item and Total <span class="hlt">HoNOS</span> score in predicting costs. However, the absence of associations for the remaining <span class="hlt">HoNOS</span> items indicates that current PbR clusters have minimal ability to predict costs, so potentially contributing to a misallocation of NHS resources across England. The findings may inform the development of mental health payment systems internationally</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_6 --> <div id="page_7" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="121"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27902745','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27902745"><span>Utility of the Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) in Predicting Mental Health Service Costs for Patients with Common Mental Health Problems: Historical Cohort Study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Twomey, Conal; Prina, A Matthew; Baldwin, David S; Das-Munshi, Jayati; Kingdon, David; Koeser, Leonardo; Prince, Martin J; Stewart, Robert; Tulloch, Alex D; Cieza, Alarcos</p> <p>2016-01-01</p> <p>Few countries have made much progress in implementing transparent and efficient systems for the allocation of mental health care resources. In England there are ongoing efforts by the National Health Service (NHS) to develop mental health 'payment by results' (PbR). The system depends on the ability of patient 'clusters' derived from the Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) to predict costs. We therefore investigated the associations of individual <span class="hlt">HoNOS</span> items and the Total <span class="hlt">HoNOS</span> score at baseline with mental health service costs at one year follow-up. An historical cohort study using secondary care patient records from the UK financial year 2012-2013. Included were 1,343 patients with 'common mental health problems', represented by ICD-10 disorders between F32-48. Costs were based on patient contacts with community-based and hospital-based mental health services. The costs outcome was transformed into 'high costs' vs 'regular costs' in main analyses. After adjustment for covariates, 11 <span class="hlt">HoNOS</span> items were not associated with costs. The exception was 'self-injury' with an odds ratio of 1.41 (95% CI 1.10-2.99). Population attributable fractions (PAFs) for the contribution of <span class="hlt">HoNOS</span> items to high costs ranged from 0.6% (physical illness) to 22.4% (self-injury). After adjustment, the Total <span class="hlt">HoNOS</span> score was not associated with costs (OR 1.03, 95% CI 0.99-1.07). However, the PAF (33.<span class="hlt">3</span>%) demonstrated that it might account for a modest proportion of the incidence of high costs. Our findings provide limited support for the utility of the self-injury item and Total <span class="hlt">HoNOS</span> score in predicting costs. However, the absence of associations for the remaining <span class="hlt">HoNOS</span> items indicates that current PbR clusters have minimal ability to predict costs, so potentially contributing to a misallocation of NHS resources across England. The findings may inform the development of mental health payment systems internationally, especially since the vast majority of countries have not progressed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ACP....1011115B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ACP....1011115B"><span>Laboratory measurements of trace gas emissions from biomass burning of fuel types from the southeastern and southwestern United States</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burling, I. R.; Yokelson, R. J.; Griffith, D. W. T.; Johnson, T. J.; Veres, P.; Roberts, J. M.; Warneke, C.; Urbanski, S. P.; Reardon, J.; Weise, D. R.; Hao, W. M.; de Gouw, J.</p> <p>2010-11-01</p> <p>Vegetation commonly managed by prescribed burning was collected from five southeastern and southwestern US military bases and burned under controlled conditions at the US Forest Service Fire Sciences Laboratory in Missoula, Montana. The smoke emissions were measured with a large suite of state-of-the-art instrumentation including an open-path Fourier transform infrared (OP-FTIR) spectrometer for measurement of gas-phase species. The OP-FTIR detected and quantified 19 gas-phase species in these fires: CO2, CO, CH4, C2H2, C2H4, C<span class="hlt">3</span>H6, <span class="hlt">HCHO</span>, HCOOH, CH<span class="hlt">3</span>OH, CH<span class="hlt">3</span>COOH, furan, H2<span class="hlt">O</span>, NO, NO2, <span class="hlt">HONO</span>, NH<span class="hlt">3</span>, HCN, HCl, and SO2. Emission factors for these species are presented for each vegetation type burned. Gas-phase nitrous acid (<span class="hlt">HONO</span>), an important OH precursor, was detected in the smoke from all fires. The <span class="hlt">HONO</span> emission factors ranged from 0.15 to 0.60 g kg-1 and were higher for the southeastern fuels. The fire-integrated molar emission ratios of <span class="hlt">HONO</span> (relative to NOx) ranged from approximately 0.03 to 0.20, with higher values also observed for the southeastern fuels. The majority of non-methane organic compound (NMOC) emissions detected by OP-FTIR were oxygenated volatile organic compounds (OVOCs) with the total identified OVOC emissions constituting 61 ± 12% of the total measured NMOC on a molar basis. These OVOC may undergo photolysis or further oxidation contributing to ozone formation. Elevated amounts of gas-phase HCl and SO2 were also detected during flaming combustion, with the amounts varying greatly depending on location and vegetation type. The fuels with the highest HCl emission factors were all located in the coastal regions, although HCl was also observed from fuels farther inland. Emission factors for HCl were generally higher for the southwestern fuels, particularly those found in the chaparral biome in the coastal regions of California.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ACPD...1016425B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ACPD...1016425B"><span>Laboratory measurements of trace gas emissions from biomass burning of fuel types from the Southeastern and Southwestern United States</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burling, I. R.; Yokelson, R. J.; Griffith, D. W. T.; Johnson, T. J.; Veres, P.; Roberts, J. M.; Warneke, C.; Urbanski, S. P.; Reardon, J.; Weise, D. R.; Hao, W. M.; de Gouw, J.</p> <p>2010-07-01</p> <p>Vegetation commonly managed by prescribed burning was collected from five southeastern and southwestern US military bases and burned under controlled conditions at the US Forest Service Fire Sciences Laboratory in Missoula, Montana. The smoke emissions were measured with a large suite of state-of-the-art instrumentation including an open-path Fourier transform infrared (OP-FTIR) spectrometer for measurement of gas-phase species. The OP-FTIR detected and quantified 19 gas-phase species in these fires: CO2, CO, CH4, C2H2, C2H4, C<span class="hlt">3</span>H6, <span class="hlt">HCHO</span>, HCOOH, CH<span class="hlt">3</span>OH, CH<span class="hlt">3</span>COOH, furan, H2<span class="hlt">O</span>, NO, NO2, <span class="hlt">HONO</span>, NH<span class="hlt">3</span>, HCN, HCl, and SO2. Emission factors for these species are presented for each vegetation type burned. Gas-phase nitrous acid (<span class="hlt">HONO</span>), an important OH precursor, was detected in the smoke from all fires. The <span class="hlt">HONO</span> emission factors ranged from 0.15 to 0.60 g kg-1 and were higher for the southeastern fuels. The fire-integrated molar emission ratios of <span class="hlt">HONO</span> (relative to NOx) ranged from approximately 0.03 to 0.20, with higher values also observed for the southeastern fuels. The majority of non-methane organic compound (NMOC) emissions detected by OP-FTIR were oxygenated volatile organic compounds (OVOCs) with the total identified OVOC emissions constituting 61±12% of the total measured NMOC on a molar basis. These OVOC may undergo photolysis or further oxidation contributing to ozone formation. Elevated amounts of gas-phase HCl and SO2 were also detected during flaming combustion, with the amounts varying greatly depending on location and vegetation type. The fuels with the highest HCl emission factors were all located in the coastal regions, although HCl was also observed from fuels farther inland. Emission factors for HCl were generally higher for the southwestern fuels, particularly those found in the chaparral biome in the coastal regions of California.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20100026672&hterms=decision+making&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Ddecision%2Bmaking','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20100026672&hterms=decision+making&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Ddecision%2Bmaking"><span>The Utility of the OMI <span class="hlt">HCHO</span>/NO2 in Air Quality Decision-Making Activities</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Duncan, Bryan</p> <p>2010-01-01</p> <p>I will discuss a novel and practical application of the OMI HCHU and NO2 data products to the "weight of evidence" in the air quality decision-making process (e.g., State Implementation Plan (SIP)) for a city, region, or state to demonstrate that it is making progress toward attainment of the National Ambient Air Quality Standard (NAAQS) for ozone. Any trend, or lack thereof, in the observed OMI <span class="hlt">HCHO</span>/NO2 may support that an emission control strategy implemented to reduce ozone is or is not occurring for a metropolitan area. In addition, the observed OMI <span class="hlt">HCHO</span>/NO2 may be used to define new emission control strategies as the photochemical environments of urban areas evolve over time. I will demonstrate the utility of the OMI <span class="hlt">HCHO</span>/NO2 over the U.S. for air quality applications with support from simulations with both a regional model and a photochemical box model. These results support mission planning of an OMI-like instrument for the proposed GEO-CAPE satellite that has as one of its objectives to study air quality from space. However, I'm attending the meeting as the Aura Deputy Project Scientist, so I don't technically need to present anything to justify the travel.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..16.6666Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..16.6666Z"><span>Rain-induced emission pulses of NOx and <span class="hlt">HCHO</span> from soils in African regions after dry spells as viewed by satellite sensors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zörner, Jan; Penning de Vries, Marloes; Beirle, Steffen; Veres, Patrick; Williams, Jonathan; Wagner, Thomas</p> <p>2014-05-01</p> <p>Outside industrial areas, soil emissions of NOx (stemming from bacterial emissions of NO) represent a considerable fraction of total NOx emissions, and may even dominate in remote tropical and agricultural areas. NOx fluxes from soils are controlled by abiotic and microbiological processes which depend on ambient environmental conditions. Rain-induced spikes in NOx have been observed by in-situ measurements and also satellite observations. However, the estimation of soil emissions over broad geographic regions remains uncertain using bottom-up approaches. Independent, global satellite measurements can help constrain emissions used in chemical models. Laboratory experiments on soil fluxes suggest that significant <span class="hlt">HCHO</span> emissions from soil can occur. However, it has not been previously attempted to detect <span class="hlt">HCHO</span> emissions from wetted soils by using satellite observations. This study investigates the evolution of tropospheric NO2 (as a proxy for NOx) and <span class="hlt">HCHO</span> column densities before and after the first rain fall event following a prolonged dry period in semi-arid regions, deserts as well as tropical regions in Africa. Tropospheric NO2 and <span class="hlt">HCHO</span> columns retrieved from OMI aboard the AURA satellite, GOME-2 aboard METOP and SCIAMACHY aboard ENVISAT are used to study and inter-compare the observed responses of the trace gases with multiple space-based instruments. The observed responses are prone to be affected by other sources like lightning, fire, influx from polluted air masses, as well measurement errors in the satellite retrieval caused by manifold reasons such as an increased cloud contamination. Thus, much care is taken verify that the observed spikes reflect enhancements in soil emissions. Total column measurements of H2<span class="hlt">O</span> from GOME-2 give further insight into the atmospheric state and help to explain the increase in humidity before the first precipitation event. The analysis is not only conducted for averages of distinct geographic regions, i.e. the Sahel, but also</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ACP....16.2803S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ACP....16.2803S"><span>Estimates of free-tropospheric NO2 and <span class="hlt">HCHO</span> mixing ratios derived from high-altitude mountain MAX-DOAS observations at midlatitudes and in the tropics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schreier, Stefan F.; Richter, Andreas; Wittrock, Folkard; Burrows, John P.</p> <p>2016-03-01</p> <p>In this study, mixing ratios of NO2 (XNO2) and <span class="hlt">HCHO</span> (XHCHO) in the free troposphere are derived from two multi-axis differential optical absorption spectroscopy (MAX-DOAS) data sets collected at Zugspitze (2650 m a.s.l., Germany) and Pico Espejo (4765 m a.s.l., Venezuela). The estimation of NO2 and <span class="hlt">HCHO</span> mixing ratios is based on the modified geometrical approach, which assumes a single-scattering geometry and a scattering point altitude close to the instrument altitude. Firstly, the horizontal optical path length (hOPL) is obtained from <span class="hlt">O</span>4 differential slant column densities (DSCDs) in the horizontal (0°) and vertical (90°) viewing directions. Secondly, XNO2 and XHCHO are estimated from the NO2 and <span class="hlt">HCHO</span> DSCDs at the 0° and 90° viewing directions and averaged along the obtained hOPLs. As the MAX-DOAS instrument was performing measurements in the ultraviolet region, wavelength ranges of 346-372 and 338-357 nm are selected for the DOAS analysis to retrieve NO2 and <span class="hlt">HCHO</span> DSCDs, respectively. In order to compare the measured <span class="hlt">O</span>4 DSCDs and moreover to perform some sensitivity tests, the radiative transfer model SCIATRAN with adapted altitude settings for mountainous terrain is operated to simulate synthetic spectra, on which the DOAS analysis is also applied. The overall agreement between measured and synthetic <span class="hlt">O</span>4 DSCDs is better for the higher Pico Espejo station than for Zugspitze. Further sensitivity analysis shows that a change in surface albedo (from 0.05 to 0.7) can influence the <span class="hlt">O</span>4 DSCDs, with a larger absolute difference observed for the horizontal viewing direction. Consequently, the hOPL can vary by about 5 % throughout the season, for example when winter snow cover fully disappears in summer. Typical values of hOPLs during clear-sky conditions are 19 km (14 km) at Zugspitze and 34 km (26.5 km) at Pico Espejo when using the 346-372 (338-357 nm) fitting window. The estimated monthly values of XNO2 (XHCHO), averaged over these hOPLs during clear-sky conditions</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ACPD...1531781S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ACPD...1531781S"><span>Estimates of free-tropospheric NO2 and <span class="hlt">HCHO</span> mixing ratios derived from high-altitude mountain MAX-DOAS observations in the mid-latitudes and tropics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schreier, S. F.; Richter, A.; Wittrock, F.; Burrows, J. P.</p> <p>2015-11-01</p> <p>In this study, mixing ratios of NO2 (XNO2) and <span class="hlt">HCHO</span> (XHCHO) in the free troposphere are derived from two Multi-AXis Differential Optical Absorption Spectroscopy (MAX-DOAS) data sets collected at Zugspitze (2650 m a.s.l., Germany) and Pico Espejo (4765 m a.s.l., Venezuela). The estimation of NO2 and <span class="hlt">HCHO</span> mixing ratios is based on the modified geometrical approach, which assumes a single-scattering geometry and a scattering point altitude close to the instrument. Firstly, the horizontal optical path length (hOPL) is obtained from <span class="hlt">O</span>4 differential slant column densities (DSCDs) in the horizontal (0°) and vertical (90°) viewing directions. Secondly, XNO2 and XHCHO are estimated from the NO2 and <span class="hlt">HCHO</span> DSCDs at the 0 and 90° viewing directions and averaged along the obtained hOPLs. As the MAX-DOAS instrument was performing measurements in the ultraviolet region, wavelength ranges of 346-372 and 338-357 nm are selected for the DOAS analysis to retrieve NO2 and <span class="hlt">HCHO</span> DSCDs, respectively. In order to compare the measured <span class="hlt">O</span>4 DSCDs and moreover to perform some sensitivity tests, the radiative transfer model SCIATRAN with adapted altitude settings for mountainous terrain is operated to simulate synthetic spectra, on which the DOAS analysis is also applied. The overall agreement between measured and synthetic <span class="hlt">O</span>4 DSCDs is better for the higher Pico Espejo station than for Zugspitze. Further sensitivity analysis shows that a change in surface albedo (from 0.05 to 0.7) can influence the <span class="hlt">O</span>4 DSCDs, with a larger absolute difference observed for the horizontal viewing direction. Consequently, the hOPL can vary by about 5 % throughout the season, for example when winter snow cover fully disappears in summer. Typical values of hOPLs during clear sky conditions are 19 km (14 km) at Zugspitze and 34 km (26.5 km) at Pico Espejo when using the 346-372 nm (338-357 nm) fitting window. The estimated monthly values of XNO2 (XHCHO), averaged over these hOPLs during clear sky conditions, are in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29198026','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29198026"><span>The synthetic evaluation of Cu<span class="hlt">O</span>-MnOx-modified pinecone biochar for simultaneous removal formaldehyde and elemental mercury from simulated flue gas.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yi, Yaoyao; Li, Caiting; Zhao, Lingkui; Du, Xueyu; Gao, Lei; Chen, Jiaqiang; Zhai, Yunbo; Zeng, Guangming</p> <p>2018-02-01</p> <p>A series of low-cost Cu-Mn-mixed oxides supported on biochar (CuMn/HBC) synthesized by an impregnation method were applied to study the simultaneous removal of formaldehyde (<span class="hlt">HCHO</span>) and elemental mercury (Hg 0 ) at 100-300° C from simulated flue gas. The metal loading value, Cu/Mn molar ratio, flue gas components, reaction mechanism, and interrelationship between <span class="hlt">HCHO</span> removal and Hg 0 removal were also investigated. Results suggested that 12%CuMn/HBC showed the highest removal efficiency of <span class="hlt">HCHO</span> and Hg 0 at 175° C corresponding to 89%and 83%, respectively. The addition of NO and SO 2 exhibited inhibitive influence on <span class="hlt">HCHO</span> removal. For the removal of Hg 0 , NO showed slightly positive influence and SO 2 had an inhibitive effect. Meanwhile, <span class="hlt">O</span> 2 had positive impact on the removal of <span class="hlt">HCHO</span> and Hg 0 . The samples were characterized by SEM, XRD, BET, XPS, ICP-AES, FTIR, and H 2 -TPR. The sample characterization illustrated that CuMn/HBC possessed the high pore volume and specific surface area. The chemisorbed oxygen (<span class="hlt">O</span> β ) and the lattice oxygen (<span class="hlt">O</span> α ) which took part in the removal reaction largely existed in CuMn/HBC. What is more, Mn<span class="hlt">O</span> 2 and Cu<span class="hlt">O</span> (or Cu 2 <span class="hlt">O</span>) were highly dispersed on the CuMn/HBC surface. The strong synergistic effect between Cu-Mn mixed oxides was critical to the removal reaction of <span class="hlt">HCHO</span> and Hg 0 via the redox equilibrium of Mn 4+ + Cu + ↔ Mn <span class="hlt">3</span>+ + Cu 2+ .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850035061&hterms=physical+chemistry&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dphysical%2Bchemistry','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850035061&hterms=physical+chemistry&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dphysical%2Bchemistry"><span>Importance of formaldehyde in cloud chemistry</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Adewuyi, Y. G.; Cho, S.-Y.; Tsay, R.-P.; Carmichael, G. R.</p> <p>1984-01-01</p> <p>A physical-chemical model which is an extension of that of Hong and Carmichael (1983) is used to investigate the role of formaldehyde in cloud chemistry. This model takes into account the mass transfer of SO2, <span class="hlt">O</span><span class="hlt">3</span>, NH<span class="hlt">3</span>, HNO<span class="hlt">3</span>, H2<span class="hlt">O</span>2, CO2, HCl, <span class="hlt">HCHO</span>, <span class="hlt">O</span>2, OH and HO2 into cloud droplets and their subsequent chemical reactions. The model is used to assess the importance of S(IV)-<span class="hlt">HCHO</span> adduct formation, the reduction of H2<span class="hlt">O</span>2 by <span class="hlt">HCHO</span>, <span class="hlt">HCHO</span>-free radical interactions, and the formation of HCOOH in the presence of <span class="hlt">HCHO</span> in cloud droplets. Illustrative calculations indicate that the presence of <span class="hlt">HCHO</span> inhibits sulfate production rate in cloud droplets. The direct inhibition of sulfate production rate in cloudwater due to nucleophilic addition of HSO<span class="hlt">3</span>(-) to <span class="hlt">HCHO</span>(aq) to form hydroxymethanesulfonate is generally low for concentrations of <span class="hlt">HCHO</span> typical of ambient air. However, inhibition of sulfate production due to formaldehyde-free radical interactions in solution can be important. These formaldehyde-free radical reactions can also generate appreciable quantities of formic acid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SSSci..63...42Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SSSci..63...42Z"><span>Synergistic effect of Fe2<span class="hlt">O</span><span class="hlt">3</span>/Ho2<span class="hlt">O</span><span class="hlt">3</span> Co-modified 2D-titanate heterojunctions on enhanced photocatalytic degradation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Xiaona; Liu, Xinzhao; Lu, Dingze; Wu, Pei; Yan, Qiuyang; Liu, Min; Fang, Pengfei</p> <p>2017-01-01</p> <p>Ti<span class="hlt">O</span>2-based nanosheets (TNSs) co-modified by Fe2<span class="hlt">O</span><span class="hlt">3</span> and Ho2<span class="hlt">O</span><span class="hlt">3</span> were synthesized by one-pot hydrothermal method using Fe(NO<span class="hlt">3)3</span> and <span class="hlt">Ho(NO</span><span class="hlt">3)3</span> as precursors compositing with Ti<span class="hlt">O</span>2. The Fe2<span class="hlt">O</span><span class="hlt">3</span>/Ho2<span class="hlt">O</span><span class="hlt">3</span>-TNSs heterojunctions possessed a thickness of approximately <span class="hlt">3</span>-4 nm, large specific surface area of 210-310 cm2/g, with Fe2<span class="hlt">O</span><span class="hlt">3</span> and Ho2<span class="hlt">O</span><span class="hlt">3</span> nanoparticles highly dispersed over the surface of the nanosheets. The crystallization of the samples gradually increased with the amount of Fe2<span class="hlt">O</span><span class="hlt">3</span> nanoparticles, which was confirmed by the XRD, BET and Raman spectra, indicating that Ho2<span class="hlt">O</span><span class="hlt">3</span> and Fe2<span class="hlt">O</span><span class="hlt">3</span> influenced the crystallinity and structure evolution of the TNSs, besides, led to an improved the visible-light absorption. Surface photocurrent and fluorescence spectral studies revealed that the photo-generated charge carrier separation efficiency could be efficiently improved by an appropriate amount of modification. The Fe2<span class="hlt">O</span><span class="hlt">3</span>/Ho2<span class="hlt">O</span><span class="hlt">3</span>-TNSs exhibited synergistic effect on photocatalytic degradation of RhB as well as MO under visible light. The highest efficiency was obtained by 0.05%-Fe2<span class="hlt">O</span><span class="hlt">3</span>/Ho2<span class="hlt">O</span><span class="hlt">3</span>-TNSs (Fe:Ho:Ti = 0.05:1:100), which was 8.86 and 6.72 times than that of individual 1.0%-Ho2<span class="hlt">O</span><span class="hlt">3</span>-TNSs (Ho:Ti = 1:100) and 0.05%-Fe2<span class="hlt">O</span><span class="hlt">3</span>-TNSs (Fe:Ti = 0.05:100), respectively. The possible mechanism for enhanced visible-light-induced photocatalytic activity was proposed. Ho2<span class="hlt">O</span><span class="hlt">3</span> introduced in the photocatalysts may act as the hole capture while Fe2<span class="hlt">O</span><span class="hlt">3</span> may share the same Fermi levels with TNSs and serve as the electron capture center in the n-n-p system, which reduced the recombination rate of photo-induced electron-hole pairs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25277736','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25277736"><span>Effect of high-fat and high-carbohydrate diets on pulmonary <span class="hlt">O</span>2 uptake kinetics during the transition to moderate-intensity exercise.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Raper, J A; Love, L K; Paterson, D H; Peters, S J; Heigenhauser, G J F; Kowalchuk, J M</p> <p>2014-12-01</p> <p>Mitochondrial pyruvate dehydrogenase (PDH) regulates the delivery of carbohydrate-derived substrate to the mitochondrial tricarboxylic acid cycle and electron transport chain. PDH activity at rest and its activation during exercise is attenuated following high-fat (HFAT) compared with high-carbohydrate (<span class="hlt">HCHO</span>) diets. Given the reliance on carbohydrate-derived substrate early in transitions to exercise, this study examined the effects of HFAT and <span class="hlt">HCHO</span> on phase II pulmonary <span class="hlt">O</span>2 uptake (V̇<span class="hlt">o</span>2 p) kinetics during transitions into the moderate-intensity (MOD) exercise domain. Eight active adult men underwent dietary manipulations consisting of 6 days of HFAT (73% fat, 22% protein, 5% carbohydrate) followed immediately by 6 days of <span class="hlt">HCHO</span> (10% fat, 10% protein, 80% carbohydrate); each dietary phase was preceded by a glycogen depletion protocol. Participants performed three MOD transitions from a 20 W cycling baseline to work rate equivalent to 80% of estimated lactate threshold on days 5 and 6 of each diet. Steady-state V̇<span class="hlt">o</span>2 p was greater (P < 0.05), and respiratory exchange ratio and carbohydrate oxidation rates were lower (P < 0.05) during HFAT. The phase II V̇<span class="hlt">o</span>2 p time constant (τV̇<span class="hlt">o</span>2 p) [HFAT 40 ± 16, <span class="hlt">HCHO</span> 32 ± 19 s (mean ± SD)] and V̇<span class="hlt">o</span>2 p gain (HFAT 10.<span class="hlt">3</span> ± 0.8, <span class="hlt">HCHO</span> 9.4 ± 0.7 ml·min(-1·)W(-1)) were greater (P < 0.05) in HFAT. The overall adjustment (effective time constant) of muscle deoxygenation (Δ[HHb]) was not different between diets (HFAT 24 ± 4 s, <span class="hlt">HCHO</span> 23 ± 4 s), which coupled with a slower τV̇<span class="hlt">o</span>2 p, indicates a slowed microvascular blood flow response. These results suggest that the slower V̇<span class="hlt">o</span>2 p kinetics associated with HFAT are consistent with inhibition and slower activation of PDH, a lower rate of pyruvate production, and/or attenuated microvascular blood flow and <span class="hlt">O</span>2 delivery. Copyright © 2014 the American Physiological Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AtmEn.130..163R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AtmEn.130..163R"><span>Aqueous-phase story of isoprene - A mini-review and reaction with <span class="hlt">HONO</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rudziński, Krzysztof J.; Szmigielski, Rafał; Kuznietsova, Inna; Wach, Paulina; Staszek, Dorota</p> <p>2016-04-01</p> <p>Isoprene is a major biogenic hydrocarbon emitted to the atmosphere and a well-recognized player in atmospheric chemistry, formation of secondary organic aerosol and air quality. Most of the scientific work on isoprene has focused on the gas-phase and smog chamber processing while direct aqueous chemistry has escaped the major attention because physical solubility of isoprene in water is low. Therefore, this work recollects the results of genuine research carried on atmospherically relevant aqueous-phase transformations of isoprene. It clearly shows that isoprene dissolves in water and reacts in aqueous solutions with common atmospheric oxidants such as hydrogen peroxide, ozone, hydroxyl radicals, sulfate radicals and sulfite radicals. The reactions take place in the bulk of solutions or on the gas-liquid interfaces and often are acid-catalyzed and/or enhanced by light. The review is appended by an experimental study of the aqueous-phase reaction of isoprene with nitrous acid (<span class="hlt">HONO</span>). The decay of isoprene and formation of new products are demonstrated. The tentative chemical mechanism of the reaction is suggested, which starts with slow decomposition of <span class="hlt">HONO</span> to NO2 and NO. The aqueous chemistry of isoprene explains the formation of a few tropospheric components identified by scientists yet considered of unknown origin. The reaction of isoprene with sulfate radicals explains formation of the MW 182 organosulfate found in ambient aerosol and rainwater while the reaction of isoprene with <span class="hlt">HONO</span> explains formation of the MW 129 and MW 229 nitroorganic compounds identified in rainwater. Thus, aqueous transformations of isoprene should not be neglected without evidence but rather considered and evaluated in modeling of atmospheric chemical processes even if alternative and apparently dominant heterogeneous pathways of isoprene transformation, dry or wet, are demonstrated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27616336','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27616336"><span>Intramolecular Oxidative <span class="hlt">O</span>-Demethylation of an Oxoferryl Porphyrin Complexed with a Per-<span class="hlt">O</span>-methylated β-Cyclodextrin Dimer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kitagishi, Hiroaki; Kurosawa, Shun; Kano, Koji</p> <p>2016-11-22</p> <p>The intramolecular oxidation of ROCH <span class="hlt">3</span> to ROCH 2 OH, where the latter compound spontaneously decomposed to ROH and <span class="hlt">HCHO</span>, was observed during the reaction of the supramolecular complex (met-hemoCD<span class="hlt">3</span>) with cumene hydroperoxide in aqueous solution. Met-hemoCD<span class="hlt">3</span> is composed of meso-tetrakis(4-sulfonatophenyl)porphinatoiron(III) (Fe III TPPS) and a per-<span class="hlt">O</span>-methylated β-cyclodextrin dimer having an -OCH 2 PyCH 2 <span class="hlt">O</span>- linker (Py=pyridine-<span class="hlt">3</span>,5-diyl). The <span class="hlt">O</span>=Fe IV TPPS complex was formed by the reaction of met-hemoCD<span class="hlt">3</span> with cumene hydroperoxide, and isolated by gel-filtration chromatography. Although the isolated <span class="hlt">O</span>=Fe IV TPPS complex in the cyclodextrin cage was stable in aqueous solution at 25 °C, it was gradually converted to Fe II TPPS (t 1/2 =7.6 h). This conversion was accompanied by oxidative <span class="hlt">O</span>-demethylation of an OCH <span class="hlt">3</span> group in the cyclodextrin dimer. The results indicated that hydrogen abstraction by <span class="hlt">O</span>=Fe IV TPPS from ROCH <span class="hlt">3</span> yields HO-Fe III TPPS and ROCH 2 . . This was followed by radical coupling to afford Fe II TPPS and ROCH 2 OH. The hemiacetal (ROCH 2 OH) immediately decomposed to ROH and <span class="hlt">HCHO</span>. This study revealed the ability of oxoferryl porphyrin to induce two-electron oxidation. © 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.A31B0041G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.A31B0041G"><span>Formaldehyde Source Attribution in Houston during TexAQS II and TRAMP</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guven, B.; Olaguer, E. P.</p> <p>2010-12-01</p> <p>To determine the relative importance of primary vs secondary formaldehyde in Houston, source apportionment was performed on continuous online measurements of VOCs, formaldehyde (<span class="hlt">HCHO</span>), CO, SO2, and <span class="hlt">HONO</span> at one urban and two industrial sites. The results of source apportionment were used in conjunction with the meteorological, emission inventory, emission event, and back trajectory data catalogued in Air Research Information Infrastructure (ARII) to determine the dominant source regions and evaluate the accuracy of reported regular and upset emissions from industrial facilities. The contribution of industrial sources such as flares from petrochemical plants and refineries to total atmospheric formaldehyde concentrations at the urban site is estimated to be 17% compared to 23% for mobile sources, amounting to 40% for the total contribution of primary <span class="hlt">HCHO</span> sources. The relative contribution of industrial sources to <span class="hlt">HCHO</span> concentration at the urban site increased to about 66% on some mornings coinciding with the <span class="hlt">HCHO</span> peak concentrations. Secondary formation of <span class="hlt">HCHO</span> during the day and night resulted from the reactions of industrial olefins and other VOCs with OH or ozone was a significant contributor to <span class="hlt">HCHO</span> concentrations at the urban site. An analysis of emission event, back trajectory and ambient concentration data in ARII showed that a large percentage of emission events were associated with trajectories that passed through the two industrial sites when peaks in concentrations were detected at those sites. Some peak <span class="hlt">HCHO</span> concentrations can also be linked to emission events of other VOCs, while a significant portion remained unexplained by the reported events. It is likely, based on the results from the SHARP campaign and our analysis, that some episodic emission events containing <span class="hlt">HCHO</span> are unreported to the TCEQ. Overlaid CPF plots for nighttime (green) and daytime (red) <span class="hlt">HCHO</span> concentrations measured at three sites and the locations of the largest emitting point</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhyE...97...38Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhyE...97...38Z"><span>Facile hydrothermal synthesis of mesoporous In2<span class="hlt">O</span><span class="hlt">3</span> nanoparticles with superior formaldehyde-sensing properties</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Su; Song, Peng; Yang, Zhongxi; Wang, Qi</p> <p>2018-03-01</p> <p>Mesoporous In2<span class="hlt">O</span><span class="hlt">3</span> nanoparticles were successfully synthesized via a facile, template free, and low-cost hydrothermal method. Their morphology and structure were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), differential thermal and thermogravimetry analysis (DSC-TG), and N2 adsorption-desorption analyses. The results reveal that mesoporous In2<span class="hlt">O</span><span class="hlt">3</span> nanoparticles with a size range of 40-60 nm, possess plenty of pores, and average pore size is about 5 nm. Importantly, the mesoporous structure, large specific surface area, and small size endow the mesoporous In2<span class="hlt">O</span><span class="hlt">3</span> nanoparticles with highly sensing performance for formaldehyde detection. The response value to 10 ppm <span class="hlt">HCHO</span> is 20 at an operating temperature of 280 °C, and the response and recovery time are 4 and 8 s, respectively. It is expected that the mesoporous In2<span class="hlt">O</span><span class="hlt">3</span> nanoparticles with large specific surface area and excellent sensing properties will become a promising functional material in monitoring and detecting formaldehyde.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhDT.......198C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhDT.......198C"><span>MAX-DOAS measurements of aerosol, <span class="hlt">HCHO</span>, and NO2 over Los Angeles from an elevated mountaintop site</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cheung, Ross</p> <p></p> <p>MAX-DOAS measurements of aerosol, <span class="hlt">HCHO</span>, and NO2 over Los Angeles from an elevated mountaintop site. By. Ross Cheung. Doctor of Philosophy in Atmospheric and Oceanic Sciences. University of California, Los Angeles, 2016. Professor Jochen Stutz, Chair. Differential Optical Absorption Spectroscopy (DOAS) has become a popular technique for measuring atmospheric trace gases using UV/Vis narrow-band absorption features along a light path through the atmosphere. The UCLA Multi-Axis DOAS instrument (MAX-DOAS) is a ground-based spectrometer currently located at Mt. Wilson, California (1700 meters above sea level) that measures solar scattered light at various viewing elevation angles. Since May of 2010, it has been taking regular measurements of atmospheric pollutants in the boundary layer of the atmosphere in and above the Los Angeles Basin. This thesis presents the experimental setup and spectral retrievals, as well as results of our observations of measurements of NO2 and <span class="hlt">HCHO</span> from Mt. Wilson. Radiative transfer modeling efforts of the deployment at Mt. Wilson will be presented, as well as our efforts to model and account for the effects of clouds and aerosols on MAX-DOAS measurements. Because of the unique challenges presented by aerosols in the ultraviolet and visible light region in a polluted urban boundary layer, new techniques were developed to account for and quantify these effects. Observations of path-integrated NO2 and <span class="hlt">HCHO</span>, some of the primary precursors to ozone formation in the lower troposphere, as well as aerosol extinctions using the UCLA MAX-DOAS will be presented, and the advantages of a mountaintop measurement strategy will be discussed in light of the amount of vertical information that can be retrieved from this approach. The techniques developed to improve the optimal estimation of vertical aerosol extinction profiles and trace gas concentration profiles will be discussed. Finally, an application of these observations uses the ratio of <span class="hlt">HCHO</span>/NO2 to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1167609-conversion-propylene-glycol-rutile-tio2','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1167609-conversion-propylene-glycol-rutile-tio2"><span>Conversion of 1,<span class="hlt">3</span>-Propylene Glycol on Rutile Ti<span class="hlt">O</span>2(110)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Chen, Long; Li, Zhenjun; Smith, R. Scott</p> <p>2014-10-09</p> <p>The adsorption of 1,<span class="hlt">3</span>-propylene glycol (1,<span class="hlt">3</span>-PG) on partially reduced Ti<span class="hlt">O</span>2(110) and its conversion to products have been studied by a combination of molecular beam dosing and temperature programmed desorption (TPD). When the Ti surface sites are saturated by 1,<span class="hlt">3</span>-PG, ~80% of the molecules undergo further reactions to yield products that are liberated during the TPD ramp. In contrast to ethylene glycol (EG) and 1,2- propylene glycol (1,2-PG) that yield only alkenes and water at very low coverages (< 0.05 ML), two additional products, <span class="hlt">HCHO</span> and C2H4, along with propylene (CH<span class="hlt">3</span>CHCH2) and water are observed for 1,<span class="hlt">3</span>-PG. Identical TPD line shapesmore » and desorption yields for <span class="hlt">HCHO</span> and C2H4 suggest that these products result from C-C bond cleavage and are coupled. At higher 1,<span class="hlt">3</span>-PG coverages (> 0.1 ML), propanal (CH<span class="hlt">3</span>CH2CHO) and two additional products, 1-propanol (CH<span class="hlt">3</span>CH2CH2OH) and acrolein (CH2CHCHO), are observed. The desorption of 1-propanol is found to be coupled with the desorption of acrolein, suggesting that these products are formed by the disproportionation of two 1,<span class="hlt">3</span>-PG molecules. The coverage dependent TPD results further show that propylene formation dominates at low coverages (< 0.<span class="hlt">3</span> ML), while the decomposition and disproportionation channels increase rapidly at higher coverages and reach yields comparable to that of propylene at the 1,<span class="hlt">3</span>-PG saturation coverage of 0.5 ML. The observed surface chemistry clearly shows how the molecular structure of glycols influences their reaction pathways on oxide surfaces.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhD...50U5105T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhD...50U5105T"><span>Sensing mechanism of Sn<span class="hlt">O</span>2/Zn<span class="hlt">O</span> nanofibers for CH<span class="hlt">3</span>OH sensors: heterojunction effects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tang, Wei</p> <p>2017-11-01</p> <p>Sn<span class="hlt">O</span>2/Zn<span class="hlt">O</span> composite nanofibers were synthesized by a simple electrospinning method. The prepared Sn<span class="hlt">O</span>2/Zn<span class="hlt">O</span> gas sensors exhibited good linear and high response to methanol. The enhanced sensing behavior of Sn<span class="hlt">O</span>2/Zn<span class="hlt">O</span> might be associated with the homotypic heterojunction effects formed in n-Sn<span class="hlt">O</span>2/n-Zn<span class="hlt">O</span> nanograins boundaries. In addition, the possible sensing mechanisms of methanol on Sn<span class="hlt">O</span>2/Zn<span class="hlt">O</span> surface were investigated by density functional theory in order to make the methanol adsorption and desorption process clear. Zn doped Sn<span class="hlt">O</span>2 model was adopted to approximate the Sn<span class="hlt">O</span>2/Zn<span class="hlt">O</span> structure because of the calculation power limitations. Calculation results showed that when exposed to methanol, the methanol would react with bridge oxygen <span class="hlt">O</span>2c , planar <span class="hlt">O</span><span class="hlt">3</span>c and pre adsorbed oxygen vacancy on the lattice surface. The -CH<span class="hlt">3</span> and -OH of methanol molecule would both lose one H atom. The lost H atoms bonded with oxygen at the adsorption sites. The final products were <span class="hlt">HCHO</span> and H2<span class="hlt">O</span>. Electrons were transferred from methanol to the lattice surface to reduce the resistance of semiconductor gas sensitive materials, which is in agreement with the experimental phenomena. More adsorption models of other interfering gases, such as ethanol, formaldehyde and acetone will be built and calculated to explain the selectivity issue from the perspective of adsorption energy, transferred charge and density of states in the future work.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22420555','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22420555"><span>BioArena studies: unique function of endogenous formaldehyde and ozone in the antibiotic effect--a review.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tyihák, Erno; Móricz, Agnes M; Ott, Péter G</p> <p>2012-01-01</p> <p>The investigations demonstrated clearly a unique function and role of endogenous formaldehyde (<span class="hlt">HCHO</span>) and ozone (<span class="hlt">O</span><span class="hlt">3</span>) in the antibiotic effect of diverse molecules having different chemical structure. Elimination of <span class="hlt">HCHO</span> and/or <span class="hlt">O</span><span class="hlt">3</span> from the layer chromatographic spots resulted in a decrease in the antimicrobial activity. On the basis of detection and measure of endogenous <span class="hlt">HCHO</span> and <span class="hlt">O</span><span class="hlt">3</span> BioArena enables to both direct isolation and biological evaluation of new bioactive compounds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ACPD...13.7503E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ACPD...13.7503E"><span>Ozone photochemistry in an oil and natural gas extraction region during winter: simulations of a snow-free season in the Uintah Basin, Utah</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Edwards, P. M.; Young, C. J.; Aikin, K.; deGouw, J. A.; Dubé, W. P.; Geiger, F.; Gilman, J. B.; Helmig, D.; Holloway, J. S.; Kercher, J.; Lerner, B.; Martin, R.; McLaren, R.; Parrish, D. D.; Peischl, J.; Roberts, J. M.; Ryerson, T. B.; Thornton, J.; Warneke, C.; Williams, E. J.; Brown, S. S.</p> <p>2013-03-01</p> <p>The Uintah Basin in northeastern Utah, a region of intense oil and gas extraction, experienced ozone (<span class="hlt">O</span><span class="hlt">3</span>) concentrations above levels harmful to human health for multiple days during the winters of 2009-2010 and 2010-2011. These wintertime <span class="hlt">O</span><span class="hlt">3</span> pollution episodes occur during cold, stable periods when the ground is snowcovered, and have been linked to emissions from the oil and gas extraction process. The Uintah Basin Winter Ozone Study (UBWOS) was a field intensive in early 2012, whose goal was to address current uncertainties in the chemical and physical processes that drive wintertime <span class="hlt">O</span><span class="hlt">3</span> production in regions of oil and gas development. Although elevated <span class="hlt">O</span><span class="hlt">3</span> concentrations were not observed during the winter of 2011-2012, the comprehensive set of observations tests of our understanding of <span class="hlt">O</span><span class="hlt">3</span> photochemistry in this unusual emissions environment. A box model, constrained to the observations and using the explicit Master Chemical Mechanism (MCM) V<span class="hlt">3</span>.2 chemistry scheme, has been used to investigate the sensitivities of <span class="hlt">O</span><span class="hlt">3</span> production during UBWOS 2012. Simulations identify the <span class="hlt">O</span><span class="hlt">3</span> production photochemistry to be highly radical limited. Production of OH from <span class="hlt">O</span><span class="hlt">3</span> photolysis (through reaction of <span class="hlt">O</span>(1D) with water vapor) contributed only 170 pptv day-1, 8% of the total primary radical source on average. Other radical sources, including the photolysis of formaldehyde (<span class="hlt">HCHO</span>, 52%), nitrous acid (<span class="hlt">HONO</span>, 26%), and nitryl chloride (ClNO2, 13%) were larger. <span class="hlt">O</span><span class="hlt">3</span> production was also found to be highly sensitive to aromatic volatile organic compound (VOC) concentrations, due to radical amplification reactions in the oxidation scheme of these species. Radical production was shown to be small in comparison to the emissions of nitrogen oxides (NOx), such that NOx acted as the primary radical sink. Consequently, the system was highly VOC sensitive, despite the much larger mixing ratio of total non-methane hydrocarbons (230 ppbv (2080 ppbC), 6 week average) relative to NOx (5.6 ppbv average</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.5869Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.5869Y"><span>Nitrous acid in a street canyon environment: sources and the contribution to local oxidation capacity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yun, Hui; Wang, Zhe; Zha, Qiaozhi; Wang, Weihao; Xue, Likun; Zhang, Li; Li, Qinyi; Cui, Long; Lee, Shuncheng; Poon, Steven; Wang, Tao</p> <p>2017-04-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) is one of the dominant sources of hydroxyl radical (OH) and plays an important role in photochemical oxidation processes in the atmosphere. Even though <span class="hlt">HONO</span> has been extensively studied in urban areas, its importance and effects in street canyon microenvironment has not been thoroughly investigated. Street canyons which suffer serious air pollution problem are widely distributed in downtown areas with paralleled high buildings and narrow roads in the center. In this study, we measured <span class="hlt">HONO</span> at a roadside of a street canyon in urban Hong Kong and applied an observation-based box model based on Master Chemical Mechanism (MCM <span class="hlt">3.3</span>) to investigate the contribution of <span class="hlt">HONO</span> to local oxidation chemistry. Higher <span class="hlt">HONO</span> mixing ratios were observed in the daytime than in the nighttime. An average emission ratio (Δ<span class="hlt">HONO</span>/ΔNOx) of 1.0% (±0.5%) was derived at this roadside site and the direct <span class="hlt">HONO</span> emission from vehicles contributed to 38% of the measured <span class="hlt">HONO</span> in the street canyon. Heterogeneous NO2 conversion on humid ground or building surfaces and the uptake of NO2 on fresh soot surfaces were the other two important <span class="hlt">HONO</span> sources in this microenvironment. OBM simulations constrained with observed <span class="hlt">HONO</span> showed that the peak concentration of OH, HO2 and RO2 is 7.9, 5.0 and 7.5 times of the result in the case with only OH+NO as <span class="hlt">HONO</span> source. Photolysis of <span class="hlt">HONO</span> contributed to 86.5% of the total primary radical production rates and can lead to efficient NO2 and <span class="hlt">O</span><span class="hlt">3</span> production under the condition of weak regional <span class="hlt">O</span><span class="hlt">3</span> transport. Our study suggests that <span class="hlt">HONO</span> could significantly increase the atmospheric oxidation capacity in a street canyon which may impact the secondary formation of aerosols and OVOCs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A13F2147K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A13F2147K"><span>High surface <span class="hlt">O</span><span class="hlt">3</span> episodes in Seoul under different meteorological regimes during KORUS-AQ campaign.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, H.; Lee, M.; Jung, J.; Cho, S.; Shin, H.; Lee, G.; Park, M.; Hong, J.</p> <p>2017-12-01</p> <p>To examine chemical characteristics of ozone (<span class="hlt">O</span><span class="hlt">3</span>) formation in Seoul Metropolitan Area (SMA), H2<span class="hlt">O</span>2, PAN, and <span class="hlt">HONO</span> were measured in conjunction with <span class="hlt">O</span><span class="hlt">3</span> and its precursors. The experiment was conducted at Olympic Park in Seoul during May 12 June 15, 2016. For the entire experiment period, the high <span class="hlt">O</span><span class="hlt">3</span> episodes of hourly mean concentration over 100 ppbv occurred on May 20, 23, 25, 29, and 30 and June 10 and 14. These episodes were different in meteorological conditions, precursor strengths, and chemical characteristics. The local influence was dominant under stagnant condition on May 20, 23 and June 10. When stagnant conditions developed over the Korean peninsula, the PBL (Planetary Boundary Layer) height often changed rapidly, leading to abrupt change in <span class="hlt">O</span><span class="hlt">3</span> and NOx. Particularly the nighttime concentrations of reactive gases such as <span class="hlt">O</span><span class="hlt">3</span> and NOx were sensitive to the change in PBL height. It is thought to be driven by land-sea breeze circulation. During May 25 28 when air was coming from the Eastern China, <span class="hlt">O</span><span class="hlt">3</span> was enhanced with aerosols and high SO2 and CO but low NOx concentration. Odd-Oxygen (<span class="hlt">O</span><span class="hlt">3</span>+NO2, OX) ratio indicates the different chemical regimes, particularly at night(8PM - 7AM). <span class="hlt">O</span><span class="hlt">3</span>/OX ratio was close to zero when local influence was dominant due to <span class="hlt">O</span><span class="hlt">3</span>-titration by NOx. In contrast, this ratio was high over 0.6 in Chinese outflow plumes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ACP....17.9733Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ACP....17.9733Z"><span>Combined impacts of nitrous acid and nitryl chloride on lower-tropospheric ozone: new module development in WRF-Chem and application to China</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Li; Li, Qinyi; Wang, Tao; Ahmadov, Ravan; Zhang, Qiang; Li, Meng; Lv, Mengyao</p> <p>2017-08-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) and nitryl chloride (ClNO2) - through their photolysis - can have profound effects on the nitrogen cycle and oxidation capacity of the lower troposphere. Previous numerical studies have separately considered and investigated the sources/processes of these compounds and their roles in the fate of reactive nitrogen and the production of ozone (<span class="hlt">O</span><span class="hlt">3</span>), but their combined impact on the chemistry of the lower part of the troposphere has not been addressed yet. In this study, we updated the WRF-Chem model with the currently known sources and chemistry of <span class="hlt">HONO</span> and chlorine in a new chemical mechanism (CBMZ_ReNOM), and applied it to a study of the combined effects of <span class="hlt">HONO</span> and ClNO2 on summertime <span class="hlt">O</span><span class="hlt">3</span> in the boundary layer over China. We simulated the spatial distributions of <span class="hlt">HONO</span>, ClNO2, and related compounds at the surface and within the lower troposphere. The results showed that the modeled <span class="hlt">HONO</span> levels reached up to 800-1800 ppt at the surface (0-30 m) over the North China Plain (NCP), the Yangtze River Delta (YRD), and the Pearl River Delta (PRD) regions and that <span class="hlt">HONO</span> was concentrated within a 0-200 m layer. In comparison, the simulated surface ClNO2 mixing ratio was around 800-1500 ppt over the NCP, YRD, and central China regions and was predominantly present in a 0-600 m layer. <span class="hlt">HONO</span> enhanced daytime ROx (OH + HO2 + RO2) and <span class="hlt">O</span><span class="hlt">3</span> at the surface (0-30 m) by 2.8-4.6 ppt (28-37 %) and 2.9-6.2 ppb (6-13 %), respectively, over the three most developed regions, whereas ClNO2 increased surface <span class="hlt">O</span><span class="hlt">3</span> in the NCP and YRD regions by 2.4-<span class="hlt">3.3</span> ppb (or 5-6 %) and it also had a significant impact (<span class="hlt">3</span>-6 %) on above-surface <span class="hlt">O</span><span class="hlt">3</span> within 200-500 m. The combined effects increased surface <span class="hlt">O</span><span class="hlt">3</span> by 11.5, 13.5, and 13.<span class="hlt">3</span> % in the NCP, YRD, and PRD regions, respectively. Over the boundary layer (0-1000 m), the <span class="hlt">HONO</span> and ClNO2 enhanced <span class="hlt">O</span><span class="hlt">3</span> by up to 5.1 and <span class="hlt">3</span>.2 %, respectively, and their combined effect increased <span class="hlt">O</span><span class="hlt">3</span> by 7.1-8.9 % in the three regions. The new module noticeably improved <span class="hlt">O</span><span class="hlt">3</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1390423-multi-year-application-wrf-cam5-over-east-asia-part-comprehensive-evaluation-formation-regimes-pm','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1390423-multi-year-application-wrf-cam5-over-east-asia-part-comprehensive-evaluation-formation-regimes-pm"><span>Multi-year application of WRF-CAM5 over East Asia-Part I: Comprehensive evaluation and formation regimes of <span class="hlt">O</span> <span class="hlt">3</span> and PM 2.5</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>He, Jian; Zhang, Yang; Wang, Kai</p> <p></p> <p>Accurate simulations of air quality and climate require robust model parameterizations on regional and global scales. The Weather Research and Forecasting model with Chemistry version <span class="hlt">3</span>.4.1 has been coupled with physics packages from the Community Atmosphere Model version 5 (CAM5) (WRF-CAM5) to assess the robustness of the CAM5 physics package for regional modeling at higher grid resolutions than typical grid resolutions used in global modeling. In this two-part study, Part I describes the application and evaluation of WRF-CAM5 over East Asia at a horizontal resolution of 36-km for six years: 2001, 2005, 2006, 2008, 2010, and 2011. The simulations aremore » evaluated comprehensively with a variety of datasets from surface networks, satellites, and aircraft. The results show that meteorology is relatively well simulated by WRF-CAM5. However, cloud variables are largely or moderately underpredicted, indicating uncertainties in the model treatments of dynamics, thermodynamics, and microphysics of clouds/ices as well as aerosol-cloud interactions. For chemical predictions, the tropospheric column abundances of CO, NO2, and <span class="hlt">O</span><span class="hlt">3</span> are well simulated, but those of SO2 and <span class="hlt">HCHO</span> are moderately overpredicted, and the column <span class="hlt">HCHO</span>/NO2 indicator is underpredicted. Large biases exist in the surface concentrations of CO, NO2, and PM10 due to uncertainties in the emissions as well as vertical mixing. The underpredictions of NO lead to insufficient <span class="hlt">O</span><span class="hlt">3</span> titration, thus <span class="hlt">O</span><span class="hlt">3</span> overpredictions. The model can generally reproduce the observed <span class="hlt">O</span><span class="hlt">3</span> and PM indicators. These indicators suggest to control NOx emissions throughout the year, and VOCs emissions in summer in big cities and in winter over North China Plain, North/South Korea, and Japan to reduce surface <span class="hlt">O</span><span class="hlt">3</span>, and to control SO2, NH<span class="hlt">3</span>, and NOx throughout the year to reduce inorganic surface PM.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JQSRT.196...69M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JQSRT.196...69M"><span>Effective line intensity measurements of trans-nitrous acid (<span class="hlt">HONO</span>) of the ν1 band near 3600 cm-1 using laser difference-frequency spectrometer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maamary, Rabih; Fertein, Eric; Fourmentin, Marc; Dewaele, Dorothée; Cazier, Fabrice; Chen, Changshui; Chen, Weidong</p> <p>2017-07-01</p> <p>We report on the measurements of the effective line intensities of the ν1 fundamental band of trans-nitrous acid (trans-<span class="hlt">HONO</span>) in the infrared near 3600 cm-1 (2.78 μm). A home-made widely tunable laser spectrometer based on difference-frequency generation (DFG) was used for this study. The strengths of 28 well-resolved absorption lines of the ν1 band were determined by scaling their absorption intensities to the well referenced absorption line intensity of the ν<span class="hlt">3</span> band of trans-<span class="hlt">HONO</span> around 1250 cm-1 recorded simultaneously with the help of a DFB quantum cascade laser (QCL) spectrometer. The maximum measurement uncertainty of 12% in the line intensities is mainly determined by the uncertainty announced in the referenced line intensities, while the measurement precision in frequency positions of the absorption lines is better than 6×10-4 cm-1. The cross-measurement carried out in the present work allows one to perform intensity calibration using well referenced line parameters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/36990','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/36990"><span>Measurement of <span class="hlt">HONO</span>, HNCO, and other inorganic acids by negative-ion proton-transfer chemical-ionization mass spectrometry (NI-PT-CIMS): application to biomass burning emissions</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>J. M. Roberts; P. Veres; C. Warneke; J. A. Neuman; R. A. Washenfelder; S. S. Brown; M. Baasandorj; J. B. Burkholder; I. R. Burling; T. J. Johnson; R. J. Yokelson; J. de Gouw</p> <p>2010-01-01</p> <p>A negative-ion proton transfer chemical ionization mass spectrometric technique (NI-PT-CIMS), using acetate as the reagent ion, was applied to the measurement of volatile inorganic acids of atmospheric interest: hydrochloric (HCl), nitrous (<span class="hlt">HONO</span>), nitric 5 (HNO<span class="hlt">3</span>), and isocyanic (HNCO) acids. Gas phase calibrations through the sampling inlet showed the method to be...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29678991','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29678991"><span>Measurement properties of the Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) family of measures: protocol for a systematic review.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Harris, Meredith G; Sparti, Claudia; Scheurer, Roman; Coombs, Tim; Pirkis, Jane; Ruud, Torleif; Kisely, Steve; Hanssen-Bauer, Ketil; Siqveland, Johan; Burgess, Philip M</p> <p>2018-04-20</p> <p>The Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) for adults, and equivalent measures for children and adolescents and older people, are widely used in clinical practice and research contexts to measure mental health and functional outcomes. Additional <span class="hlt">HoNOS</span> measures have been developed for special populations and applications. Stakeholders require synthesised information about the measurement properties of these measures to assess whether they are fit for use with intended service settings and populations and to establish performance benchmarks. This planned systematic review will critically appraise evidence on the measurement properties of the <span class="hlt">HoNOS</span> family of measures. Journal articles meeting inclusion criteria will be identified via a search of seven electronic databases: MEDLINE via EBSCOhost, PsycINFO via APA PsycNET, Embase via Elsevier, Cumulative Index to Nursing and Allied Health Literature via EBSCOhost, Web of Science via Thomson Reuters, Google Scholar and the Cochrane Library. Variants of 'Health of the Nation Outcome Scales' or '<span class="hlt">HoNOS</span>' will be searched as text words. No restrictions will be placed on setting or language of publication. Reference lists of relevant studies and reviews will be scanned for additional eligible studies. Appraisal of reliability, validity, responsiveness and interpretability will be guided by the COnsensus-based Standards for the selection of health Measurement INstruments checklist. Feasibility/utility will be appraised using definitions and criteria derived from previous reviews. For reliability studies, we will also apply the Guidelines for Reporting Reliability and Agreement Studies to assess quality of reporting. Results will be synthesised narratively, separately for each measure, and by subgroup (eg, treatment setting, rater profession/experience or training) where possible. Meta-analyses will be undertaken where data are adequate. Ethics approval is not required as no primary data will be collected. Outcomes will be</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ACPD...1030129W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ACPD...1030129W"><span>Vertical profiles of nitrous acid in the nocturnal urban atmosphere of Houston, TX</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wong, K. W.; Oh, H.-J.; Lefer, B.; Rappenglück, B.; Stutz, J.</p> <p>2010-12-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) often plays an important role in tropospheric photochemistry as a major precursor of the hydroxyl radical (OH) in early morning hours and potentially during the day. However, the processes leading to formation of <span class="hlt">HONO</span> and its vertical distribution at night, which can have a considerable impact on daytime ozone formation, are currently poorly characterized by observations and models. Long-path differential optical absorption spectroscopy (LP-DOAS) measurements of <span class="hlt">HONO</span> during the 2006 TexAQS II Radical and Aerosol Measurement Project (TRAMP), near downtown Houston, TX, show nocturnal vertical profiles of <span class="hlt">HONO</span>, with mixing ratios of up to 2.2 ppb near the surface and below 100 ppt aloft. Three nighttime periods of <span class="hlt">HONO</span>, NO2 and <span class="hlt">O</span><span class="hlt">3</span> observations during TRAMP were used to perform model simulations of vertical mixing ratio profiles. By adjusting vertical mixing and NOx emissions the modeled NO2 and <span class="hlt">O</span><span class="hlt">3</span> mixing ratios showed very good agreement with the observations. Using a simple conversion of NO2 to <span class="hlt">HONO</span> on the ground, direct <span class="hlt">HONO</span> emissions, as well as <span class="hlt">HONO</span> loss at the ground and on aerosol, the observed <span class="hlt">HONO</span> profiles were reproduced well by the model. The unobserved increase of <span class="hlt">HONO</span> to NO2 ratio (<span class="hlt">HONO</span>/NO2) with altitude that was simulated by the initial model runs was found to be due to <span class="hlt">HONO</span> uptake being too small on aerosol and too large on the ground. Refined model runs, with adjusted <span class="hlt">HONO</span> uptake coefficients, showed much better agreement of <span class="hlt">HONO</span> and <span class="hlt">HONO</span>/NO2 for two typical nights, except during morning rush hour, when other <span class="hlt">HONO</span> formation pathways are most likely active. One of the nights analyzed showed increase of <span class="hlt">HONO</span> mixing ratios together with decreasing NO2 mixing ratios that the model was unable to reproduce, most likely due to the impact of weak precipitation during this night. <span class="hlt">HONO</span> formation and removal rates averaged over the lowest 300 m of the atmosphere showed that NO2 to <span class="hlt">HONO</span> conversion on the ground was the dominant source of <span class="hlt">HONO</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ACP....11.3595W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ACP....11.3595W"><span>Vertical profiles of nitrous acid in the nocturnal urban atmosphere of Houston, TX</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wong, K. W.; Oh, H.-J.; Lefer, B. L.; Rappenglück, B.; Stutz, J.</p> <p>2011-04-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) often plays an important role in tropospheric photochemistry as a major precursor of the hydroxyl radical (OH) in early morning hours and potentially during the day. However, the processes leading to formation of <span class="hlt">HONO</span> and its vertical distribution at night, which can have a considerable impact on daytime ozone formation, are currently poorly characterized by observations and models. Long-path differential optical absorption spectroscopy (LP-DOAS) measurements of <span class="hlt">HONO</span> during the 2006 TexAQS II Radical and Aerosol Measurement Project (TRAMP), near downtown Houston, TX, show nocturnal vertical profiles of <span class="hlt">HONO</span>, with mixing ratios of up to 2.2 ppb near the surface and below 100 ppt aloft. Three nighttime periods of <span class="hlt">HONO</span>, NO2 and <span class="hlt">O</span><span class="hlt">3</span> observations during TRAMP were used to perform model simulations of vertical mixing ratio profiles. By adjusting vertical mixing and NOx emissions the modeled NO2 and <span class="hlt">O</span><span class="hlt">3</span> mixing ratios showed very good agreement with the observations. Using a simple conversion of NO2 to <span class="hlt">HONO</span> on the ground, direct <span class="hlt">HONO</span> emissions, as well as <span class="hlt">HONO</span> loss at the ground and on aerosol, the observed <span class="hlt">HONO</span> profiles were reproduced by the model for 1-2 and 7-8 September in the nocturnal boundary layer (NBL). The unobserved increase of <span class="hlt">HONO</span> to NO2 ratio (<span class="hlt">HONO</span>/NO2) with altitude that was simulated by the initial model runs was found to be due to <span class="hlt">HONO</span> uptake being too small on aerosol and too large on the ground. Refined model runs, with adjusted <span class="hlt">HONO</span> uptake coefficients, showed much better agreement of <span class="hlt">HONO</span> and <span class="hlt">HONO</span>/NO2 for two typical nights, except during morning rush hour, when other <span class="hlt">HONO</span> formation pathways are most likely active. One of the nights analyzed showed an increase of <span class="hlt">HONO</span> mixing ratios together with decreasing NO2 mixing ratios that the model was unable to reproduce, most likely due to the impact of weak precipitation during this night. <span class="hlt">HONO</span> formation and removal rates averaged over the lowest 300 m of the atmosphere showed that NO2 to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12406722','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12406722"><span>Biodegradation of hexahydro-1,<span class="hlt">3</span>,5-trinitro-1,<span class="hlt">3</span>,5-triazine and its mononitroso derivative hexahydro-1-nitroso-<span class="hlt">3</span>,5-dinitro-1,<span class="hlt">3</span>,5-triazine by Klebsiella pneumoniae strain SCZ-1 isolated from an anaerobic sludge.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, Jian-Shen; Halasz, Annamaria; Paquet, Louise; Beaulieu, Chantale; Hawari, Jalal</p> <p>2002-11-01</p> <p>In previous work, we found that an anaerobic sludge efficiently degraded hexahydro-1,<span class="hlt">3</span>,5-trinitro-1,<span class="hlt">3</span>,5-triazine (RDX), but the role of isolates in the degradation process was unknown. Recently, we isolated a facultatively anaerobic bacterium, identified as Klebsiella pneumoniae strain SCZ-1, using MIDI and the 16S rRNA method from this sludge and employed it to degrade RDX. Strain SCZ-1 degraded RDX to formaldehyde (<span class="hlt">HCHO</span>), methanol (CH<span class="hlt">3</span>OH) (12% of total C), carbon dioxide (CO(2)) (72% of total C), and nitrous oxide (N2<span class="hlt">O</span>) (60% of total N) through intermediary formation of methylenedinitramine (<span class="hlt">O</span>(2)NNHCH(2)NHNO(2)). Likewise, hexahydro-1-nitroso-<span class="hlt">3</span>,5-dinitro-1,<span class="hlt">3</span>,5-triazine (MNX) was degraded to <span class="hlt">HCHO</span>, CH<span class="hlt">3</span>OH, and N2<span class="hlt">O</span> (16.5%) with a removal rate (0.39 micromol. h(-1). g [dry weight] of cells(-1)) similar to that of RDX (0.41 micromol. h(-1). g [dry weight] of cells(-1)) (biomass, 0.91 g [dry weight] of cells. liter(-1)). These findings suggested the possible involvement of a common initial reaction, possibly denitration, followed by ring cleavage and decomposition in water. The trace amounts of MNX detected during RDX degradation and the trace amounts of hexahydro-1,<span class="hlt">3</span>-dinitroso-5-nitro-1,<span class="hlt">3</span>,5-triazine detected during MNX degradation suggested that another minor degradation pathway was also present that reduced -NO2 groups to the corresponding -NO groups.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5880299','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5880299"><span>Observing atmospheric formaldehyde (<span class="hlt">HCHO</span>) from space: validation and intercomparison of six retrievals from four satellites (OMI, GOME2A, GOME2B, OMPS) with SEAC4RS aircraft observations over the Southeast US</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zhu, Lei; Jacob, Daniel J.; Kim, Patrick S.; Fisher, Jenny A.; Yu, Karen; Travis, Katherine R.; Mickley, Loretta J.; Yantosca, Robert M.; Sulprizio, Melissa P.; De Smedt, Isabelle; Abad, Gonzalo Gonzalez; Chance, Kelly; Li, Can; Ferrare, Richard; Fried, Alan; Hair, Johnathan W.; Hanisco, Thomas F.; Richter, Dirk; Scarino, Amy Jo; Walega, James; Weibring, Petter; Wolfe, Glenn M.</p> <p>2018-01-01</p> <p>Formaldehyde (<span class="hlt">HCHO</span>) column data from satellites are widely used as a proxy for emissions of volatile organic compounds (VOCs) but validation of the data has been extremely limited. Here we use highly accurate <span class="hlt">HCHO</span> aircraft observations from the NASA SEAC4RS campaign over the Southeast US in August–September 2013 to validate and intercompare six retrievals of <span class="hlt">HCHO</span> columns from four different satellite instruments (OMI, GOME2A, GOME2B and OMPS) and three different research groups. The GEOS-Chem chemical transport model is used as a common intercomparison platform. All retrievals feature a <span class="hlt">HCHO</span> maximum over Arkansas and Louisiana, consistent with the aircraft observations and reflecting high emissions of biogenic isoprene. The retrievals are also interconsistent in their spatial variability over the Southeast US (r=0.4–0.8 on a 0.5°×0.5° grid) and in their day-to-day variability (r=0.5–0.8). However, all retrievals are biased low in the mean by 20–51%, which would lead to corresponding bias in estimates of isoprene emissions from the satellite data. The smallest bias is for OMI-BIRA, which has high corrected slant columns relative to the other retrievals and low scattering weights in its air mass factor (AMF) calculation. OMI-BIRA has systematic error in its assumed vertical <span class="hlt">HCHO</span> shape profiles for the AMF calculation and correcting this would eliminate its bias relative to the SEAC4RS data. Our results support the use of satellite <span class="hlt">HCHO</span> data as a quantitative proxy for isoprene emission after correction of the low mean bias. There is no evident pattern in the bias, suggesting that a uniform correction factor may be applied to the data until better understanding is achieved. PMID:29619044</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28942315','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28942315"><span>Identification of chemical fingerprints in long-range transport of burning induced upper tropospheric ozone from Colorado to the North Atlantic Ocean.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jeon, Wonbae; Choi, Yunsoo; Souri, Amir Hossein; Roy, Anirban; Diao, Lijun; Pan, Shuai; Lee, Hwa Woon; Lee, Soon-Hwan</p> <p>2018-02-01</p> <p>This study investigates a significant biomass burning (BB) event occurred in Colorado of the United States in 2012 using the Community Multi-scale Air Quality (CMAQ) model. The simulation reasonably reproduced the significantly high upper tropospheric <span class="hlt">O</span> <span class="hlt">3</span> concentrations (up to 145ppb) caused by BB emissions. We find the BB-induced <span class="hlt">O</span> <span class="hlt">3</span> was primarily affected by chemical reactions and dispersion during its transport. In the early period of transport, high NO x and VOCs emissions caused <span class="hlt">O</span> <span class="hlt">3</span> production due to reactions with the peroxide and hydroxyl radicals, HO 2 and OH. Here, NO x played a key role in <span class="hlt">O</span> <span class="hlt">3</span> formation in the BB plume. The results indicated that HO 2 in the BB plume primarily came from formaldehyde (<span class="hlt">HCHO</span>+hv=2HO 2 +CO), a secondary alkoxy radical (ROR=HO 2 ). CO played an important role in the production of recycled HO 2 (OH+CO=HO 2 ) because of its abundance in the BB plume. The chemically produced HO 2 was largely converted to OH by the reactions with NO (HO 2 +NO=OH+NO 2 ) from BB emissions. This is in contrast to the surface, where HO 2 and OH are strongly affected by VOC and <span class="hlt">HONO</span>, respectively. In the late stages of transport, the <span class="hlt">O</span> <span class="hlt">3</span> concentration was primarily controlled by dispersion. It stayed longer in the upper troposphere compared to the surface due to sustained depletion of NO x . Sensitivity analysis results support that <span class="hlt">O</span> <span class="hlt">3</span> in the BB plume is significantly more sensitive to NO x than VOCs. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4486460','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4486460"><span>Investigation of Ground-Level Ozone and High-Pollution Episodes in a Megacity of Eastern China</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zhao, Heng; Wang, Shanshan; Wang, Wenxin; Liu, Rui; Zhou, Bin</p> <p>2015-01-01</p> <p>Differential Optical Absorption Spectroscopy (DOAS) was used for the long-term observation of ground-level ozone (<span class="hlt">O</span><span class="hlt">3</span>) from March 2010 to March 2013 over Shanghai, China. The 1-hour average concentration of <span class="hlt">O</span><span class="hlt">3</span> was 27.2 ± 17.0 ppbv. <span class="hlt">O</span><span class="hlt">3</span> level increased during spring, reached the peak in late spring and early summer, and then decreased in autumn and finally dropped to the bottom in winter. The highest monthly average <span class="hlt">O</span><span class="hlt">3</span> concentration in June (41.1 ppbv) was nearly three times as high as the lowest level recorded in December (15.2 ppbv). In terms of pollution episodes, 56 hourly samples (on 14 separate days) in 2010 exceeded the 1-hour ozone limit of 200 μg/m<span class="hlt">3</span> specified by the Grade II of the Chinese Ambient Air Quality Standards (CAAQS, revised GB 3095-2012). Utilizing the Hybrid Single Particle Lagrangian Integrated Trajectory (HYSPLIT) model, the primary contribution to high ozone days (HODs) was identified as the regional transportation of volatile organic compounds (VOC) and high concentrations of <span class="hlt">O</span><span class="hlt">3</span> from the chemical industrial zone in the Jinshan district of Shanghai. HODs showed higher concentrations of <span class="hlt">HONO</span> and NO2 than non-episode conditions, implying that <span class="hlt">HONO</span> at high concentration during HODs was capable of increasing the <span class="hlt">O</span><span class="hlt">3</span> concentration. The photolysis rate of <span class="hlt">HONO</span> was estimated, suggesting that the larger number of OH radicals resulting from high concentrations of <span class="hlt">HONO</span> have a considerable impact on ozone concentrations. PMID:26121146</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ACP....14.9727S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ACP....14.9727S"><span>Trace gas emissions from combustion of peat, crop residue, domestic biofuels, grasses, and other fuels: configuration and Fourier transform infrared (FTIR) component of the fourth Fire Lab at Missoula Experiment (FLAME-4)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stockwell, C. E.; Yokelson, R. J.; Kreidenweis, S. M.; Robinson, A. L.; DeMott, P. J.; Sullivan, R. C.; Reardon, J.; Ryan, K. C.; Griffith, D. W. T.; Stevens, L.</p> <p>2014-09-01</p> <p>During the fourth Fire Lab at Missoula Experiment (FLAME-4, October-November 2012) a large variety of regionally and globally significant biomass fuels was burned at the US Forest Service Fire Sciences Laboratory in Missoula, Montana. The particle emissions were characterized by an extensive suite of instrumentation that measured aerosol chemistry, size distribution, optical properties, and cloud-nucleating properties. The trace gas measurements included high-resolution mass spectrometry, one- and two-dimensional gas chromatography, and open-path Fourier transform infrared (OP-FTIR) spectroscopy. This paper summarizes the overall experimental design for FLAME-4 - including the fuel properties, the nature of the burn simulations, and the instrumentation employed - and then focuses on the OP-FTIR results. The OP-FTIR was used to measure the initial emissions of 20 trace gases: CO2, CO, CH4, C2H2, C2H4, C<span class="hlt">3</span>H6, <span class="hlt">HCHO</span>, HCOOH, CH<span class="hlt">3</span>OH, CH<span class="hlt">3</span>COOH, glycolaldehyde, furan, H2<span class="hlt">O</span>, NO, NO2, <span class="hlt">HONO</span>, NH<span class="hlt">3</span>, HCN, HCl, and SO2. These species include most of the major trace gases emitted by biomass burning, and for several of these compounds, this is the first time their emissions are reported for important fuel types. The main fire types included African grasses, Asian rice straw, cooking fires (open (three-stone), rocket, and gasifier stoves), Indonesian and extratropical peat, temperate and boreal coniferous canopy fuels, US crop residue, shredded tires, and trash. Comparisons of the OP-FTIR emission factors (EFs) and emission ratios (ERs) to field measurements of biomass burning verify that the large body of FLAME-4 results can be used to enhance the understanding of global biomass burning and its representation in atmospheric chemistry models. Crop residue fires are widespread globally and account for the most burned area in the US, but their emissions were previously poorly characterized. Extensive results are presented for burning rice and wheat straw: two major global crop residues</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RSPTA.37670152L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RSPTA.37670152L"><span>Quantum chemical study of the structure, spectroscopy and reactivity of NO+.(H2<span class="hlt">O</span>)n=1-5 clusters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Linton, Kirsty A.; Wright, Timothy G.; Besley, Nicholas A.</p> <p>2018-03-01</p> <p>Quantum chemical methods including Møller-Plesset perturbation (MP2) theory and density functional theory (DFT) have been used to study the structure, spectroscopy and reactivity of NO+.(H2<span class="hlt">O</span>)n=1-5 clusters. MP2/6-311++G** calculations are shown to describe the structure and spectroscopy of the clusters well. DFT calculations with exchange-correlation functionals with a low fraction of Hartree-Fock exchange give a binding energy of NO+.(H2<span class="hlt">O</span>) that is too high and incorrectly predict the lowest energy structure of NO+.(H2<span class="hlt">O</span>)2, and this error may be associated with a delocalization of charge onto the water molecule directly binding to NO+. Ab initio molecular dynamics (AIMD) simulations were performed to study the NO+.(H2<span class="hlt">O</span>)5 H+.(H2<span class="hlt">O</span>)4 + <span class="hlt">HONO</span> reaction to investigate the formation of <span class="hlt">HONO</span> from NO+.(H2<span class="hlt">O</span>)5. Whether an intracluster reaction to form <span class="hlt">HONO</span> is observed depends on the level of electronic structure theory used. Of note is that methods that accurately describe the relative energies of the product and reactant clusters did not show reactions on the timescales studied. This suggests that in the upper atmosphere the reaction may occur owing to the energy present in the NO+.(H2<span class="hlt">O</span>)5 complex following its formation. This article is part of the theme issue `Modern theoretical chemistry'.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.899f2002L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.899f2002L"><span>Comparison between premixed and partially premixed combustion in swirling jet from PIV, OH PLIF and <span class="hlt">HCHO</span> PLIF measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lobasov, A. S.; Chikishev, L. M.; Dulin, V. M.</p> <p>2017-09-01</p> <p>The present paper reports on the investigation of fuel-rich and fuel-lean turbulent combustion in a high-swirl jet. The jet flow was featured by a breakdown of the vortex core, presence of the central recirculation zone and intensive precession of the flow. The measurements were performed by the stereo PIV, OH PLIF and <span class="hlt">HCHO</span> PLIF techniques, simultaneously. Fluorescence of OH* in the flame and combustion products was excited via transition in the (1,0) vibrational band of the A2Σ+ - X2Π electronic system. The fluorescence was detected in the spectral range of 305-320 nm. In the case of <span class="hlt">HCHO</span> PLIF measurements the A-X {4}01 transition was excited. The jet Reynolds number was fixed as 5 000 (the bulk velocity was U 0 = 5 m/s). Three cases of the equivalence ratio ϕ of methane/air mixture issued from the nozzle were considered 0.7, 1.4 and 2.5. In all cases the flame front was subjected to deformations due to large-scale vortices, which rolled-up in the inner (around the central recirculation zone) and outer (between the annular jet core and surrounding air) mixing layers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170003580','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170003580"><span>Observing Atmospheric Formaldehyde (<span class="hlt">HCHO</span>) from Space: Validation and Intercomparison of Six Retrievals from Four Satellites (OMI, GOME2A, GOME2B, OMPS) with SEAC4RS Aircraft Observations over the Southeast US</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zhu, Lei; Jacob, Daniel J.; Kim, Patrick S.; Fisher, Jenny A.; Yu, Karen; Travis, Katherine R.; Mickley, Loretta J.; Yantosca, Robert M.; Sulprizio, Melissa P.; De Smedt, Isabelle; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170003580'); toggleEditAbsImage('author_20170003580_show'); toggleEditAbsImage('author_20170003580_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170003580_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170003580_hide"></p> <p>2016-01-01</p> <p>Formaldehyde (<span class="hlt">HCHO</span>) column data from satellites are widely used as a proxy for emissions of volatile organic compounds (VOCs), but validation of the data has been extremely limited. Here we use highly accurate <span class="hlt">HCHO</span> aircraft observations from the NASA SEAC4RS (Studies of Emissions, Atmospheric Composition, Clouds and Climate Coupling by Regional Surveys) campaign over the southeast US in August-September 2013 to validate and intercompare six retrievals of <span class="hlt">HCHO</span> columns from four different satellite instruments (OMI (Ozone Monitoring Instrument), GOME (Global Ozone Monitoring Experiment) 2A, GOME (Global Ozone Monitoring Experiment) 2B and OMPS (Ozone Mapping and Profiler Suite)) and three different research groups. The GEOS (Goddard Earth Observing System)-Chem chemical transport model is used as a common intercomparison platform. All retrievals feature a <span class="hlt">HCHO</span> maximum over Arkansas and Louisiana, consistent with the aircraft observations and reflecting high emissions of biogenic isoprene. The retrievals are also interconsistent in their spatial variability over the southeast US (r equals 0.4 to 0.8 on a 0.5 degree by 0.5 degree grid) and in their day-to-day variability (r equals 0.5 to 0.8). However, all retrievals are biased low in the mean by 20 to 51 percent, which would lead to corresponding bias in estimates of isoprene emissions from the satellite data. The smallest bias is for OMI-BIRA (Ozone Monitoring Instrument - Belgian Institute for Space Aeronomy), which has high corrected slant columns relative to the other retrievals and low scattering weights in its air mass factor (AMF) calculation. OMI-BIRA has systematic error in its assumed vertical <span class="hlt">HCHO</span> shape profiles for the AMF calculation, and correcting this would eliminate its bias relative to the SEAC (sup 4) RS data. Our results support the use of satellite <span class="hlt">HCHO</span> data as a quantitative proxy for isoprene emission after correction of the low mean bias. There is no evident pattern in the bias, suggesting that</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ACP....13.8955E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ACP....13.8955E"><span>Ozone photochemistry in an oil and natural gas extraction region during winter: simulations of a snow-free season in the Uintah Basin, Utah</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Edwards, P. M.; Young, C. J.; Aikin, K.; deGouw, J.; Dubé, W. P.; Geiger, F.; Gilman, J.; Helmig, D.; Holloway, J. S.; Kercher, J.; Lerner, B.; Martin, R.; McLaren, R.; Parrish, D. D.; Peischl, J.; Roberts, J. M.; Ryerson, T. B.; Thornton, J.; Warneke, C.; Williams, E. J.; Brown, S. S.</p> <p>2013-09-01</p> <p>The Uintah Basin in northeastern Utah, a region of intense oil and gas extraction, experienced ozone (<span class="hlt">O</span><span class="hlt">3</span>) concentrations above levels harmful to human health for multiple days during the winters of 2009-2010 and 2010-2011. These wintertime <span class="hlt">O</span><span class="hlt">3</span> pollution episodes occur during cold, stable periods when the ground is snow-covered, and have been linked to emissions from the oil and gas extraction process. The Uintah Basin Winter Ozone Study (UBWOS) was a field intensive in early 2012, whose goal was to address current uncertainties in the chemical and physical processes that drive wintertime <span class="hlt">O</span><span class="hlt">3</span> production in regions of oil and gas development. Although elevated <span class="hlt">O</span><span class="hlt">3</span> concentrations were not observed during the winter of 2011-2012, the comprehensive set of observations tests our understanding of <span class="hlt">O</span><span class="hlt">3</span> photochemistry in this unusual emissions environment. A box model, constrained to the observations and using the near-explicit Master Chemical Mechanism (MCM) v<span class="hlt">3</span>.2 chemistry scheme, has been used to investigate the sensitivities of <span class="hlt">O</span><span class="hlt">3</span> production during UBWOS 2012. Simulations identify the <span class="hlt">O</span><span class="hlt">3</span> production photochemistry to be highly radical limited (with a radical production rate significantly smaller than the NOx emission rate). Production of OH from <span class="hlt">O</span><span class="hlt">3</span> photolysis (through reaction of <span class="hlt">O</span>(1D) with water vapor) contributed only 170 pptv day-1, 8% of the total primary radical source on average (primary radicals being those produced from non-radical precursors). Other radical sources, including the photolysis of formaldehyde (<span class="hlt">HCHO</span>, 52%), nitrous acid (<span class="hlt">HONO</span>, 26%), and nitryl chloride (ClNO2, 13%) were larger. <span class="hlt">O</span><span class="hlt">3</span> production was also found to be highly sensitive to aromatic volatile organic compound (VOC) concentrations, due to radical amplification reactions in the oxidation scheme of these species. Radical production was shown to be small in comparison to the emissions of nitrogen oxides (NOx), such that NOx acted as the primary radical sink. Consequently, the system was highly VOC</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRD..12211934W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRD..12211934W"><span>Sources and Potential Photochemical Roles of Formaldehyde in an Urban Atmosphere in South China</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Chuan; Huang, Xiao-Feng; Han, Yu; Zhu, Bo; He, Ling-Yan</p> <p>2017-11-01</p> <p>Formaldehyde (<span class="hlt">HCHO</span>) is an important intermediate in tropospheric photochemistry. However, study of its evolution characteristics under heavy pollution conditions in China is limited, especially for high temporal resolutions, making it difficult to analyze its sources and environmental impacts. In this study, ambient levels of <span class="hlt">HCHO</span> were monitored using a proton-transfer reaction mass spectrometer at an urban site in the Pearl River Delta of China. Continuous monitoring campaigns were conducted in the spring, summer, fall, and winter in 2016. The highest averaged <span class="hlt">HCHO</span> concentrations were observed in autumn (5.1 ± <span class="hlt">3</span>.1 ppbv) and summer (5.0 ± 4.4 ppbv), followed by winter (4.2 ± 2.2 ppbv) and spring (<span class="hlt">3</span>.4 ± 1.6 ppbv). The daily maximum of <span class="hlt">HCHO</span> occurs in the early afternoon and shows good correlations with <span class="hlt">O</span><span class="hlt">3</span> and the secondary organic aerosol tracer during the day, revealing close relationships between ambient <span class="hlt">HCHO</span> and secondary formations in Shenzhen, especially in summer and autumn. The daytime <span class="hlt">HCHO</span> is estimated to be the major contributor to <span class="hlt">O</span><span class="hlt">3</span> formation and OH radical production, indicating that <span class="hlt">HCHO</span> plays a key role in the urban atmospheric photochemical reactions. Anthropogenic secondary formation was calculated to be the dominant source of <span class="hlt">HCHO</span> using a photochemical age-based parameterization method, with an average proportion of 39%. The contributions of biogenic sources in summer (41%) and autumn (39%) are much higher than those in spring (26%) and winter (28%), while the contributions of anthropogenic primary sources in spring (20%) and winter (18%) are twice those in summer (9%) and autumn (9%).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28671745','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28671745"><span>Specially Treated Aramid Fiber Stabilized Gel-Emulsions: Preparation of Porous Polymeric Monoliths and Highly Efficient Removing of Airborne <span class="hlt">HCHO</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Jianfei; Chen, Xiangli; Wang, Pei; Fu, Xuwei; Liu, Kaiqiang; Fang, Yu</p> <p>2017-08-01</p> <p>Porous polymeric monoliths with densities as low as ≈0.060 g cm -<span class="hlt">3</span> are prepared in a gel-emulsion template way, of which the stabilizer employed is a newly discovered acidified aramid fiber that is so efficient that 0.05% (w/v, accounts for continuous phase) is enough to gel the system. The porous monoliths as obtained can be dried at ambient conditions, avoiding energy-consuming processes. Importantly, the monoliths show selective adsorption to <span class="hlt">HCHO</span>, and the corresponding adsorption capacity (M6) is ≈2700 mg g -1 , the best result that is reported until now. More importantly, the monoliths can be reused after drying. © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A51B2056R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A51B2056R"><span>Mechanistic modeling of reactive soil nitrogen emissions across agricultural management practices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rasool, Q. Z.; Miller, D. J.; Bash, J. O.; Venterea, R. T.; Cooter, E. J.; Hastings, M. G.; Cohan, D. S.</p> <p>2017-12-01</p> <p>The global reactive nitrogen (N) budget has increased by a factor of 2-<span class="hlt">3</span> from pre-industrial levels. This increase is especially pronounced in highly N fertilized agricultural regions in summer. The reactive N emissions from soil to atmosphere can be in reduced (NH<span class="hlt">3</span>) or oxidized (NO, <span class="hlt">HONO</span>, N2<span class="hlt">O</span>) forms, depending on complex biogeochemical transformations of soil N reservoirs. Air quality models like CMAQ typically neglect soil emissions of <span class="hlt">HONO</span> and N2<span class="hlt">O</span>. Previously, soil NO emissions estimated by models like CMAQ remained parametric and inconsistent with soil NH<span class="hlt">3</span> emissions. Thus, there is a need to more mechanistically and consistently represent the soil N processes that lead to reactive N emissions to the atmosphere. Our updated approach estimates soil NO, <span class="hlt">HONO</span> and N2<span class="hlt">O</span> emissions by incorporating detailed agricultural fertilizer inputs from EPIC, and CMAQ-modeled N deposition, into the soil N pool. EPIC addresses the nitrification, denitrification and volatilization rates along with soil N pools for agricultural soils. Suitable updates to account for factors like nitrite (NO2-) accumulation not addressed in EPIC, will also be made. The NO and N2<span class="hlt">O</span> emissions from nitrification and denitrification are computed mechanistically using the N sub-model of DAYCENT. These mechanistic definitions use soil water content, temperature, NH4+ and NO<span class="hlt">3</span>- concentrations, gas diffusivity and labile C availability as dependent parameters at various soil layers. Soil <span class="hlt">HONO</span> emissions found to be most probable under high NO2- availability will be based on observed ratios of <span class="hlt">HONO</span> to NO emissions under different soil moistures, pH and soil types. The updated scheme will utilize field-specific soil properties and N inputs across differing manure management practices such as tillage. Comparison of the modeled soil NO emission rates from the new mechanistic and existing schemes against field measurements will be discussed. Our updated framework will help to predict the diurnal and daily variability</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JChPh.140o4306V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JChPh.140o4306V"><span>Thermal decomposition of 1,<span class="hlt">3,3</span>-trinitroazetidine (TNAZ): A density functional theory and ab initio study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Veals, Jeffrey D.; Thompson, Donald L.</p> <p>2014-04-01</p> <p>Density functional theory and ab initio methods are employed to investigate decomposition pathways of 1,<span class="hlt">3,3</span>-trinitroazetidine initiated by unimolecular loss of NO2 or <span class="hlt">HONO</span>. Geometry optimizations are performed using M06/cc-pVTZ and coupled-cluster (CC) theory with single, double, and perturbative triple excitations, CCSD(T), is used to calculate accurate single-point energies for those geometries. The CCSD(T)/cc-pVTZ energies for NO2 elimination by N-N and C-N bond fission are, including zero-point energy (ZPE) corrections, 43.21 kcal/mol and 50.46 kcal/mol, respectively. The decomposition initiated by trans-<span class="hlt">HONO</span> elimination can occur by a concerted H-atom and nitramine NO2 group elimination or by a concerted H-atom and nitroalkyl NO2 group elimination via barriers (at the CCSD(T)/cc-pVTZ level with ZPE corrections) of 47.00 kcal/mol and 48.27 kcal/mol, respectively. Thus, at the CCSD(T)/cc-pVTZ level, the ordering of these four decomposition steps from energetically most favored to least favored is: NO2 elimination by N-N bond fission, <span class="hlt">HONO</span> elimination involving the nitramine NO2 group, <span class="hlt">HONO</span> elimination involving a nitroalkyl NO2 group, and finally NO2 elimination by C-N bond fission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29431680','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29431680"><span>Quantum chemical study of the structure, spectroscopy and reactivity of NO+.(H2<span class="hlt">O</span>) n=1-5 clusters.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Linton, Kirsty A; Wright, Timothy G; Besley, Nicholas A</p> <p>2018-03-13</p> <p>Quantum chemical methods including Møller-Plesset perturbation (MP2) theory and density functional theory (DFT) have been used to study the structure, spectroscopy and reactivity of NO + (H 2 <span class="hlt">O</span>) n =1-5 clusters. MP2/6-311++G** calculations are shown to describe the structure and spectroscopy of the clusters well. DFT calculations with exchange-correlation functionals with a low fraction of Hartree-Fock exchange give a binding energy of NO + (H 2 <span class="hlt">O</span>) that is too high and incorrectly predict the lowest energy structure of NO + (H 2 <span class="hlt">O</span>) 2 , and this error may be associated with a delocalization of charge onto the water molecule directly binding to NO + Ab initio molecular dynamics (AIMD) simulations were performed to study the NO + (H 2 <span class="hlt">O</span>) 5 [Formula: see text] H + (H 2 <span class="hlt">O</span>) 4 + <span class="hlt">HONO</span> reaction to investigate the formation of <span class="hlt">HONO</span> from NO + (H 2 <span class="hlt">O</span>) 5 Whether an intracluster reaction to form <span class="hlt">HONO</span> is observed depends on the level of electronic structure theory used. Of note is that methods that accurately describe the relative energies of the product and reactant clusters did not show reactions on the timescales studied. This suggests that in the upper atmosphere the reaction may occur owing to the energy present in the NO + (H 2 <span class="hlt">O</span>) 5 complex following its formation.This article is part of the theme issue 'Modern theoretical chemistry'. © 2018 The Author(s).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1812154P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1812154P"><span><span class="hlt">HCHO</span> and NO2 MAXDOAS retrieval strategies harmonization: Recent results from the EU FP7 project QA4ECV</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pinardi, Gaia; Peters, Enno; Hendrick, François; Gielen, Clio; Van Roozendael, Michel; Richter, Andreas; Piters, Ankie; Wagner, Thomas; Wang, Yang; Drosoglou, Theano; Bais, Alkis; Wang, Shanshan; Saiz-Lopez, Alfonso</p> <p>2016-04-01</p> <p>During the last decade, it has been extensively demonstrated that MAXDOAS is a useful and reliable technique to retrieve integrated column amounts of tropospheric trace gases and aerosols, as well as information on their vertical distributions. Since it is based on optical remote-sensing in the UV-visible region like nadir backscatter space-borne sensors, MAXDOAS is also increasingly recognized as a reference technique for validating satellite nadir observations of air quality species like NO2 and <span class="hlt">HCHO</span>. However, building up an harmonized network of MAXDOAS spectrometers requires significant efforts in terms of common retrieval strategies and best-practices definitions. Within the EU FP7 project QA4ECV (Quality Assurance for Essential Climate Variables; see http://www.qa4ecv.eu/), harmonization activities have been initiated focusing on the two main steps of the MAXDOAS retrieval, i.e. the DOAS spectral fit providing the so-called differential slant column densities (DSCDs) and the conversion of the retrieved DSCDs to vertical profiles and/or vertical column densities (VCDs). Regarding the first step, the DOAS settings for <span class="hlt">HCHO</span> and NO2 are optimized through an intercomparison exercise of slant column retrievals involving 15 groups of the MAXDOAS community including the QA4ECV partners, and based on the radiance spectra acquired during the MAD-CAT campaign held in Mainz (Germany) in June-July 2013 (see http://joseba.mpch-mainz.mpg.de/mad_cat.htm). The harmonization of the second step is done through the application of an AMF (aim mass factor) look-up table (LUT) approach on the optimized NO2 and <span class="hlt">HCHO</span> DSCDs. The AMF LUTs depend on entry parameters like SZA, elevation and relative azimuth angles, wavelength, boundary layer height, AOD, and surface albedo. The advantages and drawbacks of the LUT approach are illustrated at several stations through comparison of the derived VCDs with those retrieved using the more sophisticated Optimal-Estimation-based profiling method</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.A31I0171M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.A31I0171M"><span>Airborne In-Situ Measurements of Formaldehyde over California: One Year of Results from the Compact Formaldehyde Fluorescence Experiment (COFFEE) Instrument</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marrero, J. E.; St Clair, J. M.; Yates, E. L.; Ryoo, J. M.; Gore, W.; Swanson, A. K.; Iraci, L. T.; Hanisco, T. F.</p> <p>2016-12-01</p> <p>Formaldehyde (<span class="hlt">HCHO</span>) is one of the most abundant oxygenated volatile organic compounds (VOCs) in the atmosphere, playing a role in multiple atmospheric processes, such as ozone (<span class="hlt">O</span><span class="hlt">3</span>) production in polluted environments. Due to its short lifetime of only a few hours in daytime, <span class="hlt">HCHO</span> also serves as tracer of recent photochemical activity. While photochemical oxidation of non-methane hydrocarbons is the dominant source, <span class="hlt">HCHO</span> can also be emitted directly from fuel combustion, vegetation, and biomass burning. The Compact Formaldehyde FluorescencE Experiment (COFFEE) instrument was built for integration onto the Alpha Jet Atmospheric eXperiment (AJAX) payload, based out of NASA's Ames Research Center (Moffett Field, CA). Using Non-Resonant Laser Induced Fluorescence (NR-LIF), trace concentrations of <span class="hlt">HCHO</span> can be detected with a sensitivity of 200 parts per trillion. Since its first research flight in December 2015, COFFEE has successfully flown on more than 20 science missions throughout California and Nevada. Presented here are results from these flights, including boundary layer measurements and vertical profiles throughout the tropospheric column. California's San Joaquin Valley is a primary focus, as this region is known for its elevated levels of <span class="hlt">HCHO</span> as well as <span class="hlt">O</span><span class="hlt">3</span>. Measurements collected in wildfire plumes, urban centers, agricultural lands, and on and off shore comparisons will be presented. In addition, the correlation of <span class="hlt">HCHO</span> to other trace gases also measured by AJAX, including <span class="hlt">O</span><span class="hlt">3</span>, methane, carbon dioxide, and water vapor will also be shown. Lastly, the implications of these <span class="hlt">HCHO</span> measurements on calibration and validation of remote sensing data collected by NASA's OMI (Aura) and OMPS (SuomiNPP) satellites will be addressed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1915858P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1915858P"><span>On the use of harmonized <span class="hlt">HCHO</span> and NO2 MAXDOAS measurements for the validation of GOME-2 and OMI satellite sensors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pinardi, Gaia; Hendrick, François; Gielen, Clio; Van Roozendael, Michel; De Smedt, Isabelle; Lambert, Jean-Christopher; Granville, José; Compernolle, Steven; Richter, Andreas; Peters, Enno; Piters, Ankie; Wagner, Thomas; Wang, Yang; Drosoglou, Theano; Bais, Alkis; Wang, Shanshan; Saiz-Lopez, Alfonso</p> <p>2017-04-01</p> <p>During the last decade, the MAXDOAS technique has been increasingly recognized as a source of Fiducial Reference Measurements (FRM) suitable for the validation of satellite nadir observations of species relevant for climate and air quality like NO2 and <span class="hlt">HCHO</span>. As part of the EU FP7 QA4ECV (Quality Assurance for Essential Climate Variables; see http://www.qa4ecv.eu/) project, efforts have been recently made to harmonize a network of a dozen of MAXDOAS spectrometers in view of their use to assess the quality of satellite climate data records generated within the same project. Harmonization tasks have addressed both retrieval steps involved in MAXDOAS retrievals, i.e. the DOAS spectral fit providing the differential slant column densities (DSCDs) and the conversion of the retrieved DSCDs into vertical profiles and/or vertical column densities (VCDs). In this work, we illustrate the successive harmonization steps and present the resulting QA4ECV MAXDOAS database v2. The approach adopted for the conversion of slant to vertical columns is based on a simplified look-up-table approach. The strength and limitation of this approach are discussed using reference data retrieved using an optimal estimation scheme. The QA4ECV MAXDOAS database is then used to validate satellite data sets of NO2 and <span class="hlt">HCHO</span> columns derived from the Aura/OMI and MetOp/GOME-2 sensors. The methodology of comparison, which is also a subject of the QA4ECV project, is reviewed with respect to co-location criteria, impact of vertical and horizontal smoothing and representativeness of validation sites. We conclude by assessing the current strengths and limitations of the existing MAXDOAS datasets for NO2 and <span class="hlt">HCHO</span> satellite validation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19730009428','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19730009428"><span>The reaction of H2<span class="hlt">O</span>2 with NO2 and NO</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gray, D.; Lissi, E.; Heicklen, J.</p> <p>1972-01-01</p> <p>The reactions of NO and NO2 with H2<span class="hlt">O</span>2 have been examined at 25 C. Reaction mixtures were monitored by continuously bleeding through a pinhole into a monopole mass spectrometer. NO2 was also monitored by its optical absorption in the visible part of the spectrum. Reaction mixtures containing initially 1.5 - 2.5 torr of NO2 and 0.8 - 1.4 torr of H2<span class="hlt">O</span>2 or 1 - 12 torr of NO and 0.5 - 1.5 torr of H2<span class="hlt">O</span>2 were studied. The H2<span class="hlt">O</span>2 - NO reaction was complex. There was an induction period followed by a marked acceleration in reactant removal. The final products of the reaction, NO2, probably H2<span class="hlt">O</span>, and possibly <span class="hlt">HONO</span>2 were produced mainly after all the H2<span class="hlt">O</span>2 was removed. The <span class="hlt">HONO</span> intermediate was shown to disproportionate to NO2 + NO + H2<span class="hlt">O</span> in a relatively slow first order reaction. The acceleration in H2<span class="hlt">O</span>2 removal after the NO - H2<span class="hlt">O</span>2 reaction is started is caused by NO2 catalysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApSS..426..333X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApSS..426..333X"><span>Effect of calcination temperature on formaldehyde oxidation performance of Pt/Ti<span class="hlt">O</span>2 nanofiber composite at room temperature</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, Feiyan; Le, Yao; Cheng, Bei; Jiang, Chuanjia</p> <p>2017-12-01</p> <p>Catalytic oxidation at room temperature over well-designed catalysts is an environmentally friendly method for the abatement of indoor formaldehyde (<span class="hlt">HCHO</span>) pollution. Herein, nanocomposites of platinum (Pt) and titanium dioxide (Ti<span class="hlt">O</span>2) nanofibers with various phase compositions were prepared by calcining the electrospun Ti<span class="hlt">O</span>2 precursors at different temperatures and subsequently depositing Pt nanoparticles (NPs) on the Ti<span class="hlt">O</span>2 through a NaBH4-reduction process. The phase compositions and structures of Pt/Ti<span class="hlt">O</span>2 can be easily controlled by varying the calcination temperature. The Pt/Ti<span class="hlt">O</span>2 nanocomposites showed a phase-dependent activity towards the catalytic <span class="hlt">HCHO</span> oxidation. Pt/Ti<span class="hlt">O</span>2 containing pure rutile phase showed enhanced activity with a turnover frequency (TOF) of 16.6 min-1 (for a calcination temperature of 800 °C) as compared to those containing the anatase phase or mixed phases. Density functional theory calculation shows that Ti<span class="hlt">O</span>2 nanofibers with pure rutile phase have stronger adsorption ability to Pt atoms than anatase phase, which favors the reduction of Pt over rutile phase Ti<span class="hlt">O</span>2, leading to higher contents of metallic Pt in the nanocomposite. In addition, the Pt/Ti<span class="hlt">O</span>2 with rutile phase possesses more abundant oxygen vacancies, which is conducive to the activation of adsorbed oxygen. Consequently, the Pt/rutile-Ti<span class="hlt">O</span>2 nanocomposite exhibited better catalytic activity towards <span class="hlt">HCHO</span> oxidation at room temperature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29112803','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29112803"><span>Gas Selectivity Control in Co<span class="hlt">3</span><span class="hlt">O</span>4 Sensor via Concurrent Tuning of Gas Reforming and Gas Filtering using Nanoscale Hetero-Overlayer of Catalytic Oxides.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jeong, Hyun-Mook; Jeong, Seong-Yong; Kim, Jae-Hyeok; Kim, Bo-Young; Kim, Jun-Sik; Abdel-Hady, Faissal; Wazzan, Abdulaziz A; Al-Turaif, Hamad Ali; Jang, Ho Won; Lee, Jong-Heun</p> <p>2017-11-29</p> <p>Co <span class="hlt">3</span> <span class="hlt">O</span> 4 sensors with a nanoscale Ti<span class="hlt">O</span> 2 or Sn<span class="hlt">O</span> 2 catalytic overlayer were prepared by screen-printing of Co <span class="hlt">3</span> <span class="hlt">O</span> 4 yolk-shell spheres and subsequent e-beam evaporation of Ti<span class="hlt">O</span> 2 and Sn<span class="hlt">O</span> 2 . The Co <span class="hlt">3</span> <span class="hlt">O</span> 4 sensors with 5 nm thick Ti<span class="hlt">O</span> 2 and Sn<span class="hlt">O</span> 2 overlayers showed high responses (resistance ratios) to 5 ppm xylene (14.5 and 28.8) and toluene (11.7 and 16.2) at 250 °C with negligible responses to interference gases such as ethanol, <span class="hlt">HCHO</span>, CO, and benzene. In contrast, the pure Co <span class="hlt">3</span> <span class="hlt">O</span> 4 sensor did not show remarkable selectivity toward any specific gas. The response and selectivity to methylbenzenes and ethanol could be systematically controlled by selecting the catalytic overlayer material, varying the overlayer thickness, and tuning the sensing temperature. The significant enhancement of the selectivity for xylene and toluene was attributed to the reforming of less reactive methylbenzenes into more reactive and smaller species and oxidative filtering of other interference gases, including ubiquitous ethanol. The concurrent control of the gas reforming and oxidative filtering processes using a nanoscale overlayer of catalytic oxides provides a new, general, and powerful tool for designing highly selective and sensitive oxide semiconductor gas sensors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ACP....18.1977T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ACP....18.1977T"><span>Nitrous acid formation in a snow-free wintertime polluted rural area</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsai, Catalina; Spolaor, Max; Fedele Colosimo, Santo; Pikelnaya, Olga; Cheung, Ross; Williams, Eric; Gilman, Jessica B.; Lerner, Brian M.; Zamora, Robert J.; Warneke, Carsten; Roberts, James M.; Ahmadov, Ravan; de Gouw, Joost; Bates, Timothy; Quinn, Patricia K.; Stutz, Jochen</p> <p>2018-02-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) photolysis is an important source of hydroxyl radicals (OH) in the lower atmosphere, in particular in winter when other OH sources are less efficient. The nighttime formation of <span class="hlt">HONO</span> and its photolysis in the early morning have long been recognized as an important contributor to the OH budget in polluted environments. Over the past few decades it has become clear that the formation of <span class="hlt">HONO</span> during the day is an even larger contributor to the OH budget and additionally provides a pathway to recycle NOx. Despite the recognition of this unidentified <span class="hlt">HONO</span> daytime source, the precise chemical mechanism remains elusive. A number of mechanisms have been proposed, including gas-phase, aerosol, and ground surface processes, to explain the elevated levels of daytime <span class="hlt">HONO</span>. To identify the likely <span class="hlt">HONO</span> formation mechanisms in a wintertime polluted rural environment we present LP-DOAS observations of <span class="hlt">HONO</span>, NO2, and <span class="hlt">O</span><span class="hlt">3</span> on three absorption paths that cover altitude intervals from 2 to 31, 45, and 68 m above ground level (a.g.l.) during the UBWOS 2012 experiment in the Uintah Basin, Utah, USA. Daytime <span class="hlt">HONO</span> mixing ratios in the 2-31 m height interval were, on average, 78 ppt, which is lower than <span class="hlt">HONO</span> levels measured in most polluted urban environments with similar NO2 mixing ratios of 1-2 ppb. <span class="hlt">HONO</span> surface fluxes at 19 m a.g.l., calculated using the <span class="hlt">HONO</span> gradients from the LP-DOAS and measured eddy diffusivity coefficient, show clear upward fluxes. The hourly average vertical <span class="hlt">HONO</span> flux during sunny days followed solar irradiance, with a maximum of (4.9 ± 0.2) × 1010 molec. cm-2 s-1 at noontime. A photostationary state analysis of the <span class="hlt">HONO</span> budget shows that the surface flux closes the <span class="hlt">HONO</span> budget, accounting for 63 ± 32 % of the unidentified <span class="hlt">HONO</span> daytime source throughout the day and 90 ± 30 % near noontime. This is also supported by 1-D chemistry and transport model calculations that include the measured surface flux, thus clearly identifying chemistry at the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1812436T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1812436T"><span>Biological soil crusts emit large amounts of NO and <span class="hlt">HONO</span> affecting the nitrogen cycle in drylands</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tamm, Alexandra; Wu, Dianming; Ruckteschler, Nina; Rodríguez-Caballero, Emilio; Steinkamp, Jörg; Meusel, Hannah; Elbert, Wolfgang; Behrendt, Thomas; Sörgel, Matthias; Cheng, Yafang; Crutzen, Paul J.; Su, Hang; Pöschl, Ulrich; Weber, Bettina</p> <p>2016-04-01</p> <p>Dryland systems currently cover ˜40% of the world's land surface and are still expanding as a consequence of human impact and global change. In contrast to that, information on their role in global biochemical processes is limited, probably induced by the presumption that their sparse vegetation cover plays a negligible role in global balances. However, spaces between the sparse shrubs are not bare, but soils are mostly covered by biological soil crusts (biocrusts). These biocrust communities belong to the oldest life forms, resulting from an assembly between soil particles and cyanobacteria, lichens, bryophytes, and algae plus heterotrophic organisms in varying proportions. Depending on the dominating organism group, cyanobacteria-, lichen-, and bryophyte-dominated biocrusts are distinguished. Besides their ability to restrict soil erosion they fix atmospheric carbon and nitrogen, and by doing this they serve as a nutrient source in strongly depleted dryland ecosystems. In this study we show that a fraction of the nitrogen fixed by biocrusts is metabolized and subsequently returned to the atmosphere in the form of nitric oxide (NO) and nitrous acid (<span class="hlt">HONO</span>). These gases affect the radical formation and oxidizing capacity within the troposphere, thus being of particular interest to atmospheric chemistry. Laboratory measurements using dynamic chamber systems showed that dark cyanobacteria-dominated crusts emitted the largest amounts of NO and <span class="hlt">HONO</span>, being ˜20 times higher than trace gas fluxes of nearby bare soil. We showed that these nitrogen emissions have a biogenic origin, as emissions of formerly strongly emitting samples almost completely ceased after sterilization. By combining laboratory, field, and satellite measurement data we made a best estimate of global annual emissions amounting to ˜1.1 Tg of NO-N and ˜0.6 Tg of <span class="hlt">HONO</span>-N from biocrusts. This sum of 1.7 Tg of reactive nitrogen emissions equals ˜20% of the soil release under natural vegetation according</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26621714','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26621714"><span>Biological soil crusts accelerate the nitrogen cycle through large NO and <span class="hlt">HONO</span> emissions in drylands.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Weber, Bettina; Wu, Dianming; Tamm, Alexandra; Ruckteschler, Nina; Rodríguez-Caballero, Emilio; Steinkamp, Jörg; Meusel, Hannah; Elbert, Wolfgang; Behrendt, Thomas; Sörgel, Matthias; Cheng, Yafang; Crutzen, Paul J; Su, Hang; Pöschl, Ulrich</p> <p>2015-12-15</p> <p>Reactive nitrogen species have a strong influence on atmospheric chemistry and climate, tightly coupling the Earth's nitrogen cycle with microbial activity in the biosphere. Their sources, however, are not well constrained, especially in dryland regions accounting for a major fraction of the global land surface. Here, we show that biological soil crusts (biocrusts) are emitters of nitric oxide (NO) and nitrous acid (<span class="hlt">HONO</span>). Largest fluxes are obtained by dark cyanobacteria-dominated biocrusts, being ∼20 times higher than those of neighboring uncrusted soils. Based on laboratory, field, and satellite measurement data, we obtain a best estimate of ∼1.7 Tg per year for the global emission of reactive nitrogen from biocrusts (1.1 Tg a(-1) of NO-N and 0.6 Tg a(-1) of <span class="hlt">HONO</span>-N), corresponding to ∼20% of global nitrogen oxide emissions from soils under natural vegetation. On continental scales, emissions are highest in Africa and South America and lowest in Europe. Our results suggest that dryland emissions of reactive nitrogen are largely driven by biocrusts rather than the underlying soil. They help to explain enigmatic discrepancies between measurement and modeling approaches of global reactive nitrogen emissions. As the emissions of biocrusts strongly depend on precipitation events, climate change affecting the distribution and frequency of precipitation may have a strong impact on terrestrial emissions of reactive nitrogen and related climate feedback effects. Because biocrusts also account for a large fraction of global terrestrial biological nitrogen fixation, their impacts should be further quantified and included in regional and global models of air chemistry, biogeochemistry, and climate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AtmEn..44.2632L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AtmEn..44.2632L"><span>Variations and sources of ambient formaldehyde for the 2008 Beijing Olympic games</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Yang; Shao, Min; Lu, Sihua; Chang, Chih-Chung; Dasgupta, Purnendu K.</p> <p>2010-07-01</p> <p>As the host city of the 2008 Olympic games, Beijing implemented a series of air pollution control measures before and during the Olympic games. Ambient formaldehyde (<span class="hlt">HCHO</span>) concentrations were measured using a fluorometric instrument based on a diffusion scrubber and the Hantzsch reaction; hydrocarbons were simultaneously measured using gas chromatography-mass spectrometry (GC-MS). Meteorological parameters, CO, <span class="hlt">O</span> <span class="hlt">3</span>, and NO 2 concentrations were measured by standard commercial instrumentation. In four separate periods: (a) before the vehicle plate number control (<span class="hlt">3</span>-19 July); (b) during the Olympic Games (8-24 August); (c) during the Paralympic Games (6-17 September) and (d) after the vehicle control was ceased (21-28 September), the average <span class="hlt">HCHO</span> mixing ratios were 7.31 ± 2.67 ppbv, 5.54 ± 2.41 ppbv, 8.72 ± 2.48 ppbv, and 6.42 ± 2.79 ppbv, while the total non-methane hydrocarbons (NMHCs) measured were 30.41 ± 18.08 ppbv, 18.12 ± 9.38 ppbv, 30.50 ± 13.37 ppbv, and 33.33 ± 15.85 ppbv, respectively. Both <span class="hlt">HCHO</span> and NMHC levels were the lowest during the Olympic games, and increased again during the Paralympic games even with the same vehicle control measures operative. Similar diurnal <span class="hlt">HCHO</span> and <span class="hlt">O</span> <span class="hlt">3</span> patterns indicated that photo-oxidation of NMHCs may be the major source of <span class="hlt">HCHO</span>. The diurnal profile of total NMHCs was very similar to that of NO 2 and CO: morning and evening peaks appeared in rush hours, indicating even after strict vehicle control, automobile emission may still be the dominant source of the <span class="hlt">HCHO</span> precursors. The contributions of <span class="hlt">HCHO</span>, alkanes, alkenes, and aromatics to OH loss rates were also calculated. <span class="hlt">HCHO</span> contributed 22 ± <span class="hlt">3</span>% to the total VOCs and 24 ± 1% to the total OH loss rate. <span class="hlt">HCHO</span> was not only important in term of abundance, but also important in chemical reactivity in the air.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5914766','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5914766"><span>Measurement properties of the Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) family of measures: protocol for a systematic review</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Harris, Meredith G; Sparti, Claudia; Scheurer, Roman; Coombs, Tim; Pirkis, Jane; Ruud, Torleif; Kisely, Steve; Hanssen-Bauer, Ketil; Siqveland, Johan; Burgess, Philip M</p> <p>2018-01-01</p> <p>Introduction The Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) for adults, and equivalent measures for children and adolescents and older people, are widely used in clinical practice and research contexts to measure mental health and functional outcomes. Additional <span class="hlt">HoNOS</span> measures have been developed for special populations and applications. Stakeholders require synthesised information about the measurement properties of these measures to assess whether they are fit for use with intended service settings and populations and to establish performance benchmarks. This planned systematic review will critically appraise evidence on the measurement properties of the <span class="hlt">HoNOS</span> family of measures. Methods and analysis Journal articles meeting inclusion criteria will be identified via a search of seven electronic databases: MEDLINE via EBSCOhost, PsycINFO via APA PsycNET, Embase via Elsevier, Cumulative Index to Nursing and Allied Health Literature via EBSCOhost, Web of Science via Thomson Reuters, Google Scholar and the Cochrane Library. Variants of ‘Health of the Nation Outcome Scales’ or ‘HoNOS’ will be searched as text words. No restrictions will be placed on setting or language of publication. Reference lists of relevant studies and reviews will be scanned for additional eligible studies. Appraisal of reliability, validity, responsiveness and interpretability will be guided by the COnsensus-based Standards for the selection of health Measurement INstruments checklist. Feasibility/utility will be appraised using definitions and criteria derived from previous reviews. For reliability studies, we will also apply the Guidelines for Reporting Reliability and Agreement Studies to assess quality of reporting. Results will be synthesised narratively, separately for each measure, and by subgroup (eg, treatment setting, rater profession/experience or training) where possible. Meta-analyses will be undertaken where data are adequate. Ethics and dissemination Ethics approval is</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ApSS..274..110L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ApSS..274..110L"><span>Bio-template-assisted synthesis of hierarchically hollow Si<span class="hlt">O</span>2 microtubes and their enhanced formaldehyde adsorption performance</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Le, Yao; Guo, Daipeng; Cheng, Bei; Yu, Jiaguo</p> <p>2013-06-01</p> <p>The indoor air quality is crucial for human health, taking into account that people often spend more than 80% of their time in houses, offices and cars. Formaldehyde (<span class="hlt">HCHO</span>) is a major pollutant and long-term exposure to <span class="hlt">HCHO</span> may cause health problems such as nasal tumors and skin irritation. In this work, for the first time, hierarchically hollow silica microtubes (HHSM) were synthesized by a simple sol-gel and calcination method using cetyltrimethyl ammonium bromide (CTAB) and bio-template poplar catkin (PC) as co-templates and the PC/Si<span class="hlt">O</span>2 weight ratio R was varied from 0, 0.1, 0.<span class="hlt">3</span> and 1. The prepared samples were further modified with tetraethylenepentamine (TEPA) and characterized by scanning electron microscope (SEM), transmission electron microscopy (TEM), Fourier transform infrared (FTIR), X-ray photoelectron spectroscopy (XPS), differential thermal analysis (DTA), thermal gravimetric analysis (TGA), and N2 physisorption techniques. This was followed by formaldehyde adsorption tests at ambient temperature. The results showed that all the prepared HHSM samples contained small mesopores with peak pore size at ca. 2.5 nm and large several tens of nanometer-sized pores on the tube wall. The R exhibited an obvious influence on specific surface areas and the sample prepared at R = 0.<span class="hlt">3</span> exhibited highest specific surface area (896 m2/g). All the TEPA-modified samples exhibited enhanced formaldehyde adsorption performance. The maximum <span class="hlt">HCHO</span> adsorption capacity (20.65 mg/g adsorbent) was achieved on the sample prepared at R = 0.<span class="hlt">3</span> and modified by 50 wt.% TEPA. The present study will provide new insight for the utilization of bio-template used for the fabrication of inorganic hollow tubes with high <span class="hlt">HCHO</span> adsorption performance for indoor air purification.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4687600','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4687600"><span>Biological soil crusts accelerate the nitrogen cycle through large NO and <span class="hlt">HONO</span> emissions in drylands</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wu, Dianming; Tamm, Alexandra; Ruckteschler, Nina; Rodríguez-Caballero, Emilio; Meusel, Hannah; Elbert, Wolfgang; Behrendt, Thomas; Sörgel, Matthias; Cheng, Yafang; Crutzen, Paul J.; Su, Hang; Pöschl, Ulrich</p> <p>2015-01-01</p> <p>Reactive nitrogen species have a strong influence on atmospheric chemistry and climate, tightly coupling the Earth’s nitrogen cycle with microbial activity in the biosphere. Their sources, however, are not well constrained, especially in dryland regions accounting for a major fraction of the global land surface. Here, we show that biological soil crusts (biocrusts) are emitters of nitric oxide (NO) and nitrous acid (<span class="hlt">HONO</span>). Largest fluxes are obtained by dark cyanobacteria-dominated biocrusts, being ∼20 times higher than those of neighboring uncrusted soils. Based on laboratory, field, and satellite measurement data, we obtain a best estimate of ∼1.7 Tg per year for the global emission of reactive nitrogen from biocrusts (1.1 Tg a−1 of NO-N and 0.6 Tg a−1 of <span class="hlt">HONO</span>-N), corresponding to ∼20% of global nitrogen oxide emissions from soils under natural vegetation. On continental scales, emissions are highest in Africa and South America and lowest in Europe. Our results suggest that dryland emissions of reactive nitrogen are largely driven by biocrusts rather than the underlying soil. They help to explain enigmatic discrepancies between measurement and modeling approaches of global reactive nitrogen emissions. As the emissions of biocrusts strongly depend on precipitation events, climate change affecting the distribution and frequency of precipitation may have a strong impact on terrestrial emissions of reactive nitrogen and related climate feedback effects. Because biocrusts also account for a large fraction of global terrestrial biological nitrogen fixation, their impacts should be further quantified and included in regional and global models of air chemistry, biogeochemistry, and climate. PMID:26621714</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A13F2133C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A13F2133C"><span>PCA-based SO2, NO2, and <span class="hlt">HCHO</span> retrievals from GeoTASO airborne measurements during KORUS-AQ 2016 campaign</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chong, H.; Lee, S.; Jeong, U.; Kim, J.; Li, C.; Krotkov, N. A.; Al-Saadi, J. A.; Janz, S. J.; Kowalewski, M. G.; Nowlan, C. R.; Kang, M.; Joiner, J.; Haffner, D. P.; Koo, J. H.; Hong, H.; Lee, H.</p> <p>2017-12-01</p> <p>The Geostationary Trace gas and Aerosol Sensor Optimization (GeoTASO) is an airborne instrument measuring backscattered radiance with a spectrometer covering the spectral range between 290-695 nm. GeoTASO flew on the B-200 (UC-12B) - LARC aircraft during the KORUS-AQ campaign, of which the spatial resolution is about 250 nm x 250 m. Principal component analysis (PCA) technique is used to retrieve slant column densities (SCD) of sulfur dioxide (SO2), nitrogen dioxide (NO2), and formaldehyde (<span class="hlt">HCHO</span>). The fitting windows of SO2, NO2, and <span class="hlt">HCHO</span> are 310-325 nm, 350-380 nm, and 335-357 nm respectively. The clear PCs of each species are collected from rural areas where are found to have less SCDs of each species from prior iteration step. Using the clear sector PCs and the cross section of each species, SCDs of each trace gas are obtained using the multiple linear regression method. Air mass factors (AMF) of each species are obtained using the atmospheric profiles from chemical transport model calculations during the campaign to convert SCDs to vertical column densities (VCD). The retrieved VCDs of each species well capture small point sources on the flight paths and their plumes propagating downwind areas, which was not available from the ground-based in-situ measurements. The retrieved VCDs will be compared and/or validated against other benchmark measurements during the campaign.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AtmEn.122..521M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AtmEn.122..521M"><span>Assessment of the impact of oxidation processes on indoor air pollution using the new time-resolved INCA-Indoor model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mendez, Maxence; Blond, Nadège; Blondeau, Patrice; Schoemaecker, Coralie; Hauglustaine, Didier A.</p> <p>2015-12-01</p> <p>INCA-Indoor, a new indoor air quality (IAQ) model, has been developed to simulate the concentrations of volatile organic compounds (VOC) and oxidants considering indoor air specific processes such as: emission, ventilation, surface interactions (sorption, deposition, uptake). Based on the detailed version of SAPRC-07 chemical mechanism, INCA-Indoor is able to analyze the contribution of the production and loss pathways of key chemical species (VOCs, oxidants, radical species). The potential of this model has been tested through three complementary analyses: a comparison with the most detailed IAQ model found in the literature, focusing on oxidant species; realistic scenarios covering a large range of conditions, involving variable OH sources like <span class="hlt">HONO</span>; and the investigation of alkenes ozonolysis under a large range of indoor conditions that can increase OH and HO2 concentrations. Simulations have been run changing nitrous acid (<span class="hlt">HONO</span>) concentrations, NOx levels, photolysis rates and ventilation rates, showing that <span class="hlt">HONO</span> can be the main source of indoor OH. Cleaning events using products containing D-limonene have been simulated at different periods of the day. These scenarios show that HOX concentrations can significantly increase in specific conditions. An assessment of the impact of indoor chemistry on the potential formation of secondary species such as formaldehyde (<span class="hlt">HCHO</span>) and acetaldehyde (CH<span class="hlt">3</span>CHO) has been carried out under various room configuration scenarios and a study of the HOx budget for different realistic scenarios has been performed. It has been shown that, under the simulation conditions, formaldehyde can be affected by oxidant concentrations via chemical production which can account for more than 10% of the total production, representing 6.5 ppb/h. On the other hand, acetaldehyde production is affected more by oxidation processes. When the photolysis rates are high, chemical processes are responsible for about 50% of the total production of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170009898&hterms=ross&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D10%26Ntt%3DWill%2Bross','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170009898&hterms=ross&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D10%26Ntt%3DWill%2Bross"><span>Impact of Evolving Isoprene Mechanisms on Simulated Formaldehyde: An Inter-comparison Supported by in Situ Observations from SENEX</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Marvin, Margaret R.; Wolfe, Glenn M.; Salawitch, Ross J.; Canty, Timothy P.; Roberts, Sandra J.; Travis, Katherine R.; Aiken, Kenneth C.; de Gouw, Joost A.; Graus, Martin; Hanisco, Thomas F.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170009898'); toggleEditAbsImage('author_20170009898_show'); toggleEditAbsImage('author_20170009898_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170009898_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170009898_hide"></p> <p>2017-01-01</p> <p>Isoprene oxidation schemes vary greatly among gas-phase chemical mechanisms, with potentially significant ramifications for air quality modeling and interpretation of satellite observations in biogenic-rich regions. In this study, in situ observations from the 2013 SENEX mission are combined with a constrained <span class="hlt">O</span>-D photochemical box model to evaluate isoprene chemistry among five commonly used gas-phase chemical mechanisms: CBO5, CB6r2, MCMv<span class="hlt">3</span>.2, MCMv<span class="hlt">3.3</span>.1, and a recent version of GEOS-Chem. Mechanisms are evaluated and inter-compared with respect to formaldehyde (<span class="hlt">HCHO</span>), a high-yield product of isoprene oxidation. Though underestimated by all considered mechanisms, observed <span class="hlt">HCHO</span> mixing ratios are best reproduced by MCMv<span class="hlt">3.3</span>.1 (normalized mean bias = -15%), followed by GEOS-Chem (-17%), MCMv<span class="hlt">3</span>.2 (-25%), CB6r2 (-32%) and CB05 (-33%). Inter-comparison of <span class="hlt">HCHO</span> production rates reveals that major restructuring of the isoprene oxidation scheme in the Carbon Bond mechanism increases <span class="hlt">HCHO</span> production by only approx. 5% in CB6r2 relative to CBO5, while further refinement of the complex isoprene scheme in the Master Chemical Mechanism increases <span class="hlt">HCHO</span> production by approx. 16% in MCMv<span class="hlt">3.3</span>.1 relative to MCMv<span class="hlt">3</span>.2. The GEOS-Chem mechanism provides a good approximation of the explicit isoprene chemistry in MCMv<span class="hlt">3.3</span>.1 and generally reproduces the magnitude and source distribution of <span class="hlt">HCHO</span> production rates. We analytically derive improvements to the isoprene scheme in CB6r2 and incorporate these changes into a new mechanism called CB6r2-UMD, which is designed to preserve computational efficiency. The CB6r2-UMD mechanism mimics production of <span class="hlt">HCHO</span> in MCMv<span class="hlt">3.3</span>.1 and demonstrates good agreement with observed mixing ratios from SENEX (-14%). Improved simulation of <span class="hlt">HCHO</span> also impacts modeled ozone: at approx. 0.<span class="hlt">3</span> ppb NO, the ozone production rate increases approx. <span class="hlt">3</span>% between CB6r2 and CB6r2-UMD, and rises another approx. 4% when <span class="hlt">HCHO</span> is constrained to match observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1996GeoRL..23.1045S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1996GeoRL..23.1045S"><span>Field measurement evidence for an atmospheric chemical source of formic and acetic acids in the tropic</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sanhueza, Eugenio; Santana, Magaly; Trapp, Dorothea; de Serves, Claes; Figueroa, Luis; Romero, Rodrigo; Rondón, Alberto; Donoso, Loreto</p> <p></p> <p>The simultaneous measurements of atmospheric HCOOH, CH<span class="hlt">3</span>COOH H2<span class="hlt">O</span>2, organic peroxides, <span class="hlt">HCHO</span>, CH<span class="hlt">3</span>CHO and isoprene made in the Venezuelan savannah region, in the wet season (September, 1993) and during the period of high solar irradiation is reported. The average concentrations (in ppbv) between 10:00 and 16:00 were: HCOOH 0.75±0.32, CH<span class="hlt">3</span>COOH 0.56±0.28, H2<span class="hlt">O</span>2 1.37±0.48, the total peroxides 1.83±0.60, <span class="hlt">HCHO</span> 1.38± .43, CH<span class="hlt">3</span>CHO 0.35±0.15, and isoprene 2.18±0.78. A good correlation was observed between the concentrations (15 min averages) of both acids. The acids also correlate with isoprene (the most abundant olefin in the savannah atmosphere), H2<span class="hlt">O</span>2 and the total peroxides. HCOOH also correlates well with <span class="hlt">HCHO</span> and CH<span class="hlt">3</span>CHO. These results support the hypothesis that significant amount of formic and acetic acids are produced in the tropical atmosphere as a result of the oxidation of reactive hydrocarbons.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_9 --> <div id="page_10" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="181"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1816756L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1816756L"><span>Intercomparison of OMI NO2 and <span class="hlt">HCHO</span> air mass factor calculations: recommendations and best practices for retrievals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lorente Delgado, Alba; Klaas Boersma, Folkert; Hilboll, Andreas; Richter, Andreas; Yu, Huan; van Roozendael, Michel; Dörner, Steffen; Wagner, Thomas; Barkley, Michael; Lamsal, Lok; Lin, Jintai; Liu, Mengyao</p> <p>2016-04-01</p> <p>We present a detailed comparison of the air mass factor (AMF) calculation process used by various research groups for OMI satellite retrievals of NO2 and <span class="hlt">HCHO</span>. Although satellite retrievals have strongly improved over the last decades, there is still a need to better understand and reduce the uncertainties associated with every retrieval step of satellite data products, such as the AMF calculation. Here we compare and evaluate the different approaches used to calculate AMFs by several scientific groups (KNMI (WUR), IASB-BIRA, IUP-UNI. BREMEN, MPI-C, NASA GSFC, LEICESTER UNI. and PEKING UNI.). Each group calculated altitude dependent (box-) AMFs and clear sky and total tropospheric AMFs for several OMI orbits. First, European groups computed AMFs for one OMI orbit using common settings for the choice of surface albedo data, terrain height, cloud treatment and a priori vertical profile. Second, every group computed AMFs for two complete days in different seasons using preferred settings for the ancillary data and cloud treatment as a part of a Round Robin exercise. Box-AMFs comparison showed good consistency and underlined the importance of a correct treatment of the physical processes affecting the effective light path and the vertical discretization of the atmosphere. Using common settings, tropospheric NO2 AMFs in polluted pixels on average agreed within 4.7% whereas in remote pixels agreed within <span class="hlt">3</span>.5%. Using preferred settings relative differences between AMFs increase up to 15-30%. This increase is traced back to the different choices and assumptions made throughout the AMF calculation, which affect the final AMF values and thus the uncertainty in the AMF calculation. Differences between state of the art cloud treatment approaches highlight the importance of an accurate cloud correction: total and clear sky AMFs in polluted conditions differ by up to 40% depending on the retrieval scenario. Based on the comparison results, specific recommendations on best</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhDT........50A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhDT........50A"><span>Photochemistry and transport of tropospheric ozone and its precursors in urban and remote environments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Anderson, Daniel Craig</p> <p></p> <p>Tropospheric ozone (<span class="hlt">O</span><span class="hlt">3</span>) adversely affects human health, reduces crop yields, and contributes to climate forcing. To limit these effects, the processes controlling <span class="hlt">O</span><span class="hlt">3</span> abundance as well as that of its precursor molecules must be fully characterized. Here, I examine three facets of <span class="hlt">O</span> <span class="hlt">3</span> production, both in heavily polluted and remote environments. First, using in situ observations from the DISCOVER-AQ field campaign in the Baltimore/Washington region, I evaluate the emissions of the <span class="hlt">O</span> <span class="hlt">3</span> precursors CO and NOx (NOx = NO + NO2) in the National Emissions Inventory (NEI). I find that CO/NOx emissions ratios derived from observations are 21% higher than those predicted by the NEI. Comparisons to output from the CMAQ model suggest that CO in the NEI is accurate within 15 +/- 11%, while NOx emissions are overestimated by 51-70%, likely due to errors in mobile sources. These results imply that ambient ozone concentrations will respond more efficiently to NOx controls than current models suggest. I then investigate the source of high <span class="hlt">O</span><span class="hlt">3</span> and low H2<span class="hlt">O</span> structures in the Tropical Western Pacific (TWP). A combination of in situ observations, satellite data, and models show that the high <span class="hlt">O</span><span class="hlt">3</span> results from photochemical production in biomass burning plumes from fires in tropical Southeast Asia and Central Africa; the low relative humidity results from large-scale descent in the tropics. Because these structures have frequently been attributed to mid-latitude pollution, biomass burning in the tropics likely contributes more to the radiative forcing of climate than previously believed. Finally, I evaluate the processes controlling formaldehyde (<span class="hlt">HCHO</span>) in the TWP. Convective transport of near surface <span class="hlt">HCHO</span> leads to a 33% increase in upper tropospheric <span class="hlt">HCHO</span> mixing ratios; convection also likely increases upper tropospheric CH <span class="hlt">3</span>OOH to ~230 pptv, enough to maintain background <span class="hlt">HCHO</span> at ~75 pptv. The long-range transport of polluted air, with NO four times the convectively controlled background</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhCS.980a2033L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhCS.980a2033L"><span>Investigation of turbulent swirling jet-flames by PIV / OH PLIF / <span class="hlt">HCHO</span> PLIF</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lobasov, A. S.; Chikishev, L. M.</p> <p>2018-03-01</p> <p>The present paper reports on the investigation of fuel-lean and fuel-rich turbulent combustion in a high-swirl jet. Swirl rate of the flow exceeded a critical value for breakdown of the swirling jet’s vortex core and formation of the recirculation zone at the jet axis. The measurements were performed by the stereo PIV, OH PLIF and <span class="hlt">HCHO</span> PLIF techniques, simultaneously. The Reynolds number based on the flow rate and viscosity of the air was fixed as 5 000 (the bulk velocity was U 0 = 5 m/s). Three cases of the equivalence ratio ϕ of the mixture issuing from the nozzle-burner were considered, viz., 0.7, 1.4 and 2.5. The latter case corresponded to a lifted flame of fuel-rich swirling jet flow, partially premixed with the surrounding air. In all cases the flame front was subjected to deformations due to large-scale vortices, which rolled-up in the inner (around the central recirculation zone) and outer (between the annular jet core and surrounding air) mixing layers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3763807','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3763807"><span>Houston’s rapid ozone increases: preconditions and geographic origins</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Couzo, Evan; Jeffries, Harvey E.; Vizuete, William</p> <p>2013-01-01</p> <p>Many of Houston’s highest 8-h ozone (<span class="hlt">O</span><span class="hlt">3</span>) peaks are characterised by increases in concentrations of at least 40 ppb in 1 h, or 60 ppb in 2 h. These rapid increases are called non-typical <span class="hlt">O</span><span class="hlt">3</span> changes (NTOCs). In 2004, the Texas Commission on Environmental Quality (TCEQ) developed a novel emissions control strategy aimed at eliminating NTOCs. The strategy limited routine and short-term emissions of ethene, propene, 1,<span class="hlt">3</span>-butadiene and butene isomers, collectively called highly reactive volatile organic compounds (HRVOCs), which are released from petrochemical facilities. HRVOCs have been associated with NTOCs through field campaigns and modelling studies. This study analysed wind measurements and <span class="hlt">O</span><span class="hlt">3</span>, formaldehyde (<span class="hlt">HCHO</span>) and sulfur dioxide (SO2) concentrations from 2000 to 2011 at 25 ground monitors in Houston. NTOCs almost always occurred when monitors were downwind of petrochemical facilities. Rapid <span class="hlt">O</span><span class="hlt">3</span> increases were associated with low wind speeds; 75 % of NTOCs occurred when the <span class="hlt">3</span>-h average wind speed preceding the event was less than 6.5 km h−1. Statistically significant differences in <span class="hlt">HCHO</span> concentrations were seen between days with and without NTOCs. Early afternoon <span class="hlt">HCHO</span> concentrations were greater on NTOC days. In the morning before an observed NTOC event, however, there were no significant differences in <span class="hlt">HCHO</span> concentrations between days with and without NTOCs. Hourly SO2 concentrations also increased rapidly, exhibiting behaviour similar to NTOCs. Oftentimes, the SO2 increases preceded a NTOC. These findings show that, despite the apparent success of targeted HRVOC emission controls, further restrictions may be needed to eliminate the remaining <span class="hlt">O</span><span class="hlt">3</span> events. PMID:24014080</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23682559','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23682559"><span>Uptake of HNO<span class="hlt">3</span> on aviation kerosene and aircraft engine soot: influences of H2<span class="hlt">O</span> or/and H2SO4.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Loukhovitskaya, Ekaterina E; Talukdar, Ranajit K; Ravishankara, A R</p> <p>2013-06-13</p> <p>The uptake of HNO<span class="hlt">3</span> on aviation kerosene soot (TC-1 soot) was studied in the absence and presence of water vapor at 295 and 243 K. The influence of H2SO4 coating of the TC-1 soot surface on HNO<span class="hlt">3</span> uptake was also investigated. Only reversible uptake of HNO<span class="hlt">3</span> was observed. <span class="hlt">HONO</span> and NO2, potential products of reactive uptake of HNO<span class="hlt">3</span>, were not observed under any conditions studied here. The uptake of nitric acid increased slightly with relative humidity (RH). Coating of the TC-1 soot surface with sulfuric acid decreased the uptake of HNO<span class="hlt">3</span> and did not lead to displacement of H2SO4 from the soot surface. A limited set of measurements was carried out on soot generated by aircraft engine combustor (E-soot) with results similar to those on TC-1 soot. The influence of water on HNO<span class="hlt">3</span> uptake on E-soot appeared to be more pronounced than on TC-1 soot. Our results suggest that HNO<span class="hlt">3</span> loss in the upper troposphere due to soot is not significant except perhaps in aircraft exhaust plumes. Our results also suggest that HNO<span class="hlt">3</span> is not converted to either NO2 or <span class="hlt">HONO</span> upon its uptake on soot in the atmosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23614454','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23614454"><span>Gas phase hydrolysis of formaldehyde to form methanediol: impact of formic acid catalysis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hazra, Montu K; Francisco, Joseph S; Sinha, Amitabha</p> <p>2013-11-21</p> <p>We find that formic acid (FA) is very effective at facilitating diol formation through its ability to reduce the barrier for the formaldehyde (<span class="hlt">HCHO</span>) hydrolysis reaction. The rate limiting step in the mechanism involves the isomerization of a prereactive collision complex formed through either the HCHO···H2<span class="hlt">O</span> + FA and/or <span class="hlt">HCHO</span> + FA···H2<span class="hlt">O</span> pathways. The present study finds that the effective barrier height, defined as the difference between the zero-point vibrational energy (ZPE) corrected energy of the transition state (TS) and the HCHO···H2<span class="hlt">O</span> + FA and <span class="hlt">HCHO</span> + FA···H2<span class="hlt">O</span> starting reagents, are respectively only ∼1 and ∼4 kcal/mol. These barriers are substantially lower than the ∼17 kcal/mol barrier associated with the corresponding step in the hydrolysis of <span class="hlt">HCHO</span> catalyzed by a single water molecule (<span class="hlt">HCHO</span> + H2<span class="hlt">O</span> + H2<span class="hlt">O</span>). The significantly lower barrier heights for the formic acid catalyzed pathway reveal a new important role that organic acids play in the gas phase hydrolysis of atmospheric carbonyl compounds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26931162','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26931162"><span>A simple and sensitive flow injection method based on the catalytic activity of CdS quantum dots in an acidic permanganate chemiluminescence system for determination of formaldehyde in water and wastewater.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Khataee, Alireza; Lotfi, Roya; Hasanzadeh, Aliyeh; Iranifam, Mortaza</p> <p>2016-04-01</p> <p>A simple and sensitive flow injection chemiluminescence (CL) method in which CdS quantum dots (QDs) enhanced the CL intensity of a KMn<span class="hlt">O</span>4-formaldehyde (<span class="hlt">HCHO</span>) reaction was offered for the determination of <span class="hlt">HCHO</span>. This CL system was based on the catalytic activity of CdS QDs and their participation in the CL resonance energy transfer (CRET) phenomenon. A possible mechanism for the supplied CL system was proposed using the kinetic curves of the CL systems and the spectra of CL, photoluminescence (PL) and ultraviolet-visible (UV-Vis). The emanated CL intensity of the KMn<span class="hlt">O</span>4-CdS QDs system was amplified in the presence of a trace level of <span class="hlt">HCHO</span>. Based on this enhancement effect, a simple and sensitive flow injection CL method was suggested for the determination of <span class="hlt">HCHO</span> concentration in environmental water and wastewater samples. Under selected optimized experimental conditions, the increased CL intensity was proportional to the <span class="hlt">HCHO</span> concentration in the range of 0.03-4.5 μg L(-1) and 4.5-10.0 μg L(-1). The detection limits (<span class="hlt">3</span>σ) were 0.0003 μg L(-1) and 1.2 μg L(-1). The relative standard deviations (RSD%) for eleven replicate determinations of 4.0 μg L(-1) <span class="hlt">HCHO</span> were 2.2%. Furthermore, the feasibility of the developed method was investigated via the determination of <span class="hlt">HCHO</span> concentration in environmental water and wastewater samples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AtmEn..87..200R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AtmEn..87..200R"><span>The reduction of HNO<span class="hlt">3</span> by volatile organic compounds emitted by motor vehicles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rutter, A. P.; Malloy, Q. G. J.; Leong, Y. J.; Gutierrez, C. V.; Calzada, M.; Scheuer, E.; Dibb, J. E.; Griffin, R. J.</p> <p>2014-04-01</p> <p>Nitric acid (HNO<span class="hlt">3</span>) was reduced in a flow tube by volatile organic carbon compounds (VOCs) generated from engine oil vapor. The primary reaction product was believed to be <span class="hlt">HONO</span>. The reaction was not enhanced when Teflon® Raschig rings were added to the flow tube to increase surface area, thereby showing the reaction to be homogeneous under the conditions studied. The <span class="hlt">HONO</span> formation observed ranged between 0.1 and 0.6 ppb h-1, with a mean of 0.<span class="hlt">3</span> ± 0.1 ppb h-1, for typical HNO<span class="hlt">3</span> concentrations of 4-5 ppb and estimated concentrations of the reactive components in the engine oil vapor between 200 and 300 ppt. The observations in this study compare well to a recently published field study conducted in Houston that observed average formation rates of 0.6 ± 0.<span class="hlt">3</span> ppb h-1. Water vapor was found to decrease the <span class="hlt">HONO</span> formation rate by ˜0.1 ppb h-1 for every 1% increase in the water mixing ratio.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20075247','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20075247"><span>How the shape of an H-bonded network controls proton-coupled water activation in <span class="hlt">HONO</span> formation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Relph, Rachael A; Guasco, Timothy L; Elliott, Ben M; Kamrath, Michael Z; McCoy, Anne B; Steele, Ryan P; Schofield, Daniel P; Jordan, Kenneth D; Viggiano, Albert A; Ferguson, Eldon E; Johnson, Mark A</p> <p>2010-01-15</p> <p>Many chemical reactions in atmospheric aerosols and bulk aqueous environments are influenced by the surrounding solvation shell, but the precise molecular interactions underlying such effects have rarely been elucidated. We exploited recent advances in isomer-specific cluster vibrational spectroscopy to explore the fundamental relation between the hydrogen (H)-bonding arrangement of a set of ion-solvating water molecules and the chemical activity of this ensemble. We find that the extent to which the nitrosonium ion (NO+)and water form nitrous acid (<span class="hlt">HONO</span>) and a hydrated proton cluster in the critical trihydrate depends sensitively on the geometrical arrangement of the water molecules in the network. Theoretical analysis of these data details the role of the water network in promoting charge delocalization.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.A24A..06M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A24A..06M"><span>Fast in-situ measurements of glyoxal (CHOCHO) and nitrous acid (<span class="hlt">HONO</span>) in northern Chinese plane during CAREBEIJING - NCP2014</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Min, K. E.; Dube, W. P.; Washenfelder, R. A.; Langford, A. O.; Brown, S. S.; Broch, S.; Fuchs, H.; Gomm, S.; Hofzumahaus, A.; Holland, F.; Hu, M.; Huey, L. G.; Kubik, K.; Li, X.; Liu, X.; Lu, K.; Rohrer, F.; Shao, M.; Sjostedt, S. J.; Tan, Z.; Zhu, T.; Wahner, A.; Wang, B.; Wang, M.; Wang, Y.; Zeng, L.; Zhang, Y.</p> <p>2014-12-01</p> <p>The Northern China Plain has experienced visibility degradation and detrimental health impacts due to aerosol and photochemical pollution. To examine these air quality issues, CAREBEIJING-NCP2014 (Care Beijing - Northern China Plain 2014) was held in WangDu, Hebei province, China from 6 June to 15 July 2014. We deployed our newly developed instrument, ACES (Airborne Cavity Enhanced Spectrometer), for high time resolution in-situ measurement of glyoxal (CHOCHO), nitrous acid (<span class="hlt">HONO</span>) and other trace gases (NO2, H2<span class="hlt">O</span>) to investigate mechanisms of oxidation processes and secondary organic aerosol (SOA) formation. The in situ measurements of CHOCHO provide observational constraints on secondary organic aerosol formation and oxidation processes, since this molecule has been proposed to play a crucial role in forming aerosol due to its high water solubility, isomerization, and abundant production from the oxidation of many different volatile organic compounds (VOCs). A box model analysis incorporating secondary glyoxal sources from VOC oxidation and sinks to OH reaction, photolysis and heterogeneous uptake will be used to determine a budget and potential for SOA formation. This work was supported by the National Natural Science Foundation of China (21190052), the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB05010500) and the U.S. National Science Foundation Atmospheric (AGS-1405805).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A51B2050D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A51B2050D"><span>Below-Canopy Isoprene Nitrate Chemistry and Dynamics in a Mixed Coniferous/Deciduous Forest Canopy during the 2016 PROPHET-AMOS Summer Field Campaign.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Desrochers, S. J.; Slade, J. H., Jr.; Shepson, P. B.; Alwe, H. D.; Millet, D. B.; Kavassalis, S.; Shi, Q.; Murphy, J. G.; Bloss, W.; Wood, E.; Stevens, P. S.; Mauldin, L.; Cantrell, C. A.; Kim, T.; Zhou, X.; Helmig, D.; Shutter, J. D.; Rivera, J. C.; Keutsch, F. N.; Flynn, J. H., III; Alvarez, S. L.; Erickson, M.; Wang, W.; Griffin, R. J.; Bui, A. T.; Kim, K.; Wallace, H. W., IV</p> <p>2017-12-01</p> <p>Isoprene is the most abundant biogenic volatile organic compound (BVOC) emitted in forest ecosystems. Isoprene hydroxynitrates (IN) can be produced via OH oxidation of isoprene in the presence of NOx, thereby sequestering NOx, and limiting ozone production. Furthermore, IN can lead to the formation of secondary organic aerosols (SOA), which affect air quality and radiative forcing. Forest environments are often complex in terms of isoprene emission as a function of height through the canopy, and to date, there has been little study of chemistry below the canopy. However, the above- and below-canopy environments can be quite different, e.g. in terms of irradiance, NOx, and temperature. Thus, for example, there can be significantly more nitrate radical chemistry below canopy during daytime. Here we present and discuss IN measurements from the 2016 PROPHET-AMOS summer field campaign at the University of Michigan Biological Station (UMBS). IN was sampled from inlets at two heights, below the canopy and within the canopy, using a single quadrupole chemical ionization mass spectrometer, using I- ion chemistry. Differences in [IN] measured from the two inlets varied throughout the campaign, indicating a difference in the importance of light and dark dominated IN production pathways. Measurements of isoprene, terpenes, OH, <span class="hlt">HONO</span>, <span class="hlt">HCHO</span>, NOx, <span class="hlt">O</span><span class="hlt">3</span>, HO2, H2<span class="hlt">O</span>, j<span class="hlt">O</span><span class="hlt">3</span>, jNO2, and jNO<span class="hlt">3</span> values conducted during the campaign were used to constrain a 0-D box model to simulate IN concentrations in order to better understand the influence of local-scale chemistry on IN production and destruction, and mixing below and through the forest canopy. Here we will compare measurements with model results, and discuss the implications for IN processing and mixing under the canopy and for nitrogen exchange above and below the forest canopy environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20180002375','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20180002375"><span>Proposed Trace Gas Measurements Over the Western United States for TROPOMI Validation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Parworth, Caroline L.; Marrero, Josette E.; Yates, Emma L.; Ryoo, Ju-Mee; Iraci, Laura T.</p> <p>2018-01-01</p> <p>The Alpha Jet Atmospheric eXperiment (AJAX), located in the Bay Area of California, is a joint effort between NASA Ames Research Center and H211, LCC. AJAX makes in-situ airborne measurements of trace gases 2-4 times per month, resulting in over 216 flights since 2011. Current measurements include ozone (<span class="hlt">O</span><span class="hlt">3</span>), carbon dioxide (CO2), methane (CH4), water (H2<span class="hlt">O</span>), formaldehyde (<span class="hlt">HCHO</span>), and meteorological measurements (i.e., ambient pressure, temperature, and <span class="hlt">3</span>D winds). Currently, the AJAX team is working to incorporate nitrogen dioxide (NO2) measurements with a Cavity Attenuated Phase Shift Spectrometer (CAPS). Successful science flights coincident with satellite overpasses have been performed since 2011 by the Alpha Jet, with more than 40 flights under the Greenhouse Observing SATellite (GOSAT) and several flights under the Orbiting Carbon Observatory-2 (OCO-2). Results from these flights, which have covered a range of different surfaces and seasonal conditions, will be presented. In-situ vertical profiles of <span class="hlt">O</span><span class="hlt">3</span>, CO2, CH4, H2<span class="hlt">O</span>, <span class="hlt">HCHO</span>, and NO2 from the surface to 28,000 feet made by AJAX will also be valuable for satellite validation of data products obtained from the TROPOspheric Montoring Instrument (TROPOMI). TROPOMI is on board the Copernicus Sentinel-5 precursor (S5p) satellite, with level 2 products including <span class="hlt">O</span><span class="hlt">3</span>, CO, CH4, <span class="hlt">HCHO</span>, NO2, and aerosols.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1206591','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1206591"><span>Impact of isoprene and <span class="hlt">HONO</span> chemistry on ozone and OVOC formation in a semirural South Korean forest</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kim, S.; Kim, S. -Y.; Lee, M.</p> <p></p> <p>Rapid urbanization and economic development in East Asia in past decades has led to photochemical air pollution problems such as excess photochemical ozone and aerosol formation. Asian megacities such as Seoul, Tokyo, Shanghai, Guangzhou, and Beijing are surrounded by densely forested areas, and recent research has consistently demonstrated the importance of biogenic volatile organic compounds (VOCs) from vegetation in determining oxidation capacity in the suburban Asian megacity regions. Uncertainties in constraining tropospheric oxidation capacity, dominated by hydroxyl radical, undermine our ability to assess regional photochemical air pollution problems. We present an observational data set of CO, NO x, SO 2,more » ozone, <span class="hlt">HONO</span>, and VOCs (anthropogenic and biogenic) from Taehwa research forest (TRF) near the Seoul metropolitan area in early June 2012. The data show that TRF is influenced both by aged pollution and fresh biogenic volatile organic compound emissions. With the data set, we diagnose HO x (OH, HO 2, and RO 2) distributions calculated using the University of Washington chemical box model (UWCM v2.1) with near-explicit VOC oxidation mechanisms from MCM v<span class="hlt">3</span>.2 (Master Chemical Mechanism). Uncertainty from unconstrained <span class="hlt">HONO</span> sources and radical recycling processes highlighted in recent studies is examined using multiple model simulations with different model constraints. The results suggest that (1) different model simulation scenarios cause systematic differences in HO x distributions, especially OH levels (up to 2.5 times), and (2) radical destruction (HO 2 + HO 2 or HO 2 + RO 2) could be more efficient than radical recycling (RO 2 + NO), especially in the afternoon. Implications of the uncertainties in radical chemistry are discussed with respect to ozone–VOC–NO x sensitivity and VOC oxidation product formation rates. Overall, the NO x limited regime is assessed except for the morning hours (8 a.m. to 12 p.m. local standard time), but the degree</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1206591-impact-isoprene-hono-chemistry-ozone-ovoc-formation-semirural-south-korean-forest','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1206591-impact-isoprene-hono-chemistry-ozone-ovoc-formation-semirural-south-korean-forest"><span>Impact of isoprene and <span class="hlt">HONO</span> chemistry on ozone and OVOC formation in a semirural South Korean forest</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Kim, S.; Kim, S. -Y.; Lee, M.; ...</p> <p>2015-04-29</p> <p>Rapid urbanization and economic development in East Asia in past decades has led to photochemical air pollution problems such as excess photochemical ozone and aerosol formation. Asian megacities such as Seoul, Tokyo, Shanghai, Guangzhou, and Beijing are surrounded by densely forested areas, and recent research has consistently demonstrated the importance of biogenic volatile organic compounds (VOCs) from vegetation in determining oxidation capacity in the suburban Asian megacity regions. Uncertainties in constraining tropospheric oxidation capacity, dominated by hydroxyl radical, undermine our ability to assess regional photochemical air pollution problems. We present an observational data set of CO, NO x, SO 2,more » ozone, <span class="hlt">HONO</span>, and VOCs (anthropogenic and biogenic) from Taehwa research forest (TRF) near the Seoul metropolitan area in early June 2012. The data show that TRF is influenced both by aged pollution and fresh biogenic volatile organic compound emissions. With the data set, we diagnose HO x (OH, HO 2, and RO 2) distributions calculated using the University of Washington chemical box model (UWCM v2.1) with near-explicit VOC oxidation mechanisms from MCM v<span class="hlt">3</span>.2 (Master Chemical Mechanism). Uncertainty from unconstrained <span class="hlt">HONO</span> sources and radical recycling processes highlighted in recent studies is examined using multiple model simulations with different model constraints. The results suggest that (1) different model simulation scenarios cause systematic differences in HO x distributions, especially OH levels (up to 2.5 times), and (2) radical destruction (HO 2 + HO 2 or HO 2 + RO 2) could be more efficient than radical recycling (RO 2 + NO), especially in the afternoon. Implications of the uncertainties in radical chemistry are discussed with respect to ozone–VOC–NO x sensitivity and VOC oxidation product formation rates. Overall, the NO x limited regime is assessed except for the morning hours (8 a.m. to 12 p.m. local standard time), but the degree</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21171657','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21171657"><span>Thermodynamics of the formaldehyde-water and formaldehyde-ice systems for atmospheric applications.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Barret, Manuel; Houdier, Stephan; Domine, Florent</p> <p>2011-01-27</p> <p>Formaldehyde (<span class="hlt">HCHO</span>) is a species involved in numerous key atmospheric chemistry processes that can significantly impact the oxidative capacity of the atmosphere. Since gaseous <span class="hlt">HCHO</span> is soluble in water, the water droplets of clouds and the ice crystals of snow exchange <span class="hlt">HCHO</span> with the gas phase and the partitioning of <span class="hlt">HCHO</span> between the air, water, and ice phases must be known to understand its chemistry. This study proposes thermodynamic formulations for the partitioning of <span class="hlt">HCHO</span> between the gas phase and the ice and liquid water phases. A reanalysis of existing data on the vapor-liquid equilibrium has shown the inadequacy of the Henry's law formulation, and we instead propose the following equation to predict the mole fraction of <span class="hlt">HCHO</span> in liquid water at equilibrium, X(<span class="hlt">HCHO</span>,liq), as a function of the partial pressure P(<span class="hlt">HCHO</span>) (Pa) and temperature T (K): X(<span class="hlt">HCHO</span>,liq) = 1.700 × 10(-15) e((8014/T))(P(<span class="hlt">HCHO</span>))(1.105). Given the paucity of data on the gas-ice equilibrium, the solubility of <span class="hlt">HCHO</span> and the diffusion coefficient (D(<span class="hlt">HCHO</span>)) in ice were measured by exposing large single ice crystals to low P(<span class="hlt">HCHO</span>). Our recommended value for D(<span class="hlt">HCHO</span>) over the temperature range 243-266 K is D(<span class="hlt">HCHO</span>) = 6 × 10(-12) cm(2) s(-1). The solubility of <span class="hlt">HCHO</span> in ice follows the relationship X(<span class="hlt">HCHO</span>,ice) = 9.898 × 10(-13) e((4072/T))(P(<span class="hlt">HCHO</span>))(0.803). Extrapolation of these data yields the P(<span class="hlt">HCHO</span>) versus 1/T phase diagram for the H(2)<span class="hlt">O-HCHO</span> system. The comparison of our results to existing data on the partitioning of <span class="hlt">HCHO</span> between the snow and the atmosphere in the high arctic highlights the interplay between thermodynamic equilibrium and kinetics processes in natural systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009JGRD..11421306W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009JGRD..11421306W"><span>Emissions of NOx, SO2, CO, and <span class="hlt">HCHO</span> from commercial marine shipping during Texas Air Quality Study (TexAQS) 2006</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Williams, E. J.; Lerner, B. M.; Murphy, P. C.; Herndon, S. C.; Zahniser, M. S.</p> <p>2009-11-01</p> <p>We report measurements of NOx, SO2, CO, and <span class="hlt">HCHO</span> mass-based emission factors from more than 200 commercial vessel encounters in the Gulf of Mexico and the Houston-Galveston region of Texas during August and September, 2006. For underway ships, bulk freight carriers have the highest average NOx emissions at ˜87 g NOx (kg fuel)-1, followed by tanker ships at ˜79 g NOx (kg fuel)-1, while container carriers, passenger ships, and tugs all emit an average of about ˜60 g NOx (kg fuel)-1. Emission of NOx from stationary vessels was lower, except for container ships and tugs, and likely reflects use of medium-speed diesel engines. Overall, our mean NOx emission factors are 10-15% lower than published data. Average emission of SO2 was lower for passenger ships and tugs and tows (6-7 g SO2 (kg fuel)-1) than for larger cargo vessels (20-30 g SO2 (kg fuel)-1). Our data for large cargo ships in this region indicate an average residual fuel sulfur content of ˜1.4% which is a factor of two lower than the global average of 2.7%. Emission of CO was low for all categories (7-16 g CO (kg fuel)-1), although our mean overall CO emission factor is about 10% higher than published data. Emission of <span class="hlt">HCHO</span> was less than 5% that of CO. Despite considerable variability, no functional relationships, such as emissions changes with engine speed or load, could be discerned. Comparison of emission factors from ships to those from other sources suggests ship emissions in this region cannot be ignored.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NML....10....4W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NML....10....4W"><span>Simultaneous Detection and Removal of Formaldehyde at Room Temperature: Janus Au@Zn<span class="hlt">O</span>@ZIF-8 Nanoparticles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Dawei; Li, Zhiwei; Zhou, Jian; Fang, Hong; He, Xiang; Jena, Puru; Zeng, Jing-Bin; Wang, Wei-Ning</p> <p>2018-03-01</p> <p>The detection and removal of volatile organic compounds (VOCs) are of great importance to reduce the risk of indoor air quality concerns. This study reports the rational synthesis of a dual-functional Janus nanostructure and its feasibility for simultaneous detection and removal of VOCs. The Janus nanostructure was synthesized via an anisotropic growth method, composed of plasmonic nanoparticles, semiconductors, and metal organic frameworks (e.g., Au@Zn<span class="hlt">O</span>@ZIF-8). It exhibits excellent selective detection to formaldehyde (<span class="hlt">HCHO</span>, as a representative VOC) at room temperature over a wide range of concentrations (from 0.25 to 100 ppm), even in the presence of water and toluene molecules as interferences. In addition, <span class="hlt">HCHO</span> was also found to be partially oxidized into non-toxic formic acid simultaneously with detection. The mechanism underlying this technology was unraveled by both experimental measurements and theoretical calculations: Zn<span class="hlt">O</span> maintains the conductivity, while ZIF-8 improves the selective gas adsorption; the plasmonic effect of Au nanorods enhances the visible-light-driven photocatalysis of Zn<span class="hlt">O</span> at room temperature. [Figure not available: see fulltext.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=borderline&pg=7&id=EJ866815','ERIC'); return false;" href="https://eric.ed.gov/?q=borderline&pg=7&id=EJ866815"><span>Reliability and Validity of the HoNOS-LD and <span class="hlt">HoNOS</span> in a Sample of Individuals with Mild to Borderline Intellectual Disability and Severe Emotional and Behavior Disorders</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Tenneij, Nienke; Didden, Robert; Veltkamp, Eline; Koot, Hans M.</p> <p>2009-01-01</p> <p>In this study, psychometric properties of the Health of the Nation Outcome scales (<span class="hlt">HoNOS</span>) and Health of the Nation Outcome Scales for People with Learning Disabilities (HoNOS-LD) were investigated in a sample (n = 79) of (young) adults with mild to borderline intellectual disability (ID) and severe behavior and mental health problems who were…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000GeoRL..27.2093F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000GeoRL..27.2093F"><span>Field intercomparison of a novel optical sensor for formaldehyde quantification</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Friedfeld, Stephen; Fraser, Matthew; Lancaster, David; Leleux, Darrin; Rehle, Dirk; Tittel, Frank</p> <p>2000-08-01</p> <p>A one-week in situ intercomparison campaign was completed on the Rice University campus for measuring <span class="hlt">HCHO</span> using three different techniques, including a novel optical sensor based on difference frequency generation (DFG) operating at room temperature. Two chemical derivatization methods, 2,4-dinitrophenylhydrazine (DNPH) and <span class="hlt">o</span>-(2,<span class="hlt">3</span>,4,5,6-pentafluorobenzyl) hydroxylamine (PFBHA), were deployed during the daylight hours for three- to four-hour time-integrated samples. A real-time optical sensor based on laser absorption spectroscopy was operated simultaneously, including nighttime hours. This tunable spectroscopic source based on difference frequency mixing of two fiber-amplified diode lasers in periodically poled LiNb<span class="hlt">O</span><span class="hlt">3</span> (PPLN) was operated at <span class="hlt">3</span>.5315 µm (2831.64 cm-1) to access a strong <span class="hlt">HCHO</span> ro-vibrational transition free of interferences from other species. The results showed a bias of -1.7 and -1.2 ppbv and a gross error of 2.6 and 1.5 ppbv for DNPH and PFBHA measurements, respectively, compared with DFG measurements. These results validate the DFG sensor for time-resolved measurements of <span class="hlt">HCHO</span> in urban areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4671125','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4671125"><span>The cumulative influence of hyperoxia and hypercapnia on blood oxygenation and R2*</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Faraco, Carlos C; Strother, Megan K; Siero, Jeroen CW; Arteaga, Daniel F; Scott, Allison O; Jordan, Lori C; Donahue, Manus J</p> <p>2015-01-01</p> <p>Cerebrovascular reactivity (CVR)-weighted blood-oxygenation-level-dependent magnetic resonance imaging (BOLD-MRI) experiments are frequently used in conjunction with hyperoxia. Owing to complex interactions between hyperoxia and hypercapnia, quantitative effects of these gas mixtures on BOLD responses, blood and tissue R2*, and blood oxygenation are incompletely understood. Here we performed BOLD imaging (<span class="hlt">3</span> T; TE/TR=35/2,000 ms; spatial resolution=<span class="hlt">3</span> × <span class="hlt">3</span> × <span class="hlt">3</span>.5 mm<span class="hlt">3</span>) in healthy volunteers (n=12; age=29±4.1 years) breathing (i) room air (RA), (ii) normocapnic–hyperoxia (95% <span class="hlt">O</span>2/5% N2, HO), (iii) hypercapnic–normoxia (5% CO2/21% <span class="hlt">O</span>2/74% N2, HC-NO), and (iv) hypercapnic–hyperoxia (5% CO2/95% <span class="hlt">O</span>2, <span class="hlt">HC-HO</span>). For <span class="hlt">HC-HO</span>, experiments were performed with separate RA and HO baselines to control for changes in <span class="hlt">O</span>2. T2-relaxation-under-spin-tagging MRI was used to calculate basal venous oxygenation. Signal changes were quantified and established hemodynamic models were applied to quantify vasoactive blood oxygenation, blood–water R2*, and tissue–water R2*. In the cortex, fractional BOLD changes (stimulus/baseline) were HO/RA=0.011±0.007; HC-NO/RA=0.014±0.004; HC-HO/HO=0.020±0.008; and HC-HO/RA=0.035±0.010; for the measured basal venous oxygenation level of 0.632, this led to venous blood oxygenation levels of 0.660 (HO), 0.665 (HC-NO), and 0.712 (<span class="hlt">HC-HO</span>). Interleaving a <span class="hlt">HC-HO</span> stimulus with HO baseline provided a smaller but significantly elevated BOLD response compared with a HC-NO stimulus. Results provide an outline for how blood oxygenation differs for several gas stimuli and provides quantitative information on how hypercapnic BOLD CVR and R2* are altered during hyperoxia. PMID:26174329</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1813514D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1813514D"><span>Tropospheric NO2 and <span class="hlt">HCHO</span> columns derived from ground-based MAX-DOAS system in Guangzhou, China and comparison with satellite observations: First results within the EU FP7 project MarcoPolo</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Drosoglou, Theano; Kouremeti, Natalia; Bais, Alkis; Zyrichidou, Irene; Li, Shu; Balis, Dimitris; Huang, Zhonghui</p> <p>2016-04-01</p> <p>A miniature MAX-DOAS system, Phaethon, has been developed at the Laboratory of Atmospheric Physics of the Aristotle University of Thessaloniki, Greece, for ground-based monitoring of column densities of atmospheric gases. Simultaneous measurements with two Phaethon systems at the city centre of Thessaloniki and at a rural location about 30 km away have shown that Phaethon provides NO2 and <span class="hlt">HCHO</span> tropospheric column measurements of acceptable accuracy under both low and high air-pollution levels. Currently three systems have been deployed in areas with different pollution patterns to support air quality and satellite validation studies. In the framework of the EU FP7 Monitoring and Assessment of Regional air quality in China using space Observations, Project Of Long-term sino-european co-Operation, MarcoPolo project, one of the Phaethon systems has been installed since April 2015 in the Guangzhou region in China. Tropospheric NO2 and <span class="hlt">HCHO</span> columns derived at Guangzhou during the first 10 months of operation are compared with corresponding retrievals from OMI/Aura and GOME-2/Metop-A and /Metop-B satellite sensors. The area is characterized by humid subtropical monsoon climate and cloud-free conditions are rather rare from early March to mid-October. Despite this limitation and the short period of operation of Phaethon in Guangzhou, the agreement between ground-based and satellite observations is generally good for both NO2 and <span class="hlt">HCHO</span>. It appears that GOME-2 sensors seem to underestimate the tropospheric NO2, possibly due to their large pixel size, whereas the comparison with OMI data is better, especially when a small cloud fraction (< 0.2) is used for cloud screening.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRD..122.3672L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRD..122.3672L"><span>Radical budget and ozone chemistry during autumn in the atmosphere of an urban site in central China</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lu, Xingcheng; Chen, Nan; Wang, Yuhang; Cao, Wenxiang; Zhu, Bo; Yao, Teng; Fung, Jimmy C. H.; Lau, Alexis K. H.</p> <p>2017-03-01</p> <p>The ROx (=OH + HO2 + RO2) budget and <span class="hlt">O</span><span class="hlt">3</span> production at an urban site in central China (Wuhan) during autumn were simulated and analyzed for the first time using a UW Chemical Model 0-D box model constrained by in situ observational data. The daytime average OH, HO2, and RO2 concentrations were 2.2 × 106, 1.0 × 108, and 5.2 × 107 molecules cm-<span class="hlt">3</span>, respectively. The average daytime <span class="hlt">O</span><span class="hlt">3</span> production rate was 8.8 ppbv h-1, and alkenes were the most important VOC species for <span class="hlt">O</span><span class="hlt">3</span> formation (contributing 45%) at this site. Our sensitivity test indicated that the atmospheric environment in Wuhan during autumn belongs to the VOC-limited regime. The daily average <span class="hlt">HONO</span> concentration at this site during the study period reached 1.1 ppbv and played an important role in the oxidative capacity of the atmosphere. Without the source of excess <span class="hlt">HONO</span>, the average daytime OH, HO2, RO2, and <span class="hlt">O</span><span class="hlt">3</span> production rates decreased by 36%, 26%, 27%, and 31% respectively. A correlation between the <span class="hlt">HONO</span> to NO2 heterogeneous conversion efficiency and PM2.5 × SWR was found at this site; based on this relationship, if the PM2.5 concentration met the World Health Organization air quality standard (25 µg m-<span class="hlt">3</span>), the <span class="hlt">O</span><span class="hlt">3</span> production rate in this city would decrease by 19% during late autumn. The burning of agricultural biomass severely affected the air quality in Wuhan during summer and autumn. Agricultural burning was found to account for 18% of the <span class="hlt">O</span><span class="hlt">3</span> formation during the study period. Our results suggest that VOC control and a ban on agricultural biomass burning should be considered as high-priority measures for improving the air quality in this region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.9342J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.9342J"><span>Measurements of HCl and HNO<span class="hlt">3</span> with the new research aircraft HALO - Quantification of the stratospheric contribution to the <span class="hlt">O</span><span class="hlt">3</span> and HNO<span class="hlt">3</span> budget in the UT/LS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jurkat, Tina; Kaufmann, Stefan; Voigt, Christiane; Zahn, Andreas; Schlager, Hans; Engel, Andreas; Bönisch, Harald; Dörnbrack, Andreas</p> <p>2013-04-01</p> <p>Dynamic and chemical processes modify the ozone (<span class="hlt">O</span><span class="hlt">3</span>) budget of the upper troposphere/lower stratosphere, leading to locally variable <span class="hlt">O</span><span class="hlt">3</span> trends. In this region, <span class="hlt">O</span><span class="hlt">3</span> acts as a strong greenhouse gas with a net positive radiative forcing. It has been suggested, that the correlation of the stratospheric tracer hydrochloric acid (HCl) with <span class="hlt">O</span><span class="hlt">3</span> can be used to quantify stratospheric <span class="hlt">O</span><span class="hlt">3</span> in the UT/LS region (Marcy et al., 2004). The question is, whether the stratospheric contribution to the nitric acid (HNO<span class="hlt">3</span>) budget in the UT/LS can be determined by a similar approach in order to differentiate between tropospheric and stratospheric sources of HNO<span class="hlt">3</span>. To this end, we performed in situ measurements of HCl and HNO<span class="hlt">3</span> with a newly developed Atmospheric chemical Ionization Mass Spectrometer (AIMS) during the TACTS (Transport and Composition in the UTLS) / ESMVal (Earth System Model Validation) mission in August/September 2012. The linear quadrupole mass spectrometer deployed aboard the new German research aircraft HALO was equipped with a new discharge source generating SF5- reagent ions and an in-flight calibration allowing for accurate, spatially highly resolved trace gas measurements. In addition, sulfur dioxide (SO2), nitrous acid (<span class="hlt">HONO</span>) and chlorine nitrate (ClONO2) have been simultaneously detected with the AIMS instrument. Here, we show trace gas distributions of HCl and HNO<span class="hlt">3</span> measured during a North-South transect from Northern Europe to Antarctica (68° N to 65° S) at 8 to 15 km altitude and discuss their latitude dependence. In particular, we investigate the stratospheric ozone contribution to the ozone budget in the mid-latitude UT/LS using correlations of HCl with <span class="hlt">O</span><span class="hlt">3</span>. Differences in these correlations in the subtropical and Polar regions are discussed. A similar approach is used to quantify the HNO<span class="hlt">3</span> budget of the UT/LS. We identify unpolluted atmospheric background distributions and various tropospheric HNO<span class="hlt">3</span> sources in specific regions. Our observations can be compared to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25325182','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25325182"><span>Theoretical study of the hydrogen abstraction of substituted phenols by nitrogen dioxide as a source of <span class="hlt">HONO</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shenghur, Abraham; Weber, Kevin H; Nguyen, Nhan D; Sontising, Watit; Tao, Fu-Ming</p> <p>2014-11-20</p> <p>The mild yet promiscuous reactions of nitrogen dioxide (NO2) and phenolic derivatives to produce nitrous acid (<span class="hlt">HONO</span>) have been explored with density functional theory calculations. The reaction is found to occur via four distinct pathways with both proton coupled electron transfer (PCET) and hydrogen atom transfer (HAT) mechanisms available. While the parent reaction with phenol may not be significant in the gas phase, electron donating groups in the ortho and para positions facilitate the reduction of nitrogen dioxide by electronically stabilizing the product phenoxy radical. Hydrogen bonding groups in the ortho position may additionally stabilize the nascent resonantly stabilized radical product, thus enhancing the reaction. Catechol (ortho-hydroxy phenol) has a predicted overall free energy change ΔG(0) = -0.8 kcal mol(-1) and electronic activation energy Ea = 7.0 kcal mol(-1). Free amines at the ortho and para positions have ΔG(0) = -<span class="hlt">3</span>.8 and -1.5 kcal mol(-1); Ea = 2.<span class="hlt">3</span> and 2.1 kcal mol(-1), respectively. The results indicate that the hydrogen abstraction reactions of these substituted phenols by NO2 are fast and spontaneous. Hammett constants produce a linear correlation with bond dissociation energy (BDE) demonstrating that the BDE is the main parameter controlling the dark abstraction reaction. The implications for atmospheric chemistry and ground-level nitrous acid production are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5045382','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5045382"><span>Magnolol and honokiol exert a synergistic anti-tumor effect through autophagy and apoptosis in human glioblastomas</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Cheng, Yu-Chen; Hueng, Dueng-Yuan; Huang, Hua-Yin; Chen, Jang-Yi; Chen, Ying</p> <p>2016-01-01</p> <p>Glioblastoma (GBM) is a malignant brain tumor associated with a high mortality rate. The aim of this study is to investigate the synergistic effects of honokiol (<span class="hlt">Hono</span>) and magnolol (Mag), extracted from Magnolia officinalis, on cytotoxicity and inhibition of human GBM tumor progression in cellular and animal models. In comparison with <span class="hlt">Hono</span> or Mag alone, co-treatment with <span class="hlt">Hono</span> and Mag (<span class="hlt">Hono</span>-Mag) decreased cyclin A, D1 and cyclin-dependent kinase 2, 4, 6 significantly, leading to cell cycle arrest in U87MG and LN229 human glioma cells. In addition, phosphorylated phosphoinositide <span class="hlt">3</span>-kinase (p-PI<span class="hlt">3</span>K), p-Akt, and Ki67 were decreased after <span class="hlt">Hono</span>-Mag treatment, showing proliferation inhibition. <span class="hlt">Hono</span>-Mag treatment also reduced p-p38 and p-JNK but elevated p-ERK expression. Besides, <span class="hlt">Hono</span>-Mag treatment induced autophagy and intrinsic and extrinsic apoptosis. Both ERK and autophagy inhibitors enhanced <span class="hlt">Hono</span>-Mag-induced apoptosis in LN229 cells, indicating a rescuer role of ERK. In human GBM orthotopic xenograft model, the <span class="hlt">Hono</span>-Mag treatment inhibited the tumor progression and induced apoptosis more efficiently than Temozolomide, <span class="hlt">Hono</span>, or Mag group. In conclusion, the <span class="hlt">Hono</span>-Mag exerts a synergistic anti-tumor effect by inhibiting cell proliferation and inducing autophagy and apoptosis in human GBM cells. The <span class="hlt">Hono</span>-Mag may be applied as an adjuvant therapy to improve the therapeutic efficacy of GBM treatment. PMID:27074557</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUFM.A13E..02B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUFM.A13E..02B"><span>Mexico City's active photochemistry: conclusions from the MCMA-2003 study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brune, W.; Shirley, T.; Lesher, R.; Mao, J.; Volkamer, R.; Molina, L.; Molina, M.; Velasco, E.; Westberg, H.; Lamb, B.; Jobson, T.; Alexander, M.; Gonzalez, B. C.</p> <p>2004-12-01</p> <p>Mexico City Metropolitan Area's active photochemistry was studied using an extensive suite of measurements on the CENICA environmental laboratory's roof, as part of the MCMA-2003 field study. Intense morning sunlight photolyzed <span class="hlt">HONO</span> and <span class="hlt">HCHO</span>, producing hydrogen oxides (OH and HO2) at high rates. The HOx interacted with rush-hour volatile organic compounds (VOCs) and nitrogen oxides (NOx), amplifying the production rate of ozone and nitric acid. With typically 100 ppbv of NOx and 1 ppmC of VOCs, ozone production rates exceeded 30 ppbv/hour, routinely creating in excess of 150 ppbv of ozone, even though the midday mixed layer was more than <span class="hlt">3</span> km deep. Analyses of glyoxal, a product of VOC oxidation, and the hydroperoxyl radical (HO2) indicate that MCMA's ozone production was VOC-limited during morning rush hour, when typically 1/2 of the ozone is produced, and for a significant number of days during midday and afternoon at the site. Aspects of Mexico City's active photochemistry will be compared to the observed photochemistry in U.S. urban areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA183686','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA183686"><span>Cloud Climatology Derived from the AFGWC (Air Force Global Weather Central) <span class="hlt">3</span>D-Nephanalysis for January and July 1979.</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1987-01-15</p> <p>January 1979. 29 Coo E 1979 2~0 C NHOOO. 03 to F *N9 90 E Januar y T i a J d e 1979 ClIou d 20 0 t0i 1 31 " -M No n t h I y 90 w Fig. 14 The monthly...r- 4 -Vto C UovjS <span class="hlt">3</span>H <span class="hlt">3</span>AOV 30 IJWII H913<span class="hlt">Hono</span>-4 9__ . bases. The cloud amount from the visual satellite processor is obtained by comparing the 64...satellite analyses are combined to produce a single satellite c .. 10 1n 0 a 0 c ____ ___ ____ ___ ____ ___ a__ _ a0 0 0 "w -4 <span class="hlt">O</span> *0u mC 07 <span class="hlt">o</span> v" 0 0. cr. 4" L</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A11H1968B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A11H1968B"><span>Development of a Photo-fragmentation Laser Induced Fluorescence Instrument for the Measurement of Nitrous Acid</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Boustead, G.; Blitz, M. A.; Heard, D. E.; Ingham, T.; Stone, D. J.; Whalley, L.</p> <p>2017-12-01</p> <p>The hydroxyl radical (OH) is the primary oxidant in the atmosphere with its concentration determining the lifetime of many species and its reaction with volatile organic compounds (VOCs) leading to the production of secondary organic aerosols (SOA) and tropospheric ozone. Nitrous acid (<span class="hlt">HONO</span>) is an important source of OH as it builds up overnight and is photolysed to form OH in the morning. <span class="hlt">HONO</span> is also present during the day, with several hundred pptv observed in urban environments. Atmospheric models currently under predict the concentration of <span class="hlt">HONO</span> observed during the day, indicating a missing source. Heterogeneous reactions on ground and aerosol surfaces are now recognised as potentially important <span class="hlt">HONO</span> sources. To identify and quantify some of these missing heterogeneous sources, a photo-fragmentation laser induced fluorescence (PF-LIF) instrument has been developed to provide a fast and sensitive measurement of <span class="hlt">HONO</span>. The PF-LIF instrument detects the OH fragment at 308 nm after the photolysis of <span class="hlt">HONO</span> at 355 nm. This instrument is coupled to an aerosol flow tube and can be used to measure <span class="hlt">HONO</span> production from aerosols under illuminated conditions. We will present an overview of the instrument and its development as well as preliminary results including the production of <span class="hlt">HONO</span> from illuminated Ti<span class="hlt">O</span>2 aerosols.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.4416S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.4416S"><span>Characterization of tropospheric ozone based on lidar measurement in Hangzhou, East China during the G20 Leaders' Summit</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Su, Wenjing; Liu, Cheng; Fan, Guangqiang; Hu, Qihou; Huang, Xin; Dong, Yunsheng; Zhang, Tianshu; Liu, Jianguo</p> <p>2017-04-01</p> <p>Owing to the G20 (Group of Twenty Finance Ministers and Central Bank Governors) Leaders' Summit (Sep.5th-6th, 2016), a series of strict air quality control measures were implemented in Hangzhou and its surrounding regions from Aug.26th to Sep.6th. A differential absorption lidar was employed to monitor tropospheric ozone in urban Hangzhou during a campaign from Aug. 24th to Sep. 10th, and the satellite-based NO2 VCDs and <span class="hlt">HCHO</span> VCDs in the troposphere were also retrieved using the Ozone Monitoring Instrument (OMI). During our campaign, six <span class="hlt">O</span><span class="hlt">3</span> pollution events, which were determined according to the National Ambient Air Quality Standard of China (GB-3095-2012), and two stages with rapid reduction of <span class="hlt">O</span><span class="hlt">3</span> concentration on Aug. 26th and Sep.4-6th were observed. The temporal variation tendency of <span class="hlt">O</span><span class="hlt">3</span> concentrations was well reproduced by the Weather Research and Forecasting model coupled with chemistry (WRF-Chem). Typical cases with the abrupt rise and decline of <span class="hlt">O</span><span class="hlt">3</span> concentrations were analyzed using Hybrid Single-Particle Lagrangian Integrated Trajectory (HYSPLIT) back trajectory, satellite NO2 and <span class="hlt">HCHO</span> product and the prediction by WRF-Chem model. The transport from northern cities have an important impact on pollutants observed in Hangzhou, and the chemical sensitivity of <span class="hlt">O</span><span class="hlt">3</span> production, which were approximately evaluated using the ratio of <span class="hlt">HCHO</span> VCDs to NO2 VCDs in the troposphere, was turned from a mixed VOC-NOx-limited regime into a NOX-limited regime in Hangzhou due to the strict emission control measures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.A22B..08K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.A22B..08K"><span>Satellite-observed NO2, SO2, and <span class="hlt">HCHO</span> Vertical Column Densities in East Asia: Recent Changes and Comparisons with Regional Model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, H. C.; Lee, P.; Kim, S.; Mok, J.; Yoo, H. L.; Bae, C.; Kim, B. U.; Lim, Y. K.; Woo, J. H.; Park, R.</p> <p>2015-12-01</p> <p>This study reports the recent changes in tropospheric NO2, SO2, and <span class="hlt">HCHO</span> vertical column densities (VCD) in East Asia observed from multiple satellites, highlighting especially the annual trend changes of NO2 and SO2 over Beijing-Tianjin-Hebei (BTH) region of China since 2010. Tropospheric VCD data from Global Ozone Monitoring Experiment (GOME), SCanning Imaging Absorption spectroMeter for Atmospheric CHartographY (SCIAMACHY), Ozone Monitoring Instrument (OMI) and GOME-2, retrieved from the Royal Netherlands Meteorological Institute (KNMI) and OMI National Aeronautics and Space Administration (NASA) standard products, are utilized to investigate the annual trends of NO2, SO2, and <span class="hlt">HCHO</span> VCDs from 2001 to 2015. They are also compared with simulations from Community Multi-scale Air Quality Model (CMAQ) based forecast system by the Integrated Multi-scale Air Quality System for Korea (IMAQS-K) of Ajou University. Until 2011, the changes in NO2 VCD over East Asian countries agree well with the findings of previous research, including the impact of the economic downturn during 2008-2009 and the subsequent quick recovery in China. After peaking in 2011, the NO2 VCD observations from active instruments (OMI and GOME-2) over China started to show a slower decreasing trend, mostly led by the rapid changes in the BTH region in northern China. On the other hand, SO2 started to decline earlier, from 2007, but inclined back from 2010 to 2012, and then back to declining trend since 2012. While satellite observations show dramatic recent changes, the model could not reproduce those changes mostly due to its use of fixed emission inventory. We conclude that rapid update of latest emission inventory is necessary for an accurate forecast of regional air quality in east Asia, especially for upcoming international sports events in PyeongChang (Korea), Tokyo (Japan) and Beijing (China) in 2018, 2020 and 2022, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20100021397&hterms=Atlantic+Forest&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DAtlantic%2BForest','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20100021397&hterms=Atlantic+Forest&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DAtlantic%2BForest"><span>Quantifying VOC-Reaction Tracers, Ozone Production, and Continuing Aerosol Production Rates in Urban and Far-Downwind Atmospheres</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chatfield, Robert; Ren, X.; Brune, W.; Fried, A.; Schwab, J.</p> <p>2008-01-01</p> <p>We have found a surprisingly informative decomposition of the complex question of smoggy ozone production (basically, [HO2] in a more locally determined field of [NO]) in the process of linked investigations of modestly smoggy Eastern North America (by NASA aircraft, July 2004) and rather polluted Flushing, NYC (Queens College, July, 2001). In both rural and very polluted situations, we find that a simple contour graph parameterization of the local principal ozone production rate can be estimated using only the variables [NO] and j(sub rads) [<span class="hlt">HCHO</span>]: Po(<span class="hlt">O</span><span class="hlt">3</span>) = c (j(sub rads) [<span class="hlt">HCHO</span>])(sup a) [<span class="hlt">HCHO</span>](sup b). Here j(sub rads) is the photolysis of <span class="hlt">HCHO</span> to radicals, presumably capturing many harder-UV photolytic processes and the principle ozone production is that due to HO2; mechanisms suggest that ozone production due to RO2 is closely correlated, often suggesting a limited range of different proportionality factors. The method immediately suggests a local interpretation for concepts of VOC limitation and NOx limitation. We believe that the product j(sub rads) [<span class="hlt">HCHO</span>] guages the oxidation rate of observed VOC mixtures in a way that also provides [HO2] useful for the principle ozone production rate k [HO2] [NO], and indeed, all ozone chemical production. The success of the method suggests that dominant urban primary-<span class="hlt">HCHO</span> sources may transition to secondary plume-<span class="hlt">HCHO</span> sources in a convenient way. Are there other, simple, near-terminal oxidized VOC's which help guage ozone production and aerosol particle formation? Regarding particles, we report on, to the extent NASA Research resources allow, on appealing relationships between far-downwind (Atlantic PBL) <span class="hlt">HCHO</span> and very fine aerosol (including sulfate. Since j(sub rads) [<span class="hlt">HCHO</span>] provides a time-scale, we may understand distant-plume particle production in a more quantitative manner. Additionally we report on a statistical search in the nearer field for relationships between glyoxals (important near-terminal aromatic and isoprene</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170003443','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170003443"><span>Formaldehyde Production From Isoprene Oxidation Across NOx Regimes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wolfe, G. M.; Kaiser, J.; Hanisco, T. F.; Keutsch, F. N.; de Gouw, J. A.; Gilman, J. B.; Graus, M.; Hatch, C. D.; Holloway, J.; Horowitz, L. W.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170003443'); toggleEditAbsImage('author_20170003443_show'); toggleEditAbsImage('author_20170003443_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170003443_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170003443_hide"></p> <p>2016-01-01</p> <p>The chemical link between isoprene and formaldehyde (<span class="hlt">HCHO</span>) is a strong, non-linear function of NOx (= NO + NO2). This relationship is a linchpin for top-down isoprene emission inventory verification from orbital <span class="hlt">HCHO</span> column observations. It is also a benchmark for overall photochemical mechanism performance with regard to VOC oxidation. Using a comprehensive suite of airborne in situ observations over the southeast US, we quantify <span class="hlt">HCHO</span> production across the urban-rural spectrum. Analysis of isoprene and its major first-generation oxidation products allows us to define both a prompt yield of <span class="hlt">HCHO</span> (molecules of <span class="hlt">HCHO</span> produced per molecule of freshly emitted isoprene) and the background <span class="hlt">HCHO</span> mixing ratio (from oxidation of longer-lived hydrocarbons). Over the range of observed NOx values (roughly 0.1 - 2 ppbv), the prompt yield increases by a factor of <span class="hlt">3</span> (from 0.<span class="hlt">3</span> to 0.9 ppbv ppbv(exp. -10), while background <span class="hlt">HCHO</span> increases by a factor of 2 (from 1.6 to <span class="hlt">3.3</span> ppbv). We apply the same method to evaluate the performance of both a global chemical transport model (AM<span class="hlt">3</span>) and a measurement-constrained 0-D steady-state box model. Both models reproduce the NOx dependence of the prompt <span class="hlt">HCHO</span> yield, illustrating that models with updated isoprene oxidation mechanisms can adequately capture the link between <span class="hlt">HCHO</span> and recent isoprene emissions. On the other hand, both models underestimate background <span class="hlt">HCHO</span> mixing ratios, suggesting missing <span class="hlt">HCHO</span> precursors, inadequate representation of later-generation isoprene degradation and/or underestimated hydroxyl radical concentrations. Detailed process rates from the box model simulation demonstrate a <span class="hlt">3</span>-fold increase in <span class="hlt">HCHO</span> production across the range of observed NOx values, driven by a 100% increase in OH and a 40% increase in branching of organic peroxy radical reactions to produce <span class="hlt">HCHO</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5879783','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5879783"><span>Formaldehyde production from isoprene oxidation across NOx regimes</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wolfe, G. M.; Kaiser, J.; Hanisco, T. F.; Keutsch, F. N.; de Gouw, J. A.; Gilman, J. B.; Graus, M.; Hatch, C. D.; Holloway, J.; Horowitz, L. W.; Lee, B. H.; Lerner, B. M.; Lopez-Hilifiker, F.; Mao, J.; Marvin, M. R.; Peischl, J.; Pollack, I. B.; Roberts, J. M.; Ryerson, T. B.; Thornton, J. A.; Veres, P. R.; Warneke, C.</p> <p>2018-01-01</p> <p>The chemical link between isoprene and formaldehyde (<span class="hlt">HCHO</span>) is a strong, non-linear function of NOx (= NO + NO2). This relationship is a linchpin for top-down isoprene emission inventory verification from orbital <span class="hlt">HCHO</span> column observations. It is also a benchmark for overall photochemical mechanism performance with regard to VOC oxidation. Using a comprehensive suite of airborne in situ observations over the Southeast U.S., we quantify <span class="hlt">HCHO</span> production across the urban-rural spectrum. Analysis of isoprene and its major first-generation oxidation products allows us to define both a “prompt” yield of <span class="hlt">HCHO</span> (molecules of <span class="hlt">HCHO</span> produced per molecule of freshly-emitted isoprene) and the background <span class="hlt">HCHO</span> mixing ratio (from oxidation of longer-lived hydrocarbons). Over the range of observed NOx values (roughly 0.1 – 2 ppbv), the prompt yield increases by a factor of <span class="hlt">3</span> (from 0.<span class="hlt">3</span> to 0.9 ppbv ppbv−1), while background <span class="hlt">HCHO</span> increases by a factor of 2 (from 1.6 to <span class="hlt">3.3</span> ppbv). We apply the same method to evaluate the performance of both a global chemical transport model (AM<span class="hlt">3</span>) and a measurement-constrained 0-D steady state box model. Both models reproduce the NOx dependence of the prompt <span class="hlt">HCHO</span> yield, illustrating that models with updated isoprene oxidation mechanisms can adequately capture the link between <span class="hlt">HCHO</span> and recent isoprene emissions. On the other hand, both models under-estimate background <span class="hlt">HCHO</span> mixing ratios, suggesting missing <span class="hlt">HCHO</span> precursors, inadequate representation of later-generation isoprene degradation and/or under-estimated hydroxyl radical concentrations. Detailed process rates from the box model simulation demonstrate a <span class="hlt">3</span>-fold increase in <span class="hlt">HCHO</span> production across the range of observed NOx values, driven by a 100% increase in OH and a 40% increase in branching of organic peroxy radical reactions to produce <span class="hlt">HCHO</span>. PMID:29619046</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29619046','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29619046"><span>Formaldehyde production from isoprene oxidation across NOx regimes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wolfe, G M; Kaiser, J; Hanisco, T F; Keutsch, F N; de Gouw, J A; Gilman, J B; Graus, M; Hatch, C D; Holloway, J; Horowitz, L W; Lee, B H; Lerner, B M; Lopez-Hilifiker, F; Mao, J; Marvin, M R; Peischl, J; Pollack, I B; Roberts, J M; Ryerson, T B; Thornton, J A; Veres, P R; Warneke, C</p> <p>2016-01-01</p> <p>The chemical link between isoprene and formaldehyde (<span class="hlt">HCHO</span>) is a strong, non-linear function of NO x (= NO + NO 2 ). This relationship is a linchpin for top-down isoprene emission inventory verification from orbital <span class="hlt">HCHO</span> column observations. It is also a benchmark for overall photochemical mechanism performance with regard to VOC oxidation. Using a comprehensive suite of airborne in situ observations over the Southeast U.S., we quantify <span class="hlt">HCHO</span> production across the urban-rural spectrum. Analysis of isoprene and its major first-generation oxidation products allows us to define both a "prompt" yield of <span class="hlt">HCHO</span> (molecules of <span class="hlt">HCHO</span> produced per molecule of freshly-emitted isoprene) and the background <span class="hlt">HCHO</span> mixing ratio (from oxidation of longer-lived hydrocarbons). Over the range of observed NO x values (roughly 0.1 - 2 ppbv), the prompt yield increases by a factor of <span class="hlt">3</span> (from 0.<span class="hlt">3</span> to 0.9 ppbv ppbv -1 ), while background <span class="hlt">HCHO</span> increases by a factor of 2 (from 1.6 to <span class="hlt">3.3</span> ppbv). We apply the same method to evaluate the performance of both a global chemical transport model (AM<span class="hlt">3</span>) and a measurement-constrained 0-D steady state box model. Both models reproduce the NO x dependence of the prompt <span class="hlt">HCHO</span> yield, illustrating that models with updated isoprene oxidation mechanisms can adequately capture the link between <span class="hlt">HCHO</span> and recent isoprene emissions. On the other hand, both models under-estimate background <span class="hlt">HCHO</span> mixing ratios, suggesting missing <span class="hlt">HCHO</span> precursors, inadequate representation of later-generation isoprene degradation and/or under-estimated hydroxyl radical concentrations. Detailed process rates from the box model simulation demonstrate a <span class="hlt">3</span>-fold increase in <span class="hlt">HCHO</span> production across the range of observed NO x values, driven by a 100% increase in OH and a 40% increase in branching of organic peroxy radical reactions to produce <span class="hlt">HCHO</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A51I..05V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A51I..05V"><span>The Second Cabauw Intercomparison of Nitrogen Dioxide Measuring Instruments (CINDI-2)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Van Roozendael, M.; Hendrick, F.; Apituley, A.; Kreher, K.; Friess, U.; Richter, A.; Wagner, T.; Fehr, T.</p> <p>2017-12-01</p> <p>For the validation of space borne UV-Vis observations of air quality gases, ground based remote-sensing instruments using the MAXDOAS technique are an excellent choice, since they rely on similar retrieval techniques as the observations from orbit. Over the last decade, MAXDOAS instruments of various designs (including PANDORA systems) have been deployed worldwide forming the basis for a global ground based reference network suitable for the validation of future satellite sensors, such as TROPOMI/Sentinel-5 precursor, GEMS, TEMPO, and Sentinel 4 and 5. To ensure proper traceability of these observations, assess their accuracy and progress towards harmonized data acquisition and delivery, a thorough intercomparison campaign known as the Second Cabauw Intercomparison of Nitrogen Dioxide Measuring Instruments (CINDI-2) was held in Cabauw, The Netherlands during the month of September 2016. About 35 MAXDOAS instruments operated by 25 different groups were deployed, together with systems providing key ancillary in-situ measurements of NO2 and aerosol optical properties, as well as vertical profiles of NO2 by lidar and sonde and vertical profiles of aerosol optical properties by Raman lidar. We provide an overview of the main outcome of the campaign, which included a formal semi-blind slant column intercomparison and a number of additional exercises aiming at assessing the potential of the MAXDOAS technique for retrieving vertically-resolved information on NO2, aerosol, <span class="hlt">HCHO</span>, <span class="hlt">O</span><span class="hlt">3</span> and a few other more challenging species such as <span class="hlt">HONO</span> and glyoxal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JaJAP..54aAE04N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JaJAP..54aAE04N"><span>Experimental study of the visible-light photocatalytic activity of oxygen-deficient Ti<span class="hlt">O</span>2 prepared with Ar/H2 plasma surface treatment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakano, Takuma; Yazawa, Shota; Araki, Shota; Kogoshi, Sumio; Katayama, Noboru; Kudo, Yusuke; Nakanishi, Tetsuya</p> <p>2015-01-01</p> <p>Oxygen-deficient Ti<span class="hlt">O</span>2 (Ti<span class="hlt">O</span>2-x) has been proposed as a visible-light-responsive photocatalyst. Ti<span class="hlt">O</span>2-x thin films were prepared by Ar/H2 plasma surface treatment, applying varying levels of microwave input power and processing times. The highest visible light photocatalytic activity was observed when using an input power of 200 W, a plasma processing time of 10 min, and a 1:1 \\text{Ar}:\\text{H}2 ratio, conditions that generate an electron temperature of 5.7(±1.0) eV and an electron density of 8.5 × 1010 cm-<span class="hlt">3</span>. The maximum formaldehyde (<span class="hlt">HCHO</span>) removal rate of the Ti<span class="hlt">O</span>2-x film was 2.6 times higher than that obtained from a Ti<span class="hlt">O</span>2-xNx film under the same test conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1208741-impact-isoprene-hono-chemistry-ozone-ovoc-formation-semirural-south-korean-forest','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1208741-impact-isoprene-hono-chemistry-ozone-ovoc-formation-semirural-south-korean-forest"><span>Impact of isoprene and <span class="hlt">HONO</span> chemistry on ozone and OVOC formation in a semirural South Korean forest</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kim, Saewung; Kim, So-Young; Lee, Meehye</p> <p></p> <p>Rapid urbanization and economic development in East Asia in past decades has led to photochemical air pollution problems such as excess photochemical ozone and aerosol formation. Asian megacities such as Seoul, Tokyo, Shanghai, Gangzhou, and Beijing are surrounded by densely forested areas and recent research has consistently demonstrated the importance of biogenic volatile organic compounds from vegetation in determining oxidation capacity in the suburban Asian megacity regions. Uncertainties in constraining tropospheric oxidation capacity, dominated by hydroxyl radical concentrations, undermine our ability to assess regional photochemical air pollution problems. We present an observational dataset of CO, NOX, SO2, ozone, <span class="hlt">HONO</span>, andmore » VOCs (anthropogenic and biogenic) from Taehwa Research Forest (TRF) near the Seoul Metropolitan Area (SMA) in early June 2012. The data show that TRF is influenced both by aged pollution and fresh BVOC emissions. With the dataset, we diagnose HOx (OH, HO2, and RO2) distributions calculated with the University of Washington Chemical Box Model (UWCM v 2.1). Uncertainty from unconstrained <span class="hlt">HONO</span> sources and radical recycling processes highlighted in recent studies is examined using multiple model simulations with different model constraints. The results suggest that 1) different model simulation scenarios cause systematic differences in HOX distributions especially OH levels (up to 2.5 times) and 2) radical destruction (HO2+HO2 or HO2+RO2) could be more efficient than radical recycling (HO2+NO) especially in the afternoon. Implications of the uncertainties in radical chemistry are discussed with respect to ozone-VOC-NOX sensitivity and oxidation product formation rates. Overall, the VOC limited regime in ozone photochemistry is predicted but the degree of sensitivity can significantly vary depending on the model scenarios. The model results also suggest that RO2 levels are positively correlated with OVOCs production that is not</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/53199','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/53199"><span>Ground-level air pollution changes during a boreal wildland mega-fire</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Andrzej Bytnerowicz; Yu-Mei Hsu; Kevin Percy; Allan Legge; Mark E. Fenn; Susan Schilling; Witold Frączek; Diane Alexander</p> <p>2016-01-01</p> <p>The 2011 Richardson wildland mega-fire in the Athabasca Oil Sands Region (AOSR) in northern Alberta, Canada had large effects on air quality. At a receptor site in the center of the AOSR ambient PM2.5, <span class="hlt">O</span><span class="hlt">3</span>, NO, NO2, SO2, NH<span class="hlt">3</span>, <span class="hlt">HONO</span>, HNO<span class="hlt">3</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A43E2506C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A43E2506C"><span>Tracking nitrogen oxides, nitrous acid, and nitric acid from biomass burning</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chai, J.; Miller, D. J.; Scheuer, E. M.; Dibb, J. E.; Hastings, M. G.</p> <p>2017-12-01</p> <p>Biomass burning emissions are an important source of atmospheric nitrogen oxides (NOx = NO + NO2) and nitrous acid (<span class="hlt">HONO</span>), which play important roles in atmosphere oxidation capacity (hydroxyl radical and ozone formation) and have severe impacts on air quality and climate in the presence of sunlight and volatile organic compounds. However, tracking NOx and <span class="hlt">HONO</span> and their chemistry in the atmosphere based on concentration alone is challenging. Isotopic analysis provides a potential tracking tool. In this study, we measured the nitrogen isotopic composition (δ15N) of NOx (NO + NO2) and <span class="hlt">HONO</span>, and soluble <span class="hlt">HONO</span> and HNO<span class="hlt">3</span> during the Fire Influence on Regional and Global Environments Experiment (FIREX) laboratory experiments at the Missoula Fire Laboratory. Our newly developed and validated annular denuder system (ADS) enabled us to effectively trap <span class="hlt">HONO</span> prior to a NOx collection system in series for isotopic analysis. In total we investigated 25 "stack" fires of various biomass materials where the emissions were measured within a few seconds of production by the fire. <span class="hlt">HONO</span> concentration was measured in parallel using mist chamber/ion chromatography (MC/IC). The recovered mean <span class="hlt">HONO</span> concentrations from ADS during the burn of each fire agree well with that measured via MC/IC. δ15N-NOx ranged from -4.<span class="hlt">3</span> ‰ to + 7.0 ‰ with a median of 0.7 ‰. Combined with a similar, recent study by our group [Fibiger et al., ES&T, 2017] the δ15N-NOx follows a linear relationship with δ15N-biomass (δ15N-NOx =0.94 x δ15N-biomass +1.98; R2=0.72). δ15N-<span class="hlt">HONO</span> ranged from -5.<span class="hlt">3</span> to +8.<span class="hlt">3</span> ‰ with a median of 1.4 ‰. While both <span class="hlt">HONO</span> and NOx are sourced from N in the biomass fuel, the secondary formation of <span class="hlt">HONO</span> likely induces fractionation of the N that leads to the difference between δ15N-NOx and δ15N-<span class="hlt">HONO</span>. We found a correlation of δ15N-<span class="hlt">HONO</span>= 0.86 x δ15N-NOx + 0.52 (R2=0.55), which can potentially be used to track the chemistry of <span class="hlt">HONO</span> formation following fire emissions. The methods</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150021531','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150021531"><span>Formaldehyde Production from Isoprene Oxidation Across NOx Regimes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wolfe, G. M.; Kaiser, J.; Hanisco, T. F.; Keutsch, F. N.; de Gouw, J. A.; Gilman, J. B.; Graus, M.; Hatch, C. D.; Holloway, J.; Horowitz, L. W.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20150021531'); toggleEditAbsImage('author_20150021531_show'); toggleEditAbsImage('author_20150021531_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20150021531_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20150021531_hide"></p> <p>2015-01-01</p> <p>The chemical link between isoprene and formaldehyde (<span class="hlt">HCHO</span>) is a strong, non-linear function of NOx (= 27 NO + NO2). This relationship is a linchpin for top-down isoprene emission inventory verification from orbital <span class="hlt">HCHO</span> column observations. It is also a benchmark for overall mechanism performance with regard to VOC oxidation. Using a comprehensive suite of airborne in situ observations over the Southeast U.S., we quantify <span class="hlt">HCHO</span> production across the urban-rural spectrum. Analysis of isoprene and its major first-generation oxidation products allows us to define both a "prompt" yield of <span class="hlt">HCHO</span> (molecules of <span class="hlt">HCHO</span> produced per molecule of freshly-emitted isoprene) and the background <span class="hlt">HCHO</span> mixing ratio (from oxidation of longer-lived hydrocarbons). Over the range of observed NOx values (roughly 0.1 - 2 ppbv), the prompt yield increases by a factor of <span class="hlt">3</span> (from 0.<span class="hlt">3</span> to 0.9), while background <span class="hlt">HCHO</span> increases by more than a factor of 2 (from 1.5 to <span class="hlt">3.3</span> ppbv). We apply the same method to evaluate the performance of both a global chemical transport model (AM<span class="hlt">3</span>) and a measurement-constrained 0-D chemical box model. Both models reproduce the NOx dependence of the prompt <span class="hlt">HCHO</span> yield, illustrating that models with updated isoprene oxidation mechanisms can adequately capture the link between <span class="hlt">HCHO</span> and recent isoprene emissions. On the other hand, both models under-estimate background <span class="hlt">HCHO</span> mixing ratios, suggesting missing <span class="hlt">HCHO</span> precursors, inadequate representation of later-generation isoprene degradation and/or under-estimated hydroxyl radical concentrations. Moreover, we find that the total organic peroxy radical production rate is essentially independent of NOx, as the increase in oxidizing capacity with NOx is largely balanced by a decrease in VOC reactivity. Thus, the observed NOx dependence of <span class="hlt">HCHO</span> mainly reflects the changing fate of organic peroxy radicals.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.7521G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.7521G"><span>Aircraft-borne DOAS limb observations of iodine monoxide around Borneo</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Großmann, Katja; Hossaini, Ryan; Mantle, Hannah; Chipperfield, Martyn; Wittrock, Folkard; Peters, Enno; Lampel, Johannes; Walker, Hannah; Heard, Dwayne; Krystofiak, Gisèle; Catoire, Valéry; Dorf, Marcel; Werner, Bodo; Pfeilsticker, Klaus</p> <p>2015-04-01</p> <p>Iodine monoxide (IO) has a major impact on the photochemistry of the troposphere. It can for example catalytically destroy ozone, influence the atmospheric oxidation capacity by changing the partitioning of the HOx and NOx species, or contribute to the formation of ultrafine particles. Information regarding the vertical distribution of IO is still sparse since only few vertical profiles of IO exist for the troposphere. Spectroscopic measurements were carried out from aboard the research aircraft DLR-Falcon during the SHIVA (Stratospheric ozone: Halogen Impacts in a Varying Atmosphere) campaign at Malaysian Borneo in November and December 2011 to study the abundance and transport of trace gases in the lower atmosphere. Sixteen research flights were performed covering legs near the surface in the marine boundary layer (MBL) as well as in the free troposphere (FT) up to an altitude of 13 km. The spectroscopic measurements were evaluated using the Differential Optical Absorption Spectroscopy (DOAS) technique in limb geometry, which supports observations of UV/visible absorbing trace gases, such as <span class="hlt">O</span>4, Br<span class="hlt">O</span>, IO, NO2, <span class="hlt">HCHO</span>, CHOCHO, <span class="hlt">HONO</span> and H2<span class="hlt">O</span>, and altitude information was gained via the <span class="hlt">O</span>4 scaling technique and/or full inversion. The inferred vertical profiles of IO showed mixing ratios of 0.5-1.5 ppt in the MBL, which decreased to 0.1-0.<span class="hlt">3</span> ppt in the FT. Occasionally, the IO observed in the FT of the marine environment coincided with elevated amounts of CO, but no IO was observed over land, neither in the boundary layer, nor in the FT. This behavior strongly indicated that the major sources for IO were organic and inorganic precursor molecules emitted from the ocean, which during daytime rapidly formed a sizable amount of IO in the MBL that was occasionally transported into the FT where efficient loss processes for IO must exist. The inferred vertical profiles of IO are compared to simulations using the global <span class="hlt">3</span>-D chemistry transport model TOMCAT including recent fluxes</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21243897','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21243897"><span>Performance evaluation of an in situ nitrous acid measurement system and continuous measurement of nitrous acid in an indoor environment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Park, Seung Shik; Cho, Sung Yong</p> <p>2010-12-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) may cause adverse effects to mucous membranes and lung function when people are exposed to higher <span class="hlt">HONO</span> concentrations than those present in typical indoor residential environments. Therefore, determination of <span class="hlt">HONO</span> concentration in indoor environments is required to investigate occurrences of high <span class="hlt">HONO</span> levels. In this work, a high-time-resolution measurement system was utilized to better understand the levels and dynamic behavior of <span class="hlt">HONO</span> in an indoor environment. The performance of the in situ <span class="hlt">HONO</span> analyzer applied to this work was evaluated using a 12-hr integrated annular denuder technique under ambient conditions. Both methods for the measurements of <span class="hlt">HONO</span> were in good agreement, with a regression slope of 0.84, an intercept of 0.09, and correlation coefficient (r2) of 0.67. Indoor <span class="hlt">HONO</span> and nitrogen oxide concentrations were also observed for approximately 5 days in winter in the living room of an apartment that had a gas range for cooking in the kitchen. Investigation of the relationships among nitric oxide (NO), nitrite (NO2), and <span class="hlt">HONO</span> concentrations suggests that <span class="hlt">HONO</span> production during combustion could be the result of direct emission, whereas the heterogeneous NO2 chemistry during the background period and after combustion was the possible pathway of <span class="hlt">HONO</span> production. Controlled combustion experiments, performed at a burning rate of 50% valve setting, show peak <span class="hlt">HONO</span> concentrations during the unvented combustion to be approximately 8-10 times higher than background levels depending on the time of day. At a burning rate setting of 50%, the peak concentration of <span class="hlt">HONO</span> during unvented combustion was found to be 33-37% higher than those from "weak" (airflow = 340 m<span class="hlt">3</span>/hr) and "strong" (airflow = 540 m<span class="hlt">3</span>/hr) vented combustions. The decay rate of the <span class="hlt">HONO</span> concentrations for the unvented combustion conditions was approximately 2-fold higher in the daytime than in the nighttime and significantly less than those of NO and NO2.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12141490','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12141490"><span>Biotransformation of hexahydro-1,<span class="hlt">3</span>,5-trinitro-1,<span class="hlt">3</span>,5-tiazine catalyzed by a NAD(P)H: nitrate oxidoreductase from Aspergillus niger.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bhushan, Bharat; Halasz, Annamaria; Spain, Jim; Thiboutot, Sonia; Ampleman, Guy; Hawari, Jalal</p> <p>2002-07-15</p> <p>Hexahydro-1,<span class="hlt">3</span>,5-trinitro-1,<span class="hlt">3</span>,5-triazine (RDX) can be efficiently mineralized with anaerobic domestic sludge, but the initial enzymatic processes involved in its transformation are unknown. To test the hypothesis that the initial reaction involves reduction of nitro group(s), we designed experiments to test the ability of a nitrate reductase (EC 1.6.6.2) to catalyze the initial reaction leading to ring cleavage and subsequent decomposition. A nitrate reductase from Aspergillus niger catalyzed the biotransformation of RDX most effectively at pH 7.0 and 30 degrees C under anaerobic conditions using NADPH as electron donor. LC/MS (ES-) chromatograms showed the formation of hexahydro-1-nitroso-<span class="hlt">3</span>,5-dinitro-1,<span class="hlt">3</span>,5-triazine (MNX) and methylenedinitramine as key initial products of RDX, but neither the dinitroso neither (DNX) nor trinitroso (TNX) derivatives were observed. None of the above detected products persisted, and their disappearance was accompanied by the accumulation of nitrous oxide (N20), formaldehyde (<span class="hlt">HCHO</span>), and ammonium ion (NH4+). Stoichiometric studies showed that three NADPH molecules were consumed, and one molecule of methylenedinitramine was produced per RDX molecule. The carbon and nitrogen mass balances were 96.14% and 82.10%, respectively. The stoichiometries and mass balance measurements supported a mechanism involving initial transformation of RDX to MNX via a two-electron reduction mechanism. Subsequent reduction of MNX followed by rapid ring cleavage gave methylenedinitramine which in turn decomposed in water to produce quantitatively N2<span class="hlt">O</span> and <span class="hlt">HCHO</span>. The results clearly indicate that an initial reduction of a nitro group by nitrate reductase is sufficient for the decomposition of RDX.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RaPC..144..419S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RaPC..144..419S"><span>Investigation of radiation shielding properties for Me<span class="hlt">O</span>-PbCl2-Te<span class="hlt">O</span>2 (Me<span class="hlt">O</span> = Bi2<span class="hlt">O</span><span class="hlt">3</span>, Mo<span class="hlt">O</span><span class="hlt">3</span>, Sb2<span class="hlt">O</span><span class="hlt">3</span>, WO<span class="hlt">3</span>, Zn<span class="hlt">O</span>) glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sayyed, M. I.; Çelikbilek Ersundu, M.; Ersundu, A. E.; Lakshminarayana, G.; Kostka, P.</p> <p>2018-03-01</p> <p>In this work, glasses in the Me<span class="hlt">O</span>-PbCl2-Te<span class="hlt">O</span>2 (Me<span class="hlt">O</span> = Bi2<span class="hlt">O</span><span class="hlt">3</span>, Mo<span class="hlt">O</span><span class="hlt">3</span>, Sb2<span class="hlt">O</span><span class="hlt">3</span>, WO<span class="hlt">3</span>, Zn<span class="hlt">O</span>) system, which show a great potential for optoelectronic applications, were used to evaluate their resistance under high energy ionizing radiations. The basic shielding quantities for determining the penetration of radiation in glass, such as mass attenuation coefficient (μ/ρ), half value layer (HVL), mean free path (MFP) and exposure buildup factor (EBF) values were investigated within the energy range 0.015 MeV ‒ 15 MeV using XCOM program and variation of shielding parameters were compared with different glass systems and ordinary concrete. From the derived results, it was determined that Me<span class="hlt">O</span>-PbCl2-Te<span class="hlt">O</span>2 (Me<span class="hlt">O</span> = Bi2<span class="hlt">O</span><span class="hlt">3</span>, Mo<span class="hlt">O</span><span class="hlt">3</span>, Sb2<span class="hlt">O</span><span class="hlt">3</span>, WO<span class="hlt">3</span>, Zn<span class="hlt">O</span>) glasses show great potentiality to be used under high energy radiations. Among the studied glass compositions, Bi2<span class="hlt">O</span><span class="hlt">3</span> and WO<span class="hlt">3</span> containing glasses were found to possess superior gamma-ray shielding effectiveness.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1816768H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1816768H"><span>Assessment of the Impact of The East Asian Summer Monsoon on the Air Quality Over China</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hao, Nan; Ding, Aijun; Safieddine, Sarah; Valks, Pieter; Clerbaux, Cathy; Trautmann, Thomas</p> <p>2016-04-01</p> <p>Air pollution is one of the most important environmental problems in developing Asian countries like China. In this region, studies showed that the East Asian monsoon plays a significant role in characterizing the temporal variation and spatial patterns of air pollution, since monsoon is a major atmospheric system affecting air mass transport, convection, and precipitation. Knowledge gaps still exist in the understanding of Asian monsoon impact on the air quality in China under the background of global climate change. For the first time satellite observations of tropospheric ozone and its precursors will be integrated with the ground-based, aircraft measurements of air pollutants and model simulations to study the impact of the East Asian monsoon on air quality in China. We apply multi-platform satellite observations by the GOME-2, IASI, and MOPITT instruments to analyze tropospheric ozone and CO, precursors of ozone (NO2, <span class="hlt">HCHO</span> and CHOCHO) and other related trace gases over China. Two years measurements of air pollutants including NO2, <span class="hlt">HONO</span>, SO2, <span class="hlt">HCHO</span> and CHOCHO at a regional back-ground site in the western part of the Yangtze River Delta (YRD) in eastern China will be presented. The potential of using the current generation of satellite instruments, ground-based instruments and aircraft to monitor air quality changes caused by the East Asian monsoon circulation will be presented. Preliminary comparison results between satellite measurement and limited but valuable ground-based and aircraft measurements will also be showed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23057084','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23057084"><span>Photochemistry of formaldoxime−nitrous acid complexes in an argon matrix: identification of formaldoxime nitrite.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Golec, Barbara; Bil, Andrzej; Mielke, Zofia</p> <p>2009-08-27</p> <p>We have studied the structure and photochemistry of the formaldoxime−nitrous acid system (CH2NOH−<span class="hlt">HONO</span>) by help of FTIR matrix isolation spectroscopy and ab initio methods. The MP2/6-311++G(2d,2p) calculations show stability of six isomeric CH2NOH···<span class="hlt">HONO</span> complexes. The FTIR spectra evidence formation of two hydrogen bonded complexes in an argon matrix whose structures are determined by comparison of the experimental spectra with the calculated ones for the six stable complexes. In the matrix there is present the most stable cyclic complex with two O−H···N bonds; a strong bond is formed between the OH group of <span class="hlt">HONO</span> and the N atom of CH2NOH and the weaker one between the OH group of CH2NOH and the N atom of <span class="hlt">HONO</span>. In the other complex present in the matrix the OH group of formaldoxime is attached to the OH group of <span class="hlt">HONO</span> forming an O−H···<span class="hlt">O</span> bond. The irradiation of the CH2NOH···<span class="hlt">HONO</span> complexes with the filtered output of the mercury lamp (λ > 345 nm) leads to the formation of formaldoxime nitrite, CH2NONO, and its two isomeric complexes with water. The main product is the CH2NONO···H2<span class="hlt">O</span> complex in which water is hydrogen bonded to the N atom of the C═N group. The identity of the photoproducts is confirmed by both FTIR spectroscopy and MP2 or QCISD(full) calculations with the 6-311++G(2d,2p) basis set. The intermediate in this reaction is iminoxyl radical that is formed by abstraction of hydrogen atom from formaldoxime OH group by an OH radical originating from <span class="hlt">HONO</span> photolysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28688374','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28688374"><span>Luminescence and energy transfer of Tb<span class="hlt">3</span>+-doped Ba<span class="hlt">O</span>-Gd2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zuo, Chenggang; Huang, Jinze; Liu, Shaoyou; Xiao, Anguo; Shen, Youming; Zhang, Xiangyang; Zhou, Zhihua; Zhu, Ligang</p> <p>2017-12-05</p> <p>Transparent Tb <span class="hlt">3</span>+ -doped Ba<span class="hlt">O</span>-Gd 2 <span class="hlt">O</span> <span class="hlt">3</span> -Al 2 <span class="hlt">O</span> <span class="hlt">3</span> -B 2 <span class="hlt">O</span> <span class="hlt">3</span> -Si<span class="hlt">O</span> 2 glasses with the greater than 4g/cm <span class="hlt">3</span> were prepared by high temperature melting method and its luminescent properties have been investigated by measured UV-vis transmission, excitation, emission and luminescence decay spectra. The transmission spectrum shows there are three weak absorption bands locate at about 312, 378 and 484nm in the glasses and it has good transmittance in the visible spectrum region. Intense green emission can be observed under UV excitation. The effective energy transfer from Gd <span class="hlt">3</span>+ ion to Tb <span class="hlt">3</span>+ ion could occur and sensitize the luminescence of Tb <span class="hlt">3</span>+ ion. The green emission intensity of Tb <span class="hlt">3</span>+ ion could change with the increasing Si<span class="hlt">O</span> 2 /B 2 <span class="hlt">O</span> <span class="hlt">3</span> ratio in the borosilicate glass matrix. With the increasing concentration of Tb <span class="hlt">3</span>+ ion, 5 D 4 → 7 F J transitions could be enhanced through the cross relaxation between the two nearby Tb <span class="hlt">3</span>+ ions. Luminescence decay time of 2.12ms from 546nm emission is obtained. The results indicate that Tb <span class="hlt">3</span>+ -doped Ba<span class="hlt">O</span>-Gd 2 <span class="hlt">O</span> <span class="hlt">3</span> -Al 2 <span class="hlt">O</span> <span class="hlt">3</span> -B 2 <span class="hlt">O</span> <span class="hlt">3</span> -Si<span class="hlt">O</span> 2 glasses would be potential scintillating material for applications in X-ray imaging. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AtmEn.106..382W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AtmEn.106..382W"><span>Volatile organic compound conversion by ozone, hydroxyl radicals, and nitrate radicals in residential indoor air: Magnitudes and impacts of oxidant sources</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Waring, Michael S.; Wells, J. Raymond</p> <p>2015-04-01</p> <p>Indoor chemistry may be initiated by reactions of ozone (<span class="hlt">O</span><span class="hlt">3</span>), the hydroxyl radical (OH), or the nitrate radical (NO<span class="hlt">3</span>) with volatile organic compounds (VOC). The principal indoor source of <span class="hlt">O</span><span class="hlt">3</span> is air exchange, while OH and NO<span class="hlt">3</span> formation are considered as primarily from <span class="hlt">O</span><span class="hlt">3</span> reactions with alkenes and nitrogen dioxide (NO2), respectively. Herein, we used time-averaged models for residences to predict <span class="hlt">O</span><span class="hlt">3</span>, OH, and NO<span class="hlt">3</span> concentrations and their impacts on conversion of typical residential VOC profiles, within a Monte Carlo framework that varied inputs probabilistically. We accounted for established oxidant sources, as well as explored the importance of two newly realized indoor sources: (i) the photolysis of nitrous acid (<span class="hlt">HONO</span>) indoors to generate OH and (ii) the reaction of stabilized Criegee intermediates (SCI) with NO2 to generate NO<span class="hlt">3</span>. We found total VOC conversion to be dominated by reactions both with <span class="hlt">O</span><span class="hlt">3</span>, which almost solely reacted with D-limonene, and also with OH, which reacted with D-limonene, other terpenes, alcohols, aldehydes, and aromatics. VOC oxidation rates increased with air exchange, outdoor <span class="hlt">O</span><span class="hlt">3</span>, NO2 and D-limonene sources, and indoor photolysis rates; and they decreased with <span class="hlt">O</span><span class="hlt">3</span> deposition and nitric oxide (NO) sources. Photolysis was a strong OH formation mechanism for high NO, NO2, and <span class="hlt">HONO</span> settings, but SCI/NO2 reactions weakly generated NO<span class="hlt">3</span> except for only a few cases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AtmEn.164...61W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AtmEn.164...61W"><span>The OH-initiated oxidation of atmospheric peroxyacetic acid: Experimental and model studies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Huihui; Wang, Yin; Li, Huan; Huang, Liubin; Huang, Dao; Shen, Hengqing; Xing, Yanan; Chen, Zhongming</p> <p>2017-09-01</p> <p>Peroxyacetic acid (PAA, CH<span class="hlt">3</span>C(<span class="hlt">O</span>)OOH) plays an important role in atmospheric chemistry, serving as reactive oxidant and affecting radical recycling. However, previous studies revealed an obvious gap between modelled and observed concentrations of atmospheric PAA, which may be partly ascribed to the uncertainty in the kinetics and mechanism of OH-oxidation. In this study, we measured the rate constant of OH radical reaction with PAA (kPAA+OH) and investigated the products in order to develop a more robust atmospheric PAA chemistry. Using the relative rates technique and employing toluene and meta-xylene as reference compounds, the kPAA+OH was determined to be (9.4-11.9) × 10-12 cm<span class="hlt">3</span> molecule-1 s-1 at 298 K and 1 atm, which is about (2.5-<span class="hlt">3</span>.2) times larger than that parameter used in Master Chemical Mechanism v<span class="hlt">3.3</span>.1 (MCM v<span class="hlt">3.3</span>.1) (<span class="hlt">3</span>.70 × 10-12 cm<span class="hlt">3</span> molecule-1 s-1). Incorporation of a box model and MCM v<span class="hlt">3.3</span>.1 with revised PAA chemistry represented a better simulation of atmospheric PAA observed during Wangdu Campaign 2014, a rural site in North China Plain. It is found that OH-oxidation is an important sink of atmospheric PAA in this rural area, accounting for ∼30% of the total loss. Moreover, the major terminal products of PAA-OH reaction were identified as formaldehyde (<span class="hlt">HCHO</span>) and formic acid (HC(<span class="hlt">O</span>)OH). The modelled results show that both primary and secondary chemistry play an important role in the large <span class="hlt">HCHO</span> and HC(<span class="hlt">O</span>)OH formation under experimental conditions. There should exist the channel of methyl H-abstraction for PAA-OH reaction, which may also provide routes to <span class="hlt">HCHO</span> and HC(<span class="hlt">O</span>)OH formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/837401','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/837401"><span>Selective oxidation of methanol and ethanol on supported ruthenium oxide clusters at low temperatures</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Liu, Haichao; Iglesia, Enrique</p> <p></p> <p>Ru<span class="hlt">O</span>2 domains supported on Sn<span class="hlt">O</span>2, Zr<span class="hlt">O</span>2, Ti<span class="hlt">O</span>2, Al2<span class="hlt">O</span><span class="hlt">3</span>, and Si<span class="hlt">O</span>2 catalyze the oxidative conversion of methanol to formaldehyde, methylformate, and dimethoxymethane with unprecedented rates and high combined selectivity (>99 percent) and yield at low temperatures (300-400 K). Supports influence turnover rates and the ability of Ru<span class="hlt">O</span>2 domains to undergo redox cycles required for oxidation turnovers. Oxidative dehydrogenation turnover rates and rates of stoichiometric reduction of Ru<span class="hlt">O</span>2 in H2 increased in parallel when Ru<span class="hlt">O</span>2 domains were dispersed on more reducible supports. These support effects, the kinetic effects of CH<span class="hlt">3</span>OH and <span class="hlt">O</span>2 on reaction rates, and the observed kinetic isotope effects withmore » CH<span class="hlt">3</span>OD and CD<span class="hlt">3</span>OD reactants are consistent with a sequence of elementary steps involving kinetically relevant H-abstraction from adsorbed methoxide species using lattice oxygen atoms and with methoxide formation in quasi-equilibrated CH<span class="hlt">3</span>OH dissociation on nearly stoichiometric Ru<span class="hlt">O</span>2 surfaces. Anaerobic transient experiments confirmed that CH<span class="hlt">3</span>OH oxidation to <span class="hlt">HCHO</span> requires lattice oxygen atoms and that selectivities are not influenced by the presence of <span class="hlt">O</span>2. Residence time effects on selectivity indicate that secondary <span class="hlt">HCHO</span>-CH<span class="hlt">3</span>OH acetalization reactions lead to hemiacetal or methoxymethanol intermediates that convert to dimethoxymethane in reactions with CH<span class="hlt">3</span>OH on support acid sites or dehydrogenate to form methylformate on Ru<span class="hlt">O</span>2 and support redox sites. These conclusions are consistent with the tendency of Al2<span class="hlt">O</span><span class="hlt">3</span> and Si<span class="hlt">O</span>2 supports to favor dimethoxymethane formation, while Sn<span class="hlt">O</span>2, Zr<span class="hlt">O</span>2, and Ti<span class="hlt">O</span>2 preferentially form methylformate. These support effects on secondary reactions were confirmed by measured CH<span class="hlt">3</span>OH oxidation rates and selectivities on physical mixtures of supported Ru<span class="hlt">O</span>2 catalysts and pure supports. Ethanol also reacts on supported Ru<span class="hlt">O</span>2 domains to form predominately acetaldehyde and diethoxyethane at 300-400 K. The bifunctional nature of these reaction pathways and the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AtmEn.139...37T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AtmEn.139...37T"><span>Photoactive roadways: Determination of CO, NO and VOC uptake coefficients and photolabile side product yields on Ti<span class="hlt">O</span>2 treated asphalt and concrete</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Toro, C.; Jobson, B. T.; Haselbach, L.; Shen, S.; Chung, S. H.</p> <p>2016-08-01</p> <p>This work reports uptake coefficients and by-product yields of ozone precursors onto two photocatalytic paving materials (asphalt and concrete) treated with a commercial Ti<span class="hlt">O</span>2 surface application product. The experimental approach used a continuously stirred tank reactor (CSTR) and allowed for testing large samples with the same surface morphology encountered with real urban surfaces. The measured uptake coefficient (γgeo) and surface resistances are useful for parametrizing dry deposition velocities in air quality model evaluation of the impact of photoactive surfaces on urban air chemistry. At 46% relative humidity, the surface resistance to NO uptake was ∼1 s cm-1 for concrete and ∼2 s cm-1 for a freshly coated older roadway asphalt sample. <span class="hlt">HONO</span> and NO2 were detected as side products from NO uptake to asphalt, with NO2 molar yields on the order of 20% and <span class="hlt">HONO</span> molar yields ranging between 14 and 33%. For concrete samples, the NO2 molar yields increased with the increase of water vapor, ranging from 1% to 35% and <span class="hlt">HONO</span> was not detected as a by-product. Uptake of monoaromatic VOCs to the asphalt sample set displayed a dependence on the compound vapor pressure, and was influenced by competitive adsorption from less volatile VOCs. Formaldehyde and acetaldehyde were detected as byproducts, with molar yields ranging from 5 to 32%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRD..12210439J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRD..12210439J"><span>Evaluating a Space-Based Indicator of Surface Ozone-NOx-VOC Sensitivity Over Midlatitude Source Regions and Application to Decadal Trends</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jin, Xiaomeng; Fiore, Arlene M.; Murray, Lee T.; Valin, Lukas C.; Lamsal, Lok N.; Duncan, Bryan; Folkert Boersma, K.; De Smedt, Isabelle; Abad, Gonzalo Gonzalez; Chance, Kelly; Tonnesen, Gail S.</p> <p>2017-10-01</p> <p>Determining effective strategies for mitigating surface ozone (<span class="hlt">O</span><span class="hlt">3</span>) pollution requires knowledge of the relative ambient concentrations of its precursors, NOx, and VOCs. The space-based tropospheric column ratio of formaldehyde to NO2 (FNR) has been used as an indicator to identify NOx-limited versus NOx-saturated <span class="hlt">O</span><span class="hlt">3</span> formation regimes. Quantitative use of this indicator ratio is subject to three major uncertainties: (1) the split between NOx-limited and NOx-saturated conditions may shift in space and time, (2) the ratio of the vertically integrated column may not represent the near-surface environment, and (<span class="hlt">3</span>) satellite products contain errors. We use the GEOS-Chem global chemical transport model to evaluate the quantitative utility of FNR observed from the Ozone Monitoring Instrument over three northern midlatitude source regions. We find that FNR in the model surface layer is a robust predictor of the simulated near-surface <span class="hlt">O</span><span class="hlt">3</span> production regime. Extending this surface-based predictor to a column-based FNR requires accounting for differences in the <span class="hlt">HCHO</span> and NO2 vertical profiles. We compare four combinations of two OMI <span class="hlt">HCHO</span> and NO2 retrievals with modeled FNR. The spatial and temporal correlations between the modeled and satellite-derived FNR vary with the choice of NO2 product, while the mean offset depends on the choice of <span class="hlt">HCHO</span> product. Space-based FNR indicates that the spring transition to NOx-limited regimes has shifted at least a month earlier over major cities (e.g., New York, London, and Seoul) between 2005 and 2015. This increase in NOx sensitivity implies that NOx emission controls will improve <span class="hlt">O</span><span class="hlt">3</span> air quality more now than it would have a decade ago.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4284574','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4284574"><span>Soil surface acidity plays a determining role in the atmospheric-terrestrial exchange of nitrous acid</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Donaldson, Melissa A.; Bish, David L.; Raff, Jonathan D.</p> <p>2014-01-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) is an important hydroxyl (OH) radical source that is formed on both ground and aerosol surfaces in the well-mixed boundary layer. Recent studies report the release of <span class="hlt">HONO</span> from nonacidic soils, although it is unclear how soil that is more basic than the pKa of <span class="hlt">HONO</span> (∼<span class="hlt">3</span>) is capable of protonating soil nitrite to serve as an atmospheric <span class="hlt">HONO</span> source. Here, we used a coated-wall flow tube and chemical ionization mass spectrometry (CIMS) to study the pH dependence of <span class="hlt">HONO</span> uptake onto agricultural soil and model substrates under atmospherically relevant conditions (1 atm and 30% relative humidity). Experiments measuring the evolution of <span class="hlt">HONO</span> from pH-adjusted surfaces treated with nitrite and potentiometric titrations of the substrates show, to our knowledge for the first time, that surface acidity rather than bulk aqueous pH determines <span class="hlt">HONO</span> uptake and desorption efficiency on soil, in a process controlled by amphoteric aluminum and iron (hydr)oxides present. The results have important implications for predicting when soil nitrite, whether microbially derived or atmospherically deposited, will act as a net source or sink of atmospheric <span class="hlt">HONO</span>. This process represents an unrecognized mechanism of <span class="hlt">HONO</span> release from soil that will contribute to <span class="hlt">HONO</span> emissions throughout the day. PMID:25512517</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25512517','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25512517"><span>Soil surface acidity plays a determining role in the atmospheric-terrestrial exchange of nitrous acid.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Donaldson, Melissa A; Bish, David L; Raff, Jonathan D</p> <p>2014-12-30</p> <p>Nitrous acid (<span class="hlt">HONO</span>) is an important hydroxyl (OH) radical source that is formed on both ground and aerosol surfaces in the well-mixed boundary layer. Recent studies report the release of <span class="hlt">HONO</span> from nonacidic soils, although it is unclear how soil that is more basic than the pKa of <span class="hlt">HONO</span> (∼ <span class="hlt">3</span>) is capable of protonating soil nitrite to serve as an atmospheric <span class="hlt">HONO</span> source. Here, we used a coated-wall flow tube and chemical ionization mass spectrometry (CIMS) to study the pH dependence of <span class="hlt">HONO</span> uptake onto agricultural soil and model substrates under atmospherically relevant conditions (1 atm and 30% relative humidity). Experiments measuring the evolution of <span class="hlt">HONO</span> from pH-adjusted surfaces treated with nitrite and potentiometric titrations of the substrates show, to our knowledge for the first time, that surface acidity rather than bulk aqueous pH determines <span class="hlt">HONO</span> uptake and desorption efficiency on soil, in a process controlled by amphoteric aluminum and iron (hydr)oxides present. The results have important implications for predicting when soil nitrite, whether microbially derived or atmospherically deposited, will act as a net source or sink of atmospheric <span class="hlt">HONO</span>. This process represents an unrecognized mechanism of <span class="hlt">HONO</span> release from soil that will contribute to <span class="hlt">HONO</span> emissions throughout the day.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.A23A0272J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.A23A0272J"><span>Expected trace gas and aerosol retrieval accuracy of the Geostationary Environment Monitoring Spectrometer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jeong, U.; Kim, J.; Liu, X.; Lee, K. H.; Chance, K.; Song, C. H.</p> <p>2015-12-01</p> <p>The predicted accuracy of the trace gases and aerosol retrievals from the geostationary environment monitoring spectrometer (GEMS) was investigated. The GEMS is one of the first sensors to monitor NO2, SO2, <span class="hlt">HCHO</span>, <span class="hlt">O</span><span class="hlt">3</span>, and aerosols onboard geostationary earth orbit (GEO) over Asia. Since the GEMS is not launched yet, the simulated measurements and its precision were used in this study. The random and systematic component of the measurement error was estimated based on the instrument design. The atmospheric profiles were obtained from Model for Ozone And Related chemical Tracers (MOZART) simulations and surface reflectances were obtained from climatology of OMI Lambertian equivalent reflectance. The uncertainties of the GEMS trace gas and aerosol products were estimated based on the OE method using the atmospheric profile and surface reflectance. Most of the estimated uncertainties of NO2, <span class="hlt">HCHO</span>, stratospheric and total <span class="hlt">O</span><span class="hlt">3</span> products satisfied the user's requirements with sufficient margin. However, about 26% of the estimated uncertainties of SO2 and about 30% of the estimated uncertainties of tropospheric <span class="hlt">O</span><span class="hlt">3</span> do not meet the required precision. Particularly the estimated uncertainty of SO2 is high in winter, when the emission is strong in East Asia. Further efforts are necessary in order to improve the retrieval accuracy of SO2 and tropospheric <span class="hlt">O</span><span class="hlt">3</span> in order to reach the scientific goal of GEMS. Random measurement error of GEMS was important for the NO2, SO2, and <span class="hlt">HCHO</span> retrieval, while both the random and systematic measurement errors were important for the <span class="hlt">O</span><span class="hlt">3</span> retrievals. The degree of freedom for signal of tropospheric <span class="hlt">O</span><span class="hlt">3</span> was 0.8 ± 0.2 and that for stratospheric <span class="hlt">O</span><span class="hlt">3</span> was 2.9 ± 0.5. The estimated uncertainties of the aerosol retrieval from GEMS measurements were predicted to be lower than the required precision for the SZA range of the trace gas retrievals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.A51C0035T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.A51C0035T"><span>Potential Impacts from Using Photoactive Roads as AN Air Quality Mitigation Strategy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Toro, C.; Jobson, B. T.; Shen, S.; Chung, S. H.; Haselbach, L.</p> <p>2013-12-01</p> <p>Mobile sources are major contributors to photochemical air pollution in urban areas. It has been proposed that the use of Ti<span class="hlt">O</span>2 coated roadways ('photoactive roads') could be an effective approach to reduce mobile source emissions by oxidizing NOx and VOC emissions at the roadway surface. However, studies have shown that formation of <span class="hlt">HONO</span> and aldehydes can occur from some Ti<span class="hlt">O</span>2 treated surfaces during the photocatalytic oxidation of NOx and VOC, respectively. By changing the NOx-to-VOC ratio and generating photolabile HOx radical precursors, photoactive roads may enhance ozone formation rates in urban areas. In this work we present results that quantify NOx and VOC loss rates onto Ti<span class="hlt">O</span>2 treated asphalt and concrete samples, as well as <span class="hlt">HONO</span> and aldehydes yields that result from the photocatalytic process. The treatment used a commercially available product. These objectives are relevant considering that the quantification of pollutant loss rates and yields of byproducts have not been determined for asphalt and that in the US more than 90% of the roadway surface is made of this material. Surface reaction probabilities (γ) and byproduct yields were determined using a CSTR photochemical chamber under varying conditions of water vapor and UV-A light intensity. Our results indicate that asphalt surfaces have a significantly higher molar yield of <span class="hlt">HONO</span> compared to concrete surfaces with similar Ti<span class="hlt">O</span>2 loading. Concrete surfaces have reaction probabilities with NO one order of magnitude higher than asphalt samples. Fresh asphalt samples showed negligible photocatalytic activity, presumably due to absorption of Ti<span class="hlt">O</span>2 into the bitumen substrate. Laboratory-prepared asphalt samples with a higher degree of exposed aggregates showed increased <span class="hlt">HONO</span> molar yields when compared to real-road asphalt samples, whose <span class="hlt">HONO</span> molar yield was ~1%. Preliminary results for aldehydes formation showed similar molar yields between aged asphalt and concrete, even though aged asphalt samples had twice</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1012857-reactivity-hydrogen-methanol-surfaces-wo3-reo3-wo3-reo3-reo3-wo3','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1012857-reactivity-hydrogen-methanol-surfaces-wo3-reo3-wo3-reo3-reo3-wo3"><span>Reactivity of Hydrogen and Methanol on (001) Surfaces of WO<span class="hlt">3</span>, Re<span class="hlt">O</span><span class="hlt">3</span>, WO<span class="hlt">3</span>/Re<span class="hlt">O</span><span class="hlt">3</span> and Re<span class="hlt">O</span><span class="hlt">3</span>/WO<span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Ling, Sanliang; Mei, Donghai; Gutowski, Maciej S.</p> <p>2011-05-16</p> <p>Bulk tungsten trioxide (WO<span class="hlt">3</span>) and rhenium trioxide (Re<span class="hlt">O</span><span class="hlt">3</span>) share very similar structures but display different electronic properties. WO<span class="hlt">3</span> is a wide bandgap semiconductor while Re<span class="hlt">O</span><span class="hlt">3</span> is an electronic conductor. With the advanced molecular beam epitaxy techniques, it is possible to make heterostructures comprised of layers of WO<span class="hlt">3</span> and Re<span class="hlt">O</span><span class="hlt">3</span>. These heterostructures might display reactivity different than pure WO<span class="hlt">3</span> and Re<span class="hlt">O</span><span class="hlt">3</span>. The interactions of two probe molecules (hydrogen and methanol) with the (001) surfaces of WO<span class="hlt">3</span>, Re<span class="hlt">O</span><span class="hlt">3</span>, and two heterostructures Re<span class="hlt">O</span><span class="hlt">3</span>/WO<span class="hlt">3</span> and WO<span class="hlt">3</span>/Re<span class="hlt">O</span><span class="hlt">3</span> were investigated at the density functional theory level. Atomic hydrogen prefers to adsorb at the terminal <span class="hlt">O</span>1C sitesmore » forming a surface hydroxyl on four surfaces. Dissociative adsorption of a hydrogen molecule at the <span class="hlt">O</span>1C site leads to formation of a water molecule adsorbed at the surface M5C site. This is thermodynamically the most stable state. A thermodynamically less stable dissociative state involves two surface hydroxyl groups <span class="hlt">O</span>1CH and <span class="hlt">O</span>2CH. The interaction of molecular hydrogen and methanol with pure Re<span class="hlt">O</span><span class="hlt">3</span> is stronger than with pure WO<span class="hlt">3</span> and the strength of the interaction substantially changes on the WO<span class="hlt">3</span>/Re<span class="hlt">O</span><span class="hlt">3</span> and Re<span class="hlt">O</span><span class="hlt">3</span>/WO<span class="hlt">3</span> heterostructures. The reaction barriers for decomposition and recombination reactions are sensitive to the nature of heterostructure. The calculated adsorption energy of methanol on WO<span class="hlt">3</span>(001) of -65.6 kJ/mol is consistent with the previous experimental estimation of -67 kJ/mol. This material is based upon work supported as part of the Center for Molecular Electrocatalysis, an Energy Frontier Research Center funded by the US Department of Energy, Office of Science, Office of Basic Energy Sciences.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/15006071-bands','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/15006071-bands"><span>The ν <span class="hlt">3</span> and 2ν <span class="hlt">3</span> bands of 32S 16<span class="hlt">O</span> <span class="hlt">3</span>, 32S 18<span class="hlt">O</span> <span class="hlt">3</span>, 34S 16<span class="hlt">O</span> <span class="hlt">3</span>, and 34S 18<span class="hlt">O</span> <span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Sharpe, Steven W.; Blake, Thomas A.; Sams, Robert L.</p> <p>2003-12-01</p> <p>The fifth of a series of publications on the high resolution rotation-vibration spectra of sulfur trioxide reports the results of a systematic study of the v<span class="hlt">3</span>(é) and 2v<span class="hlt">3</span>(A1'+E') infrared bands of the four symmetric top isotopomers 32S 16<span class="hlt">O</span> <span class="hlt">3</span>, 34S 16<span class="hlt">O</span> <span class="hlt">3</span>, 32S 18<span class="hlt">O</span> <span class="hlt">3</span>, and 34S 18<span class="hlt">O</span> <span class="hlt">3</span>. An internal coupling between the l = 0 and l = +2 levels of the 2v<span class="hlt">3</span> (A1'+E') states was observed. This small perturbation results in a level crossing between K-l = 9 and 12, in consequence of which the band origins of the A1', l=0 “ghost” states could be determined tomore » a high degree of accuracy. Ground and upper state rotational as well as vibrational anharmonicity constants are reported. The constants for the center-of-mass substituted species 32S 16<span class="hlt">O</span> <span class="hlt">3</span> and 34S 16<span class="hlt">O</span> <span class="hlt">3</span> vary only slightly, as do the constants for the 32S 18<span class="hlt">O</span> <span class="hlt">3</span>, 34S 18<span class="hlt">O</span> <span class="hlt">3</span> pair. The S-<span class="hlt">O</span> bond lengths for the vibrational ground states of the species 32S 16<span class="hlt">O</span> <span class="hlt">3</span>, 34S 16<span class="hlt">O</span> <span class="hlt">3</span>, 32S 18<span class="hlt">O</span> <span class="hlt">3</span> and 34S 18<span class="hlt">O</span> <span class="hlt">3</span>, are, respectively, 141.981992(6), 141.979412(20), 150.605240(73), and 150.602348(73) pm, where the uncertainties, given in parentheses, are two standard deviations and refer to the last digits of the associated quantity.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.A34B..01S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.A34B..01S"><span>Nitrous acid chemistry in Los Angeles during the CalNex-LA experiment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stutz, J.; Tsai, C.; Wong, K.; Pikelnaya, O.; Hurlock, S. C.; Young, C. J.; Veres, P. R.; Washenfelder, R. A.; Brown, S. S.; Flynn, J. H.; Grossberg, N.; Lefer, B. L.</p> <p>2011-12-01</p> <p>The role of nitrous acid, <span class="hlt">HONO</span>, as an OH radical precursor during the early morning and during the day has received considerable attention over the past decade. Several studies have reported that <span class="hlt">HONO</span> photolysis in the polluted boundary layer is the dominant source of OH in the early morning, and that it contributes up to 30% to the diurnally averaged primary OH formation. Despite the importance of <span class="hlt">HONO</span> for boundary layer chemistry, our understanding of <span class="hlt">HONO</span> sources is still incomplete. Laboratory studies suggest that <span class="hlt">HONO</span> is formed by the conversion of NO2 on humid surfaces at night. As this process is too slow during the day, several photo-enhanced processes have been proposed that either accelerate the NO2 conversion, or involve other nitrogen species, such as HNO<span class="hlt">3</span>. Field observations of vertical <span class="hlt">HONO</span> and NO2 concentration profiles, together with accurate measurements of other nitrogen species and actinic flux measurements offer a unique opportunity to constrain the proposed <span class="hlt">HONO</span> formation mechanisms. Here we present observations of <span class="hlt">HONO</span>, NO2, and other parameters made by various instruments during the 2010 CalNex experiment on the east side of the Los Angeles basin. We will discuss the vertical profiles of <span class="hlt">HONO</span> and NO2 measured by LP-DOAS, CEAS, CIMS, and photolytic conversion + CL with regard to the formation of <span class="hlt">HONO</span>. The observations will be compared to 1D chemistry and transport model calculations to test various proposed formation mechanisms. We will discuss the most likely formation pathway of <span class="hlt">HONO</span> and the potential impact of <span class="hlt">HONO</span> on atmospheric chemistry in Los Angeles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AdAtS..23..605X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AdAtS..23..605X"><span>Numerical study on the impacts of heterogeneous reactions on ozone formation in the Beijing urban area</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, Jun; Zhang, Yuanhang; Wang, Wei</p> <p>2006-12-01</p> <p>The air quality model CMAQ-MADRID (Community Multiscale Air Quality-Model of Aerosol Dynamics, Reaction, Ionization and Dissolution) was employed to simulate summer <span class="hlt">O</span><span class="hlt">3</span> formation in Beijing China, in order to explore the impacts of four heterogeneous reactions on <span class="hlt">O</span><span class="hlt">3</span> formation in an urban area. The results showed that the impacts were obvious and exhibited the characteristics of a typical response of a VOC-limited regime in the urban area. For the four heterogeneous reactions considered, the NO2 and HO2 heterogeneous reactions have the most severe impacts on <span class="hlt">O</span><span class="hlt">3</span> formation. During the <span class="hlt">O</span><span class="hlt">3</span> formation period, the NO2 heterogeneous reaction increased new radical creation by 30%, raising the atmospheric activity as more NO→NO2 conversion occurred, thus causing the <span class="hlt">O</span><span class="hlt">3</span> to rise. The increase of <span class="hlt">O</span><span class="hlt">3</span> peak concentration reached a maximum value of 67 ppb in the urban area. In the morning hours, high NO titration reduced the effect of the photolysis of <span class="hlt">HONO</span>, which was produced heterogeneously at night in the surface layer. The NO2 heterogeneous reaction in the daytime is likely one of the major reasons causing the <span class="hlt">O</span><span class="hlt">3</span> increase in the Beijing urban area. The HO2 heterogeneous reaction accelerated radical termination, resulting in a decrease of the radical concentration by 44% at the most. <span class="hlt">O</span><span class="hlt">3</span> peak concentration decreased by a maximum amount of 24 ppb in the urban area. The simulation results were improved when the heterogeneous reactions were included, with the <span class="hlt">O</span><span class="hlt">3</span> and <span class="hlt">HONO</span> model results close to the observations.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_12 --> <div id="page_13" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="241"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1812850G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1812850G"><span>Characterisation of Central-African emissions based on MAX-DOAS measurements, satellite observations and model simulations over Bujumbura, Burundi.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gielen, Clio; Hendrick, Francois; Pinardi, Gaia; De Smedt, Isabelle; Stavrakou, Trissevgeni; Yu, Huan; Fayt, Caroline; Hermans, Christian; Bauwens, Maité; Ndenzako, Eugene; Nzohabonayo, Pierre; Akimana, Rachel; Niyonzima, Sébastien; Müller, Jean-Francois; Van Roozendael, Michel</p> <p>2016-04-01</p> <p>Central Africa is known for its strong biogenic, pyrogenic, and to a lesser extent anthropogenic emissions. Satellite observations of species like nitrogen dioxide (NO2) and formaldehyde (<span class="hlt">HCHO</span>), as well as inverse modelling results have shown that there are large uncertainties associated with the emissions in this region. There is thus a need for additional measurements, especially from the ground, in order to better characterise the biomass-burning and biogenic products emitted in this area. We present MAX-DOAS measurements of NO2, <span class="hlt">HCHO</span>, and aerosols performed in Central Africa, in the city of Bujumbura, Burundi (<span class="hlt">3</span>°S, 29°E, 850m). A MAX-DOAS instrument has been operating at this location by BIRA-IASB since late 2013. Aerosol-extinction and trace-gases vertical profiles are retrieved by applying the optimal-estimation-based profiling tool bePRO to the measured <span class="hlt">O</span>4, NO2 and <span class="hlt">HCHO</span> slant-column densities. The MAX-DOAS vertical columns and profiles are used for investigating the diurnal and seasonal cycles of NO2, <span class="hlt">HCHO</span>, and aerosols. Regarding the aerosols, the retrieved AODs are compared to co-located AERONET sun photometer measurements for verification purpose, while in the case of NO2 and <span class="hlt">HCHO</span>, the MAX-DOAS vertical columns and profiles are used for validating GOME-2 and OMI satellite observations. To characterise the biomass-burning and biogenic emissions in the Bujumbura region, the trace gases and aerosol MAX-DOAS retrievals are used in combination to MODIS fire counts/radiative-power and GOME-2/OMI NO2 and <span class="hlt">HCHO</span> satellite data, as well as simulations from the NOAA backward trajectory model HYSPLIT. First results show that <span class="hlt">HCHO</span> seasonal variation around local noon is driven by the alternation of rain and dry periods, the latter being associated with intense biomass-burning agricultural activities and forest fires in the south/south-east and transport from this region to Bujumbura. In contrast, NO2 is seen to depend mainly on local emissions close to the city, due</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.A41A0034H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.A41A0034H"><span>Evaluation of biogenic emission flux and its impact on oxidants and inorganic aerosols in East Asia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Han, K. M.; Song, C. H.; Park, R. S.; Woo, J.; Kim, H.</p> <p>2010-12-01</p> <p>As a major precursor during the summer season, biogenic species are of primary importance in the ozone and SOAs (secondary organic aerosols) formations. Isoprene and mono-terpene also influence the level of inorganic aerosols (i.e. sulfate and nitrate) by controlling OH radicals. However, biogenic emission fluxes are highly uncertain in East Asia. While isoprene emission fluxes from the GEIA (Global Emissions Inventory Activity) and POET (Precursors of Ozone and their Effects in the Troposphere) inventories estimate approximately 20 Tg yr-1 in East Asia, those from the MEGAN (Model of Emissions of Gases and Aerosols from Nature) and MOHYCAN (MOdel for Hydrocarbon emissions by the CANopy) estimate approximately 10 Tg yr-1 and 5 Tg yr-1, respectively. In order to evaluate and/or quantify the magnitude of biogenic emission fluxes over East Asia, the tropospheric <span class="hlt">HCHO</span> columns obtained from the GOME (Global Ozone Monitoring Experiment) observations were compared with the <span class="hlt">HCHO</span> columns from the CMAQ (Community Multi-scale Air Quality) simulations over East Asia. In this study, US EPA Models-<span class="hlt">3</span>/CMAQ v4.5.1 model simulation using the ACE-ASIA (Asia Pacific Regional Aerosol Characterization Experiment) emission inventory for anthropogenic pollutants and GEIA, POET, MEGAN, and MOHYCAN emission inventories for biogenic species was carried out in conjunction with the Meteorological fields generated from the PSU/NCAR MM5 (Pennsylvania state University/National Center for Atmospheric Research Meso-scale Model 5) model for the summer episodes of the year 2002. In addition to an evaluation of the biogenic emission flux, we investigated the impact of the uncertainty in biogenic emission inventory on inorganic aerosol formations and variations of oxidants (OH, <span class="hlt">O</span><span class="hlt">3</span>, and H2<span class="hlt">O</span>2) in East Asia. In this study, when the GEIA and POET emission inventories are used, the CMAQ-derived <span class="hlt">HCHO</span> columns are highly overestimated over East Asia, particularly South China compared with GOME-derived <span class="hlt">HCHO</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.A31A0007T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.A31A0007T"><span>Formaldehyde and Glyoxal Measurements as Tracers of Oxidation Chemistry in the Amazon Basin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thayer, M. P.; Dorris, M. R.; Keutsch, F. N.; Springston, S. R.; Jimenez, J. L.; Palm, B. B.; Seco, R.; Kim, S.; Yee, L.; Wernis, R. A.; Goldstein, A. H.; Isaacman-VanWertz, G. A.; Liu, Y.; Martin, S. T.</p> <p>2015-12-01</p> <p>Formaldehyde (<span class="hlt">HCHO</span>) and glyoxal (CHOCHO) are important tracers for oxidative processes in the atmosphere such as oxidation of volatile organic compounds (VOCs) and production of HO2 radicals by photolysis or reaction with OH. Products of VOC oxidation and radical cycling, such as aerosols and tropospheric ozone, have direct impacts on human health. During the Green Ocean Amazon campaign (GoAmazon2014/5), <span class="hlt">HCHO</span> and CHOCHO measurements were obtained together with OH, RO2+HO2, CO, CO2, <span class="hlt">O</span><span class="hlt">3</span>, NOx, (<span class="hlt">o</span>)VOCs, and aerosol particle size distribution. <span class="hlt">HCHO</span> concentration was measured by the Madison FIber Laser-Induced Fluorescence (FILIF) instrument, while CHOCHO concentrations were collected by the Madison Laser-Induced Phosphorescence (Mad-LIP) instrument. Here we present data collected during 2014 at the T<span class="hlt">3</span> field site, 60 km to the west of Manaus, Brazil (<span class="hlt">3</span>°12'47.82"S, 60°35'55.32"W). The T<span class="hlt">3</span> GoAmazon site varies between sampling strictly pristine (biogenic) emissions and influence from anthropogenic emissions from Manaus, depending on meteorological conditions. Here we present overall trends and regimes observed during the campaign, with a focus on <span class="hlt">HCHO</span>, CHOCHO, and related species within the context of VOC oxidation and secondary pollutant production. We acknowledge the support from the Central Office of the Large Scale Biosphere Atmosphere Experiment in Amazonia (LBA), the Instituto Nacional de Pesquisas da Amazonia (INPA), and the Universidade do Estado do Amazonia (UEA). The work was conducted under 001030/2012-4 of the Brazilian National Council for Scientific and Technological Development (CNPq). Data were collected from the Atmospheric Radiation Measurement (ARM) Climate Research Facility, a U.S. Department of Energy Office of Science user facility sponsored by the Office of Biological and Environmental Research. Additionally, we acknowledge logistical support from the ARM Climate Research Facility. Additional funding from: NSF GRFP DGE-1256259, and NSF AGS-1051338</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/14968860','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14968860"><span>Modeling the oxidative capacity of the atmosphere of the south coast air basin of California. 2. HOx radical production.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Griffin, Robert J</p> <p>2004-02-01</p> <p>The production of HOx radicals in the South Coast Air Basin of California is investigated during the smog episode of September 9, 1993 using the California Institute of Technology (CIT) air-quality model. Sources of HOx(hydroxyl, hydroperoxy, and organic peroxy radicals) incorporated into the associated gas-phase chemical mechanism include the combination of excited-state singlet oxygen (formed from ozone (<span class="hlt">O</span><span class="hlt">3</span>) photolysis (hv)) with water, the photolysis of nitrous acid, hydrogen peroxide (H2<span class="hlt">O</span>2), and carbonyl compounds (formaldehyde (<span class="hlt">HCHO</span>) or higher aldehydes and ketones), the consumption of aldehydes and alkenes (ALK) by the nitrate radical, and the consumption of alkenes by <span class="hlt">O</span><span class="hlt">3</span> and the oxygen atom (<span class="hlt">O</span>). At a given time or location for surface cells and vertical averages, each route of HOx formation may be the greatest contributor to overall formation except <span class="hlt">HCHO</span>-hv, H2<span class="hlt">O</span>2-hv, and ALK-<span class="hlt">O</span>, the latter two of which are insignificant pathways in general. The contribution of the ALK-<span class="hlt">O</span><span class="hlt">3</span> pathway is dependent on the stoichiometric yield of OH, but this pathway, at least for the studied smog episode, may not be as generally significant as previous research suggests. Future emissions scenarios yield lower total HOx production rates and a shift in the relative importance of individual pathways.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=75594&keyword=chemiluminescence&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=75594&keyword=chemiluminescence&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>EVALUATION AND USE OF STAND-ALONE COMMERCIAL PHOTOLYTIC CONVERTERS FOR NO2 TO NO CONVERSION</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Two types of stand-alone commercial photolytic converters of NO2 to NO are now available for use with NO, <span class="hlt">O</span><span class="hlt">3</span> chemiluminescence monitors for the measurement of NO2. Both units have been tested for interferences resulting from photolysis of <span class="hlt">HONO</span> or from decomposition of PAN. On...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A53B2225X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A53B2225X"><span>Sources of secondary organic aerosols over North China Plain in winter</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xing, L.; Li, G.; Tie, X.; Junji, C.; Long, X.</p> <p>2017-12-01</p> <p>Organic aerosol (OA) concentrations are simulated over the North China Plain (NCP) from 10th to 26th January, 2014 using the Weather Research and Forecasting model coupled to chemistry (WRF-CHEM), with the goal of examining the impact of heterogeneous <span class="hlt">HONO</span> sources on atmospheric oxidation capacity and consequently on SOA formation and SOA formation from different pathways in winter. Generally, the model well reproduced the spatial and temporal distribution of PM2.5, SO2, NO2, and <span class="hlt">O</span><span class="hlt">3</span> concentrations. The heterogeneous <span class="hlt">HONO</span> formation contributed a major part of atmospheric <span class="hlt">HONO</span> concentrations in Beijing. The heterogeneous <span class="hlt">HONO</span> sources significantly increased the daily maximum OH concentrations by 260% on average in Beijing, which enhanced the atmospheric oxidation capacity and consequently SOA concentrations by 80% in Beijing on average. Under severe haze pollution on January 16th 2014, the regional average <span class="hlt">HONO</span> concentration over NCP was 0.86 ppb, which increased SOA concentration by 68% on average. The average mass fractions of ASOA (SOA from oxidation of anthropogenic VOCs), BSOA (SOA from oxidation of biogenic VOCs), PSOA (SOA from oxidation of evaporated POA), and GSOA (SOA from irreversible uptake of glyoxal and methylglyoxal) during the simulation period over NCP were 24%, 5%, 26% and 45%, respectively. GSOA contributed most to the total SOA mass over NCP in winter. The model sensitivity simulation revealed that GSOA in winter was mainly from primary residential sources. The regional average of GSOA from primary residential sources constituted 87% of total GSOA mass.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..MARU12012I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..MARU12012I"><span>Thickness dependent thermoelectric properties of SrTi<span class="hlt">O</span><span class="hlt">3</span>/SrLaTi<span class="hlt">O</span><span class="hlt">3</span> and SrZr<span class="hlt">O</span><span class="hlt">3</span>/SrLaTi<span class="hlt">O</span><span class="hlt">3</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ishii, Masatoshi; Baniecki, John; Schafranek, Robert; Kerman, Kian; Kurihara, Kazuaki</p> <p>2013-03-01</p> <p>Thermoelectric power generators will be required for future sensor network systems. SrTi<span class="hlt">O</span><span class="hlt">3</span> (STO) is one candidate thermoelectric material due to its non-toxicity and comparable power factor to Bismuth telluride. The energy conversion efficiency of SrTi<span class="hlt">O</span><span class="hlt">3</span>-based thermoelectric energy conversion elements has been reported to be enhanced by quantum size effects, such as the two dimensional (2D) electron gas in SrTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi0.8Nb0.2<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span>. Nevertheless, a complete understanding of the mechanisms for the reported increase in efficiency are missing owing to a lack of understanding of the thickness dependence of the transport properties. In the talk, we will present a study of the thickness dependence of the transport properties of SrTi<span class="hlt">O</span><span class="hlt">3</span>/SrLaTi<span class="hlt">O</span><span class="hlt">3</span> and SrZr<span class="hlt">O</span><span class="hlt">3</span>/SrLaTi<span class="hlt">O</span><span class="hlt">3</span> heterostructures. The SrZr<span class="hlt">O</span><span class="hlt">3</span>/SrLaTi<span class="hlt">O</span><span class="hlt">3</span> interface has a large conduction band off-set of 1.9 eV which can be utilized to confine electrons in a 2D quantum well. Characterization of the thermopower, conductivity, and Hall effect will be presented as a function of the SrLaTi<span class="hlt">O</span><span class="hlt">3</span> thickness down to a few unit cells and the implications of the thickness dependence of the transport properties on carrier confinement and increasing the efficiency STO-based 2DEG quantum well structures will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29682438','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29682438"><span>Evaluating a Space-Based Indicator of Surface Ozone-NO x -VOC Sensitivity Over Midlatitude Source Regions and Application to Decadal Trends.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jin, Xiaomeng; Fiore, Arlene M; Murray, Lee T; Valin, Lukas C; Lamsal, Lok N; Duncan, Bryan; Boersma, K Folkert; De Smedt, Isabelle; Abad, Gonzalo Gonzalez; Chance, Kelly; Tonnesen, Gail S</p> <p>2017-10-16</p> <p>Determining effective strategies for mitigating surface ozone (<span class="hlt">O</span> <span class="hlt">3</span> ) pollution requires knowledge of the relative ambient concentrations of its precursors, NO x , and VOCs. The space-based tropospheric column ratio of formaldehyde to NO 2 (FNR) has been used as an indicator to identify NO x -limited versus NO x -saturated <span class="hlt">O</span> <span class="hlt">3</span> formation regimes. Quantitative use of this indicator ratio is subject to three major uncertainties: (1) the split between NO x -limited and NO x -saturated conditions may shift in space and time, (2) the ratio of the vertically integrated column may not represent the near-surface environment, and (<span class="hlt">3</span>) satellite products contain errors. We use the GEOS-Chem global chemical transport model to evaluate the quantitative utility of FNR observed from the Ozone Monitoring Instrument over three northern midlatitude source regions. We find that FNR in the model surface layer is a robust predictor of the simulated near-surface <span class="hlt">O</span> <span class="hlt">3</span> production regime. Extending this surface-based predictor to a column-based FNR requires accounting for differences in the <span class="hlt">HCHO</span> and NO 2 vertical profiles. We compare four combinations of two OMI <span class="hlt">HCHO</span> and NO 2 retrievals with modeled FNR. The spatial and temporal correlations between the modeled and satellite-derived FNR vary with the choice of NO 2 product, while the mean offset depends on the choice of <span class="hlt">HCHO</span> product. Space-based FNR indicates that the spring transition to NO x -limited regimes has shifted at least a month earlier over major cities (e.g., New York, London, and Seoul) between 2005 and 2015. This increase in NO x sensitivity implies that NO x emission controls will improve <span class="hlt">O</span> <span class="hlt">3</span> air quality more now than it would have a decade ago.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040088590&hterms=hydroxylamine&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dhydroxylamine','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040088590&hterms=hydroxylamine&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dhydroxylamine"><span>Field intercomparison of a novel optical sensor for formaldehyde quantification</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Friedfeld, S.; Fraser, M.; Lancaster, D.; Leleux, D.; Rehle, D.; Tittel, F.</p> <p>2000-01-01</p> <p>A one-week in situ intercomparison campaign was completed on the Rice University campus for measuring <span class="hlt">HCHO</span> using three different techniques, including a novel optical sensor based on difference frequency generation (DFG) operating at room temperature. Two chemical derivatization methods, 2,4-dinitrophenylhydrazine (DNPH) and <span class="hlt">o</span>-(2,<span class="hlt">3</span>,4,5,6-pentafluorobenzyl) hydroxylamine (PFBHA), were deployed during the daylight hours for three- to four-hour time-integrated samples. A real-time optical sensor based on laser absorption spectroscopy was operated simultaneously, including nighttime hours. This tunable spectroscopic source based on difference frequency mixing of two fiber-amplified diode lasers in periodically poled LiNb03 (PPLN) was operated at <span class="hlt">3</span>.5315 micrometers (2831.64 cm 1) to access a strong <span class="hlt">HCHO</span> ro-vibrational transition free of interferences from other species. The results showed a bias of -1.7 and -1.2 ppbv and a gross error of 2.6 and 1.5 ppbv for DNPH and PFBHA measurements, respectively, compared with DFG measurements. These results validate the DFG sensor for time-resolved measurements of <span class="hlt">HCHO</span> in urban areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.A31B0042O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.A31B0042O"><span>Beyond SHARP-- Primary Formaldehyde from Oil and Gas Exploration and Production in the Gulf of Mexico Region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Olaguer, E. P.</p> <p>2010-12-01</p> <p>Formaldehyde has been named by the EPA as a hazardous air pollutant that may be carcinogenic and also cause irritation to the eyes, nose, throat and lung. Moreover, it is a powerful radical and ozone precursor. The 2009 Study of Houston Atmospheric Radical Precursors (SHARP) was conceived by the Houston Advanced Research Center (HARC) on behalf of the Texas Environmental Research Consortium (TERC) to examine the relative importance of primary and secondary formaldehyde (<span class="hlt">HCHO</span>) and nitrous acid (<span class="hlt">HONO</span>) in ozone formation. SHARP confirmed that primary combustion sources of <span class="hlt">HCHO</span>, such as flares end engines, may be underestimated (by an order of magnitude or more) in official emission inventories used for the purpose of air quality modeling in highly industrialized areas such as Houston. This presentation provides recently generated modeling and observational evidence that the same may be true in both rural and urban areas with oil and gas exploration and production (E&P) activities, such as the Upper Green River Basin of Wyoming and the Barnett Shale of Texas. Oil and gas E&P is increasing in the Gulf of Mexico region, particularly in the Barnett, Haynesville, Eagle Ford, Cana-Woodford, and Fayetteville shale basins. In the Barnett Shale, E&P activities are moving into urban neighborhoods, and may affect the ability to bring the Dallas-Ft. Worth region into attainment of the federal ozone standard. Data concerning formaldehyde emissions from drill rig and pipeline compressor engines, flares, and glycol or amine reboilers, should be obtained in order to more accurately model air quality in the Gulf of Mexico region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JGRD..118.1525Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JGRD..118.1525Z"><span>Source apportionment of formaldehyde during TexAQS 2006 using a source-oriented chemical transport model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Hongliang; Li, Jingyi; Ying, Qi; Guven, Birnur Buzcu; Olaguer, Eduardo P.</p> <p>2013-02-01</p> <p>In this study, a source-oriented version of the Community Multiscale Air Quality (CMAQ) model was developed and used to quantify the contributions of five major local emission source types in Southeast Texas (vehicles, industry, natural gas combustion, wildfires, biogenic sources), as well as upwind sources, to regional primary and secondary formaldehyde (<span class="hlt">HCHO</span>) concentrations. Predicted <span class="hlt">HCHO</span> concentrations agree well with observations at two urban sites (the Moody Tower [MT] site at the University of Houston and the Haden Road #<span class="hlt">3</span> [HRM-<span class="hlt">3</span>] site operated by Texas Commission on Environmental Quality). However, the model underestimates concentrations at an industrial site (Lynchburg Ferry). Throughout most of Southeast Texas, primary <span class="hlt">HCHO</span> accounts for approximately 20-30% of total <span class="hlt">HCHO</span>, while the remaining portion is due to secondary <span class="hlt">HCHO</span> (30-50%) and upwind sources (20-50%). Biogenic sources, natural gas combustion, and vehicles are important sources of primary <span class="hlt">HCHO</span> in the urban Houston area, respectively, accounting for 10-20%, 10-30%, and 20-60% of total primary <span class="hlt">HCHO</span>. Biogenic sources, industry, and vehicles are the top three sources of secondary <span class="hlt">HCHO</span>, respectively, accounting for 30-50%, 10-30%, and 5-15% of overall secondary <span class="hlt">HCHO</span>. It was also found that over 70% of PAN in the Houston area is due to upwind sources, and only 30% is formed locally. The model-predicted source contributions to <span class="hlt">HCHO</span> at the MT generally agree with source apportionment results obtained from the Positive Matrix Factorization (PMF) technique.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001JAP....89.5053B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001JAP....89.5053B"><span>Microstructure and dielectric parameters of epitaxial SrRu<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span>/SrRu<span class="hlt">O</span><span class="hlt">3</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Boikov, Yu. A.; Claeson, T.</p> <p>2001-05-01</p> <p>Epitaxial films of ferroelectric barium titanate are desirable in a number of applications but their properties are inferior to those of bulk material. Relations between microstructure and dielectric properties may give better understanding of limitations. Trilayer heterostructures SrRu<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span>/SrRu<span class="hlt">O</span><span class="hlt">3</span> were grown by laser ablation on (100)LaAl<span class="hlt">O</span><span class="hlt">3</span> and (100)Mg<span class="hlt">O</span> substrates. The BaTi<span class="hlt">O</span><span class="hlt">3</span> layer was granular in structure. When grown on (100)SrRu<span class="hlt">O</span><span class="hlt">3</span>/(100)LaAl<span class="hlt">O</span><span class="hlt">3</span>, it was preferentially a-axis oriented due to tensile mechanical stress. Using (100)Mg<span class="hlt">O</span> as a substrate, on the other hand, produced a mixture of about equal value of a-axis and c-axis oriented grains of BaTi<span class="hlt">O</span><span class="hlt">3</span>. The dielectric permittivity, ɛ, of the BaTi<span class="hlt">O</span><span class="hlt">3</span> layer was almost twice as large, at T>200 K and f=100 kHz, for the LaAl<span class="hlt">O</span><span class="hlt">3</span> substrate as compared to the Mg<span class="hlt">O</span> one. Its maximum value (ɛ/ɛ0≈6200) depended on temperature of growth, grain size, and electric field and compares well with optimal values commonly used for ceramic material. The maximum in the ɛ(T) shifted from about 370 to 320 K when the grain size in the BaTi<span class="hlt">O</span><span class="hlt">3</span> film decreased from 100 to 40 nm. At T<300 K, hysteresis loops in polarization versus electric field were roughly symmetric. The BaTi<span class="hlt">O</span><span class="hlt">3</span> films grown on (100)SrRu<span class="hlt">O</span><span class="hlt">3</span>/(100)Mg<span class="hlt">O</span> exhibit the largest remnant polarizations and coercive fields in the temperature range 100-380 K.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2692320','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2692320"><span>Analysis of non-enzymatically glycated peptides: neutral-loss-triggered MS<span class="hlt">3</span> versus multi-stage activation tandem mass spectrometry</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zhang, Qibin; Petyuk, Vladislav A.; Schepmoes, Athena A.; Orton, Daniel J.; Monroe, Matthew E.; Yang, Feng; Smith, Richard D.; Metz, Thomas O.</p> <p>2009-01-01</p> <p>Non-enzymatic glycation of tissue proteins has important implications in the development of complications of diabetes mellitus. While electron transfer dissociation (ETD) has been shown to outperform collision-induced dissociation (CID) in sequencing glycated peptides by tandem mass spectrometry, ETD instrumentation is not yet widely available and often suffers from significantly lower sensitivity than CID. In this study, we evaluated different advanced CID techniques (i.e., neutral-loss-triggered MS<span class="hlt">3</span> and multi-stage activation) during liquid chromatography/multi-stage mass spectrometric (LC/MSn) analyses of Amadori-modified peptides enriched from human serum glycated in vitro. During neutral-loss-triggered MS<span class="hlt">3</span> experiments, MS<span class="hlt">3</span> scans triggered by neutral losses of <span class="hlt">3</span> H2<span class="hlt">O</span> or <span class="hlt">3</span> H2<span class="hlt">O</span> + <span class="hlt">HCHO</span> produced similar results in terms of glycated peptide identifications. However, neutral losses of <span class="hlt">3</span> H2<span class="hlt">O</span> resulted in significantly more glycated peptide identifications during multi-stage activation experiments. Overall, the multi-stage activation approach produced more glycated peptide identifications, while the neutral-loss-triggered MS<span class="hlt">3</span> approach resulted in much higher specificity. Both techniques are viable alternatives to ETD for identifying glycated peptides. PMID:18763275</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AtmEn.183..135B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AtmEn.183..135B"><span>Insights into the degradation of (CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span> and its oxidative product (CF<span class="hlt">3</span>)2CHOCHO & the formation and catalytic degradation of organic nitrates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bai, Feng-Yang; Jia, Zi-Man; Pan, Xiu-Mei</p> <p>2018-06-01</p> <p>In this work, a systematic investigation of the atmospheric oxidation mechanism of (CF<span class="hlt">3</span>)2CXOCH<span class="hlt">3</span> and their oxidative products (CF<span class="hlt">3</span>)2CXOCHO (X = H, F) initiated by OH radical or Cl atom is performed by density functional theory. This study reveals that the introduction of NO and <span class="hlt">O</span>2 promotes the formation of organic nitrates, which are hygroscopic and are inclined to form secondary organic aerosols (SOA) and can affect the air quality. The rate constants of the individual reactions are found to be in agreement with the experimental results. One of the intriguing findings of this work is that the peroxynitrite of (CF<span class="hlt">3</span>)2CHOCH2OONO formed from the subsequent reactions of (CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span> is more favorable to isomerize to organic nitrate (CF<span class="hlt">3</span>)2CHOCH2ONO2 than to dissociate into alkoxy radical (CF<span class="hlt">3</span>)2CHOCH2<span class="hlt">O</span> and NO2 because of the lower energy barrier of isomerization. The second significant observation is that the organic nitrate can be degraded more favorably with the presence of NH<span class="hlt">3</span>, CH<span class="hlt">3</span>NH2, and CH<span class="hlt">3</span>NHCH<span class="hlt">3</span> than its naked decomposition reaction (CF<span class="hlt">3</span>)2CHOCH2ONO2→(CF<span class="hlt">3</span>)2CHOCHO + <span class="hlt">HONO</span>. The ammonium salt, a vital part of haze, is harmful to human health and can be formed in the existence of the NH<span class="hlt">3</span>, CH<span class="hlt">3</span>NH2, and CH<span class="hlt">3</span>NHCH<span class="hlt">3</span>. In addition, the toxic substance of peroxyalkyl nitrate (CF<span class="hlt">3</span>)2CHOC(<span class="hlt">O</span>)ONO2 which can reduce the visibility of the atmosphere is produced as the primary subsequent oxidation product of (CF<span class="hlt">3</span>)2CHOCHO in a NO-rich environment. The main species detected experimentally are confirmed by this study. The computational results are crucial to risk assessment and pollution prevention of the volatile organic compounds (VOCs).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA602237','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA602237"><span>Advanced Chemical Measurements of Smoke from DoD-prescribed Burns</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2014-04-01</p> <p>and in the Yucatan Peninsula of Mexico. Keene et al. [2006] 72 observed Δ<span class="hlt">HONO</span>/ΔNOx ratios (50th percentile) for African samples of grass (0.048...h from a plume in the Yucatan . In context, <span class="hlt">O</span><span class="hlt">3</span> formation is probably ubiquitous in tropical biomass burning plumes, but <span class="hlt">O</span><span class="hlt">3</span> destruction, as well as...via two dimensional heteronuclear NMR spectroscopy,” Phys. Chem. Chem. Phys., 11, 7810–7818, (2009). Marschner, H. 1986. Mineral Nutrition of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999MPLB...13..879S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999MPLB...13..879S"><span>Spectroscopic Properties of B2<span class="hlt">O</span><span class="hlt">3</span>-Pb<span class="hlt">O</span>-Nd2<span class="hlt">O</span><span class="hlt">3</span> Glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Simon, V.; Ardelean, I.; Milea, I.; Peteanu, M.; Simon, S.</p> <p></p> <p>Samples belonging to xNd2<span class="hlt">O</span><span class="hlt">3</span>(100-x) [2B2<span class="hlt">O</span><span class="hlt">3</span>·Pb<span class="hlt">O</span>] glass system, with 0≤ x≤ 40 mol%, are investigated by IR and UV-VIS spectroscopies in order to obtain evidence for the influence of Nd2<span class="hlt">O</span><span class="hlt">3</span> on the local order from 2B2<span class="hlt">O</span><span class="hlt">3</span>·Pb<span class="hlt">O</span> glass matrix. Besides the IR absorption bands characteristic to lead and boron arrangements, typical absorption lines of Nd<span class="hlt">3</span>+ ions around 4000 cm-1 and 6000 cm-1 are recorded. The 6000 cm-1 band appears only for the samples with x≥25 mol% Nd2<span class="hlt">O</span><span class="hlt">3</span>. The split of some UV-VIS absorption bands arising from transitions of neodymium ions in doublet lines as well as the shift of the absorption bands as the Nd2<span class="hlt">O</span><span class="hlt">3</span> content increases denote the influence of the lead-borate matrix on the radiative transitions of the lanthanide ion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title40-vol19/pdf/CFR-2012-title40-vol19-sec86-127-12.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title40-vol19/pdf/CFR-2012-title40-vol19-sec86-127-12.pdf"><span>40 CFR 86.127-12 - Test procedures; overview.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2012&page.go=Go">Code of Federal Regulations, 2012 CFR</a></p> <p></p> <p>2012-07-01</p> <p>...: (1) Gaseous exhaust THC, NMHC, NMOG, CO, NOX, CO2, N2<span class="hlt">O</span>, CH4, CH<span class="hlt">3</span>OH, C2H5OH, C2H4<span class="hlt">O</span>, and <span class="hlt">HCHO</span>. (2... exhaust emission test is designed to determine gaseous THC, NMHC, NMOG, CO, CO2, CH4, NOX, N2<span class="hlt">O</span>, and... THC using a heated sample line and analyzer; the other gaseous emissions (CH4, CO, CO2, N2<span class="hlt">O</span>, and NOX...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title40-vol19/pdf/CFR-2013-title40-vol19-sec86-127-12.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title40-vol19/pdf/CFR-2013-title40-vol19-sec86-127-12.pdf"><span>40 CFR 86.127-12 - Test procedures; overview.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2013&page.go=Go">Code of Federal Regulations, 2013 CFR</a></p> <p></p> <p>2013-07-01</p> <p>...: (1) Gaseous exhaust THC, NMHC, NMOG, CO, NOX, CO2, N2<span class="hlt">O</span>, CH4, CH<span class="hlt">3</span>OH, C2H5OH, C2H4<span class="hlt">O</span>, and <span class="hlt">HCHO</span>. (2... exhaust emission test is designed to determine gaseous THC, NMHC, NMOG, CO, CO2, CH4, NOX, N2<span class="hlt">O</span>, and... THC using a heated sample line and analyzer; the other gaseous emissions (CH4, CO, CO2, N2<span class="hlt">O</span>, and NOX...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title40-vol19/pdf/CFR-2014-title40-vol19-sec86-127-12.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title40-vol19/pdf/CFR-2014-title40-vol19-sec86-127-12.pdf"><span>40 CFR 86.127-12 - Test procedures; overview.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-07-01</p> <p>...: (1) Gaseous exhaust THC, NMHC, NMOG, CO, NOX, CO2, N2<span class="hlt">O</span>, CH4, CH<span class="hlt">3</span>OH, C2H5OH, C2H4<span class="hlt">O</span>, and <span class="hlt">HCHO</span>. (2... exhaust emission test is designed to determine gaseous THC, NMHC, NMOG, CO, CO2, CH4, NOX, N2<span class="hlt">O</span>, and... THC using a heated sample line and analyzer; the other gaseous emissions (CH4, CO, CO2, N2<span class="hlt">O</span>, and NOX...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MARX30005F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MARX30005F"><span>Quantifying the electronic reconstruction in LaTi<span class="hlt">O</span><span class="hlt">3</span>/LaNi<span class="hlt">O</span><span class="hlt">3</span>/(LaAl<span class="hlt">O</span><span class="hlt">3)3</span> heterostructures using RIXS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fabbris, Gilberto; Disa, Ankit S.; Ismail-Beigi, Sohab; Walker, Frederick J.; Ahn, Charles H.; Pelliciari, Jonathan; Huang, Yaobo; Schmitt, Thorsten; Xu, Lei; Hozoi, Liviu; van den Brink, Jeroen; Dean, Mark</p> <p></p> <p>A novel approach for manipulating the <span class="hlt">3</span>d state in transition metal oxide heterostructures has emerged with the growth of trilayer nickelate LaTi<span class="hlt">O</span><span class="hlt">3</span>/LaNi<span class="hlt">O</span><span class="hlt">3</span>/(LaAl<span class="hlt">O</span><span class="hlt">3)3</span> (LTNAO). This heterostructure induces a striking reconstruction of the LaNi<span class="hlt">O</span><span class="hlt">3</span> electronic structure, which is due to a combination of charge transfer from Ti's <span class="hlt">3</span>d state and octahedral elongation along the c axis. We use resonant inelastic x-ray scattering (RIXS) experiments at Ni L2,<span class="hlt">3</span> and <span class="hlt">O</span> K edges to spectroscopically resolve the LTNAO electronic structure. Surprisingly, our results show that the octahedral elongation generates minor changes in crystal fields at Ni's <span class="hlt">3</span>d state compared to bulk LaNi<span class="hlt">O</span><span class="hlt">3</span>. Instead, heterostructuring creates an anisotropic reconstruction of the Ni <span class="hlt">3</span>d - <span class="hlt">O</span> 2p hybridization. The x2-y2 orbital is significantly more hybridized with <span class="hlt">O</span> p, leading to a <span class="hlt">3</span>z2-r2/x2-y2 hole ratio of ~0.55 and large orbital polarization as measured by x-ray absorption spectroscopy. This work establishes RIXS as an ultra-sensitive probe of complex oxide heterostructures. Work at BNL was supported by the US Department of Energy under Award No DEAC02-98CH10886 and under Early Career Award No 20878.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_13 --> <div id="page_14" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="261"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..442..195S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..442..195S"><span>Solid strong base K-Pt/NaY zeolite nano-catalytic system for completed elimination of formaldehyde at room temperature</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Song, Shaoqing; Wu, Xi; Lu, Changhai; Wen, Meicheng; Le, Zhanggao; Jiang, Shujuan</p> <p>2018-06-01</p> <p>Solid strong base nano-catalytic system of K-modification NaY zeolite supported 0.08% Pt (K-Pt/NaY) were constructed for eliminating <span class="hlt">HCHO</span> at room temperature. In the catalytic process, activation energy over K-Pt/NaY nano-catalytic system was greatly decreased along with the enhanced reaction rate. Characterization and catalytic tests revealed the surface electron structure of K-Pt/NaY was improved, as reflected by the enhanced <span class="hlt">HCHO</span> adsorption capability, high sbnd OH concentration, and low-temperature reducibility. Therefore, the optimal K-Pt/NaY showed high catalytic efficiency and strong H2<span class="hlt">O</span> tolerance for <span class="hlt">HCHO</span> elimination by directly promoting the reaction between active sbnd OH and formate species. These results may suggest a new way for probing the advanced solid strong base nano-catalytic system for the catalytic elimination of indoor <span class="hlt">HCHO</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1987AtmEn..21.1529S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1987AtmEn..21.1529S"><span>Kinetics of the reaction between nitrogen dioxide and water vapour</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Svensson, R.; Ljungström, E.; Lindqvist, O.</p> <p></p> <p>The rate of disappearance of nitrogen dioxide (NO 2) with water vapour and formation of nitrous acid (<span class="hlt">HONO</span>) in the dark has been investigated in batch experiments. IR spectroscopy was used to determine the concentrations of NO 2, <span class="hlt">HONO</span> and NO. The reaction is first order both with respect to NO 2 and water vapour and proceeds heterogenously on most unpoisoned surfaces. Initially, the amount of <span class="hlt">HONO</span> formed is close to half the NO 2 which has disappeared. When the surface in the present reactor (surface to volume ratio = 14 m -1) has reached its limiting state of poisoning, the reaction is still active and the NO 2 disappearance follows the expression: -d[NO 2] /dt = 2k 1[NO 2] [H 2<span class="hlt">O</span>] where k1 = 4.1 (± 0.8) 10 -8 ppm -1 min -1 (22°C). The S/V ratio dependence of the rate shows that a heterogenous reaction proceeds but the existing evidence is not conclusive about a possible homogenous contribution to the remaining activity. A rate expression which describes the overall reaction at temperatures around 25°C, when the surface present is made passive, is: -d[NO 2] /dt = ( S/V5.6(±0.9)10 -9 + 2.<span class="hlt">3</span>(±6.5)10 -9)[NO 2][H 2<span class="hlt">O</span>] .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MMTB...48.1721K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MMTB...48.1721K"><span>Critical Evaluations and Thermodynamic Optimizations of the Mn<span class="hlt">O</span>-Mn2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 and Fe<span class="hlt">O</span>-Fe2<span class="hlt">O</span><span class="hlt">3</span>-Mn<span class="hlt">O</span>-Mn2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 Systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kang, Youn-Bae; Jung, In-Ho</p> <p>2017-06-01</p> <p>A critical evaluation and thermodynamic modeling for thermodynamic properties of all oxide phases and phase diagrams in the Fe-Mn-Si-<span class="hlt">O</span> system (Mn<span class="hlt">O</span>-Mn2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 and Fe<span class="hlt">O</span>-Fe2<span class="hlt">O</span><span class="hlt">3</span>-Mn<span class="hlt">O</span>-Mn2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 systems) are presented. Optimized Gibbs energy parameters for the thermodynamic models of the oxide phases were obtained which reproduce all available and reliable experimental data within error limits from 298 K (25°C) to above the liquidus temperatures at all compositions covering from known oxide phases, and oxygen partial pressure from metal saturation to 0.21 bar. The optimized thermodynamic properties and phase diagrams are believed to be the best estimates presently available. Slag (molten oxide) was modeled using the modified quasichemical model in the pair approximation. Olivine (Fe2Si<span class="hlt">O</span>4-Mn2Si<span class="hlt">O</span>4) was modeled using two-sublattice model in the framework of the compound energy formalism (CEF), while rhodonite (MnSi<span class="hlt">O</span><span class="hlt">3</span>-FeSi<span class="hlt">O</span><span class="hlt">3</span>) and braunite (Mn7Si<span class="hlt">O</span>_{12} with excess Mn2<span class="hlt">O</span><span class="hlt">3</span>) were modeled as simple Henrian solutions. It is shown that the already developed models and databases of two spinel phases (cubic- and tetragonal-(Fe, Mn)<span class="hlt">3</span><span class="hlt">O</span>4) using CEF [Kang and Jung, J. Phys. Chem. Solids (2016), vol. 98, pp. 237-246] can successfully be integrated into a larger thermodynamic database to be used in practically important higher order system such as silicate. The database of the model parameters can be used along with a software for Gibbs energy minimization in order to calculate any type of phase diagram section and thermodynamic properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.A53E0348K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.A53E0348K"><span>The State of Ambient Air Quality of a Mega City in Southeast Asia (Karachi, Pakistan)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Khwaja, H. A.; Hussain, M. M.; Naqvi, I.; Malik, A.; Siddiqui, S. A.; Khan, A.</p> <p>2016-12-01</p> <p>Outdoor air pollution is a serious public health problem. Studies indicate that in recent years exposure levels have increased considerably in some parts of the world, particularly in developing countries of Asia with large populations. Simultaneous measurements of PM2.5 and gaseous pollutants NO, NO2 , SO2 , <span class="hlt">O</span><span class="hlt">3</span> , <span class="hlt">HONO</span>, HNO<span class="hlt">3</span>, HF, and HCl were carried out in the city of Karachi. This is the first systematic study of this kind carried out in a mega city of Pakistan. Mean concentration of PM2.5 was 186 µg/m<span class="hlt">3</span>. Concentrations of NO, NO2 , SO2 , <span class="hlt">O</span><span class="hlt">3</span> , <span class="hlt">HONO</span> , HNO<span class="hlt">3</span>, HF, and HCl varied from 8.6 - 194 ppb, 15.7 - 131 ppb, 7.9 - 60 ppb, 5.0 - 218 ppb, 0.05 - 6.6 ppb, 0.1 - 10.8 ppb, 0.1 - 2.8 ppb, and 0.<span class="hlt">3</span> - 568 ppb, respectively. Daily patterns were observed. The 24 h mean PM2.5 on weekdays was significantly higher than the weekend value, indicating that vehicular pollution is one of the important source of PM2.5. The diurnal variations of both NO and NO2 showed higher concentrations during morning and evening rush-hours and lower concentrations at night, indicating that vehicular traffic is the principal source of NOx . Peak <span class="hlt">HONO</span> concentration of 6.6 ppb was observed in the morning hours. The highest SO2 , HNO<span class="hlt">3</span> , HF, and HCl values occurred during the daytime when general pollution levels, particularly those of suspended particulate matter, were also high. Concentrations of <span class="hlt">O</span><span class="hlt">3</span> are observed to increase during the daytime, consistent with its formation by photochemical reactions. The present findings are compared with similar measurements worldwide. Results have demonstrated that WHO air quality standard for PM2.5 (20 µg/m<span class="hlt">3</span>) were exceeded by a factor of 5 - 13. Concentrations of NO2 , SO2 and <span class="hlt">O</span><span class="hlt">3</span> were found to be significantly higher than the WHO air quality guidelines. The reported high levels were attributed to vehicular traffic and industrial activity. It has been concluded that air pollution levels in Karachi are extremely high and can be considered an alarming indicator</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AtmEn.127..272G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AtmEn.127..272G"><span>Evaluation of nitrous acid sources and sinks in urban outflow</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gall, Elliott T.; Griffin, Robert J.; Steiner, Allison L.; Dibb, Jack; Scheuer, Eric; Gong, Longwen; Rutter, Andrew P.; Cevik, Basak K.; Kim, Saewung; Lefer, Barry; Flynn, James</p> <p>2016-02-01</p> <p>Intensive air quality measurements made from June 22-25, 2011 in the outflow of the Dallas-Fort Worth (DFW) metropolitan area are used to evaluate nitrous acid (<span class="hlt">HONO</span>) sources and sinks. A two-layer box model was developed to assess the ability of established and recently identified <span class="hlt">HONO</span> sources and sinks to reproduce observations of <span class="hlt">HONO</span> mixing ratios. A baseline model scenario includes sources and sinks established in the literature and is compared to scenarios including three recently identified sources: volatile organic compound-mediated conversion of nitric acid to <span class="hlt">HONO</span> (S1), biotic emission from the ground (S2), and re-emission from a surface nitrite reservoir (S<span class="hlt">3</span>). For all mechanisms, ranges of parametric values span lower- and upper-limit values. Model outcomes for 'likely' estimates of sources and sinks generally show under-prediction of <span class="hlt">HONO</span> observations, implying the need to evaluate additional sources and variability in estimates of parameterizations, particularly during daylight hours. Monte Carlo simulation is applied to model scenarios constructed with sources S1-S<span class="hlt">3</span> added independently and in combination, generally showing improved model outcomes. Adding sources S2 and S<span class="hlt">3</span> (scenario S2/S<span class="hlt">3</span>) appears to best replicate observed <span class="hlt">HONO</span>, as determined by the model coefficient of determination and residual sum of squared errors (r2 = 0.55 ± 0.03, SSE = 4.6 × 106 ± 7.6 × 105 ppt2). In scenario S2/S<span class="hlt">3</span>, source S2 is shown to account for 25% and 6.7% of the nighttime and daytime budget, respectively, while source S<span class="hlt">3</span> accounts for 19% and 11% of the nighttime and daytime budget, respectively. However, despite improved model fit, there remains significant underestimation of daytime <span class="hlt">HONO</span>; on average, a 0.15 ppt/s unknown daytime <span class="hlt">HONO</span> source, or 67% of the total daytime source, is needed to bring scenario S2/S<span class="hlt">3</span> into agreement with observation. Estimates of 'best fit' parameterizations across lower to upper-limit values results in a moderate reduction of the unknown</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AIPC.1512..638R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AIPC.1512..638R"><span>A comparative study of photoconductivity in LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> and LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> 2-DEG heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rastogi, A.; Hossain, Z.; Budhani, R. C.</p> <p>2013-02-01</p> <p>Here we compare the growth temperature dependence of the response of LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> and LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> 2D-electron gas (2-DEG) field effect structure to the optical radiation of near ultraviolet frequency and electrostatic gate field. For both the films the resistance of the channel increases significantly as growth temperature is lowered from 800 to 700 °C. These heterostructures show the photoconductivity (PC) simulated by UV light of λ ≤ 400 nm. The PC follows the stretched exponential dynamics. It is found that photo-response of the LaTi<span class="hlt">O</span><span class="hlt">3</span> films is prominent and has larger decay time constant as compare to LaAl<span class="hlt">O</span><span class="hlt">3</span> films. The effect of electric field on the photo-induced conducting state is also studied.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27612644','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27612644"><span>Removal of nitrate and phosphate using chitosan/Al2<span class="hlt">O</span><span class="hlt">3</span>/Fe<span class="hlt">3</span><span class="hlt">O</span>4 composite nanofibrous adsorbent: Comparison with chitosan/Al2<span class="hlt">O</span><span class="hlt">3</span>/Fe<span class="hlt">3</span><span class="hlt">O</span>4 beads.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bozorgpour, Farahnaz; Ramandi, Hossein Fasih; Jafari, Pooya; Samadi, Saman; Yazd, Shabnam Sharif; Aliabadi, Majid</p> <p>2016-12-01</p> <p>In the present study the chitosan/Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 composite nanofibrous adsorbent was prepared by electrospinning process and its application for the removal of nitrate and phosphate were compared with chitosan/Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 composite bead adsorbent. The influence of Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 composite content, pH, contact time, nitrate and phosphate initial concentrations and temperature on the nitrate and phosphate sorption using synthesized bead and nanofibrous adsorbents was investigated in a single system. The reusability of chitosan/Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 composite beads and nanofibers after five sorption-desorption cycles were carried out. The Box-Behnken design was used to investigate the interaction effects of adsorbent dosage, nitrate and phosphate initial concentrations on the nitrate and phosphate removal efficiency. The pseudo-second-order kinetic model and known Freundlich and Langmuir isotherm models were used to describe the kinetic and equilibrium data of nitrate and phosphate sorption using chitosan/Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 composite beads and nanofibers. The influence of other anions including chloride, fluoride and sulphate on the sorption efficiency of nitrate and phosphate was examined. The obtained results revealed the higher potential of chitosan/Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 composite nanofibers for nitrate and phosphate compared with chitosan/Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 composite beads. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..433..904L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..433..904L"><span>Catalytic dehydrofluorination of 1,1,1,<span class="hlt">3,3</span>-pentafluoropropane to 1,<span class="hlt">3,3,3</span>-tetrafluoropropene over fluorinated Ni<span class="hlt">O</span>/Cr2<span class="hlt">O</span><span class="hlt">3</span> catalysts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Luo, Jian-Wei; Song, Jian-Dong; Jia, Wen-Zhi; Pu, Zhi-Ying; Lu, Ji-Qing; Luo, Meng-Fei</p> <p>2018-03-01</p> <p>Catalytic dehydrofluorination of 1,1,1,<span class="hlt">3,3</span>-pentafluoropropane to 1,<span class="hlt">3,3,3</span>-tetrafluoropropene was performed on a series of fluorinated Ni<span class="hlt">O</span>/Cr2<span class="hlt">O</span><span class="hlt">3</span> catalysts. The Ni<span class="hlt">O</span>/Cr2<span class="hlt">O</span><span class="hlt">3</span> catalysts were more active than the Cr2<span class="hlt">O</span><span class="hlt">3</span> because the new acid sites provided by NiF2 had higher turnover frequencies (9.43 × 10-<span class="hlt">3</span> - 12.08 × 10-<span class="hlt">3</span> s-1) than those on the Cr2<span class="hlt">O</span><span class="hlt">3</span> (4.55 × 10-<span class="hlt">3</span> s-1). Also, the Ni<span class="hlt">O</span>/Cr2<span class="hlt">O</span><span class="hlt">3</span> was more stable than the Cr2<span class="hlt">O</span><span class="hlt">3</span> due to its lower density of surface acid sites, which alleviated the coke deposition on the catalyst as evidenced by the Raman spectroscopic results. The kinetic results revealed that the15Ni<span class="hlt">O</span>/Cr2<span class="hlt">O</span><span class="hlt">3</span> had much lower activation energy (63.6 ± 4.5 kJ mol-1) than the Cr2<span class="hlt">O</span><span class="hlt">3</span> (127.6 ± <span class="hlt">3</span>.8 kJ mol-1). Accordingly, different reaction pathways on the two catalysts were proposed, which involved the cleavage of the Csbnd F and Csbnd H bonds on the surface acid and base sites, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.A33I3315S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A33I3315S"><span>OMI Global Tropospheric Bromine Oxide (Br<span class="hlt">O</span>) Column Densities: Algorithm, Retrieval and Initial Validation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Suleiman, R. M.; Chance, K.; Liu, X.; Kurosu, T. P.; Gonzalez Abad, G.</p> <p>2014-12-01</p> <p>We present and discuss a detailed description of the retrieval algorithms for the OMI Br<span class="hlt">O</span> product. The Br<span class="hlt">O</span> algorithms are based on direct fitting of radiances from 319.0-347.5 nm. Radiances are modeled from the solar irradiance, attenuated and adjusted by contributions from the target gas and interfering gases, rotational Raman scattering, undersampling, additive and multiplicative closure polynomials and a common mode spectrum. The version of the algorithm used for both Br<span class="hlt">O</span> includes relevant changes with respect to the operational code, including the fit of the <span class="hlt">O</span>2-<span class="hlt">O</span>2 collisional complex, updates in the high resolution solar reference spectrum, updates in spectroscopy, an updated Air Mass Factor (AMF) calculation scheme, and the inclusion of scattering weights and vertical profiles in the level 2 products. Updates to the algorithms include accurate scattering weights and air mass factor calculations, scattering weights and profiles in outputs and available cross sections. We include retrieval parameter and window optimization to reduce the interference from <span class="hlt">O</span><span class="hlt">3</span>, <span class="hlt">HCHO</span>, <span class="hlt">O</span>2-<span class="hlt">O</span>2, SO2, improve fitting accuracy and uncertainty, reduce striping, and improve the long-term stability. We validate OMI Br<span class="hlt">O</span> with ground-based measurements from Harestua and with chemical transport model simulations. We analyze the global distribution and seasonal variation of Br<span class="hlt">O</span> and investigate Br<span class="hlt">O</span> emissions from volcanoes and salt lakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27981481','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27981481"><span>Methanol-enhanced removal and metabolic conversion of formaldehyde by a black soybean from formaldehyde solutions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tan, Hao; Xiong, Yun; Li, Kun-Zhi; Chen, Li-Mei</p> <p>2017-02-01</p> <p>Methanol regulation of some biochemical and physiological characteristics in plants has been documented in several references. This study showed that the pretreatment of methanol with an appropriate concentration could stimulate the <span class="hlt">HCHO</span> uptake by black soybean (BS) plants. The process of methanol-stimulated <span class="hlt">HCHO</span> uptake by BS plants was optimized using the Central Composite Design and response surface methodology for the three variables, methanol concentration, <span class="hlt">HCHO</span> concentration, and treatment time. Under optimized conditions, the best stimulation effect of methanol on <span class="hlt">HCHO</span> uptake was obtained. 13 C-NMR analysis indicated that the H 13 CHO metabolism produced H 13 COOH, [2- 13 C]Gly, and [<span class="hlt">3</span>- 13 C]Ser in BS plant roots. Methanol pretreatment enhanced the metabolic conversion of H 13 CHO in BS plant roots, which consequently increased <span class="hlt">HCHO</span> uptake by BS plants. Therefore, methanol pretreatment might be used to increase <span class="hlt">HCHO</span> uptake by plants in the phytoremediation of <span class="hlt">HCHO</span>-polluted solutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015GMD.....8.3929G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015GMD.....8.3929G"><span>Singular vector-based targeted observations of chemical constituents: description and first application of the EURAD-IM-SVA v1.0</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Goris, N.; Elbern, H.</p> <p>2015-12-01</p> <p>Measurements of the large-dimensional chemical state of the atmosphere provide only sparse snapshots of the state of the system due to their typically insufficient temporal and spatial density. In order to optimize the measurement configurations despite those limitations, the present work describes the identification of sensitive states of the chemical system as optimal target areas for adaptive observations. For this purpose, the technique of singular vector analysis (SVA), which has proven effective for targeted observations in numerical weather prediction, is implemented in the EURAD-IM (EURopean Air pollution and Dispersion - Inverse Model) chemical transport model, yielding the EURAD-IM-SVA v1.0. Besides initial values, emissions are investigated as critical simulation controlling targeting variables. For both variants, singular vectors are applied to determine the optimal placement for observations and moreover to quantify which chemical compounds have to be observed with preference. Based on measurements of the airship based ZEPTER-2 campaign, the EURAD-IM-SVA v1.0 has been evaluated by conducting a comprehensive set of model runs involving different initial states and simulation lengths. For the sake of brevity, we concentrate our attention on the following chemical compounds, <span class="hlt">O</span><span class="hlt">3</span>, NO, NO2, <span class="hlt">HCHO</span>, CO, <span class="hlt">HONO</span>, and OH, and focus on their influence on selected <span class="hlt">O</span><span class="hlt">3</span> profiles. Our analysis shows that the optimal placement for observations of chemical species is not entirely determined by mere transport and mixing processes. Rather, a combination of initial chemical concentrations, chemical conversions, and meteorological processes determines the influence of chemical compounds and regions. We furthermore demonstrate that the optimal placement of observations of emission strengths is highly dependent on the location of emission sources and that the benefit of including emissions as target variables outperforms the value of initial value optimization with growing</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15510857','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15510857"><span>Development of Al2<span class="hlt">O</span><span class="hlt">3</span> fiber-reinforced Al2<span class="hlt">O</span><span class="hlt">3</span>-based ceramics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tanimoto, Yasuhiro; Nemoto, Kimiya</p> <p>2004-09-01</p> <p>The purpose of this study was to use a tape casting technique to develop an Al2<span class="hlt">O</span><span class="hlt">3</span> fiber-reinforced Al2<span class="hlt">O</span><span class="hlt">3</span>-based ceramic material (Al2<span class="hlt">O</span><span class="hlt">3</span>-fiber/Al2<span class="hlt">O</span><span class="hlt">3</span> composite) into a new type of dental ceramic. The Al2<span class="hlt">O</span><span class="hlt">3</span>-based ceramic used a matrix consisting of 60 wt% Al2<span class="hlt">O</span><span class="hlt">3</span> powder and 40 wt% Si<span class="hlt">O</span>2-B2<span class="hlt">O</span><span class="hlt">3</span> powder. The prepreg sheets of Al2<span class="hlt">O</span><span class="hlt">3</span>-fiber/Al2<span class="hlt">O</span><span class="hlt">3</span> composite (in which uniaxially aligned Al2<span class="hlt">O</span><span class="hlt">3</span> fibers were infiltrated with the Al2<span class="hlt">O</span><span class="hlt">3</span>-based matrix) were fabricated continuously using tape casting technique with a doctor blade system. Multilayer preforms of Al2<span class="hlt">O</span><span class="hlt">3</span>-fiber/Al2<span class="hlt">O</span><span class="hlt">3</span> composite sheets were then sintered at a maximum temperature of 1000 degrees C under an atmospheric pressure in a furnace. The results showed that the shrinkage and bending properties of Al2<span class="hlt">O</span><span class="hlt">3</span>-fiber/Al2<span class="hlt">O</span><span class="hlt">3</span> composite exceeded those of unreinforced Al2<span class="hlt">O</span><span class="hlt">3</span>--hence demonstrating the positive effects of fiber reinforcement. In conclusion, the tape casting technique has been utilized to successfully develop a new type of dental ceramic material.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JSSCh.258..809G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JSSCh.258..809G"><span>Real structure of (Sb1/<span class="hlt">3</span>Zn2/<span class="hlt">3</span>)Ga<span class="hlt">O</span><span class="hlt">3</span>(Zn<span class="hlt">O</span>)<span class="hlt">3</span>, a member of the homologous series ARO<span class="hlt">3</span>(Zn<span class="hlt">O</span>)m with ordered site occupation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Garling, Jennifer; Assenmacher, Wilfried; Schmid, Herbert; Longo, Paolo; Mader, Werner</p> <p>2018-02-01</p> <p>The hitherto unknown compound (Sb1/<span class="hlt">3</span>Zn2/<span class="hlt">3</span>)Ga<span class="hlt">O</span><span class="hlt">3</span>(Zn<span class="hlt">O</span>)<span class="hlt">3</span>, a member of the homologous series with general formula ARO<span class="hlt">3</span>(Zn<span class="hlt">O</span>)m (A,R = trivalent metal cation), was prepared by solid state methods from the binary oxides in sealed Pt-tubes. The structure of (Sb1/<span class="hlt">3</span>Zn2/<span class="hlt">3</span>)Ga<span class="hlt">O</span><span class="hlt">3</span>(Zn<span class="hlt">O</span>)<span class="hlt">3</span> has been determined by X-ray diffraction from flux-grown single crystals (R <span class="hlt">3</span> ̅ m , Z = <span class="hlt">3</span>, aR = <span class="hlt">3</span>.2387(7) Å, cR = 41.78(1) Å. The analysis revealed that (Sb1/<span class="hlt">3</span>Zn2/<span class="hlt">3</span>)Ga<span class="hlt">O</span><span class="hlt">3</span>(Zn<span class="hlt">O</span>)m is isostructural with InGa<span class="hlt">O</span><span class="hlt">3</span>(Zn<span class="hlt">O</span>)m, where In<span class="hlt">3</span>+ on octahedral sites is replaced by Sb5+ and Zn2+ in a ratio of 1:2, preserving an average charge of <span class="hlt">3</span>+. (Sb1/<span class="hlt">3</span>Zn2/<span class="hlt">3</span>)Ga<span class="hlt">O</span><span class="hlt">3</span>(Zn<span class="hlt">O</span>)<span class="hlt">3</span> was furthermore analyzed by electron diffraction, High Angle Annular Dark Field (HAADF) scanning TEM, and high precision EELS spectroscopic imaging, where a periodic ordering of Sb<span class="hlt">O</span>6 octahedra connected via edge sharing to six Zn<span class="hlt">O</span>6 octahedra in the octahedral layers in a honeycomb motif is found. Due to the large lateral distance of ca. 1.4 nm between adjacent octahedral layers, electrostatic interaction might hardly dictate Sb and Zn positions in neighbouring layers, and hence is a characteristic of the real structure of (Sb1/<span class="hlt">3</span>Zn2/<span class="hlt">3</span>)Ga<span class="hlt">O</span><span class="hlt">3</span>(Zn<span class="hlt">O</span>)<span class="hlt">3</span>. A structure model of the compound in space group P3112 (Nr. 151) with strictly ordered and discrete Sb and Zn positions is derived by crystallographic transformations as closest approximant for the real structure of (Sb1/<span class="hlt">3</span>Zn2/<span class="hlt">3</span>)Ga<span class="hlt">O</span><span class="hlt">3</span>(Zn<span class="hlt">O</span>)<span class="hlt">3</span>. UV-vis measurements confirm this compound to be a transparent oxide with an optical band gap in the UV region with Eg = <span class="hlt">3</span>.15 eV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21142162','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21142162"><span>Effects of defects on thermal decomposition of HMX via ReaxFF molecular dynamics simulations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhou, Ting-Ting; Huang, Feng-Lei</p> <p>2011-01-20</p> <p>Effects of molecular vacancies on the decomposition mechanisms and reaction dynamics of condensed-phase β-HMX at various temperatures were studied using ReaxFF molecular dynamics simulations. Results show that three primary initial decomposition mechanisms, namely, N-NO(2) bond dissociation, <span class="hlt">HONO</span> elimination, and concerted ring fission, exist at both high and lower temperatures. The contribution of the three mechanisms to the initial decomposition of HMX is influenced by molecular vacancies, and the effects vary with temperature. At high temperature (2500 K), molecular vacancies remarkably promote N-N bond cleavage and concerted ring breaking but hinder <span class="hlt">HONO</span> formation. N-N bond dissociation and <span class="hlt">HONO</span> elimination are two primary competing reaction mechanisms, and the former is dominant in the initial decomposition. Concerted ring breaking of condensed-phase HMX is not favored at high temperature. At lower temperature (1500 K), the most preferential initial decomposition pathway is N-N bond dissociation followed by the formation of NO(<span class="hlt">3</span>) (<span class="hlt">O</span> migration), although all three mechanisms are promoted by molecular vacancies. The promotion effect on concerted ring breaking is considerable at lower temperature. Products resulting from concerted ring breaking appear in the defective system but not in the perfect crystal. The mechanism of <span class="hlt">HONO</span> elimination is less important at lower temperature. We also estimated the reaction rate constant and activation barriers of initial decomposition with different vacancy concentrations. Molecular vacancies accelerate the decomposition of condensed-phase HMX by increasing the reaction rate constant and reducing activation barriers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19960016945','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19960016945"><span>Interhemispheric differences in polar stratospheric HNO<span class="hlt">3</span>, H2<span class="hlt">O</span>, Cl<span class="hlt">O</span>, and <span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Santee, M. L.; Read, W. G.; Waters, J. W.; Froidevaux, L.; Manney, G. L.; Flower, D. A.; Jarnot, R. F.; Harwood, R. S.; Peckham, G. E.</p> <p>1995-01-01</p> <p>Simultaneous global measurements of nitric acid (HNO<span class="hlt">3</span>), water (H2<span class="hlt">O</span>), chlorine monoxide (Cl<span class="hlt">O</span>), and ozone (<span class="hlt">O</span><span class="hlt">3</span>) in the stratosphere have been obtained over complete annual cycles in both hemispheres by the Microwave Limb Sounder on the Upper Atmosphere Research Satellite. A sizeable decrease in gas-phase HNO<span class="hlt">3</span> was evident in the lower stratospheric vortex over Antarctica by early June 1992, followed by a significant reduction in gas-phase H2<span class="hlt">O</span> after mid-July. By mid-August, near the time of peak Cl<span class="hlt">O</span>, abundances of gas-phase HNO<span class="hlt">3</span> and H2<span class="hlt">O</span> were extremely low. The concentrations of HNO<span class="hlt">3</span> and H2<span class="hlt">O</span> over Antarctica remained depressed into November, well after temperatures in the lower stratosphere had risen above the evaporation threshold for polar stratospheric clouds, implying that denitrification and dehydration had occurred. No large decreases in either gas-phase HNO<span class="hlt">3</span> or H2<span class="hlt">O</span> were observed in the 1992-1993 Arctic winter vortex. Although Cl<span class="hlt">O</span> was enhanced over the Arctic as it was over the Antarctic, Arctic <span class="hlt">O</span><span class="hlt">3</span> depletion was substantially smaller than that over Antarctica. A major factor currently limiting the formation of an Arctic ozone 'hole' is the lack of denitrification in the northern polar vortex, but future cooling of the lower stratosphere could lead to more intense denitrification and consequently larger losses of Arctic ozone.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21956589','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21956589"><span>Methanesulfonates of high-valent metals: syntheses and structural features of Mo<span class="hlt">O</span>2(CH<span class="hlt">3</span>SO<span class="hlt">3</span>)2, UO2(CH<span class="hlt">3</span>SO<span class="hlt">3</span>)2, Re<span class="hlt">O</span><span class="hlt">3</span>(CH<span class="hlt">3</span>SO<span class="hlt">3</span>), VO(CH<span class="hlt">3</span>SO<span class="hlt">3</span>)2, and V2<span class="hlt">O</span><span class="hlt">3</span>(CH<span class="hlt">3</span>SO<span class="hlt">3</span>)4 and their thermal decomposition under N2 and <span class="hlt">O</span>2 atmosphere.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Betke, Ulf; Neuschulz, Kai; Wickleder, Mathias S</p> <p>2011-11-04</p> <p>Oxide methanesulfonates of Mo, U, Re, and V have been prepared by reaction of Mo<span class="hlt">O</span>(<span class="hlt">3</span>), UO(2)(CH(<span class="hlt">3</span>)COO)(2)·2H(2)<span class="hlt">O</span>, Re(2)<span class="hlt">O</span>(7)(H(2)<span class="hlt">O</span>)(2), and V(2)<span class="hlt">O</span>(5) with CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>)H or mixtures thereof with its anhydride. These compounds are the first examples of solvent-free oxide methanesulfonates of these elements. Mo<span class="hlt">O</span>(2)(CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>))(2) (Pbca, a=1487.05(4), b=752.55(2), c=1549.61(5) pm, V=1.73414(9) nm(<span class="hlt">3</span>), Z=8) contains [Mo<span class="hlt">O</span>(2)] moieties connected by [CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>)] ions to form layers parallel to (100). UO(2)(CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>))(2) (P2(1)/c, a=1320.4(1), b=1014.41(6), c=1533.7(1) pm, β=112.80(1)°, V=1.8937(<span class="hlt">3</span>) nm(<span class="hlt">3</span>), Z=8) consists of linear UO(2)(2+) ions coordinated by five [CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>)] ions, forming a layer structure. VO(CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>))(2) (P2(1)/c, a=1136.5(1), b=869.87(7), c=915.5(1) pm, β=113.66(1)°, V=0.8290(2) nm(<span class="hlt">3</span>), Z=4) contains [VO] units connected by methanesulfonate anions to form corrugated layers parallel to (100). In Re<span class="hlt">O</span>(<span class="hlt">3</span>)(CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>)) (P1, a=574.0(1), b=1279.6(<span class="hlt">3</span>), c=1641.9(<span class="hlt">3</span>) pm, α=102.08(2), β=96.11(2), γ=99.04(2)°, V=1.1523(4) nm(<span class="hlt">3</span>), Z=8) a chain structure exhibiting infinite <span class="hlt">O-[ReO</span>(2)]-<span class="hlt">O-[ReO</span>(2)]-<span class="hlt">O</span> chains is formed. Each [Re<span class="hlt">O</span>(2)]-<span class="hlt">O-[ReO</span>(2)] unit is coordinated by two bidentate [CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>)] ions. V(2)<span class="hlt">O</span>(<span class="hlt">3</span>)(CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>))(4) (I2/a, a=1645.2(<span class="hlt">3</span>), b=583.1(1), c=1670.2(<span class="hlt">3</span>) pm, β=102.58(<span class="hlt">3</span>), V=1.5637(5) pm(<span class="hlt">3</span>), Z=4) adopts a chain structure, too, but contains discrete [VO]-<span class="hlt">O</span>-[VO] moieties, each coordinated by two bidentate [CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>)] ligands. Additional methanesulfonate ions connect the [V(2)<span class="hlt">O</span>(<span class="hlt">3</span>)] groups along [001]. Thermal decomposition of the compounds was monitored under N(2) and <span class="hlt">O</span>(2) atmosphere by thermogravimetric/differential thermal analysis and XRD measurements. Under N(2) the decomposition proceeds with reduction of the metal leading to the oxides Mo<span class="hlt">O</span>(2), U(<span class="hlt">3</span>)<span class="hlt">O</span>(7), V(4)<span class="hlt">O</span>(7), and VO(2); for Mo<span class="hlt">O</span>(2)(CH(<span class="hlt">3</span>)SO(<span class="hlt">3</span>))(2), a small amount of MoS(2) is formed. If the thermal decomposition is carried out in a atmosphere of <span class="hlt">O</span>(2) the oxides Mo<span class="hlt">O</span>(<span class="hlt">3</span>) and V(2)<span class="hlt">O</span>(5) are formed. Copyright </p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUFM.A11A0003F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUFM.A11A0003F"><span>Open-path Emission Factors Derived from DOAS and FTIR Measurements in the Mexico City Metropolitan Area</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Flores, E.; Grutter, M.; Galle, B.; Mellqvist, J.; Samuelsson, J.; Knighton, B.; Jobson, B. T.; Volkamer, R.; Molina, L. T.; Molina, M. J.</p> <p>2004-12-01</p> <p>Mobile sources are responsible for about 50% of VOC (volatile organic compounds) and about 70% of NOx emissions in the Mexico City Metropolitan Area (MCMA). A novel approach has been developed to derive emission factors for mobile sources that are representative of the overall vehicle fleet, using collocated open-path Differential Optical Absorption Spectroscopy (DOAS) and Fourier Transform Infrared (FTIR) spectroscopic measurements. Measurements were recorded at two sites within the MCMA: (1) research-grade DOAS and FTIR systems were operated at the Mexican National Research and Training Center (CENICA) in Iztapalapa, (2) a research grade FTIR was operated at La Merced. In addition, point-sampling with a proton transfer reaction mass spectrometer (PTR-MS) was performed on the same location and the calibration standards for the PTR-MS and the DOAS instruments were cross-calibrated. The DOAS measured speciated aromatic hydrocarbons, including benzene, toluene, m-xylene, p-xylene, ethylbenzene (and mono-substituted alkylbenzenes), benzaldehyde, phenol, and p-cresol. The DOAS detection of aromatic hydrocarbons in the UV/vis spectral range between 250 to 310 nm suffers from the interference of molecular oxygen, and a novel approach is being presented that enables measurement of absolute concentrations of the above species. Further, <span class="hlt">HONO</span>, NO2, SO2 and <span class="hlt">HCHO</span> were measured at longer wavelengths. In combination with FTIR measurements of CO, CO2, NO, <span class="hlt">HCHO</span>, ethylene, ethene, and total alkane, average emission factors for NOx, SO2 and numerous hydrocarbons were derived and scaled with fuel sales data to estimate total emissions of the vehicle fleet in the MCMA. The advantages and limitations of this low-cost emission inventory for mobile sources are decsribed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1097919-chemically-activated-formation-organic-acids-reactions-criegee-intermediate-aldehydes-ketones','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1097919-chemically-activated-formation-organic-acids-reactions-criegee-intermediate-aldehydes-ketones"><span>Chemically Activated Formation of Organic Acids in Reactions of the Criegee Intermediate with Aldehydes and Ketones</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Jalan, Amrit; Allen, Joshua W.; Green, William H.</p> <p></p> <p>Reactions of the Criegee intermediate (CI, .CH2OO.) are important in atmospheric ozonolysis models. In this work, we compute the rates for reactions between .CH2OO. and <span class="hlt">HCHO</span>, CH<span class="hlt">3</span>CHO and CH<span class="hlt">3</span>COCH<span class="hlt">3</span> leading to the formation of secondary ozonides (SOZ) and organic acids. Relative to infinitely separated reactants, the SOZ in all three cases is found to be 48–51 kcal mol-1 lower in energy, formed via 1,<span class="hlt">3</span>- cycloaddition of .CH2OO. across the CQO bond. The lowest energy pathway found for SOZ decomposition is intramolecular disproportionation of the singlet biradical intermediate formed from cleavage of the O–<span class="hlt">O</span> bond to form hydroxyalkyl esters. These hydroxyalkylmore » esters undergo concerted decomposition providing a low energy pathway from SOZ to acids. Geometries and frequencies of all stationary points were obtained using the B<span class="hlt">3</span>LYP/MG<span class="hlt">3</span>S DFT model chemistry, and energies were refined using RCCSD(T)-F12a/cc-pVTZ-F12 single-point calculations. RRKM calculations were used to obtain microcanonical rate coefficients (k(E)) and the reservoir state method was used to obtain temperature and pressure dependent rate coefficients (k(T, P)) and product branching ratios. At atmospheric pressure, the yield of collisionally stabilized SOZ was found to increase in the order <span class="hlt">HCHO</span> <span class="hlt">o</span> CH<span class="hlt">3</span>CHO <span class="hlt">o</span> CH<span class="hlt">3</span>COCH<span class="hlt">3</span> (the highest yield being 10-4 times lower than the initial .CH2OO. concentration). At low pressures, chemically activated formation of organic acids (formic acid in the case of <span class="hlt">HCHO</span> and CH<span class="hlt">3</span>COCH<span class="hlt">3</span>, formic and acetic acid in the case of CH<span class="hlt">3</span>CHO) was found to be the major product channel in agreement with recent direct measurements. Collisional energy transfer parameters and the barrier heights for SOZ reactions were found to be the most sensitive parameters determining SOZ and organic acid yield.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1911027T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1911027T"><span>Observation and modelling of the OH, HO_{2} and RO_{2} radicals at a suburban site of Beijing (Huairou) in winter 2016</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tan, Zhaofeng; Lu, Keding; Ma, Xuefei; Birger, Bohn; Broch, Sebastian; Fuchs, Hendrik; Hofzumahaus, Andreas; Holland, Frank; Li, Xin; Liu, Yuhan; Novelli, Anna; Rohrer, Franz; Shao, Min; Wang, Haichao; Wu, Yusheng; Zeng, Limin; Kiendler-Scharr, Astrid; Wahner, Andreas; Zhang, Yuanhang</p> <p>2017-04-01</p> <p>A comprehensive field campaign was carried out in winter 2016 in Huairou, a small town located 60 km northeast of Beijing downtown. Concentrations of OH, HO2and RO2 radicals were measured by a laser induced fluorescence instrument. Radical concentrations were smaller than during summer because of reduced solar radiation. Maximum hourly averaged OH, HO2 and RO2 radical concentrations were (<span class="hlt">3</span>±2)×106cm-<span class="hlt">3</span>, (8±6)×107 cm-<span class="hlt">3</span> and (7±5)×107 cm-<span class="hlt">3</span>, respectively. Chemical modulation measurements were applied on a few days showing no significant OH interference for different chemical conditions. <span class="hlt">HONO</span> and <span class="hlt">HCHO</span> photolysis were found to be the most important primary source of ROx radicals. OH reactivity, the inverse of the OH radical lifetime, was also measured by a laser-photolysis and laser induced fluorescence instrument. In general, CO and NOx were the dominated OH reactants which contributed more than half of the total OH reactivity. The relative high OH concentrations in polluted episode enabled a fast oxidation of fresh emitted pollutants and the formation of secondary products. The observed radical concentrations were compared with the results from a chemical box model. The model is capable of reproducing radical concentrations in the moderate NOx conditions but has difficulty in both the low and high NOx regimes. The underestimation of RO2 radical concentrations in the high NOx conditions indicate a missing RO2 source.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1177331-interfacial-ferromagnetism-lanio3-camno3-superlattices','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1177331-interfacial-ferromagnetism-lanio3-camno3-superlattices"><span>Interfacial Ferromagnetism in LaNi<span class="hlt">O</span><span class="hlt">3</span>/CaMn<span class="hlt">O</span><span class="hlt">3</span> Superlattices</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Grutter, Alexander J.; Yang, Hao; Kirby, B. J.</p> <p>2013-08-01</p> <p>We observe interfacial ferromagnetism in superlattices of the paramagnetic metal LaNi<span class="hlt">O</span><span class="hlt">3</span> and the antiferromagnetic insulator CaMn<span class="hlt">O</span><span class="hlt">3</span>. LaNi<span class="hlt">O</span><span class="hlt">3</span> exhibits a thickness dependent metal-insulator transition and we find the emergence of ferromagnetism to be coincident with the conducting state of LaNi<span class="hlt">O</span><span class="hlt">3</span>. That is, only superlattices in which the LaNi<span class="hlt">O</span><span class="hlt">3</span> layers are metallic exhibit ferromagnetism. Using several magnetic probes, we have determined that the ferromagnetism arises in a single unit cell of CaMn<span class="hlt">O</span><span class="hlt">3</span> at the interface. Together these results suggest that ferromagnetism can be attributed to a double exchange interaction among Mn ions mediated by the adjacent itinerant metal.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ApPhL..98x1903G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ApPhL..98x1903G"><span>SmNi<span class="hlt">O</span><span class="hlt">3</span>/NdNi<span class="hlt">O</span><span class="hlt">3</span> thin film multilayers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Girardot, C.; Pignard, S.; Weiss, F.; Kreisel, J.</p> <p>2011-06-01</p> <p>Rare earth nickelates RENi<span class="hlt">O</span><span class="hlt">3</span> (RE =rare earth), which attract interest due to their sharp metal-insulator phase transition, are instable in bulk form due to the necessity of an important oxygen pressure to stabilize Ni in its <span class="hlt">3</span>+ state of oxidation. Here, we report the stabilization of RE nickelates in [(SmNi<span class="hlt">O</span><span class="hlt">3</span>)t/(NdNi<span class="hlt">O</span><span class="hlt">3</span>)t]n thin film multilayers, t being the thickness of layers alternated n times. Both bilayers and multilayers have been deposited by metal-organic chemical vapor deposition. The multilayer structure and the presence of the metastable phases SmNi<span class="hlt">O</span><span class="hlt">3</span> and NdNi<span class="hlt">O</span><span class="hlt">3</span> are evidenced from by x-ray and Raman scattering. Electric measurements of a bilayer structure further support the structural quality of the embedded RE nickelate layers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013MPLB...2750207S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013MPLB...2750207S"><span>OPTICAL AND SPECTROSCOPIC STUDIES OF Fe2<span class="hlt">O</span><span class="hlt">3</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>:V2<span class="hlt">O</span>5 GLASSES</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sanjay; Kishore, N.; Agarwal, A.; Dahiya, S.; Pal, Inder; Kumar, Navin</p> <p>2013-11-01</p> <p>The glasses of compositions xFe2<span class="hlt">O</span><span class="hlt">3</span>ṡ (40 - x)Bi2<span class="hlt">O</span><span class="hlt">3</span>ṡ60B2<span class="hlt">O</span><span class="hlt">3</span>ṡ2V2<span class="hlt">O</span>5 have been prepared by the standard melt-quenching technique. Amorphous nature of these samples is ascertained by XRD patterns. The presence of BO<span class="hlt">3</span> and BO4 units is identified by IR spectra of glass samples. The absorption edge (λcut-off) shifts toward longer wavelengths with an increase in Fe2<span class="hlt">O</span><span class="hlt">3</span> content in the glass matrix. The values of optical band gap energy for indirect allowed and forbidden transitions have been determined and it is found to decrease with increase in transition metal ions. The Urbach's energy is used to characterize the degree of disorder in amorphous solids.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22299875','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22299875"><span>17<span class="hlt">O</span> excess transfer during the NO2 + <span class="hlt">O</span><span class="hlt">3</span> → NO<span class="hlt">3</span> + <span class="hlt">O</span>2 reaction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Berhanu, Tesfaye Ayalneh; Savarino, Joël; Bhattacharya, S K; Vicars, Willliam C</p> <p>2012-01-28</p> <p>The ozone molecule possesses a unique and distinctive (17)<span class="hlt">O</span> excess (Δ(17)<span class="hlt">O</span>), which can be transferred to some of the atmospheric molecules via oxidation. This isotopic signal can be used to trace oxidation reactions in the atmosphere. However, such an approach depends on a robust and quantitative understanding of the oxygen transfer mechanism, which is currently lacking for the gas-phase NO(2) + <span class="hlt">O</span>(<span class="hlt">3</span>) reaction, an important step in the nocturnal production of atmospheric nitrate. In the present study, the transfer of Δ(17)<span class="hlt">O</span> from ozone to nitrate radical (NO(<span class="hlt">3</span>)) during the gas-phase NO(2) + <span class="hlt">O</span>(<span class="hlt">3</span>) → NO(<span class="hlt">3</span>) + <span class="hlt">O</span>(2) reaction was investigated in a series of laboratory experiments. The isotopic composition (δ(17)<span class="hlt">O</span>, δ(18)<span class="hlt">O</span>) of the bulk ozone and the oxygen gas produced in the reaction was determined via isotope ratio mass spectrometry. The Δ(17)<span class="hlt">O</span> transfer function for the NO(2) + <span class="hlt">O</span>(<span class="hlt">3</span>) reaction was determined to be: Δ(17)<span class="hlt">O(O</span>(<span class="hlt">3</span>)∗) = (1.23 ± 0.19) × Δ(17)<span class="hlt">O(O</span>(<span class="hlt">3</span>))(bulk) + (9.02 ± 0.99). The intramolecular oxygen isotope distribution of ozone was evaluated and results suggest that the excess enrichment resides predominantly on the terminal oxygen atoms of ozone. The results obtained in this study will be useful in the interpretation of high Δ(17)<span class="hlt">O</span> values measured for atmospheric nitrate, thus leading to a better understanding of the natural cycling of atmospheric reactive nitrogen. © 2012 American Institute of Physics</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26768534','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26768534"><span>The Molybdenum(V) and Tungsten(VI) Oxoazides [Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )<span class="hlt">3</span> ], [Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )<span class="hlt">3</span> ⋅2 CH<span class="hlt">3</span> CN], [(bipy)Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )<span class="hlt">3</span> ], [Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )5 ](2-) , [WO(N<span class="hlt">3</span> )4 ], and [WO(N<span class="hlt">3</span> )4 ⋅CH<span class="hlt">3</span> CN].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Haiges, Ralf; Skotnitzki, Juri; Fang, Zongtang; Dixon, David A; Christe, Karl O</p> <p>2015-12-14</p> <p>A series of novel molybdenum(V) and tungsten(VI) oxoazides was prepared starting from [MOF4 ] (M=Mo, W) and Me<span class="hlt">3</span> SiN<span class="hlt">3</span> . While [WO(N<span class="hlt">3</span> )4 ] was formed through fluoride-azide exchange in the reaction of Me<span class="hlt">3</span> SiN<span class="hlt">3</span> with WOF4 in SO2 solution, the reaction with MoOF4 resulted in a reduction of Mo(VI) to Mo(V) and formation of [Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )<span class="hlt">3</span> ]. Carried out in acetonitrile solution, these reactions resulted in the isolation of the corresponding adducts [Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )<span class="hlt">3</span> ⋅2 CH<span class="hlt">3</span> CN] and [WO(N<span class="hlt">3</span> )4 ⋅CH<span class="hlt">3</span> CN]. Subsequent reactions of [Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )<span class="hlt">3</span> ] with 2,2'-bipyridine and [PPh4 ][N<span class="hlt">3</span> ] resulted in the formation and isolation of [(bipy)Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )<span class="hlt">3</span> ] and [PPh4 ]2 [Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )5 ], respectively. Most molybdenum(V) and tungsten(VI) oxoazides were fully characterized by their vibrational spectra, impact, friction and thermal sensitivity data and, in the case of [WO(N<span class="hlt">3</span> )4 ⋅CH<span class="hlt">3</span> CN], [(bipy)Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )<span class="hlt">3</span> ], and [PPh4 ]2 [Mo<span class="hlt">O</span>(N<span class="hlt">3</span> )5 ], by their X-ray crystal structures. © 2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23013316','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23013316"><span>Vertically resolved measurements of nighttime radical reservoirs in Los Angeles and their contribution to the urban radical budget.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Young, Cora J; Washenfelder, Rebecca A; Roberts, James M; Mielke, Levi H; Osthoff, Hans D; Tsai, Catalina; Pikelnaya, Olga; Stutz, Jochen; Veres, Patrick R; Cochran, Anthony K; VandenBoer, Trevor C; Flynn, James; Grossberg, Nicole; Haman, Christine L; Lefer, Barry; Stark, Harald; Graus, Martin; de Gouw, Joost; Gilman, Jessica B; Kuster, William C; Brown, Steven S</p> <p>2012-10-16</p> <p>Photolabile nighttime radical reservoirs, such as nitrous acid (<span class="hlt">HONO</span>) and nitryl chloride (ClNO(2)), contribute to the oxidizing potential of the atmosphere, particularly in early morning. We present the first vertically resolved measurements of ClNO(2), together with vertically resolved measurements of <span class="hlt">HONO</span>. These measurements were acquired during the California Nexus (CalNex) campaign in the Los Angeles basin in spring 2010. Average profiles of ClNO(2) exhibited no significant dependence on height within the boundary layer and residual layer, although individual vertical profiles did show variability. By contrast, nitrous acid was strongly enhanced near the ground surface with much smaller concentrations aloft. These observations are consistent with a ClNO(2) source from aerosol uptake of N(2)<span class="hlt">O</span>(5) throughout the boundary layer and a <span class="hlt">HONO</span> source from dry deposition of NO(2) to the ground surface and subsequent chemical conversion. At ground level, daytime radical formation calculated from nighttime-accumulated <span class="hlt">HONO</span> and ClNO(2) was approximately equal. Incorporating the different vertical distributions by integrating through the boundary and residual layers demonstrated that nighttime-accumulated ClNO(2) produced nine times as many radicals as nighttime-accumulated <span class="hlt">HONO</span>. A comprehensive radical budget at ground level demonstrated that nighttime radical reservoirs accounted for 8% of total radicals formed and that they were the dominant radical source between sunrise and 09:00 Pacific daylight time (PDT). These data show that vertical gradients of radical precursors should be taken into account in radical budgets, particularly with respect to <span class="hlt">HONO</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24787773','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24787773"><span>Violet-green excitation for NIR luminescence of Yb<span class="hlt">3</span>+ ions in Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Weiwei; Cheng, Jimeng; Zhao, Guoying; Chen, Wei; Hu, Lili; Guzik, Malgorzata; Boulon, Georges</p> <p>2014-04-21</p> <p>60Bi(2)<span class="hlt">O</span>(<span class="hlt">3</span>)-20B(2)<span class="hlt">O</span>(<span class="hlt">3</span>)-10Si<span class="hlt">O</span>(2)-10Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>) glasses doped with 1-9 mol% Yb(2)<span class="hlt">O</span>(<span class="hlt">3</span>) were prepared and investigated mainly on their violet-green excitation for the typical NIR emission of Yb(<span class="hlt">3</span>+), generally excited in the NIR. Two violet excitation bands at 365 nm and 405 nm are related to Yb(2+) and Bi(<span class="hlt">3</span>+). 465 nm excitation band and 480 nm absorption band in the blue-green are assigned to Bi(0) metal nanoparticles/grains. Yb-content-dependence of the excitation and absorption means that Bi(0) is the reduced product of Bi(<span class="hlt">3</span>+), but greatly competed by the redox reaction of Yb(2+) ↔ Yb(<span class="hlt">3</span>+). It is proved that the violet-green excitations result in the NIR emission of Yb(<span class="hlt">3</span>+). On the energy transfer, the virtual level of Yb(<span class="hlt">3</span>+)-Yb(<span class="hlt">3</span>+) as well as Bi(0) dimers probably plays an important role. An effective and controllable way is suggested to achieve nano-optical applications by Bi(0) metal nanoparticles/grains and Yb(<span class="hlt">3</span>+).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ApPhL.102i1601H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ApPhL.102i1601H"><span>Compositional and gate tuning of the interfacial conductivity in LaAl<span class="hlt">O</span><span class="hlt">3</span>/LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hosoda, Masayuki; Bell, Christopher; Hikita, Yasuyuki; Hwang, Harold Y.</p> <p>2013-03-01</p> <p>We investigate the effect of LaTi<span class="hlt">O</span><span class="hlt">3</span> insertion at the interface between LaAl<span class="hlt">O</span><span class="hlt">3</span> and Ti<span class="hlt">O</span>2 terminated {100} SrTi<span class="hlt">O</span><span class="hlt">3</span> for a series of LaAl<span class="hlt">O</span><span class="hlt">3</span> and LaTi<span class="hlt">O</span><span class="hlt">3</span> thicknesses. A clear increase of the carrier density was observed while the Hall mobility was largely unchanged. In structures with LaAl<span class="hlt">O</span><span class="hlt">3</span> thickness ˜<span class="hlt">3</span> unit cells, close to the critical thickness for conductivity, as little as 0.25 unit cells of LaTi<span class="hlt">O</span><span class="hlt">3</span> drives an insulator-to-metal transition. These samples show a strong dependence of the conductivity on voltage with electrostatic back-gating, which can be understood in a two-carrier picture, and dominated by the change in carrier density at the interface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ApPhL.103u1601K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ApPhL.103u1601K"><span>Pinhole mediated electrical transport across LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> and LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> oxide hetero-structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kumar, Pramod; Dogra, Anjana; Toutam, Vijaykumar</p> <p>2013-11-01</p> <p>Metal-insulator-metal configuration of LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> and LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> hetero-structures between two dimensional electron gas formed at the interface and different area top electrodes is employed for Conductive Atomic force microscopy (CAFM) imaging, Current-Voltage (I-V), and Capacitance-Voltage (C-V) spectroscopy. Electrode area dependent I-V characteristics are observed for these oxide hetero-structures. With small area electrodes, rectifying I-V characteristics are observed, compared to, both tunneling and leakage current characteristics for large area electrodes. CAFM mapping confirmed the presence of pinholes on both surfaces. Resultant I-V characteristics have a contribution from both tunneling and leakage due to pinholes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994JaJAP..33.5332K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994JaJAP..33.5332K"><span>Electric Properties of Pb(Sb1/2Nb1/2)<span class="hlt">O</span><span class="hlt">3</span> PbTi<span class="hlt">O</span><span class="hlt">3</span> PbZr<span class="hlt">O</span><span class="hlt">3</span> Ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kawamura, Yasushi; Ohuchi, Hiromu</p> <p>1994-09-01</p> <p>Solid-solution ceramics of ternary system xPb(Sb1/2Nb1/2)<span class="hlt">O</span><span class="hlt">3</span> yPbTi<span class="hlt">O</span><span class="hlt">3</span> zPbZr<span class="hlt">O</span><span class="hlt">3</span> were prepared by the solid-state reaction of powder materials. Ceramic, electric, dielectric and piezoelectric properties and crystal structures of the system were studied. Sintering of the system xPb(Sb1/2Nb1/2)<span class="hlt">O</span><span class="hlt">3</span> yPbTi<span class="hlt">O</span><span class="hlt">3</span> zPbZr<span class="hlt">O</span><span class="hlt">3</span> is much easier than that of each end composition, and well-sintered high-density ceramics were obtained for the compositions near the morphotropic transformation. Piezoelectric ceramics with high relative dielectric constants, high radial coupling coefficient and low resonant resistance were obtained for the composition near the morphotropic transformation. The composition Pb(Sb1/2Nb1/2)0.075Ti0.45Zr0.475<span class="hlt">O</span><span class="hlt">3</span> showed the highest dielectric constant (ɛr=1690), and the composition Pb(Sb1/2Nb1/2)0.05Ti0.45Zr0.5<span class="hlt">O</span><span class="hlt">3</span> showed the highest radial coupling coefficient (kp=64%).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1942n0005P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1942n0005P"><span>Energy transfer mechanism of Sm<span class="hlt">3</span>+/Eu<span class="hlt">3</span>+ co-doped 2Ca<span class="hlt">O</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-P2<span class="hlt">O</span>5 phosphors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prasad, V. Reddy; Damodaraiah, S.; Ratnakaram, Y. C.</p> <p>2018-04-01</p> <p>Sm<span class="hlt">3</span>+/Eu<span class="hlt">3</span>+ co-doped calcium borophosphate phosphors were synthesized by solid state reaction method. 2Ca<span class="hlt">O</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-P2<span class="hlt">O</span>5: Sm<span class="hlt">3</span>+/Eu<span class="hlt">3</span>+ co-doped phosphors were characterized by XRD, SEM, 31P solid state NMR, excitation, photoluminescence (PL) and decay profiles.. XRD profiles showed that the prepared phosphors exhibit a hexagonal phase in crystal structure and SEM results showed that the particles are more irregular morphologies. From 31P NMR spectra of Sm<span class="hlt">3</span>+/Eu<span class="hlt">3</span>+ co-doped 2Ca<span class="hlt">O</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-P2<span class="hlt">O</span>5 phosphors, the chemical shifts located in the positive frequency region indicating the presence of mono-phosphate complexes Q0-(PO43 - ) . Photoluminescence spectra of Sm<span class="hlt">3</span>+/Eu<span class="hlt">3</span>+ co-doped 2Ca<span class="hlt">O</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-P2<span class="hlt">O</span>5 phosphors show enhancement in emission intensity of Eu<span class="hlt">3</span>+ ion due to co-doping with Sm<span class="hlt">3</span>+ ions through energy transfer process. The energy level mechanism between Sm<span class="hlt">3</span>+ and Eu<span class="hlt">3</span>+ ions has been clearly explained. The energy transfer process has also been evidenced by lifetime decay profiles. These results suggest that the prepared phosphors are potential red luminescent optical materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MRE.....5d6309S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MRE.....5d6309S"><span>SAW propagation characteristics of Te<span class="hlt">O</span><span class="hlt">3/3</span>C-SiC/LiNb<span class="hlt">O</span><span class="hlt">3</span> layered structure</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Soni, Namrata D.</p> <p>2018-04-01</p> <p>Surface acoustic wave (SAW) devices based on Lithium Niobate (LiNb<span class="hlt">O</span><span class="hlt">3</span>) single crystal are advantageous because of its high SAW phase velocity, electromechanical coupling coefficient and cost effectiveness. In the present work a new multi-layered Te<span class="hlt">O</span><span class="hlt">3/3</span>C-SiC/128° Y-X LiNb<span class="hlt">O</span><span class="hlt">3</span> SAW device has been proposed. SAW propagation properties such as phase velocity, coupling coefficient and temperature coefficient of delay (TCD) of the Te<span class="hlt">O</span><span class="hlt">3</span>/SiC/128° Y-X LiNb<span class="hlt">O</span><span class="hlt">3</span> multi layered structure is examined using theoretical calculations. It is found that the integration of 0.09λ thick <span class="hlt">3</span>C-SiC over layer on 128° Y-X LiNb<span class="hlt">O</span><span class="hlt">3</span> increases its electromechanical coupling coefficient from 5.<span class="hlt">3</span>% to 9.77% and SAW velocity from 3800 ms‑1 to 4394 ms‑1. The SiC/128° Y-X LiNb<span class="hlt">O</span><span class="hlt">3</span> bilayer SAW structure exhibits a high positive TCD value. A temperature stable layered SAW device could be obtained with introduction of 0.007λ Te<span class="hlt">O</span><span class="hlt">3</span> over layer on SiC/128° Y-X LiNb<span class="hlt">O</span><span class="hlt">3</span> bilayer structure without sacrificing the efficiency of the device. The proposed Te<span class="hlt">O</span><span class="hlt">3/3</span>C-SiC/128° Y-X LiNb<span class="hlt">O</span><span class="hlt">3</span> multi-layered SAW structure is found to be cost effective, efficient, temperature stable and suitable for high frequency application in harsh environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ACP....10.2413R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ACP....10.2413R"><span>Formaldehyde and its relation to CO, PAN, and SO2 in the Houston-Galveston airshed</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rappenglück, B.; Dasgupta, P. K.; Leuchner, M.; Li, Q.; Luke, W.</p> <p>2010-03-01</p> <p>The Houston-Galveston Airshed (HGA) is one of the major metropolitan areas in the US that is classified as a nonattainment area of federal ozone standards. Formaldehyde (<span class="hlt">HCHO</span>) is a key species in understanding ozone related air pollution; some of the highest <span class="hlt">HCHO</span> concentrations in North America have been reported for the HGA. We report on <span class="hlt">HCHO</span> measurements in the HGA from summer 2006. Among several sites, maximum <span class="hlt">HCHO</span> mixing ratios were observed in the Houston Ship Channel (HSC), a region with a very high density of industrial/petrochemical operations. <span class="hlt">HCHO</span> levels at the Moody Tower (MT) site close to downtown were dependent on the wind direction: southerly maritime winds brought in background levels (0.5-1 ppbv) while trajectories originating in the HSC resulted in high <span class="hlt">HCHO</span> (up to 31.5 ppbv). Based on the best multiparametric linear regression model fit, the <span class="hlt">HCHO</span> levels at the MT site can be accounted for as follows: 38.5±12.<span class="hlt">3</span>% from primary vehicular emissions (using CO as an index of vehicular emission), 24.1±17.7% formed photochemically (using peroxyacetic nitric anhydride (PAN) as an index of photochemical activity) and 8.9±11.2% from industrial emissions (using SO2 as an index of industrial emissions). The balance 28.5±12.7% constituted the residual which cannot be easily ascribed to the above categories and/or which is transported into the HGA. The CO related <span class="hlt">HCHO</span> fraction is dominant during the morning rush hour (06:00-09:00 h, all times are given in CDT); on a carbon basis, <span class="hlt">HCHO</span> emissions are up to 0.7% of the CO emissions. The SO2 related <span class="hlt">HCHO</span> fraction is significant between 09:00-12:00 h. After 12:00 h <span class="hlt">HCHO</span> is largely formed through secondary processes. The <span class="hlt">HCHO</span>/PAN ratios are dependent on the SO2 levels. The SO2 related <span class="hlt">HCHO</span> fraction at the downtown site originates in the ship channel. Aside from traffic-related primary <span class="hlt">HCHO</span> emissions, <span class="hlt">HCHO</span> of industrial origin serves as an appreciable source for OH in the morning.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890000753','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890000753"><span>The effect of Al2<span class="hlt">O</span><span class="hlt">3</span>, Ca<span class="hlt">O</span>, Cr2<span class="hlt">O</span><span class="hlt">3</span> and Mg<span class="hlt">O</span> on devitrification of silica</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zaplatynsky, Isidor</p> <p>1988-01-01</p> <p>The effect of doping on devitrification of vitreous silica was studied at 1100, 1200, and 1300 C. Dispersion of dopants on a molecular scale was accomplished via a sol-gel technique. All dopants accelerated the devitrification of silica but to different degrees. The most active was Ca<span class="hlt">O</span> followed by Mg<span class="hlt">O</span>, Al2<span class="hlt">O</span><span class="hlt">3</span>, and Cr2<span class="hlt">O</span><span class="hlt">3</span>. Pure silica and silica containing Cr2<span class="hlt">O</span><span class="hlt">3</span> and Al2<span class="hlt">O</span><span class="hlt">3</span> devitrified to alpha-cristobalite only, whereas silica doped with Ca<span class="hlt">O</span> and Mg<span class="hlt">O</span> produced alpha-quartz and alpha-cristobalite. It appears that prolonged heat treatment would cause alpha-quartz to transform to alpha-cristobalite.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018CPL...696...92R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018CPL...696...92R"><span>Novel Zr<span class="hlt">O</span>2 based ceramics stabilized by Fe2<span class="hlt">O</span><span class="hlt">3</span>, Si<span class="hlt">O</span>2 and Y2<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rada, S.; Culea, E.; Rada, M.</p> <p>2018-03-01</p> <p>Samples in the 5Fe2<span class="hlt">O</span><span class="hlt">3</span>·10Si<span class="hlt">O</span>2·xY2<span class="hlt">O</span><span class="hlt">3</span>·(85-x)Zr<span class="hlt">O</span>2 composition where x = 5, 10 and 15 mol% Y2<span class="hlt">O</span><span class="hlt">3</span> were synthesized and investigated by XRD, SEM, density measurements, FTIR, UV-Vis, EPR and PL spectroscopies. X-ray diffraction patterns confirm the presence of the tetragonal and cubic Zr<span class="hlt">O</span>2 crystalline phases in all samples. The IR data show the overlaps of absorption bands assigned to Zrsbnd Osbnd Zr and Sisbnd Osbnd linkages in samples. UV-Vis and PL data indicate higher concentrations of intrinsic defects by doping with Y2<span class="hlt">O</span><span class="hlt">3</span> concentrations. The EPR spectra are characterized by two resonance lines situated at about g ∼ 4.<span class="hlt">3</span> and g ∼ 2 for lower Y2<span class="hlt">O</span><span class="hlt">3</span> contents.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015NatSR...5E9093X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015NatSR...5E9093X"><span>Magnetic interactions in BiFe0.5Mn0.5<span class="hlt">O</span><span class="hlt">3</span> films and BiFe<span class="hlt">O</span><span class="hlt">3</span>/BiMn<span class="hlt">O</span><span class="hlt">3</span> superlattices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, Qingyu; Sheng, Yan; Khalid, M.; Cao, Yanqiang; Wang, Yutian; Qiu, Xiangbiao; Zhang, Wen; He, Maocheng; Wang, Shuangbao; Zhou, Shengqiang; Li, Qi; Wu, Di; Zhai, Ya; Liu, Wenqing; Wang, Peng; Xu, Y. B.; Du, Jun</p> <p>2015-03-01</p> <p>The clear understanding of exchange interactions between magnetic ions in substituted BiFe<span class="hlt">O</span><span class="hlt">3</span> is the prerequisite for the comprehensive studies on magnetic properties. BiFe0.5Mn0.5<span class="hlt">O</span><span class="hlt">3</span> films and BiFe<span class="hlt">O</span><span class="hlt">3</span>/BiMn<span class="hlt">O</span><span class="hlt">3</span> superlattices have been fabricated by pulsed laser deposition on (001) SrTi<span class="hlt">O</span><span class="hlt">3</span> substrates. Using piezoresponse force microscopy (PFM), the ferroelectricity at room temperature has been inferred from the observation of PFM hysteresis loops and electrical writing of ferroelectric domains for both samples. Spin glass behavior has been observed in both samples by temperature dependent magnetization curves and decay of thermo-remnant magnetization with time. The magnetic ordering has been studied by X-ray magnetic circular dichroism measurements, and Fe-<span class="hlt">O</span>-Mn interaction has been confirmed to be antiferromagnetic (AF). The observed spin glass in BiFe0.5Mn0.5<span class="hlt">O</span><span class="hlt">3</span> films has been attributed to cluster spin glass due to Mn-rich ferromagnetic (FM) clusters in AF matrix, while spin glass in BiFe<span class="hlt">O</span><span class="hlt">3</span>/BiMn<span class="hlt">O</span><span class="hlt">3</span> superlattices is due to competition between AF Fe-<span class="hlt">O</span>-Fe, AF Fe-<span class="hlt">O</span>-Mn and FM Mn-<span class="hlt">O</span>-Mn interactions in the well ordered square lattice with two Fe ions in BiFe<span class="hlt">O</span><span class="hlt">3</span> layer and two Mn ions in BiMn<span class="hlt">O</span><span class="hlt">3</span> layer at interfaces.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17026141','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17026141"><span>Photoemission from buried interfaces in SrTi<span class="hlt">O</span><span class="hlt">3</span>/LaTi<span class="hlt">O</span><span class="hlt">3</span> superlattices.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Takizawa, M; Wadati, H; Tanaka, K; Hashimoto, M; Yoshida, T; Fujimori, A; Chikamatsu, A; Kumigashira, H; Oshima, M; Shibuya, K; Mihara, T; Ohnishi, T; Lippmaa, M; Kawasaki, M; Koinuma, H; Okamoto, S; Millis, A J</p> <p>2006-08-04</p> <p>We have measured photoemission spectra of SrTi<span class="hlt">O</span><span class="hlt">3</span>/LaTi<span class="hlt">O</span><span class="hlt">3</span> superlattices with a topmost SrTi<span class="hlt">O</span><span class="hlt">3</span> layer of variable thickness. A finite coherent spectral weight with a clear Fermi cutoff was observed at chemically abrupt SrTi<span class="hlt">O</span><span class="hlt">3</span>/LaTi<span class="hlt">O</span><span class="hlt">3</span> interfaces, indicating that an "electronic reconstruction" occurs at the interface between the Mott insulator LaTi<span class="hlt">O</span><span class="hlt">3</span> and the band insulator SrTi<span class="hlt">O</span><span class="hlt">3</span>. For SrTi<span class="hlt">O</span><span class="hlt">3</span>/LaTi<span class="hlt">O</span><span class="hlt">3</span> interfaces annealed at high temperatures (approximately 1000 degrees C), which leads to Sr/La atomic interdiffusion and hence to the formation of La(1-x)Sr(x)Ti<span class="hlt">O</span><span class="hlt">3</span>-like material, the intensity of the incoherent part was found to be dramatically reduced whereas the coherent part with a sharp Fermi cutoff was enhanced due to the spread of charge. These important experimental features are well reproduced by layer dynamical-mean-field-theory calculation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28316822','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28316822"><span>Crystal structure of bis-(μ-<span class="hlt">3</span>-nitro-benzoato)-κ<span class="hlt">3</span><span class="hlt">O,O':O</span>;κ<span class="hlt">3</span><span class="hlt">O:O,O</span>'-bis-[bis-(<span class="hlt">3</span>-cyano-pyridine-κN1)(<span class="hlt">3</span>-nitro-benzoato-κ2<span class="hlt">O,O</span>')cadmium].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hökelek, Tuncer; Akduran, Nurcan; Özen, Azer; Uğurlu, Güventürk; Necefoğlu, Hacali</p> <p>2017-03-01</p> <p>The asymmetric unit of the title compound, [Cd 2 (C 7 H 4 NO 4 ) 4 (C 6 H 4 N 2 ) 4 ], contains one Cd II atom, two <span class="hlt">3</span>-nitro-benzoate (NB) anions and two <span class="hlt">3</span>-cyano-pyridine (CPy) ligands. The two CPy ligands act as monodentate N(pyridine)-bonding ligands, while the two NB anions act as bidentate ligands through the carboxyl-ate <span class="hlt">O</span> atoms. The centrosymmetric dinuclear complex is generated by application of inversion symmetry, whereby the Cd II atoms are bridged by the carboxyl-ate <span class="hlt">O</span> atoms of two symmetry-related NB anions, thus completing the distorted N 2 <span class="hlt">O</span> 5 penta-gonal-bipyramidal coordination sphere of each Cd II atom. The benzene and pyridine rings are oriented at dihedral angles of 10.02 (7) and 5.76 (9)°, respectively. In the crystal, C-H⋯N hydrogen bonds link the mol-ecules, enclosing R 2 2 (26) ring motifs, in which they are further linked via C-H⋯<span class="hlt">O</span> hydrogen bonds, resulting in a three-dimensional network. In addition, π-π stacking inter-actions between parallel benzene rings and between parallel pyridine rings of adjacent mol-ecules [shortest centroid-to-centroid distances = <span class="hlt">3</span>.885 (1) and <span class="hlt">3</span>.712 (1) Å, respectively], as well as a weak C-H⋯π inter-action, may further stabilize the crystal structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MS%26E..292a2046W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MS%26E..292a2046W"><span>Influence of B2<span class="hlt">O</span><span class="hlt">3</span> content on sintering behaviour and dielectric properties of La2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Ca<span class="hlt">O</span>/Al2<span class="hlt">O</span><span class="hlt">3</span> glass-ceramic composites for LTCC applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, F. L.; Zhang, Y. W.; Chen, X. Y.; Mao, H. J.; Zhang, W. J.</p> <p>2018-01-01</p> <p>La2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Ca<span class="hlt">O</span> glasses with different B2<span class="hlt">O</span><span class="hlt">3</span> content were synthesized by melting method to produce glass/ceramic composites in this work. XRD and DSC results revealed that the diminution of B2<span class="hlt">O</span><span class="hlt">3</span> content was beneficial to increase the crystallization tendency of glass and improve the quality of crystalline phase, while decreasing the effect of glass during sintering process as sintering aids. The choice of glass/ceramic mass ratio was also influenced by the B2<span class="hlt">O</span><span class="hlt">3</span> content of glass. Dense samples sintered at 875 ºC showed good dielectric properties which meet the requirement of LTCC applications: moderate dielectric constant (7.8-9.4) and low dielectric loss (2.0×10-<span class="hlt">3</span>).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22103106','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22103106"><span>Fabrication of Si<span class="hlt">O</span>2@Zr<span class="hlt">O</span>2@Y2<span class="hlt">O</span><span class="hlt">3</span>:Eu<span class="hlt">3</span>+ core-multi-shell structured phosphor.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gao, Xuan; He, Diping; Jiao, Huan; Chen, Juan; Meng, Xin</p> <p>2011-08-01</p> <p>Zr<span class="hlt">O</span>2 interface was designed to block the reaction between Si<span class="hlt">O</span>2 and Y2<span class="hlt">O</span><span class="hlt">3</span> in Si<span class="hlt">O</span>2@Y2<span class="hlt">O</span><span class="hlt">3</span>:Eu coreshell structure phosphor. Si<span class="hlt">O</span>2@Zr<span class="hlt">O</span>2@Y2<span class="hlt">O</span><span class="hlt">3</span>:Eu core-multi-shell phosphors were successfully synthesized by combing an LBL method with a Sol-gel process. Based on electron microscopy, X-ray diffraction, and spectroscopy experiments, compelling evidence for the formation of the Y2<span class="hlt">O</span><span class="hlt">3</span>:Eu outer shell on Zr<span class="hlt">O</span>2 were presented. The presence of Zr<span class="hlt">O</span>2 layer on Si<span class="hlt">O</span>2 core can block the reaction of Si<span class="hlt">O</span>2 core and Y2<span class="hlt">O</span><span class="hlt">3</span> shell effectively. By this kind of structure, the reaction temperature of the Si<span class="hlt">O</span>2 core and Y2<span class="hlt">O</span><span class="hlt">3</span> shell in the Si<span class="hlt">O</span>2@Y2<span class="hlt">O</span><span class="hlt">3</span>:Eu core-shell structure phosphor can be increased about 200-300 degrees C and the luminescent intensity of this structure phosphor can be improved obviously. Under the excitation of ultraviolet (254 nm), the Eu<span class="hlt">3</span>+ ion mainly shows its characteristic red (611 nm, 5D0-7F2) emissions in the core-multi-shell particles from Y2<span class="hlt">O</span><span class="hlt">3</span>:Eu<span class="hlt">3</span>+ shells. The emission intensity of Eu<span class="hlt">3</span>+ ions can be tuned by the annealing temperatures, the number of coating times, and the thickness of Zr<span class="hlt">O</span>2 interface, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JSSCh.237..183T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JSSCh.237..183T"><span>Material properties of perovskites in the quasi-ternary system LaFe<span class="hlt">O</span><span class="hlt">3</span>-LaCo<span class="hlt">O</span><span class="hlt">3</span>-LaNi<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tietz, F.; Arul Raj, I.; Ma, Q.; Baumann, S.; Mahmoud, A.; Hermann, R. P.</p> <p>2016-05-01</p> <p>An overview is presented on the variation of electrical conductivity, oxygen permeation, and thermal expansion coefficient as a function of the composition of perovskites in the quasi-ternary system LaFe<span class="hlt">O</span><span class="hlt">3</span>-LaCo<span class="hlt">O</span><span class="hlt">3</span>-LaNi<span class="hlt">O</span><span class="hlt">3</span>. Powders of thirteen nominal perovskite compositions were synthesized under identical conditions by the Pechini method. The powder X-ray diffraction data of two series, namely La(Ni0.5Fe0.5)1-xCox<span class="hlt">O</span><span class="hlt">3</span> and LaNi0.5-xFexCo0.5<span class="hlt">O</span><span class="hlt">3</span>, are presented after the powders had been sintered at 1100 °C for 6 h in air. The measurements revealed a rhombohedral structure for all compositions except LaNi0.5Fe0.5<span class="hlt">O</span><span class="hlt">3</span> for which 60% rhombohedral and 40% orthorhombic phase was found. The maximum DC electrical conductivity value of the perovskites at 800 °C was 1229 S cm-1 for the composition LaCo<span class="hlt">O</span><span class="hlt">3</span> and the minimum was 91 S cm-1 for the composition LaCo0.5Fe0.5<span class="hlt">O</span><span class="hlt">3</span>. The oxygen permeation of samples with promising conductivities at 800 °C was one order of magnitude lower than that of La0.6Sr0.4Co0.8Fe0.2<span class="hlt">O</span><span class="hlt">3</span> (LSCF). The highest value of 0.017 ml cm-2 min-1 at 950 °C was obtained with LaNi0.5Co0.5<span class="hlt">O</span><span class="hlt">3</span>. The coefficients of thermal expansion varied in the range of 13.2×10-6 K-1 and 21.9×10-6 K-1 for LaNi0.5Fe0.5<span class="hlt">O</span><span class="hlt">3</span> and LaCo<span class="hlt">O</span><span class="hlt">3</span>, respectively. 57Fe Mössbauer spectroscopy was used as probe for the oxidation states, local environment and magnetic properties of iron ions as a function of chemical composition. The substitution had a great influence on the chemical properties of the materials.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_15 --> <div id="page_16" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="301"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29125165','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29125165"><span>Syntheses and characterization of phosphonates and diphosphonates of molybdenum, A4[(Mo<span class="hlt">O</span><span class="hlt">3</span>)5(<span class="hlt">O</span><span class="hlt">3</span>PR)2]·xH2<span class="hlt">O</span>, A2[Mo2<span class="hlt">O</span>5(<span class="hlt">O</span><span class="hlt">3</span>PR)2] and A2[Mo2<span class="hlt">O</span>5(<span class="hlt">O</span><span class="hlt">3</span>P-R-PO<span class="hlt">3</span>)] (A = K, Rb, Cs, Tl, NH4).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Elias Jesu Packiam, D; Vidyasagar, Kanamaluru</p> <p>2017-11-28</p> <p>Twenty new molybdenum phosphonates and diphosphonates have been synthesized and structurally characterized by single crystal and powder X-ray diffraction, CHN analyses, spectroscopic and thermal studies. Four of them are molecular phenyl- and benzyl-phosphonates containing discrete [(Mo<span class="hlt">O</span> <span class="hlt">3</span> ) 5 (<span class="hlt">O</span> <span class="hlt">3</span> PR) 2 ] 4- (R = Ph or CH 2 Ph) cyclic anions. The sixteen non-molecular compounds are layered isostructural phenylphosphonates, A 2 [Mo 2 <span class="hlt">O</span> 5 (<span class="hlt">O</span> <span class="hlt">3</span> PPh) 2 ] (A = NH 4 , Tl, Rb, Cs) and K 1.5 (H <span class="hlt">3</span> <span class="hlt">O</span>) 0.5 [Mo 2 <span class="hlt">O</span> 5 (<span class="hlt">O</span> <span class="hlt">3</span> PPh) 2 ] and the corresponding diphosphonate compounds with pillared anionic layers, A 2 [Mo 2 <span class="hlt">O</span> 5 (<span class="hlt">O</span> <span class="hlt">3</span> P(CH 2 ) <span class="hlt">3</span> PO <span class="hlt">3</span> )], A 2 [Mo 2 <span class="hlt">O</span> 5 (<span class="hlt">O</span> <span class="hlt">3</span> P(CH 2 ) 4 PO <span class="hlt">3</span> )] and A 2 [Mo 2 <span class="hlt">O</span> 5 (<span class="hlt">O</span> <span class="hlt">3</span> P(C 6 H 4 )PO <span class="hlt">3</span> )]. The A + ions reside in the interlayer region as well as in the cavities within the anionic layers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APLM....3c6104Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APLM....3c6104Z"><span>LaTi<span class="hlt">O</span><span class="hlt">3</span>/KTa<span class="hlt">O</span><span class="hlt">3</span> interfaces: A new two-dimensional electron gas system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zou, K.; Ismail-Beigi, Sohrab; Kisslinger, Kim; Shen, Xuan; Su, Dong; Walker, F. J.; Ahn, C. H.</p> <p>2015-03-01</p> <p>We report a new 2D electron gas (2DEG) system at the interface between a Mott insulator, LaTi<span class="hlt">O</span><span class="hlt">3</span>, and a band insulator, KTa<span class="hlt">O</span><span class="hlt">3</span>. For LaTi<span class="hlt">O</span><span class="hlt">3</span>/KTa<span class="hlt">O</span><span class="hlt">3</span> interfaces, we observe metallic conduction from 2 K to 300 K. One serious technological limitation of SrTi<span class="hlt">O</span><span class="hlt">3</span>-based conducting oxide interfaces for electronics applications is the relatively low carrier mobility (0.5-10 cm2/V s) of SrTi<span class="hlt">O</span><span class="hlt">3</span> at room temperature. By using KTa<span class="hlt">O</span><span class="hlt">3</span>, we achieve mobilities in LaTi<span class="hlt">O</span><span class="hlt">3</span>/KTa<span class="hlt">O</span><span class="hlt">3</span> interfaces as high as 21 cm2/V s at room temperature, over a factor of <span class="hlt">3</span> higher than observed in doped bulk SrTi<span class="hlt">O</span><span class="hlt">3</span>. By density functional theory, we attribute the higher mobility in KTa<span class="hlt">O</span><span class="hlt">3</span> 2DEGs to the smaller effective mass for electrons in KTa<span class="hlt">O</span><span class="hlt">3</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020024444','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020024444"><span>Calculation of Phase Equilibria in the Y2<span class="hlt">O</span><span class="hlt">3</span>-Yb2<span class="hlt">O</span><span class="hlt">3</span>-Zr<span class="hlt">O</span>2 System</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jacobson, Nathan S.; Liu, Zi-Kui; Kaufman, Larry; Zhang, Fan</p> <p>2001-01-01</p> <p>Rare earth oxide stabilized zirconias find a wide range of applications. An understanding of phase equilibria is essential to all applications. In this study, the available phase boundary data and thermodynamic data is collected and assessed. Calphad-type databases are developed to completely describe the Y2<span class="hlt">O</span><span class="hlt">3</span>-Zr<span class="hlt">O</span>2, Yb2<span class="hlt">O</span><span class="hlt">3</span>-Zr<span class="hlt">O</span>2, and Y2<span class="hlt">O</span><span class="hlt">3</span>-Yb2<span class="hlt">O</span><span class="hlt">3</span> systems. The oxide units are treated as components and regular and subregular solution models are used. The resultant calculated phase diagrams show good agreement with the experimental data. Then the binaries are combined to form the database for the Y2<span class="hlt">O</span><span class="hlt">3</span>-Yb2<span class="hlt">O</span><span class="hlt">3</span>-Zr<span class="hlt">O</span>2 psuedo-ternary.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013SSSci..26...72A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013SSSci..26...72A"><span>Impedance spectroscopy of V2<span class="hlt">O</span>5-Bi2<span class="hlt">O</span><span class="hlt">3</span>-BaTi<span class="hlt">O</span><span class="hlt">3</span> glass-ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Al-syadi, Aref M.; Yousef, El Sayed; El-Desoky, M. M.; Al-Assiri, M. S.</p> <p>2013-12-01</p> <p>The glasses within composition as: (80 - x)V2<span class="hlt">O</span>5/20Bi2<span class="hlt">O</span><span class="hlt">3</span>/xBaTi<span class="hlt">O</span><span class="hlt">3</span> with x = 2.5, 5, 7.5 and 10 mol% have been prepared. The glass transition (Tg) increases with increasing BaTi<span class="hlt">O</span><span class="hlt">3</span> content. Synthesized glasses ceramic containing BaTi4<span class="hlt">O</span>9, Ba<span class="hlt">3</span>TiV4<span class="hlt">O</span>15 nanoparticles of the order of 25-35 nm and 30-46 nm, respectively were estimated using XRD. The dielectric properties over wide ranges of frequencies and temperatures were investigated as a function of BaTi<span class="hlt">O</span><span class="hlt">3</span> content by impedance spectroscopy measurements. The hopping frequency, ωh, dielectric constant, ɛ', activation energies for the DC conduction, Eσ, the relaxation process, Ec, and stretched exponential parameter β of the glasses samples have been estimated. The, ωh,β, decrease from 51.63 to 0.31 × 106 (s-1), 0.84 to 0.79 with increasing BaTi<span class="hlt">O</span><span class="hlt">3</span> respectively. Otherwise, the Eσ, increase from 0.279 to 0.306 eV with increasing BaTi<span class="hlt">O</span><span class="hlt">3</span>. The value of dielectric constant equal 9.5·103 for the 2.5BaTi<span class="hlt">O</span><span class="hlt">3</span>/77.5V2<span class="hlt">O</span>5/20Bi2<span class="hlt">O</span><span class="hlt">3</span> glasses-ceramic at 330 K for 1 KHz which is ten times larger than that of same glasses composition. Finally the relaxation properties of the investigated glasses are presented in the electric modulus formalism, where the relaxation time and the respective activation energy were determined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16987701','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16987701"><span>Spectroscopic properties of Er(<span class="hlt">3</span>+)/Yb(<span class="hlt">3</span>+) co-doped Bi(2)<span class="hlt">O</span>(<span class="hlt">3</span>)-B(2)<span class="hlt">O</span>(<span class="hlt">3</span>)-Ge<span class="hlt">O</span>(2) glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Xudong; Xu, Tiefeng; Nie, Qiuhua; Dai, Shixun; Shen, Xiang; Zhang, Xianghua</p> <p>2007-05-01</p> <p>Er(<span class="hlt">3</span>+)/Yb(<span class="hlt">3</span>+) co-doped 60Bi(2)<span class="hlt">O</span>(<span class="hlt">3</span>)-(40 - x)B(2)<span class="hlt">O</span>(<span class="hlt">3</span>)-xGe<span class="hlt">O</span>(2) (BBG; x=0, 5, 10, 15 mol%) glasses that are suitable for fiber lasers, amplifiers have been fabricated and characterized. The absorption spectra, emission spectra, and lifetime of the (4)I(13/2) level and quantum efficiency of Er(<span class="hlt">3</span>+):(4)I(13/2) --> (4)I(15/2) transition were measured and calculated. With the substitution of Ge<span class="hlt">O</span>(2) for B(2)<span class="hlt">O</span>(<span class="hlt">3</span>), both Delta lambda(eff) and sigma(e) decrease from 75 to 71 nm and 9.88 to 8.12 x 10(-21) cm(2), respectively. The measured lifetime of the (4)I(13/2) level and quantum efficiency of Er(<span class="hlt">3</span>+):(4)I(13/2) --> (4)I(15/2) transition increase from 1.18 to 1.5 ms and 36.2% to 43.2%, respectively. The emission spectra of Er(<span class="hlt">3</span>+):(4)I(13/2) --> (4)I(15/2) transition was also analyzed using a peak-fit routine, and an equivalent four-level system was proposed to estimate the stark splitting for the (4)I(15/2) and (4)I(13/2) levels of Er(<span class="hlt">3</span>+) in the BBG glasses. The results indicate that the (4)I(13/2) --> (4)I(15/2) emission of Er(<span class="hlt">3</span>+) can be exhibit a considerable broadening due to a significant enhance the peak A, and D emission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RJPCA..91.1939A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RJPCA..91.1939A"><span>Synthesis and properties of γ-Ga2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> solid solutions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Afonasenko, T. N.; Leont'eva, N. N.; Talzi, V. P.; Smirnova, N. S.; Savel'eva, G. G.; Shilova, A. V.; Tsyrul'nikov, P. G.</p> <p>2017-10-01</p> <p>The textural and structural properties of mixed oxides Ga2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>, obtained via impregnating γ-Al2<span class="hlt">O</span><span class="hlt">3</span> with a solution of Ga(NO<span class="hlt">3)3</span> and subsequent heat treatment, are studied. According to the results from X-ray powder diffraction, gallium ions are incorporated into the structure of aluminum oxide to form a solid solution of spinel-type γ-Ga2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> up to a Ga2<span class="hlt">O</span><span class="hlt">3</span> content of 50 wt % of the total weight of the sample, accompanied by a reduction in the specific surface area, volume, and average pore diameter. It is concluded that when the Ga2<span class="hlt">O</span><span class="hlt">3</span> content exceeds 50 wt %, the β-Ga2<span class="hlt">O</span><span class="hlt">3</span> phase is observed along with γ-Ga2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> solid solution. 71Ga and 27Al NMR spectroscopy shows that gallium replaces aluminum atoms from the tetrahedral position to the octahedral coordination in the structure of γ-Ga2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AtmEn.182..296L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AtmEn.182..296L"><span>Characteristics and sources of nitrous acid in an urban atmosphere of northern China: Results from 1-yr continuous observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Dandan; Xue, Likun; Wen, Liang; Wang, Xinfeng; Chen, Tianshu; Mellouki, Abdelwahid; Chen, Jianmin; Wang, Wenxing</p> <p>2018-06-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) is a key reservoir of the hydroxyl radical (OH) and plays a central role in the atmospheric chemistry. To understand the sources and impact of <span class="hlt">HONO</span> in the polluted atmosphere of northern China, continuous measurements of <span class="hlt">HONO</span> and related parameters were conducted from September 2015 to August 2016 at an urban site in Ji'nan, the capital city of Shandong province. <span class="hlt">HONO</span> showed well-defined seasonal and diurnal variation patterns with clear wintertime and nighttime concentration peaks. Elevated <span class="hlt">HONO</span> concentrations (e.g., over 5 ppbv) were frequently observed with a maximum value of 8.36 ppbv. The <span class="hlt">HONO</span>/NOX ratios of direct vehicle emissions varied in the range of 0.29%-0.87%, with a mean value of 0.53%. An average NO2-to-<span class="hlt">HONO</span> nighttime conversion frequency (khet) was derived to be 0.0068 ± 0.0045 h-1 from 107 <span class="hlt">HONO</span> formation cases. A detailed <span class="hlt">HONO</span> budget analysis suggests an unexplained daytime missing source of 2.95 ppb h-1 in summer, which is about seven times larger than the homogeneous reaction of NO with OH. The effect of <span class="hlt">HONO</span> on OH production was also quantified. <span class="hlt">HONO</span> photolysis was the uppermost source of local OH radical throughout the daytime. This study provides the year-round continuous record of ambient <span class="hlt">HONO</span> in the North China Plain, and offers some insights into the characteristics, sources and impacts of <span class="hlt">HONO</span> in the polluted atmospheres of China.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19036928','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19036928"><span>Postperovskite phase equilibria in the MgSi<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> system.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tsuchiya, Jun; Tsuchiya, Taku</p> <p>2008-12-09</p> <p>We investigate high-P,T phase equilibria of the MgSi<span class="hlt">O</span>(<span class="hlt">3</span>)-Al(2)<span class="hlt">O</span>(<span class="hlt">3</span>) system by means of the density functional ab initio computation methods with multiconfiguration sampling. Being different from earlier studies based on the static substitution properties with no consideration of Rh(2)<span class="hlt">O</span>(<span class="hlt">3</span>)(II) phase, present calculations demonstrate that (i) dissolving Al(2)<span class="hlt">O</span>(<span class="hlt">3</span>) tends to decrease the postperovskite transition pressure of MgSi<span class="hlt">O</span>(<span class="hlt">3</span>) but the effect is not significant ( approximately -0.2 GPa/mol% Al(2)<span class="hlt">O</span>(<span class="hlt">3</span>)); (ii) Al(2)<span class="hlt">O</span>(<span class="hlt">3</span>) produces the narrow perovskite+postperovskite coexisting P,T area (approximately 1 GPa) for the pyrolitic concentration (x(Al2<span class="hlt">O</span><span class="hlt">3</span>) approximately 6 mol%), which is sufficiently responsible to the deep-mantle D'' seismic discontinuity; (iii) the transition would be smeared (approximately 4 GPa) for the basaltic Al-rich composition (x(Al2<span class="hlt">O</span><span class="hlt">3</span>) approximately 20 mol%), which is still seismically visible unless iron has significant effects; and last (iv) the perovskite structure spontaneously changes to the Rh(2)<span class="hlt">O</span>(<span class="hlt">3</span>)(II) with increasing the Al concentration involving small displacements of the Mg-site cations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009ACPD....924193R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009ACPD....924193R"><span>Formaldehyde and its relation to CO, PAN, and SO2 in the Houston-Galveston airshed</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rappenglück, B.; Dasgupta, P. K.; Leuchner, M.; Li, Q.; Luke, W.</p> <p>2009-11-01</p> <p>The Houston-Galveston Airshed (HGA) is one of the major metropolitan areas in the US that is classified as a nonattainment area of Federal ozone standards. Formaldehyde (<span class="hlt">HCHO</span>) is a key species in understanding ozone related air pollution; some of the highest <span class="hlt">HCHO</span> concentrations in North America have been reported for the HGA. We report on <span class="hlt">HCHO</span> measurements in the HGA from summer 2006. Among several sites, maximum <span class="hlt">HCHO</span> mixing ratios were observed in the Houston Ship Channel (HSC), a region with a very high density of industrial/petrochemical operations. <span class="hlt">HCHO</span> levels at the Moody Tower (MT) site close to downtown were dependent on the wind direction: southerly maritime winds brought in background levels (0.5-1 ppbv) while trajectories originating in the HSC resulted in high HCH (up to 31.5 ppbv). Based on the best multiparametric linear regression model fit, the <span class="hlt">HCHO</span> levels at the MT site can be accounted for as follows: 38.5±12.<span class="hlt">3</span>% from primary vehicular emissions (using CO as an index of vehicular emission), 24.1±17.7% formed photochemically (using peroxyacetic nitric anhydride (PAN) as an index of photochemical activity) and 8.9±11.2% from industrial emissions (using SO2 as an index of industrial emissions). The balance 28.5±12.7% constituted the residual which cannot be easily ascribed to the above categories and/or which is transported into the HGA. The CO related <span class="hlt">HCHO</span> fraction is dominant during the morning rush hour (06:00-09:00 h, all times are given in CDT); on a carbon basis, <span class="hlt">HCHO</span> emissions are up to 0.7% of the CO emissions. The SO2 related <span class="hlt">HCHO</span> fraction is significant between 09:00-12:00 h. After 12:00 h <span class="hlt">HCHO</span> is largely formed through secondary processes. The <span class="hlt">HCHO</span>/PAN ratios are dependent on the SO2 levels. The SO2 related <span class="hlt">HCHO</span> fraction at the downtown site originates in the ship channel. Aside from traffic-related primary <span class="hlt">HCHO</span> emissions, <span class="hlt">HCHO</span> of industrial origin serves as an appreciable source for OH in the morning.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhyB..521..376Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhyB..521..376Z"><span>Preparation and characterization of BiFe<span class="hlt">O</span><span class="hlt">3</span>/La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> heterostructure grown on SrTi<span class="hlt">O</span><span class="hlt">3</span> substrate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Chenwei; Zhou, Chaochao; Chen, Changle</p> <p>2017-09-01</p> <p>In this paper, BiFe<span class="hlt">O</span><span class="hlt">3</span>/La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> heterostructure is fabricated on the SrTi<span class="hlt">O</span> (100) substrate using the pulsed laser deposition method (PLD). Magnetization hystersis loops of the BiFe<span class="hlt">O</span><span class="hlt">3</span>/La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> heterostructure are obtained at 300 K and 80 K. The heterostructure exhibits evident ferromagnetic characteristic at both room temperature and 80 K. At 80 K, magnetization of the heterostructure is stronger than room temperature magnetic measure. The temperature dependence of resistance of the heterostructure with different currents is also studied. With different currents, there appears to be a peak resistance about 180 K. When I is 50 uA, ΔR is 68.4%. And when I is 100 uA, ΔR is 79.<span class="hlt">3</span>%. The BiFe<span class="hlt">O</span><span class="hlt">3</span>/La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> heterostructure exhibits a positive colossal magnetoresistance (MR) effect over a temperature range of 80-300 K. In our heterostructure, maximum magnetic resistance appears in 210 K, and MR = 44.34%. Mechanism analysis of the leakage current at room temperature shows that the leakage current is the interface-limited Schottky emission, but not dominated by the Poole-Frenkel emission or SCLC.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvP...9e4004C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvP...9e4004C"><span>Electron Mobility in γ -Al2<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Christensen, D. V.; Frenkel, Y.; Schütz, P.; Trier, F.; Wissberg, S.; Claessen, R.; Kalisky, B.; Smith, A.; Chen, Y. Z.; Pryds, N.</p> <p>2018-05-01</p> <p>One of the key issues in engineering oxide interfaces for electronic devices is achieving high electron mobility. SrTi<span class="hlt">O</span><span class="hlt">3</span> -based interfaces with high electron mobility have gained a lot of interest due to the possibility of combining quantum phenomena with the many functionalities exhibited by SrTi<span class="hlt">O</span><span class="hlt">3</span> . To date, the highest electron mobility (140 000 cm2/V s at 2 K) is obtained by interfacing perovskite SrTi<span class="hlt">O</span><span class="hlt">3</span> with spinel γ -Al2<span class="hlt">O</span><span class="hlt">3</span> . The origin of the high mobility, however, remains poorly understood. Here, we investigate the scattering mechanisms limiting the mobility in γ -Al2<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> at temperatures between 2 and 300 K and over a wide range of sheet carrier densities. For T >150 K , we find that the mobility is limited by longitudinal optical phonon scattering. For large sheet carrier densities (>8 ×1013 cm-2 ), the screened electron-phonon coupling leads to room-temperature mobilities up to μ ˜12 cm2/V s . For 5 K <T <150 K , the mobility scales as approximately T-2 , consistent with electron-electron scattering limiting the electron mobility. For T <5 K and at an optimal sheet carrier density of approximately 4 ×1014 cm-2 , the electron mobility is found to exceed 100 000 cm2/V s . At sheet carrier densities less than the optimum, the electron mobility decreases rapidly, and the current flow becomes highly influenced by domain walls and defects in the near-interface region of SrTi<span class="hlt">O</span><span class="hlt">3</span> . At carrier densities higher than the optimum, the SrTi<span class="hlt">O</span><span class="hlt">3</span> heterostructure gradually becomes bulk conducting, and the electron mobility decreases to approximately 20 000 cm2/V s . We argue that the high electron mobility observed arises from a spatial separation of donors and electrons with oxygen-vacancy donors preferentially forming at the interface, whereas the itinerant electrons extend deeper into SrTi<span class="hlt">O</span><span class="hlt">3</span> . Understanding the scattering mechanism in γ -Al2<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> paves the way for creation of high-mobility nanoscale electronic devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12775055','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12775055"><span>Alkaline hydrolysis of the cyclic nitramine explosives RDX, HMX, and CL-20: new insights into degradation pathways obtained by the observation of novel intermediates.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Balakrishnan, Vimal K; Halasz, Annamaria; Hawari, Jalal</p> <p>2003-05-01</p> <p>Hexahydro-1,<span class="hlt">3</span>,5-trinitro-1,<span class="hlt">3</span>,5-triazine (RDX, I) and octahydro-1,<span class="hlt">3</span>,5,7-tetranitro-1,<span class="hlt">3</span>,5,7-tetrazocine (HMX) hydrolyze at pH > 10 to form end products including NO2-, <span class="hlt">HCHO</span>, HCOOH, NH<span class="hlt">3</span>, and N2<span class="hlt">O</span>, but little information is available on intermediates, apart from the tentatively identified pentahydro-<span class="hlt">3</span>,5-dinitro-1,<span class="hlt">3</span>,5-triazacyclohex-1-ene (II). Despite suggestions that RDX and HMX contaminated groundwater could be economically treated via alkaline hydrolysis, the optimization of such a process requires more detailed knowledge of intermediates and degradation pathways. In this study, we hydrolyzed the monocyclic nitramines RDX, MNX (hexahydro-1-nitroso-<span class="hlt">3</span>,5-dinitro-1,<span class="hlt">3</span>,5-triazine), and HMX in aqueous solution (pH 10-12.<span class="hlt">3</span>) and found that nitramine removal was accompanied by formation of 1 molar equiv of nitrite and the accumulation of the key ring cleavage product 4-nitro-2,4-diazabutanal (4-NDAB, <span class="hlt">O</span>2NNHCH2NHCHO). Most of the remaining C and N content of RDX, MNX, and HMX was found in <span class="hlt">HCHO</span>, N2<span class="hlt">O</span>, HCOOH, and NH<span class="hlt">3</span>. Consequently, we selected RDX as a model compound and hydrolyzed it in aqueous acetonitrile solutions (pH 12.<span class="hlt">3</span>) in the presence and absence of hydroxypropyl-beta-cyclodextrin (HP-beta-CD) to explore other early intermediates in more detail. We observed a transient LC-MS peak with a [M-H] at 192 Da that was tentatively identified as 4,6-dinitro-2,4,6-triaza-hexanal (<span class="hlt">O</span>2NNHCH2NNO2CH2NHCHO, III) considered as the hydrolyzed product of II. In addition, we detected another novel intermediate with a [M-H] at 148 Da that was tentatively identified as a hydrolyzed product of III, namely, 5-hydroxy-4-nitro-2,4-diaza-pentanal (HOCH2NNO2CH2NHCHO, IV). Both III and IV can act as precursors to 4-NDAB. In the case of the polycyclic nitramine 2,4,6,8,10,12-hexanitro-2,4,6,8,10,12-hexaazaisowurtzitane (CL-20), denitration (two NO2-) also led to the formation of HCOOH, NH<span class="hlt">3</span>, and N2<span class="hlt">O</span>, but neither <span class="hlt">HCHO</span> nor 4-NDAB were detected. The results provide strong evidence that initial denitration</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EPJWC.16201052F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EPJWC.16201052F"><span>Physical and electrical properties of SrTi<span class="hlt">O</span><span class="hlt">3</span> and SrZr<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fashren Muhamad, Norhizatol; Aina Maulat Osman, Rozana; Sobri Idris, Mohd; Yasin, Mohd Najib Mohd</p> <p>2017-11-01</p> <p>Perovskite type oxide strontium titanate (SrTi<span class="hlt">O</span><span class="hlt">3</span>) and strontium zirconate (SrZr<span class="hlt">O</span><span class="hlt">3</span>) ceramic powder has been synthesized using conventional solid state reaction method. The powders were mixed and ground undergone calcinations at 1400°C for 12 h and sintered at 1550°C for 5h. X-ray Diffraction exposes physical properties SrTi<span class="hlt">O</span><span class="hlt">3</span> which exhibit cubic phase (space group: pm-<span class="hlt">3</span>m) at room temperature meanwhile SrZr<span class="hlt">O</span><span class="hlt">3</span> has Orthorhombic phase (space group: pnma). The electrical properties such as dielectric constant (ɛr), dielectric loss (tan δ), and conductivity (σ) were studied in variation temperature and frequency. High dielectric constant of SrTi<span class="hlt">O</span><span class="hlt">3</span> and SrZr<span class="hlt">O</span><span class="hlt">3</span> were observed at 10 kHz for both samples about 240 and 21 respectively at room temperature. The dielectric loss of SrTi<span class="hlt">O</span><span class="hlt">3</span> and SrZr<span class="hlt">O</span><span class="hlt">3</span> is very low loss value approximately 0.00076 and 0.67512 indicates very good dielectric.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1164884','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1164884"><span><span class="hlt">3</span>-<span class="hlt">O</span>-methyl sugars as constituents of glycoproteins. Identification of <span class="hlt">3</span>-<span class="hlt">O</span>-methylgalactose and <span class="hlt">3</span>-<span class="hlt">O</span>-methylmannose in pulmonate gastropod haemocyanins.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hall, R L; Wood, E J; Kamberling, J P; Gerwig, G J; Vliegenthart, F G</p> <p>1977-01-01</p> <p>In addition to the already knownonosaccharides fucose, xylose, mannose, galactose, glucose, N-acetylgalactosamine and N-acetylglucosamine, the carbohydrate part of the haemocyanin from Helix pomatia (Roman snail) contains <span class="hlt">3</span>-<span class="hlt">O</span>-methylgalactose, and that from Lymnaea stagnalis (a freshwater snail) <span class="hlt">3</span>-<span class="hlt">O</span>-methylgalactose and <span class="hlt">3</span>-<span class="hlt">O</span>-methylmannose. The <span class="hlt">3</span>-<span class="hlt">O</span>-methyl sugars were identified by g.l.c.-mas spectrometry of the corresponding trimethylsilyl methyl glycosides and the alditol acetates, and by co-chromatography with the synthetic reference substances. PMID:889564</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..MARW49008Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..MARW49008Z"><span>Growth and Transport Studies of LaTi<span class="hlt">O</span><span class="hlt">3</span> / KTa<span class="hlt">O</span><span class="hlt">3</span> Heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zou, K.; Walker, F. J.; Ahn, C. H.</p> <p>2014-03-01</p> <p>Perovskite oxide heterostructures provide a rich platform for exploring emergent electronic properties, such as 2D electron gases (2DEGs) at interfaces. In this talk, we present results on the growth of LaTi<span class="hlt">O</span><span class="hlt">3</span> / KTa<span class="hlt">O</span><span class="hlt">3</span> heterostructures by molecular beam epitaxy and subsequent measurements of transport properties. Although both oxide materials are insulating in the bulk, metallic conduction is observed from T = 2 - 300 K. We achieve a room temperature carrier mobility of ~ 25 cm2 /Vs at a carrier density of ~ 1014 /cm2. By comparison, 2DEGs in LaTi<span class="hlt">O</span><span class="hlt">3</span> / SrTi<span class="hlt">O</span><span class="hlt">3</span> and LaAl<span class="hlt">O</span><span class="hlt">3</span> / SrTi<span class="hlt">O</span><span class="hlt">3</span> have lower carrier mobility, but the same carrier density. We attribute some of the increase in mobility to the smaller band effective mass of the Ta 4d electrons compared to the Ti <span class="hlt">3</span>d electrons.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20058930','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20058930"><span>Development of a mechanism for nitrate photochemistry in snow.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bock, Josué; Jacobi, Hans-Werner</p> <p>2010-02-04</p> <p>A reaction mechanism to reproduce photochemical processes in the snow is reported. We developed a box model to represent snow chemistry. Constrained by laboratory experiments carried out with artificial snow, we deduced first a reaction mechanism for N-containing species including 13 reactions. An optimization tool was developed to adjust systematically unknown photolysis rates of nitrate and nitrite (NO(2)(-)) and transfer rates of nitrogen oxides from the snow to the gas phase resulting in an optimum fit with respect to the experimental data. Further experiments with natural snow samples are presented, indicating that NO(2)(-) concentrations were much lower than in the artificial snow experiments. These observations were used to extend the reaction mechanism into a more general scheme including hydrogen peroxide (H(2)<span class="hlt">O</span>(2)) and formaldehyde (<span class="hlt">HCHO</span>) chemistry leading to a set of 18 reactions. The simulations indicate the importance of H(2)<span class="hlt">O</span>(2) and <span class="hlt">HCHO</span> as either a source or sink of hydroxyl radicals in the snow photochemistry mechanism. The addition of H(2)<span class="hlt">O</span>(2) and <span class="hlt">HCHO</span> in the mechanism allows the reproduction of the observed low NO(2)(-) concentration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002EGSGA..27.6606R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002EGSGA..27.6606R"><span>Off-axis Doas Measurements At Observatoire De Haute Provence During 2001</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Roozendael, M. Van; de Mazière, M.; Fayt, C.; Hendrick, F.; Hermans, C.</p> <p></p> <p>Since December 2000, a ground-based off-axis DOAS spectrometer has been operated by BIRA-IASB in the South of France at the Observatoire de Haute Provence (OHP, 44°N). The design of the instrument allows automated observations of the scattered light alternatively in the zenith direction and at 10° above the horizon (off-axis geometry). The spectrometer is polarised and follows the azimuth of the sun. Its temperature is regulated and it is equipped with a Princeton Instruments/ Hammamatsu cooled diode array detector. Observations are made every 5 minutes in the 320-390 nm range. The analysis of the spectra recorded between January and December 2001 demonstrates the sensitivity of the measurements to tropospheric contents of NO2, <span class="hlt">HCHO</span>, <span class="hlt">O</span><span class="hlt">3</span> and Br<span class="hlt">O</span>. Results show a large seasonality in the <span class="hlt">HCHO</span> content with maximum values in summer. The tropospheric Br<span class="hlt">O</span> column is found to be stable over the year in the range of approximately 1.5-2 x1013 molec/cm2, roughly consistent with GOME observations at Northern mid-latitudes. Large increases of the Br<span class="hlt">O</span> concentration are observed in summer likely due to local pollution in the vicinity of the station.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29306155','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29306155"><span>Interaction mechanisms between α-Fe2<span class="hlt">O</span><span class="hlt">3</span>, γ-Fe2<span class="hlt">O</span><span class="hlt">3</span> and Fe<span class="hlt">3</span><span class="hlt">O</span>4 nanoparticles and Citrus maxima seedlings.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Junli; Hu, Jing; Xiao, Lian; Wang, Yunqiang; Wang, Xilong</p> <p>2018-06-01</p> <p>The interactions between α-Fe 2 <span class="hlt">O</span> <span class="hlt">3</span> , γ-Fe 2 <span class="hlt">O</span> <span class="hlt">3</span> , and Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 nanoparticles (NPs) and Citrus maxima seedlings were examined so as to better understand possible particle applications as an Fe source for crop plants. NPs toxicity to the exposed plant was investigated as well. The α- and γ-Fe 2 <span class="hlt">O</span> <span class="hlt">3</span> NPs were accumulated by plant root cells through diapirism and endocytosis, respectively, but translocation to the shoots was negligible. Analysis of malondialdehyde (MDA), soluble protein content, and antioxidant enzyme activity revealed that Fe deficiency induced strong oxidative stress in Citrus maxima seedlings, which followed an order of Fe deficiency>Fe <span class="hlt">3</span>+ >α-Fe 2 <span class="hlt">O</span> <span class="hlt">3</span> , γ-Fe 2 <span class="hlt">O</span> <span class="hlt">3</span> NPs>Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 NPs. However, the chlorophyll leaf content of plants exposed to α-Fe 2 <span class="hlt">O</span> <span class="hlt">3</span> , γ-Fe 2 <span class="hlt">O</span> <span class="hlt">3</span> , Fe <span class="hlt">3</span> <span class="hlt">O</span> 4 NPs and Fe <span class="hlt">3</span>+ were significantly reduced by 31.1%, 14.8%, 18.8% and 22.0%, respectively, relative to the control. Furthermore, RT-PCR analysis revealed no up-regulation of AHA and Nramp<span class="hlt">3</span> genes in Citrus maxima roots; however, the relative FRO2 gene expression upon exposure to iron oxide NPs was 1.4-2.8-fold higher than the control. Ferric reductase activity was consistently enhanced upon iron oxide NPs exposure. These findings advance understanding of the interaction mechanisms between metal oxide NPs and plants, and provide important knowledge need for the possible application of these materials in agriculture. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ApPhL.103h2906Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ApPhL.103h2906Y"><span>Fatigue mechanism of textured Pb(Mg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yan, Yongke; Zhou, Yuan; Gupta, Shashaank; Priya, Shashank</p> <p>2013-08-01</p> <p>Grain orientation, BaTi<span class="hlt">O</span><span class="hlt">3</span> heterogeneous template content, and electrode materials are expected to play an important role in controlling the polarization fatigue behavior of ⟨001⟩ textured Pb(Mg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> ceramics. A comparative analysis with randomly oriented ceramics showed that ⟨001⟩ grain orientation/texture exhibits improved fatigue characteristics due to the reduced switching activation energy and high domain mobility. The hypothesis was validated from the systematic characterization of polarization—electric field behavior and domain wall density. The defect accumulation at the grain boundary and clamping effect arising from the presence of BaTi<span class="hlt">O</span><span class="hlt">3</span> heterogeneous template in the final microstructure was found to be the main cause for polarization degradation in textured ceramic.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006ApPhL..89u2906Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006ApPhL..89u2906Y"><span>Ferroelectric enhancement in heterostructured Zn<span class="hlt">O</span> /BiFe<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> film</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Shengwen; Chen, Rui; Zhang, Guanjun; Cheng, Jinrong; Meng, Zhongyan</p> <p>2006-11-01</p> <p>The authors have prepared heterostructured Zn<span class="hlt">O</span> /BiFe<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> (BFO-PT) composite film and BFO-PT film on Pt /Ti/Si<span class="hlt">O</span>2/Si substrates by pulsed-laser deposition. The structure and morphologies of the films were characterized by x-ray diffraction (XRD) and scanning electron microscope. XRD results show that both films are perovskite structured last with different orientations. The leakage current density in the Zn<span class="hlt">O</span> /BFO-PT film was found to be nearly two orders of magnitude lower. This could be due to the introduced Zn<span class="hlt">O</span> layer behaving as a Schottky barrier between the BFO-PT film and top electrodes. The dramatic ferroelectric enhancement in Zn<span class="hlt">O</span> /BFO-PT film is mostly ascribed to the improved insulation.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19498708','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19498708"><span>Broadband infrared luminescence from Li2<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Zn<span class="hlt">O-SiO</span>2 glasses doped with Bi2<span class="hlt">O</span><span class="hlt">3</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Peng, Mingying; Qiu, Jianrong; Chen, Danping; Meng, Xiangeng; Zhu, Congshan</p> <p>2005-09-05</p> <p>The broadband emission in the 1.2~1.6mum region from Li2<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Zn<span class="hlt">O-SiO</span>2 ( LAZS ) glass codoped with 0.01mol.%Cr2<span class="hlt">O</span><span class="hlt">3</span> and 1.0mol.%Bi2<span class="hlt">O</span><span class="hlt">3</span> when pumped by the 808nm laser at room temperature is not initiated from Cr4+ ions, but from bismuth, which is remarkably different from the results reported by Batchelor et al. The broad ~1300nm emission from Bi2<span class="hlt">O</span><span class="hlt">3</span>-containing LAZS glasses possesses a FWHM ( Full Width at Half Maximum ) more than 250nm and a fluorescent lifetime longer than 500mus when excited by the 808nm laser. These glasses might have the potential applications in the broadly tunable lasers and the broadband fiber amplifiers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol30/pdf/CFR-2011-title40-vol30-sec600-113-08.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol30/pdf/CFR-2011-title40-vol30-sec600-113-08.pdf"><span>40 CFR 600.113-08 - Fuel economy calculations for FTP, HFET, US06, SC03 and cold temperature FTP tests.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-07-01</p> <p>... paragraph (g) of this section. CH<span class="hlt">3</span>OH = Grams/mile CH<span class="hlt">3</span>OH (methanol) as obtained in paragraph (d) of this..., additionally for methanol-fueled automobiles, methanol (CH<span class="hlt">3</span>OH) and formaldehyde (<span class="hlt">HCHO</span>); and additionally for... for HC, CO and CO2; and, additionally for methanol-fueled automobiles, CH<span class="hlt">3</span>OH and <span class="hlt">HCHO</span>; and...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20419263','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20419263"><span>A study of the composition of the products of laser-induced breakdown of hexogen, octogen, pentrite and trinitrotoluene using selected ion flow tube mass spectrometry and UV-Vis spectrometry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sovová, Kristýna; Dryahina, Kseniya; Spanel, Patrik; Kyncl, Martin; Civis, Svatopluk</p> <p>2010-05-01</p> <p>Four types of explosives were studied using a combination of Laser Induced Breakdown Spectroscopy (LIBS) and Selected Ion Flow Tube Mass Spectrometry (SIFT-MS). The LIBS technique uses short laser pulses (ArF excimer laser) as the energy source to convert small amounts samples into plasma and to produce the emission from their molecular fragments or atoms. SIFT-MS is a novel method for absolute quantification based on chemical ionization using three precursor ions, with the capability to determine concentrations of trace gases and vapours of volatile organic compounds in real time. This is the first time that SIFT-MS has been used to study the release of NO, NO(2), HCN, HNO(<span class="hlt">3</span>), <span class="hlt">HONO</span>, <span class="hlt">HCHO</span> and C(2)H(2) after a laser-induced breakdown of pure explosive compounds HMX (1,<span class="hlt">3</span>,5,7-tetranitro-1,<span class="hlt">3</span>,5,7-tetraazacyclo-octane), RDX (1,<span class="hlt">3</span>,5-trinitro-2-oxo-1,<span class="hlt">3</span>,5-triazacyclo-hexane), PETN (pentaerithrityl-tetranitrate) and TNT (2,4,6-trinitrotoluene) in solid form. The radiation emitted after excitation was analysed using a time resolving UV-Vis spectrometer with a ICCD detector. Electronic bands of the CN radical (388 nm), the Swan system of the C(2) radical (512 nm), the NH radical (336 nm), the OH radical (308.4 nm) and atomic lines of oxygen, nitrogen and hydrogen were identified. Vibrational and excitation temperatures were determined from the intensity distributions and a scheme of chemical reactions responsible for the formation of the observed species was proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..441..429S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..441..429S"><span>Ultrathin Bi2WO6 nanosheet decorated with Pt nanoparticles for efficient formaldehyde removal at room temperature</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sun, Dong; Le, Yao; Jiang, Chuanjia; Cheng, Bei</p> <p>2018-05-01</p> <p>Two-dimensional (2D) ultrathin bismuth tungstate (Bi2WO6) nanosheets (BWO-NS) with a thickness of approximately 4.0 nm were synthesized by a one-step hydrothermal method, and decorated with platinum (Pt) nanoparticles (NPs) via an impregnation/borohydride-reduction approach. The as-prepared ultrathin Pt-BWO-NS exhibited superior catalytic activity for removing gaseous formaldehyde (<span class="hlt">HCHO</span>) at ambient temperature, in comparison with bulk counterpart with Bi2WO6 sheet thickness of tens of nanometers. The ultrathin structure endowed the Pt-BWO-NS sample with larger specific surface area, which can provide abundant surface active sites for <span class="hlt">HCHO</span> adsorption and facilitate the homogeneous dispersion of Pt NPs. X-ray photoelectron spectroscopy and hydrogen temperature-programmed reduction analyses revealed the interaction between the Bi2WO6 support and Pt species, which is crucial for activating surface oxygen atoms to participate in the catalytic <span class="hlt">HCHO</span> oxidation process. By conducting in situ diffuse reflectance infrared Fourier transform spectroscopy under different atmospheres, i.e., gaseous <span class="hlt">HCHO</span> in nitrogen or oxygen (<span class="hlt">O</span>2), the reaction mechanism and the role of <span class="hlt">O</span>2 were elucidated, with dioxymethylene, formate and linearly adsorbed carbon monoxide identified as the main reaction intermediates. This study may provide new enlightenment on fabricating novel 2D nanomaterials for efficient indoor air purification and potentially other environmental applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001ApPhL..79.1324T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001ApPhL..79.1324T"><span>Interface ferromagnetism in oxide superlattices of CaMn<span class="hlt">O</span><span class="hlt">3</span>/CaRu<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takahashi, K. S.; Kawasaki, M.; Tokura, Y.</p> <p>2001-08-01</p> <p>Oxide superlattices composed of antiferromagnetic insulator layers of CaMn<span class="hlt">O</span><span class="hlt">3</span> (10 unit cells) and paramagnetic metal layers of CaRu<span class="hlt">O</span><span class="hlt">3</span> (N unit cells) were fabricated on LaAl<span class="hlt">O</span><span class="hlt">3</span> substrates by pulsed-laser deposition. All the superlattices show ferromagnetic transitions at an almost identical temperature (TC˜95 K) and negative magnetoresistance below TC. Each magnetization and magnetoconductance of the whole superlattice at 5 K is constant and independent of CaRu<span class="hlt">O</span><span class="hlt">3</span> layer thickness when normalized by the number of the interfaces between CaMn<span class="hlt">O</span><span class="hlt">3</span> and CaRu<span class="hlt">O</span><span class="hlt">3</span>. These results indicate that the ferromagnetism shows up only at the interface and is responsible for the magnetoresistance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007JaJAP..46.7089K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007JaJAP..46.7089K"><span>Processing of Piezoelectric (Li,Na,K)Nb<span class="hlt">O</span><span class="hlt">3</span> Porous Ceramics and (Li,Na,K)Nb<span class="hlt">O</span><span class="hlt">3</span>/KNb<span class="hlt">O</span><span class="hlt">3</span> Composites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kakimoto, Ken-ichi; Imura, Tomoya; Fukui, Yasuchika; Kuno, Masami; Yamagiwa, Katsuya; Mitsuoka, Takeshi; Ohbayashi, Kazushige</p> <p>2007-10-01</p> <p>Porous Li0.06(Na0.5K0.5)0.94Nb<span class="hlt">O</span><span class="hlt">3</span> (LNKN-6) ceramics with different pore volumes have been prepared using preceramic powder and phenol resin fiber (KynolTM) as a pore former. It was confirmed that the porous ceramics synthesized by the “two-stage firing method” suppressed the loss of alkali elements from the porous body during heat treatment. The porous LNKN-6 ceramics were then converted to LNKN-6/KNb<span class="hlt">O</span><span class="hlt">3</span> composites through soaking and heat treatment using a sol-gel precursor source composed of KNb<span class="hlt">O</span><span class="hlt">3</span> to form <span class="hlt">3-3</span>-type composites. The microstructure, dielectric, and piezoelectric properties of the porous LNKN-6 ceramics and LNKN-6/KNb<span class="hlt">O</span><span class="hlt">3</span> composites were characterized and compared. The LNKN-6/KNb<span class="hlt">O</span><span class="hlt">3</span> composites had a hollow structure whose pores in the region near the surface were filled and coated with KNb<span class="hlt">O</span><span class="hlt">3</span> precipitates; however, a large amount of residual air was trapped in the pores inside the composites. As a result, the LNKN-6/KNb<span class="hlt">O</span><span class="hlt">3</span> composites fabricated using 30 vol % KynolTM showed an enhanced piezoelectric voltage output coefficient (g33) of 63.0× 10-<span class="hlt">3</span> V\\cdotm/N, compared with monolithic LNKN-6 ceramics having a g33 of 30.2× 10-<span class="hlt">3</span> V\\cdotm/N.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3953397','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3953397"><span>2DEGs at Perovskite Interfaces between KTa<span class="hlt">O</span><span class="hlt">3</span> or KNb<span class="hlt">O</span><span class="hlt">3</span> and Stannates</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Fan, Xiaofeng; Zheng, Weitao; Chen, Xin; Singh, David J.</p> <p>2014-01-01</p> <p>We report density functional studies of electron rich interfaces between KTa<span class="hlt">O</span><span class="hlt">3</span> or KNb<span class="hlt">O</span><span class="hlt">3</span> and CaSn<span class="hlt">O</span><span class="hlt">3</span> or ZnSn<span class="hlt">O</span><span class="hlt">3</span> and in particular the nature of the interfacial electron gasses that can be formed. We find that depending on the details these may occur on either the transition metal or stannate sides of the interface and in the later case can be shifted away from the interface by ferroelectricity. We also present calculations for bulk KNb<span class="hlt">O</span><span class="hlt">3</span>, KTa<span class="hlt">O</span><span class="hlt">3</span>, CaSn<span class="hlt">O</span><span class="hlt">3</span>, BaSn<span class="hlt">O</span><span class="hlt">3</span> and ZnSn<span class="hlt">O</span><span class="hlt">3</span>, showing the different transport and optical properties that may be expected on the two sides of such interfaces. The results suggest that these interfaces may display a wide range of behaviors depending on conditions, and in particular the interplay with ferroelectricity suggests that electrical control of these properties may be possible. PMID:24626191</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3902489','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3902489"><span>Ferromagnetic CaRu<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Tripathi, Shivendra; Rana, Rakesh; Kumar, Sanjay; Pandey, Parul; Singh, R. S.; Rana, D. S.</p> <p>2014-01-01</p> <p>The non-magnetic and non-Fermi-liquid CaRu<span class="hlt">O</span><span class="hlt">3</span> is the iso-structural analog of the ferromagnetic (FM) and Fermi-liquid SrRu<span class="hlt">O</span><span class="hlt">3</span>. We show that an FM order in the orthorhombic CaRu<span class="hlt">O</span><span class="hlt">3</span> can be established by the means of tensile epitaxial strain. The structural and magnetic property correlations in the CaRu<span class="hlt">O</span><span class="hlt">3</span> films formed on SrTi<span class="hlt">O</span><span class="hlt">3</span> (100) substrate establish a scaling relation between the FM moment and the tensile strain. The strain dependent crossover from non-magnetic to FM CaRu<span class="hlt">O</span><span class="hlt">3</span> was observed to be associated with switching of non-Fermi liquid to Fermi-liquid behavior. The intrinsic nature of this strain-induced FM order manifests in the Hall resistivity too; the anomalous Hall component realizes in FM tensile-strained CaRu<span class="hlt">O</span><span class="hlt">3</span> films on SrTi<span class="hlt">O</span><span class="hlt">3</span> (100) whereas the non-magnetic compressive-strained films on LaAl<span class="hlt">O</span><span class="hlt">3</span> (100) exhibit only the ordinary Hall effect. These observations of an elusive FM order are consistent with the theoretical predictions of scaling of the tensile epitaxial strain and the magnetic order in tensile CaRu<span class="hlt">O</span><span class="hlt">3</span>. We further establish that the tensile strain is more efficient than the chemical route to induce FM order in CaRu<span class="hlt">O</span><span class="hlt">3</span>. PMID:24464302</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013MPLB...2750128Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013MPLB...2750128Z"><span>PREPARATION AND ELECTRICAL PROPERTIES OF BiFe<span class="hlt">O</span><span class="hlt">3</span>/La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> MULTILAYERS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhu, Huiwen; Wang, Shunli; Li, Xiaoyun</p> <p>2013-07-01</p> <p>(La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> 12 nm/BiFe<span class="hlt">O</span><span class="hlt">3</span> 12 nm)10 was grown on SrTi<span class="hlt">O</span><span class="hlt">3</span> (001) substrate using rf magnetron sputtering. The structure analysis indicated that BiFe<span class="hlt">O</span><span class="hlt">3</span>/La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> multilayers were highly (001)-oriented. Compared with bottom La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> electrode, the top La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> electrode displayed a rougher surface. The electric transport characteristics of the sample were investigated mainly at low temperature, and it was found that the sample exhibited resistance-temperature curves similar to those of La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> with the exception of an upturn at lower temperature region. Furthermore, a nonlinear I-V curve, which is characteristic of a tunneling conduction mechanism, was observed at 50 K. At higher temperature, the I-V curves were found to be diode-like. When the temperature was further increased to 300 K, the sample showed a space charge limited conduction (SCLC) characteristic.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26315344','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26315344"><span>Polarization Rotation in Ferroelectric Tricolor PbTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span>/PbZr0.2Ti0.8<span class="hlt">O</span><span class="hlt">3</span> Superlattices.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lemée, Nathalie; Infante, Ingrid C; Hubault, Cécile; Boulle, Alexandre; Blanc, Nils; Boudet, Nathalie; Demange, Valérie; Karkut, Michael G</p> <p>2015-09-16</p> <p>In ferroelectric thin films, controlling the orientation of the polarization is a key element to controlling their physical properties. We use laboratory and synchrotron X-ray diffraction to investigate ferroelectric bicolor PbTi<span class="hlt">O</span><span class="hlt">3</span>/PbZr0.2Ti0.8<span class="hlt">O</span><span class="hlt">3</span> and tricolor PbTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span>/PbZr0.2Ti0.8<span class="hlt">O</span><span class="hlt">3</span> superlattices and to study the role of the SrTi<span class="hlt">O</span><span class="hlt">3</span> layers on the domain structure. In the tricolor superlattices, we demonstrate the existence of 180° ferroelectric stripe nanodomains, induced by the depolarization field produced by the SrTi<span class="hlt">O</span><span class="hlt">3</span> layers. Each ultrathin SrTi<span class="hlt">O</span><span class="hlt">3</span> layer modifies the electrostatic boundary conditions between the ferroelectric layers compared to the corresponding bicolor structures, leading to the suppression of the a/c polydomain states. Combined with the electrostatic effect, the tensile strain induced by PbZr0.2Ti0.8<span class="hlt">O</span><span class="hlt">3</span> in the PbTi<span class="hlt">O</span><span class="hlt">3</span> layers leads to polarization rotation in the system as evidenced by grazing incidence X-ray measurements. This polarization rotation is associated with the monoclinic Mc phase as revealed by the splitting of the (HHL) and (H0L) reciprocal lattice points. This work demonstrates that the tricolor paraelectric/ferroelectric superlattices constitute a tunable system to investigate the concomitant effects of strains and depolarizing fields. Our studies provide a pathway to stabilize a monoclinic symmetry in ferroelectric layers, which is of particular interest for the enhancement of the piezoelectric properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12856580','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12856580"><span>[Influences of R2<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 system glass and superfine alpha-Al2<span class="hlt">O</span><span class="hlt">3</span> on the sintering and phase transition of hydroxyapatite ceramics].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Zhiqiang; Chen, Xiaoxu; Cai, Yingji; Lü, Bingling</p> <p>2003-06-01</p> <p>The effects of R2<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 system glass and superfine alpha-Al2<span class="hlt">O</span><span class="hlt">3</span> on the sintering and phase transition of hydroxyapatite (HAP) ceramics were assessed. The results showed that alpha-Al2<span class="hlt">O</span><span class="hlt">3</span> impeded the sintering of HAP and raised the sintering temperature. When glass and alpha-Al2<span class="hlt">O</span><span class="hlt">3</span> were used together to reinforce HAP ceramics, better results could be obtained; the bending strength of multiphase HAP ceramics approached 106 MPa when 10% (wt) alpha-Al2<span class="hlt">O</span><span class="hlt">3</span> and 20%(wt) glass were used and sintered at 1200 for 1 h.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1261328-material-properties-perovskites-quasi-ternary-system-lafeo3lacoo3lanio3','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1261328-material-properties-perovskites-quasi-ternary-system-lafeo3lacoo3lanio3"><span>Material properties of perovskites in the quasi-ternary system LaFe<span class="hlt">O</span> <span class="hlt">3</span>–LaCo<span class="hlt">O</span> <span class="hlt">3</span>–LaNi<span class="hlt">O</span> <span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Tietz, F.; Arul Raj, I.; Ma, Q.; ...</p> <p>2016-02-02</p> <p>We present an overview on the variation of electrical conductivity, oxygen permeation, oxygen surface exchange and thermal expansion coefficient as a function of the composition of perovskites in the quasi-ternary system LaFe<span class="hlt">O</span> <span class="hlt">3</span>–LaCo<span class="hlt">O</span> <span class="hlt">3</span>–LaNi<span class="hlt">O</span> <span class="hlt">3</span>. Powders of thirteen nominal perovskite compositions were synthesized under identical conditions by the Pechini method. The powder X-ray diffraction data of two series, namely La(Ni 0.5Fe 0.5) 1-xCo x<span class="hlt">O</span> <span class="hlt">3</span> and LaNi 0.5- xFe xCo 0.5<span class="hlt">O</span> <span class="hlt">3</span>, are presented after the powders had been sintered at 1100 C for 6 h in air. The measurements revealed a rhombohedral structure for all compositions except LaNi 0.5Femore » 0.5<span class="hlt">O</span> <span class="hlt">3</span> for which 60% rhombohedral and 40% orthorhombic phase was found. Moreover, the maximum DC electrical conductivity value of the perovskites at 800 C was 1229 S cm-1 for the composition LaCo<span class="hlt">O</span> <span class="hlt">3</span> and the minimum was 91 S cm-1 for the composition LaCo 0.5Fe 0.5<span class="hlt">O</span> <span class="hlt">3</span>. The oxygen permeation of samples with promising conductivities at 800 C was one order of magnitude lower than that of La 0.6Sr 0.4Co 0.8Fe 0.2<span class="hlt">O</span> <span class="hlt">3</span> (LSCF). The highest value of 0.017 ml cm -2 min-1 at 950 C was obtained with LaNi 0.5Co 0.5<span class="hlt">O</span> <span class="hlt">3</span>. The coefficients of thermal expansion varied in the range of 13.2 x 10 -6 K -1 and 21.9 x 10 -6 K -1 for LaNi 0.5Fe 0.5<span class="hlt">O</span> <span class="hlt">3</span> and LaCo<span class="hlt">O</span> <span class="hlt">3</span>, respectively. 57Fe M ssbauer spectroscopy was used as probe for the oxidation states, local environment and magnetic properties of iron ions as a function of chemical composition. Ultimately, the substitution had a great influence on the chemical properties of the materials.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020027145','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020027145"><span>Thermodynamic Assessment of the Y2<span class="hlt">o</span><span class="hlt">3</span>-yb2<span class="hlt">o</span><span class="hlt">3</span>-zro2 System</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jacobson, Nathan S.; Liu, Zi-Kui; Kaufman, Larry; Zhang, Fan</p> <p>2002-01-01</p> <p>Yttria-zirconia (Y2<span class="hlt">O</span><span class="hlt">3</span>-Zr<span class="hlt">O</span>2) is the most widely used of the rare earth oxide-zirconia systems. There are numerous experimental studies of the phase boundaries in this system. In this paper, we assess these data and derive parameters for the solution models in this system. There is current interest in other rare earth oxide-zirconia systems as well as systems with several rare earth oxides and zirconia, which may offer improved properties over the Y2<span class="hlt">O</span><span class="hlt">3</span>-Zr<span class="hlt">O</span>2 system. For this reason, we also assess the ytterbia-zirconia (Yb2<span class="hlt">O</span><span class="hlt">3</span>-Zr<span class="hlt">O</span>2) and Y2<span class="hlt">O</span><span class="hlt">3</span>-Yb2<span class="hlt">O</span><span class="hlt">3</span>-Zr<span class="hlt">O</span>2 system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JAP...109gD729H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JAP...109gD729H"><span>Interfacial magnetism in CaRu<span class="hlt">O</span><span class="hlt">3</span>/CaMn<span class="hlt">O</span><span class="hlt">3</span> superlattices grown on (001) SrTi<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>He, C.; Zhai, X.; Mehta, V. V.; Wong, F. J.; Suzuki, Y.</p> <p>2011-04-01</p> <p>We have studied epitaxially grown superlattices of CaRu<span class="hlt">O</span><span class="hlt">3</span>/CaMn<span class="hlt">O</span><span class="hlt">3</span> as well as an alloy film of CaMn0.5Ru0.5<span class="hlt">O</span><span class="hlt">3</span> on (001) SrTi<span class="hlt">O</span><span class="hlt">3</span> substrates. In contrast to previous experiments, we have studied CRO/CMO superlattices with a constant CRO thickness and variable CMO thickness. All superlattices exhibit Curie temperatures (TC) of 110 K. The saturated magnetization per interfacial Mn cation has been found to be 1.1 μB/Mn ion. The TC's of the superlattices are much lower than the TC of the alloy film while the saturated magnetization values are larger than that of the alloy film. These observations suggest that interdiffusion alone cannot account for ferromagnetism in the superlattices and that double exchange induced FM must play a role at the interfaces.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2614732','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2614732"><span>Postperovskite phase equilibria in the MgSi<span class="hlt">O</span><span class="hlt">3</span>–Al2<span class="hlt">O</span><span class="hlt">3</span> system</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Tsuchiya, Jun; Tsuchiya, Taku</p> <p>2008-01-01</p> <p>We investigate high-P,T phase equilibria of the MgSi<span class="hlt">O</span><span class="hlt">3</span>–Al2<span class="hlt">O</span><span class="hlt">3</span> system by means of the density functional ab initio computation methods with multiconfiguration sampling. Being different from earlier studies based on the static substitution properties with no consideration of Rh2<span class="hlt">O</span><span class="hlt">3</span>(II) phase, present calculations demonstrate that (i) dissolving Al2<span class="hlt">O</span><span class="hlt">3</span> tends to decrease the postperovskite transition pressure of MgSi<span class="hlt">O</span><span class="hlt">3</span> but the effect is not significant (≈-0.2 GPa/mol% Al2<span class="hlt">O</span><span class="hlt">3</span>); (ii) Al2<span class="hlt">O</span><span class="hlt">3</span> produces the narrow perovskite+postperovskite coexisting P,T area (≈1 GPa) for the pyrolitic concentration (xAl2<span class="hlt">O</span><span class="hlt">3</span> ≈6 mol%), which is sufficiently responsible to the deep-mantle D″ seismic discontinuity; (iii) the transition would be smeared (≈4 GPa) for the basaltic Al-rich composition (xAl2<span class="hlt">O</span><span class="hlt">3</span> ≈20 mol%), which is still seismically visible unless iron has significant effects; and last (iv) the perovskite structure spontaneously changes to the Rh2<span class="hlt">O</span><span class="hlt">3</span>(II) with increasing the Al concentration involving small displacements of the Mg-site cations. PMID:19036928</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.A51A3011L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A51A3011L"><span>Improvement and validation of trace gas retrieval from ACAM aircraft observation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, C.; Liu, X.; Kowalewski, M. G.; Janz, S. J.; Gonzalez Abad, G.; Pickering, K. E.; Chance, K.; Lamsal, L. N.</p> <p>2014-12-01</p> <p>The ACAM (Airborne Compact Atmospheric Mapper) instrument, flown on board the NASA UC-12 aircraft during the DISCOVER-AQ (Deriving Information on Surface Conditions from Column and Vertically Resolved Observations Relevant to Air Quality) campaigns, was designed to provide remote sensing observations of tropospheric and boundary layer pollutants and help understand some of the most important pollutants that directly affect the health of the population. In this study, slant column densities (SCD) of trace gases (<span class="hlt">O</span><span class="hlt">3</span>, NO2, <span class="hlt">HCHO</span>) are retrieved from ACAM measurements during the Baltimore-Washington D.C. 2011 campaign by the Basic Optical Absorption Spectroscopy (BOAS) trace gas fitting algorithm using a nonlinear least-squares (NLLS) inversion technique, and then are converted to vertical column densities (VCDs) using the Air Mass Factors (AMF) calculated with the VLIDORT (Vector Linearized Discrete Ordinate Radiative Transfer) model and CMAQ (Community Multi-scale Air Quality) model simulations of trace gas profiles. For surface treatment in the AMF, we use high-resolution MODIS climatological BRDF product (Bidirectional Reflectance Distribution Function) at 470 nm for NO2, and use high-resolution surface albedo derived by combining MODIS and OMI albedo databases for <span class="hlt">HCHO</span> and <span class="hlt">O</span><span class="hlt">3</span>. We validate ACAM results with coincident ground-based PANDORA, aircraft (P<span class="hlt">3</span>B) spiral and satellite (OMI) measurements and find out generally good agreement especially for NO2 and <span class="hlt">O</span><span class="hlt">3</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhSS...60..520P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhSS...60..520P"><span>Magnetostriction of Hexagonal HoMn<span class="hlt">O</span><span class="hlt">3</span> and YMn<span class="hlt">O</span><span class="hlt">3</span> Single Crystals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pavlovskii, N. S.; Dubrovskii, A. A.; Nikitin, S. E.; Semenov, S. V.; Terent'ev, K. Yu.; Shaikhutdinov, K. A.</p> <p>2018-03-01</p> <p>We report on the magnetostriction of hexagonal HoMn<span class="hlt">O</span><span class="hlt">3</span> and YMn<span class="hlt">O</span><span class="hlt">3</span> single crystals in a wide range of applied magnetic fields (up to H = 14 T) at all possible combinations of the mutual orientations of magnetic field H and magnetostriction Δ L/L. The measured Δ L/L( H, T) data agree well with the magnetic phase diagram of the HoMn<span class="hlt">O</span><span class="hlt">3</span> single crystal reported previously by other authors. It is shown that the nonmonotonic behavior of magnetostriction of the HoMn<span class="hlt">O</span><span class="hlt">3</span> crystal is caused by the Ho<span class="hlt">3</span>+ ion; the magnetic moment of the Mn<span class="hlt">3</span>+ ion parallel to the hexagonal crystal axis. The anomalies established from the magnetostriction measurements of HoMn<span class="hlt">O</span><span class="hlt">3</span> are consistent with the phase diagram of these compounds. For the isostructural YMn<span class="hlt">O</span><span class="hlt">3</span> single crystal with a nonmagnetic rare-earth ion, the Δ L/L( H, T) dependences are described well by a conventional quadratic law in a wide temperature range (4-100 K). In addition, the magnetostriction effect is qualitatively estimated with regard to the effect of the crystal electric field on the holmium ion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ApSS..285..267J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ApSS..285..267J"><span>A comparative study of Ce<span class="hlt">O</span>2-Al2<span class="hlt">O</span><span class="hlt">3</span> support prepared with different methods and its application on Mo<span class="hlt">O</span><span class="hlt">3</span>/Ce<span class="hlt">O</span>2-Al2<span class="hlt">O</span><span class="hlt">3</span> catalyst for sulfur-resistant methanation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jiang, Minhong; Wang, Baowei; Yao, Yuqin; Li, Zhenhua; Ma, Xinbin; Qin, Shaodong; Sun, Qi</p> <p>2013-11-01</p> <p>The Ce<span class="hlt">O</span>2-Al2<span class="hlt">O</span><span class="hlt">3</span> supports prepared with impregnation (IM), deposition precipitation (DP), and solution combustion (SC) methods for Mo<span class="hlt">O</span><span class="hlt">3</span>/Ce<span class="hlt">O</span>2-Al2<span class="hlt">O</span><span class="hlt">3</span> catalyst were investigated in the sulfur-resistant methanation. The supports and catalysts were characterized by N2-physisorption, transmission electron microscopy (TEM), X-ray diffraction (XRD), Raman spectroscopy (RS), and temperature-programmed reduction (TPR). The N2-physisorption results indicated that the DP method was favorable for obtaining better textural properties. The TEM and RS results suggested that there is a Ce<span class="hlt">O</span>2 layer on the surface of the support prepared with DP method. This Ce<span class="hlt">O</span>2 layer not only prevented the interaction between Mo<span class="hlt">O</span><span class="hlt">3</span> and γ-Al2<span class="hlt">O</span><span class="hlt">3</span> to form Al2(Mo<span class="hlt">O</span>4)<span class="hlt">3</span> species, but also improved the dispersion of Mo<span class="hlt">O</span><span class="hlt">3</span> in the catalyst. Accordingly, the catalysts whose supports were prepared with DP method exhibited the best catalytic activity. The catalysts whose supports were prepared with SC method had the worst catalytic activity. This was caused by the formation of Al2(Mo<span class="hlt">O</span>4)<span class="hlt">3</span> and crystalline Mo<span class="hlt">O</span><span class="hlt">3</span>. Additionally, the Ce<span class="hlt">O</span>2 layer resulted in the instability of catalysts in reaction process. The increasing of calcination temperature of supports reduced the catalytic activity of all catalysts. The decrease extent of the catalysts whose supports were prepared with DP method was the lowest as the Ce<span class="hlt">O</span>2 layer prevented the interaction between Mo<span class="hlt">O</span><span class="hlt">3</span> and γ-Al2<span class="hlt">O</span><span class="hlt">3</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004MPLB...18..275A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004MPLB...18..275A"><span>Structural Investigations of the Mn<span class="hlt">O</span>-Bi<span class="hlt">3</span><span class="hlt">O</span><span class="hlt">3</span>-Cd<span class="hlt">O</span> Glass System by IR and Raman Spectroscopies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ardelean, I.; Todor, Ioana; PǍŞCUŢǍ, P.</p> <p></p> <p>Homogeneous glasses are formed in the Mn<span class="hlt">O</span>-Bi<span class="hlt">3</span><span class="hlt">O</span><span class="hlt">3</span>-Cd<span class="hlt">O</span> system, up to 50 mol% Mn<span class="hlt">O</span>. For these glasses, IR and Raman spectral measurements are carried out in order to elucidate the local structure. We identify by IR spectroscopy both the structural units Bi<span class="hlt">O</span><span class="hlt">3</span> and Bi<span class="hlt">O</span>6. The Raman investigation confirms the prevalence of Bi<span class="hlt">O</span>6 groups in the glass network for all concentrations. The number of these structural groups progressively increases with Mn<span class="hlt">O</span> content.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApSS..412..290W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApSS..412..290W"><span>In situ DRIFTS study of <span class="hlt">O</span><span class="hlt">3</span> adsorption on Ca<span class="hlt">O</span>, γ-Al2<span class="hlt">O</span><span class="hlt">3</span>, Cu<span class="hlt">O</span>, α-Fe2<span class="hlt">O</span><span class="hlt">3</span> and Zn<span class="hlt">O</span> at room temperature for the catalytic ozonation of cinnamaldehyde</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Jianfeng; Su, Tongming; Jiang, Yuexiu; Xie, Xinling; Qin, Zuzeng; Ji, Hongbing</p> <p>2017-08-01</p> <p>In situ DRIFTS were conducted to identify adsorbed ozone and/or adsorbed oxygen species on Ca<span class="hlt">O</span>, Zn<span class="hlt">O</span>, γ-Al2<span class="hlt">O</span><span class="hlt">3</span>, Cu<span class="hlt">O</span> and α-Fe2<span class="hlt">O</span><span class="hlt">3</span> surfaces at room temperature. Samples were characterized by means of TG, XRD, N2 adsorption-desorption, pyridine-IR, nitrobenzene-IR, chloroform-IR, and CO2-TPD. Pyridine-DRIFTS measurements evidence two kinds of acid sites in all the samples. Nitrobenzene, chloroform-DRIFTS, and CO2-TPD reveal that there are large amounts of medium-strength base sites on all the metal oxides, and only Ca<span class="hlt">O</span>, Zn<span class="hlt">O</span>, and γ-Al2<span class="hlt">O</span><span class="hlt">3</span> have strong base sites. And the benzaldehyde selectivity was increased in the same order of the alkalinity of the metal oxides. With weaker sites, ozone molecules form coordinative complexes bound via the terminal oxygen atom, observed by vibrational frequencies at 2095-2122 and 1026-1054 cm-1. The formation of ozonide <span class="hlt">O</span><span class="hlt">3</span>- at 790 cm-1, atomic oxygen at 1317 cm-1, and superoxide <span class="hlt">O</span>2- at 1124 cm-1 was detected; these species are believed to be intermediates of <span class="hlt">O</span><span class="hlt">3</span> decomposition on strong acid/base sites. The adsorption of ozone on metal oxides is a weak adsorption, and other gases, such as CO2, will compete with <span class="hlt">O</span><span class="hlt">3</span> adsorption. The mechanism of cinnamaldehyde ozonation at room temperature over Ca<span class="hlt">O</span> shows that cinnamaldehyde can not only be oxidized into cinnamic acid, but also be further oxidized into benzaldehyde, benzoic acid, maleic anhydride, and ultimately mineralized to CO2 in the presence of <span class="hlt">O</span><span class="hlt">3</span>.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_17 --> <div id="page_18" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="341"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApPhL.111w2901C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApPhL.111w2901C"><span>Enhanced texture evolution and piezoelectric properties in Cu<span class="hlt">O</span>-doped Pb(In1/2Nb1/2)<span class="hlt">O</span><span class="hlt">3</span>-Pb(Mg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> grain-oriented ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chang, Yunfei; Watson, Beecher; Fanton, Mark; Meyer, Richard J.; Messing, Gary L.</p> <p>2017-12-01</p> <p>In this work, both crystallographic texture and doping engineering strategies were integrated to develop relaxor-PbTi<span class="hlt">O</span><span class="hlt">3</span> (PT) based ternary ferroelectric ceramics with enhanced texture evolution and superior electromechanical properties. Cu<span class="hlt">O</span>-doped Pb(In1/2Nb1/2)<span class="hlt">O</span><span class="hlt">3</span>-Pb(Mg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> (PIN-PMN-PT) piezoelectric ceramics with [001]c texture fraction ≥97% were synthesized by templated grain growth. The addition of Cu<span class="hlt">O</span> significantly promotes densification and oriented grain growth in the templated ceramics, leading to full texture development at dramatically reduced times and temperatures. Moreover, the Cu<span class="hlt">O</span> dopant remarkably enhances the piezoelectric properties of the textured ceramics while maintaining high phase transition temperatures and large coercive fields. Doping 0.125 wt. % Cu<span class="hlt">O</span> yields the electromechanical properties of d33 = 927 pC/N, d33* = 1510 pm/V, g33 = 43.2 × 10-<span class="hlt">3</span> Vm/N, Kp = 0.87, Ec=8.8 kV/cm, and tan δ = 1.<span class="hlt">3</span>%, which are the best values reported so far in PIN-PMN-PT based ceramics. The high piezoelectric coefficient is mainly from the reversible piezoelectric response, with the irreversible contribution being on the order of 13.1%. We believe that this work not only facilitates closing the performance gap between ceramics and single crystals but also can expand relaxor-PT based piezoelectric application fields.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JaJAP..52jMB22K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JaJAP..52jMB22K"><span>Effect of the Molar Ratio of B2<span class="hlt">O</span><span class="hlt">3</span> to Bi2<span class="hlt">O</span><span class="hlt">3</span> in Al Paste with Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Zn<span class="hlt">O</span> Glass on Screen Printed Contact Formation and Si Solar Cell Performance</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Bit-Na; Kim, Hyeong Jun; Chang, Hyo Sik; Hong, Hyun Seon; Ryu, Sung-Soo; Lee, Heon</p> <p>2013-10-01</p> <p>In this study, eco-friendly Pb-free Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Zn<span class="hlt">O</span> glass frits were chosen as an inorganic additive for the Al paste used in Si solar cells. The effects of the molar ratio of Bi2<span class="hlt">O</span><span class="hlt">3</span> to B2<span class="hlt">O</span><span class="hlt">3</span> in the glass composition on the electrical resistance of the Al electrode and on the cell performance were investigated. The results showed that as the molar ratio of Bi2<span class="hlt">O</span><span class="hlt">3</span> to B2<span class="hlt">O</span><span class="hlt">3</span> increased, the glass transition temperature and softening temperature decreased because of the reduced glass viscosity. In Al screen-printed Si solar cells, as the molar ratio of Bi2<span class="hlt">O</span><span class="hlt">3</span> to B2<span class="hlt">O</span><span class="hlt">3</span> increased, the sheet electrical resistance of the Al electrode decreased and the cell efficiency increased. The uniformity and thickness of the back-surface field was significantly influenced by the glass composition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NIMPB.392...36Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NIMPB.392...36Y"><span>Thermoluminescence and optically stimulated luminescence properties of Dy<span class="hlt">3</span>+-doped Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-based glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yahaba, T.; Fujimoto, Y.; Yanagida, T.; Koshimizu, M.; Tanaka, H.; Saeki, K.; Asai, K.</p> <p>2017-02-01</p> <p>We developed Dy<span class="hlt">3</span>+-doped Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span> based glasses with Dy concentrations of 0.5, 1.0, and 2.0 mol% using a melt-quenching technique. The as-synthesized glasses were applicable as materials exhibiting thermoluminescence (TL) and optically stimulated luminescence (OSL). The optical and radiation response properties of the glasses were characterized. In the photoluminescence (PL) spectra, two emission bands due to the 4F9/2 → 6H15/2 and 4F9/2 → 6H13/2 transitions of Dy<span class="hlt">3</span>+ were observed at 480 and 580 nm. In the OSL spectra, the emission band due to the 4F9/2 → 6H15/2 transition of Dy<span class="hlt">3</span>+ was observed. Excellent TL and OSL responses were observed for dose ranges of 0.1-90 Gy. In addition, TL fading behavior was better than that of OSL in term of the long-time storage. These results indicate that the Dy<span class="hlt">3</span>+-doped Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-based glasses are applicable as TL materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28359141','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28359141"><span>The Study of Electrical Properties for Multilayer La2<span class="hlt">O</span><span class="hlt">3</span>/Al2<span class="hlt">O</span><span class="hlt">3</span> Dielectric Stacks and LaAl<span class="hlt">O</span><span class="hlt">3</span> Dielectric Film Deposited by ALD.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Feng, Xing-Yao; Liu, Hong-Xia; Wang, Xing; Zhao, Lu; Fei, Chen-Xi; Liu, He-Lei</p> <p>2017-12-01</p> <p>The capacitance and leakage current properties of multilayer La 2 <span class="hlt">O</span> <span class="hlt">3</span> /Al 2 <span class="hlt">O</span> <span class="hlt">3</span> dielectric stacks and LaAl<span class="hlt">O</span> <span class="hlt">3</span> dielectric film are investigated in this paper. A clear promotion of capacitance properties is observed for multilayer La 2 <span class="hlt">O</span> <span class="hlt">3</span> /Al 2 <span class="hlt">O</span> <span class="hlt">3</span> stacks after post-deposition annealing (PDA) at 800 °C compared with PDA at 600 °C, which indicated the recombination of defects and dangling bonds performs better at the high-k/Si substrate interface for a higher annealing temperature. For LaAl<span class="hlt">O</span> <span class="hlt">3</span> dielectric film, compared with multilayer La 2 <span class="hlt">O</span> <span class="hlt">3</span> /Al 2 <span class="hlt">O</span> <span class="hlt">3</span> dielectric stacks, a clear promotion of trapped charges density (N ot ) and a degradation of interface trap density (D it ) can be obtained simultaneously. In addition, a significant improvement about leakage current property is observed for LaAl<span class="hlt">O</span> <span class="hlt">3</span> dielectric film compared with multilayer La 2 <span class="hlt">O</span> <span class="hlt">3</span> /Al 2 <span class="hlt">O</span> <span class="hlt">3</span> stacks at the same annealing condition. We also noticed that a better breakdown behavior for multilayer La 2 <span class="hlt">O</span> <span class="hlt">3</span> /Al 2 <span class="hlt">O</span> <span class="hlt">3</span> stack is achieved after annealing at a higher temperature for its less defects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27750006','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27750006"><span>Lead-Free KNb<span class="hlt">O</span><span class="hlt">3</span>:xZn<span class="hlt">O</span> Composite Ceramics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lv, Xiang; Li, Zhuoyun; Wu, Jiagang; Xiao, Dingquan; Zhu, Jianguo</p> <p>2016-11-09</p> <p>It is a tough issue to develop dense and water resistant KNb<span class="hlt">O</span> <span class="hlt">3</span> ceramics due to high evaporation and hygroscopicity of K 2 <span class="hlt">O</span>. Here, KNb<span class="hlt">O</span> <span class="hlt">3</span> :xZn<span class="hlt">O</span> composite ceramics were used to successfully solve this problem, where Zn<span class="hlt">O</span> particles were randomly distributed into a KNb<span class="hlt">O</span> <span class="hlt">3</span> matrix. The addition of Zn<span class="hlt">O</span> hardly affects the phase structure of KNb<span class="hlt">O</span> <span class="hlt">3</span> , and moreover, the enhancement of electrical properties, thermal stability, and aging characteristics was observed in KNb<span class="hlt">O</span> <span class="hlt">3</span> :xZn<span class="hlt">O</span> composite ceramics. The composites possessed the maximum d 33 of 120 ± 5 pC/N, which is superior to that of pure KNb<span class="hlt">O</span> <span class="hlt">3</span> (d 33 = 80 pC/N). More importantly, a strong water resistance and an aging-free characteristic were observed in KNb<span class="hlt">O</span> <span class="hlt">3</span> :0.4Zn<span class="hlt">O</span>. This is the first time for KNb<span class="hlt">O</span> <span class="hlt">3</span> ceramics to simultaneously improve electrical properties and resolve the water-absorbing properties. We believe that these composite ceramics are promising for practical applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/197351-performance-measurements-sub-sub-carbonyl-compounds-using-dnph-coated-silica-gel-sub-cartridges','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/197351-performance-measurements-sub-sub-carbonyl-compounds-using-dnph-coated-silica-gel-sub-cartridges"><span>Performance measurements of C{sub 1}-C{sub <span class="hlt">3</span>} carbonyl compounds using DNPH-coated silica gel and C{sub 18} cartridges</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kleindienst, T.E.; Corse, E.W.; Blanchard, F.T.</p> <p>1994-12-31</p> <p>Measurements have been conducted to examine the performance of formaldehyde, acetaldehyde, and acetone using silica and C{sub 18} cartridges coated with 2,4-dinitrophenylhydrazine. Laboratory measurements for formaldehyde were conducted using a paraformaldehyde generator to produce reproducible and constant concentrations of the compound. For acetaldehyde and acetone, known concentrations were generated in Teflon chambers. The compounds were routed into a sampling manifold where simultaneous measurements could be made with multiple cartridges. Typical concentrations employed in the study were as follows. <span class="hlt">HCHO</span>: 0.5--25 ppbv; CH{sub <span class="hlt">3</span>}CHO; 0.5--10 ppbv; CH{sub <span class="hlt">3</span>}C(<span class="hlt">O</span>)CH{sub <span class="hlt">3</span>}: 0.5--10 ppbv. Additional measurements were conducted for these compounds in the presencemore » of potentially interfering compounds such as ozone and water vapor. Serial cartridge collections were periodically used to investigate breakthrough of the carbonyl compounds.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SSSci..79....6B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SSSci..79....6B"><span>Investigating the local structure of B-site cations in (1-x)BaTi<span class="hlt">O</span><span class="hlt">3</span>-xBiSc<span class="hlt">O</span><span class="hlt">3</span> and (1-x)PbTi<span class="hlt">O</span><span class="hlt">3</span>-xBiSc<span class="hlt">O</span><span class="hlt">3</span> using X-ray absorption spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Blanchard, Peter E. R.; Grosvenor, Andrew P.</p> <p>2018-05-01</p> <p>The structural properties of (1-x)BaTi<span class="hlt">O</span><span class="hlt">3</span>-xBiSc<span class="hlt">O</span><span class="hlt">3</span> and (1-x)PbTi<span class="hlt">O</span><span class="hlt">3</span>-xBiSc<span class="hlt">O</span><span class="hlt">3</span> were investigated using powder X-ray diffraction and X-ray absorption spectroscopy. Diffraction measurements confirmed that substituting small amounts of BiSc<span class="hlt">O</span><span class="hlt">3</span> into BaTi<span class="hlt">O</span><span class="hlt">3</span> initially stabilizes a cubic phase at x = 0.2 before impurity phases begin to form at x = 0.5. BiSc<span class="hlt">O</span><span class="hlt">3</span> substitution also resulted in noticeable changes in the local coordination environment of Ti4+. X-ray absorption near-edge spectroscopy (XANES) analysis showed that replacing Ti4+ with Sc<span class="hlt">3</span>+ results in an increase in the off-centre displacement of Ti4+ cations. Surprisingly, BiSc<span class="hlt">O</span><span class="hlt">3</span> substitution has no effect on the displacement of the Ti4+ cation in the (1-x)PbTi<span class="hlt">O</span><span class="hlt">3</span>-xBiSc<span class="hlt">O</span><span class="hlt">3</span> solid solution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018APExp..11f1105C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018APExp..11f1105C"><span>Twin-induced phase transition from β-Ga2<span class="hlt">O</span><span class="hlt">3</span> to α-Ga2<span class="hlt">O</span><span class="hlt">3</span> in Ga2<span class="hlt">O</span><span class="hlt">3</span> thin films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Choi, Byeongdae; Allabergenov, Bunyod; Lyu, Hong-Kun; Lee, Seong Eui</p> <p>2018-06-01</p> <p>We deposited a 300-nm-thick Ga2<span class="hlt">O</span><span class="hlt">3</span> thin film on an amorphous Si<span class="hlt">O</span>2/Si substrate via pulsed laser deposition. X-ray diffraction patterns revealed the formation of β-Ga2<span class="hlt">O</span><span class="hlt">3</span> phase at a substrate temperature of 700 °C. X-ray photoelectron spectra indicated that the degree of oxidation increased after annealing at 700 °C. Further annealings at higher temperatures led to a transition of the β-Ga2<span class="hlt">O</span><span class="hlt">3</span> phase to the α-Ga2<span class="hlt">O</span><span class="hlt">3</span> phase; this transition was caused by the twin structure formed during the crystallinity improvement process. In addition, we discuss the mechanism of the transition from the β phase to the α phase in the β-Ga2<span class="hlt">O</span><span class="hlt">3</span> thin films.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JNuM..493..219D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JNuM..493..219D"><span>Thermodynamic characterization of Ni<span class="hlt">3</span>Te<span class="hlt">O</span>6, Ni2Te<span class="hlt">3</span><span class="hlt">O</span>8 and NiTe2<span class="hlt">O</span>5</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dawar, Rimpi; Babu, R.; Ananthasivan, K.; Anthonysamy, S.</p> <p>2017-09-01</p> <p>Measurement of vapour pressure of Te<span class="hlt">O</span>2(g) over the biphasic mixture Ni<span class="hlt">3</span>Te<span class="hlt">O</span>6 (s) + Ni<span class="hlt">O</span>(s) in the temperature range 1143-1272 K was carried out using transpiration-thermogravimetric technique (TTG). Gibbs energy of formation of Ni<span class="hlt">3</span>Te<span class="hlt">O</span>6 was obtained from the temperature dependence of vapour pressure of Te<span class="hlt">O</span>2 (g) generated by the incongruent vapourisation reaction, Ni<span class="hlt">3</span>Te<span class="hlt">O</span>6 (s) → Ni<span class="hlt">O</span>(s) + Te<span class="hlt">O</span>2 (g) + 1/2 <span class="hlt">O</span>2 in the temperature range 1143-1272 K. An isoperibol type drop calorimeter was used to measure the enthalpy increments of Ni<span class="hlt">3</span>Te<span class="hlt">O</span>6, Ni2Te<span class="hlt">3</span><span class="hlt">O</span>8 and NiTe2<span class="hlt">O</span>5. Thermodynamic functions viz., heat capacity, entropy and Gibbs energy functions of these compounds were derived from the experimentally measured enthalpy increment values. Third-law analysis was carried out to ascertain absence of temperature dependent systematic errors in the measurement of vapour pressure of Te<span class="hlt">O</span>2 (g). A value of -1265.1 ± 1.5 kJ mol-1 was obtained for Δ Hf,298K <span class="hlt">o</span> (Ni<span class="hlt">3</span>Te<span class="hlt">O</span>6) using third-law analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ApPhL.100s2905Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ApPhL.100s2905Y"><span>Electromechanical behavior of [001]-textured Pb(Mg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yan, Yongke; Wang, Yu. U.; Priya, Shashank</p> <p>2012-05-01</p> <p>[001]-textured Pb(Mg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> (PMN-PT) ceramics were synthesized by using templated grain growth method. Significantly high [001] texture degree corresponding to 0.98 Lotgering factor was achieved at 1 vol. % BaTi<span class="hlt">O</span><span class="hlt">3</span> template. Electromechanical properties for [001]-textured PMN-PT ceramics with 1 vol. % BaTi<span class="hlt">O</span><span class="hlt">3</span> were found to be d33 = 1000 pC/N, d31 = 371 pC/N, ɛr = 2591, and tanδ = ˜0.6%. Elastoelectric composite based modeling results showed that higher volume fraction of template reduces the overall dielectric constant and thus has adverse effect on the piezoelectric response. Clamping effect was modeled by deriving the changes in free energy as a function of applied electric field and microstructural boundary condition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22584143-synthesis-structure-characterization-two-new-bismuth-iii-selenite-tellurite-nitrates-bi-sub-sub-seo-sub-sub-sub-bi-teo-sub-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22584143-synthesis-structure-characterization-two-new-bismuth-iii-selenite-tellurite-nitrates-bi-sub-sub-seo-sub-sub-sub-bi-teo-sub-sub"><span>Synthesis, structure, and characterization of two new bismuth(III) selenite/tellurite nitrates: [(Bi{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 2})(Se<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}](NO{sub <span class="hlt">3</span>}) and [Bi(Te<span class="hlt">O</span>{sub <span class="hlt">3</span>})](NO{sub <span class="hlt">3</span>})</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Meng, Chang-Yu; Wei, Ming-Fang; Geng, Lei, E-mail: lgeng.cn@gmail.com</p> <p></p> <p>Two new bismuth(III) selenite/tellurite nitrates, [(Bi{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 2})(Se<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}](NO{sub <span class="hlt">3</span>}) and [Bi(Te<span class="hlt">O</span>{sub <span class="hlt">3</span>})](NO{sub <span class="hlt">3</span>}), have been synthesized by conventional facile hydrothermal method at middle temperature 200 °C and characterized by single-crystal X-ray diffraction, powder diffraction, UV–vis–NIR optical absorption spectrum, infrared spectrum and thermal analylsis. Both [(Bi{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 2})(Se<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}](NO{sub <span class="hlt">3</span>}) and [Bi(Te<span class="hlt">O</span><span class="hlt">3</span>)](NO<span class="hlt">3</span>) crystallize in the monoclinic centronsymmetric space group P2{sub 1}/c with a=9.9403(4) Å, b=9.6857(4) Å, c=10.6864(5) Å, β=93.1150(10)° for [(Bi{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 2})(Se<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}](NO{sub <span class="hlt">3</span>}) and a=8.1489(<span class="hlt">3</span>) Å, b=9.0663(4) Å, c=7.4729(<span class="hlt">3</span>) Å, β=114.899(2)° for Bi(Te<span class="hlt">O</span><span class="hlt">3</span>)(NO<span class="hlt">3</span>), respectively. The two compounds, whose structures are composed of three different asymmetricmore » building units, exhibit two different types of structures. The structure of [(Bi{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 2})(Se<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}](NO{sub <span class="hlt">3</span>}) features a three-dimensional (<span class="hlt">3</span>D) bismuth(III) selenite cationic tunnel structure [(Bi{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 2})(Se<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}] {sup <span class="hlt">3</span>}{sub ∞} with NO{sub <span class="hlt">3</span>}{sup −} anion group filling in the 1D tunnel along b axis. The structure of [Bi(Te<span class="hlt">O</span>{sub <span class="hlt">3</span>})](NO{sub <span class="hlt">3</span>}) features 2D bismuth(III) tellurite [Bi(Te<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}]{sup 2}{sub ∞} layers separated by NO{sub <span class="hlt">3</span>}{sup −} anion groups. The results of optical diffuse-reflectance spectrum measurements and electronic structure calculations based on density functional theory methods show that the two compounds are wide band-gap semiconductors. - Graphical abstract: Two novel bismuth{sup III} selenite/tellurite nitrates [(Bi{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 2})(Se<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}](NO{sub <span class="hlt">3</span>}) with <span class="hlt">3</span>D tunnel structure and [Bi(Te<span class="hlt">O</span>{sub <span class="hlt">3</span>})](NO{sub <span class="hlt">3</span>}) with 2D layer structure have been firstly synthesized and characterized. Display Omitted - Highlights: • Two novel bismuth{sup III} nitrates [(Bi{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 2})(Se<span class="hlt">O</span>{sub <span class="hlt">3</span>}){sub 2}](NO{sub <span class="hlt">3</span>}) and [Bi(Te<span class="hlt">O</span>{sub <span class="hlt">3</span>})](NO{sub <span class="hlt">3</span>}) were</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A53B2236T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A53B2236T"><span>Observation and Modelling of the OH, HO2 and RO2 Radicals at a Regional Site of Beijing in Winter 2016.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tan, Z.; Lu, K.; Ma, X.; Bohn, B.; Hofzumahaus, A.; Broch, S.; Fuchs, H.; Holland, F.; Liu, Y.; Li, X.; Novelli, A.; Rohrer, F.; Wang, H.; Wu, Y.; Shao, M.; Zeng, L.; Kiendler-Scharr, A.; Wahner, A.; Zhang, Y.</p> <p>2017-12-01</p> <p>A comprehensive field campaign was carried out in winter 2016 in the campus of UCAS (University of Chinese Academy of Science), located in a small town 60 km northeast of urban Beijing. Concentrations of OH, HO2 and RO2 radicals as well as the total OH reactivity were measured by a laser induced fluorescence instrument. Maximum hourly averaged OH, HO2 and RO2 radical concentrations were (<span class="hlt">3</span>±2)×106cm-<span class="hlt">3</span>, (8±6)×107 cm-<span class="hlt">3</span> and (7±5)×107 cm-<span class="hlt">3</span>, respectively. These radical concentrations were smaller than those observed during summer because of the reduced solar radiation. A chemical modulation device to separate atmospheric OH radicals from any interfering species was applied for few days showing negligible interference for both clean and polluted air masses.<span class="hlt">HONO</span> and <span class="hlt">HCHO</span> photolysis were found to be the most important primary sources of ROx radicals. CO and NOx were the important OH reactants which contributed more than half of the total OH reactivity. The relative high OH concentrations in polluted episode enabled a fast oxidation of fresh emitted pollutants and the formation of secondary air products. The observed radical concentrations were compared with the results from a chemical box model. The model is capable of reproducing radical concentrations for moderate NOx conditions but larger discrepancies are observed for both low and high NOx regimes for the peroxy radical concentrations. The underestimation of RO2 radical concentrations for high NOx conditions is discussed in the context of recent campaigns.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20204685','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20204685"><span>Spectroscopic properties and energy transfer analysis of Tm<span class="hlt">3</span>+-doped BaF2-Ga2<span class="hlt">O</span><span class="hlt">3</span>-Ge<span class="hlt">O</span>2-La2<span class="hlt">O</span><span class="hlt">3</span> glass.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yu, Shenglei; Yang, Zhongmin; Xu, Shanhui</p> <p>2010-05-01</p> <p>This paper reports on the spectroscopic properties and energy transfer analysis of Tm(<span class="hlt">3</span>+)-doped BaF(2)-Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>)-Ge<span class="hlt">O</span>(2)-La(2)<span class="hlt">O</span>(<span class="hlt">3</span>) glasses with different Tm(2)<span class="hlt">O</span>(<span class="hlt">3</span>) doping concentrations (0.2, 0.5, 2.0, 2.5, <span class="hlt">3</span>.0, <span class="hlt">3</span>.5, <span class="hlt">3</span>.5, 4.0 wt%). Mid-IR fluorescence intensities in the range of 1,300 nm-2,200 nm have been measured when excited under an 808 nm LD for all the samples with the same pump power. Energy level structure and Judd-Ofelt parameters have been calculated based on the absorption spectra of Tm(<span class="hlt">3</span>+), cross-relaxation rates and multi-phonon relaxation rates have been estimated with different Tm(2)<span class="hlt">O</span>(<span class="hlt">3</span>) doping concentrations. The maximum fluorescence intensity at around 1.8 mum has been obtained in Tm(2)<span class="hlt">O</span>(<span class="hlt">3)-3</span> wt% sample and the maximum value of calculated stimulated emission cross-section of Tm(<span class="hlt">3</span>+) in this sample is about 0.48 x 10(-20) cm(2) at 1,793 nm, and there is not any crystallization peak in the DSC curve of this sample, which indicate the potential utility of Tm(<span class="hlt">3</span>+)-doped BaF(2)-Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>)-Ge<span class="hlt">O</span>(2)- La(2)<span class="hlt">O</span>(<span class="hlt">3</span>) glass for 2.0-microm optical fiber laser.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5844055','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5844055"><span>Electronic properties and surface reactivity of Sr<span class="hlt">O</span>-terminated SrTi<span class="hlt">O</span><span class="hlt">3</span> and Sr<span class="hlt">O</span>-terminated iron-doped SrTi<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Staykov, Aleksandar; Tellez, Helena; Druce, John; Wu, Ji; Ishihara, Tatsumi; Kilner, John</p> <p>2018-01-01</p> <p>Abstract Surface reactivity and near-surface electronic properties of Sr<span class="hlt">O</span>-terminated SrTi<span class="hlt">O</span><span class="hlt">3</span> and iron doped SrTi<span class="hlt">O</span><span class="hlt">3</span> were studied with first principle methods. We have investigated the density of states (DOS) of bulk SrTi<span class="hlt">O</span><span class="hlt">3</span> and compared it to DOS of iron-doped SrTi<span class="hlt">O</span><span class="hlt">3</span> with different oxidation states of iron corresponding to varying oxygen vacancy content within the bulk material. The obtained bulk DOS was compared to near-surface DOS, i.e. surface states, for both Sr<span class="hlt">O</span>-terminated surface of SrTi<span class="hlt">O</span><span class="hlt">3</span> and iron-doped SrTi<span class="hlt">O</span><span class="hlt">3</span>. Electron density plots and electron density distribution through the entire slab models were investigated in order to understand the origin of surface electrons that can participate in oxygen reduction reaction. Furthermore, we have compared oxygen reduction reactions at elevated temperatures for Sr<span class="hlt">O</span> surfaces with and without oxygen vacancies. Our calculations demonstrate that the conduction band, which is formed mainly by the d-states of Ti, and Fe-induced states within the band gap of SrTi<span class="hlt">O</span><span class="hlt">3</span>, are accessible only on Ti<span class="hlt">O</span>2 terminated SrTi<span class="hlt">O</span><span class="hlt">3</span> surface while the Sr<span class="hlt">O</span>-terminated surface introduces a tunneling barrier for the electrons populating the conductance band. First principle molecular dynamics demonstrated that at elevated temperatures the surface oxygen vacancies are essential for the oxygen reduction reaction. PMID:29535797</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120000846','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120000846"><span>The Vaporization of B2<span class="hlt">O</span><span class="hlt">3</span>(l) to B2<span class="hlt">O</span><span class="hlt">3</span>(g) and B2<span class="hlt">O</span>2(g)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jacobson, Nathan S.; Myers, Dwight L.</p> <p>2011-01-01</p> <p>The vaporization of B2<span class="hlt">O</span><span class="hlt">3</span> in a reducing environment leads to formation of both B2<span class="hlt">O</span><span class="hlt">3</span>(g) and B2<span class="hlt">O</span>2(g). While formation of B2<span class="hlt">O</span><span class="hlt">3</span>(g) is well understood, many questions about the formation of B2<span class="hlt">O</span>2(g) remain. Previous studies using B(s) + B2<span class="hlt">O</span><span class="hlt">3</span>(l) have led to inconsistent thermodynamic data. In this study, it was found that after heating, B(s) and B2<span class="hlt">O</span><span class="hlt">3</span>(l) appear to separate and variations in contact area likely led to the inconsistent vapor pressures of B2<span class="hlt">O</span>2(g). To circumvent this problem, an activity of boron is fixed with a two-phase mixture of FeB and Fe2B. Both second and third law enthalpies of formation were measured for B2<span class="hlt">O</span>2(g) and B2<span class="hlt">O</span><span class="hlt">3</span>(g). From these the enthalpies of formation at 298.15 K are calculated to be -479.9 +/- 41.5 kJ/mol for B2<span class="hlt">O</span>2(g) and -833.4 +/- 13.1 kJ/mol for B2<span class="hlt">O</span><span class="hlt">3</span>(g). Ab initio calculations to determine the enthalpies of formation of B2<span class="hlt">O</span>2(g) and B2<span class="hlt">O</span><span class="hlt">3</span>(g) were conducted using the W1BD composite method and show good agreement with the experimental values.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950054972&hterms=perovskite&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dperovskite','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950054972&hterms=perovskite&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dperovskite"><span>MgSi<span class="hlt">O</span><span class="hlt">3</span>-FeSi<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> in the Earth's lower mantle: Perovskite and garnet at 1200 km depth</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>O'Neill, Bridget; Jeanloz, Raymond</p> <p>1994-01-01</p> <p>Natural pyroxene and garnet starting material are used to study the effects of joint Fe and Al substitution into MgSi<span class="hlt">O</span><span class="hlt">3</span> perovskite at approxmiately 50 GPa. Garnet is found to coexist with perovskite in samples containing both Fe and Al to pressures occurring deep into the lower mantel (approximately 1200 km depth). The volume of the perovskite unit cell is V(sub <span class="hlt">o</span>(Angstrom(exp <span class="hlt">3</span>)) = 162.59 + 5.95x(sub FeSi<span class="hlt">O</span><span class="hlt">3</span>) + 10.80x(sub Al2<span class="hlt">O</span><span class="hlt">3</span>) with aluminum causing a significant increase in the distortion from the ideal cubic cell. On the basis of a proposed extension of the MgSi<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> high-pressure phase diagram toward FeSi<span class="hlt">O</span><span class="hlt">3</span>, Fe is shown to partition preferentially into the garnet phase. The stability of garnet deep into the lower mantel may hinder the penetration of subducted slabs below the transition zone.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JAP...122i4901S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JAP...122i4901S"><span>Photocatalytic self-cleaning transparent 2Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span> glass ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sharma, Sumeet Kumar; Singh, V. P.; Chauhan, Vishal S.; Kushwaha, H. S.; Vaish, Rahul</p> <p>2017-09-01</p> <p>Photocatalytic response of as-quenched and heat-treated 2Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span> glasses was studied. X ray diffraction reveals that the controlled heat treatment of glasses at 380 °C for 1 h, 2 h, and <span class="hlt">3</span> h shows the formation of Bi4B2<span class="hlt">O</span>9 crystals embedded in 2Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span> the host glass matrix. Scanning electron microscopic images reveal the presence of nanocrystallization in as-quenched glass. Significant photocatalytic activities were observed in as-quenched transparent glass. Photocatalytic activities were studied using the degradation of Resazurin as well as pharmaceutical 17 β-Estradiol under UV irradiation. Measurement of contact angle shows enhanced hydrophilicity with the increase in crystallization of the samples. Further, for as quenched 2Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span> glass ceramic, under UV irradiation, the water contact angle decreased from 92.7° to 39.5° and the sample surface transformed from hydrophobic to hydrophilic. Effective photocatalytic performance along with photoinduced hydrophilicity promotes 2Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span> glass ceramics in self-cleaning applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1377858-interface-control-ferroelectricity-srruo3-batio3-srruo3-capacitor-its-critical-thickness','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1377858-interface-control-ferroelectricity-srruo3-batio3-srruo3-capacitor-its-critical-thickness"><span>Interface Control of Ferroelectricity in an SrRu<span class="hlt">O</span> <span class="hlt">3</span>/BaTi<span class="hlt">O</span> <span class="hlt">3</span>/SrRu<span class="hlt">O</span> <span class="hlt">3</span> Capacitor and its Critical Thickness</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Shin, Yeong Jae; Kim, Yoonkoo; Kang, Sung -Jin; ...</p> <p>2017-03-03</p> <p>Here, the atomic-scale synthesis of artificial oxide heterostructures offers new opportunities to create novel states that do not occur in nature. The main challenge related to synthesizing these structures is obtaining atomically sharp interfaces with designed termination sequences. In this study, it is demonstrated that the oxygen pressure (P <span class="hlt">O</span>2) during growth plays an important role in controlling the interfacial terminations of SrRu<span class="hlt">O</span> <span class="hlt">3</span>/BaTi<span class="hlt">O</span> <span class="hlt">3</span>/SrRu<span class="hlt">O</span> <span class="hlt">3</span> (SRO/BTO/SRO) ferroelectric (FE) capacitors. The SRO/BTO/SRO heterostructures are grown by a pulsed laser deposition method. The top SRO/BTO interface, grown at high P <span class="hlt">O</span>2 (around 150 mTorr), usually exhibits a mixture of Ru<span class="hlt">O</span> 2-BaOmore » and Sr<span class="hlt">O-TiO</span> 2 terminations. By reducing P <span class="hlt">O</span>2, the authors obtain atomically sharp SRO/BTO top interfaces with uniform Sr<span class="hlt">O-TiO</span> 2 termination. Using capacitor devices with symmetric and uniform interfacial termination, it is demonstrated for the first time that the FE critical thickness can reach the theoretical limit of <span class="hlt">3</span>.5 unit cells.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24620399','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24620399"><span>Improving the indoor air quality of respiratory type of medical facility by zeolite filtering.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shen, Jyun-Hong; Wang, Yeoung-Sheng; Lin, Jhan-Ping; Wu, Sheng-Hung; Horng, Jao-Jia</p> <p>2014-01-01</p> <p>This study investigated the indoor air quality (IAQ) conditions of carbon dioxide (CO2), carbon monoxide (CO), ozone (<span class="hlt">O</span><span class="hlt">3</span>), formaldehyde (<span class="hlt">HCHO</span>), total volatile organic compounds (TVOCs), and bio-aerosols (bacteria and fungi) in a respiratory type of medical facility in Chia-Yi County in southern Taiwan. Among those IAQ conditions, the concentrations of CO, <span class="hlt">O</span><span class="hlt">3</span>, and <span class="hlt">HCHO</span> exceeded the regulation values of the Taiwan Environmental Protection Administration (EPA) mostly in the morning. The concentrations of bacteria and fungi did not exceed the regulation values but still posed potential health and environment problems for workers, patients, and visitors. Therefore, self-made silver-coated zeolite (AgZ) was used as a filter material in air cleaners to remove bio-aerosols in the respiratory care ward (RCW), and the removals were still effective after 120 hr. The cumulative bio-aerosol removals for bacteria and fungi were 900 and 1,088 colony-forming units (CFU) g(-1) after 24 hr and were above <span class="hlt">3</span>,100 and 2,700 CFU g(-1) after 120 hr. From the research results, it is suggested that AgZ filtering could be used as a feasible engineering measure for hospitals to control their bacteria and fungi parameters in IAQ management. Hospitals should maintain their environmental management and monitoring programs and use different engineering measures to improve different IAQ parameters. This study investigated the IAQ conditions in the field at a hospital in Chia-Yi County in southern Taiwan. Although concentrations of most parameters were still within the regulation values, the concentrations of CO, <span class="hlt">O</span><span class="hlt">3</span>, and <span class="hlt">HCHO</span> were partially exceeded. We propose a method using an air cleaner with silver-coated zeolite (AgZ) as a possible engineering measure, and there were effective reductions of bacteria and fungi to lower levels with antibacterial effects after 120 hr. Furthermore, this study implies that hospitals should continuously maintain environmental monitoring programs and adopt optimal</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21421488-ion-beam-irradiation-lanthanum-compounds-systems-la-sub-sub-al-sub-sub-la-sub-sub-tio-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21421488-ion-beam-irradiation-lanthanum-compounds-systems-la-sub-sub-al-sub-sub-la-sub-sub-tio-sub"><span>Ion-beam irradiation of lanthanum compounds in the systems La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}-Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} and La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}-Ti<span class="hlt">O</span>{sub 2}</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Whittle, Karl R., E-mail: karl.whittle@ansto.gov.a; Lumpkin, Gregory R.; Blackford, Mark G.</p> <p>2010-10-15</p> <p>Thin crystals of La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}, LaAl<span class="hlt">O</span>{sub <span class="hlt">3</span>}, La{sub 2/<span class="hlt">3</span>}Ti<span class="hlt">O</span>{sub <span class="hlt">3</span>}, La{sub 2}Ti<span class="hlt">O</span>{sub 5}, and La{sub 2}Ti{sub 2}<span class="hlt">O</span>{sub 7} have been irradiated in situ using 1 MeV Kr{sup 2+} ions at the Intermediate Voltage Electron Microscope-Tandem User Facility (IVEM-Tandem), Argonne National Laboratory (ANL). We observed that La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} remained crystalline to a fluence greater than <span class="hlt">3</span>.1x10{sup 16} ions cm{sup -2} at a temperature of 50 K. The four binary oxide compounds in the two systems were observed through the crystalline-amorphous transition as a function of ion fluence and temperature. Results from the ion irradiations give critical temperatures for amorphisationmore » (T{sub c}) of 647 K for LaAl<span class="hlt">O</span>{sub <span class="hlt">3</span>}, 840 K for La{sub 2}Ti{sub 2}<span class="hlt">O</span>{sub 7}, 865 K for La{sub 2/<span class="hlt">3</span>}Ti<span class="hlt">O</span>{sub <span class="hlt">3</span>}, and 1027 K for La{sub 2}Ti<span class="hlt">O</span>{sub 5}. The T{sub c} values observed in this study, together with previous data for Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} and Ti<span class="hlt">O</span>{sub 2}, are discussed with reference to the melting points for the La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}-Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} and La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}-Ti<span class="hlt">O</span>{sub 2} systems and the different local environments within the four crystal structures. Results suggest that there is an observable inverse correlation between T{sub c} and melting temperature (T{sub m}) in the two systems. More complex relationships exist between T{sub c} and crystal structure, with the stoichiometric perovskite LaAl<span class="hlt">O</span>{sub <span class="hlt">3</span>} being the most resistant to amorphisation. - Graphical abstract: La{sub 2}Ti<span class="hlt">O</span>{sub 5} with atypical co-ordination for Ti, Ti<span class="hlt">O</span>{sub 5} is found to be different in radiation resistance to La{sub 2}Ti{sub 2}<span class="hlt">O</span>{sub 7} and La{sub 2/<span class="hlt">3</span>}Ti<span class="hlt">O</span>{sub <span class="hlt">3</span>}. Irradiation of La-Ti-<span class="hlt">O</span>, and La-Al-<span class="hlt">O</span> based systems has found that radiation damage resistance is related to the ability of the system to disorder.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ApSS..256.6801L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ApSS..256.6801L"><span>Chemical quenching of positronium in Fe 2<span class="hlt">O</span> <span class="hlt">3</span>/Al 2<span class="hlt">O</span> <span class="hlt">3</span> catalysts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, C.; Zhang, H. J.; Chen, Z. Q.</p> <p>2010-09-01</p> <p>Fe 2<span class="hlt">O</span> <span class="hlt">3</span>/Al 2<span class="hlt">O</span> <span class="hlt">3</span> catalysts were prepared by solid state reaction method using α-Fe 2<span class="hlt">O</span> <span class="hlt">3</span> and γ-Al 2<span class="hlt">O</span> <span class="hlt">3</span> nano powders. The microstructure and surface properties of the catalyst were studied using positron lifetime and coincidence Doppler broadening annihilation radiation measurements. The positron lifetime spectrum shows four components. The two long lifetimes τ<span class="hlt">3</span> and τ4 are attributed to positronium annihilation in two types of pores distributed inside Al 2<span class="hlt">O</span> <span class="hlt">3</span> grain and between the grains, respectively. With increasing Fe 2<span class="hlt">O</span> <span class="hlt">3</span> content from <span class="hlt">3</span> wt% to 40 wt%, the lifetime τ<span class="hlt">3</span> keeps nearly unchanged, while the longest lifetime τ4 shows decrease from 96 ns to 64 ns. Its intensity decreases drastically from 24% to less than 8%. The Doppler broadening S parameter shows also a continuous decrease. Further analysis of the Doppler broadening spectra reveals a decrease in the p-Ps intensity with increasing Fe 2<span class="hlt">O</span> <span class="hlt">3</span> content, which rules out the possibility of spin-conversion of positronium. Therefore the decrease of τ4 is most probably due to the chemical quenching reaction of positronium with Fe ions on the surface of the large pores.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29773388','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29773388"><span>A unique polysaccharide containing <span class="hlt">3</span>-<span class="hlt">O</span>-methylarabinose and <span class="hlt">3</span>-<span class="hlt">O</span>-methylgalactose from Tinospora sinensis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nagar, Shipra; Hensel, Andreas; Mischnick, Petra; Kumar, Vineet</p> <p>2018-08-01</p> <p>Tinospora sinensis (Lour.) Merrill is of great therapeutic significance in Indian traditional medicine. Crude polysaccharides were isolated from methanol pre-extracted stems of dried material by successive extractions with cold water, hot water and NaOH (0.25 mol/L) in 0.98, 0.55 and 0.70 % yields respectively. Cold water soluble polysaccharides (CWSP) were purified and fractionated by ion exchange chromatography on DEAE-Sephacel. Neutral polysaccharides were further fractionated on Sepharose CL6B to yield three fractions TW1, TW2, TW<span class="hlt">3</span>. The study further focuses on structural elucidation of TW1. TW1 was obtained in 0.8 % yield relative to CWSP, with MW of 1.6 × 10 5  Da. It was composed of <span class="hlt">3</span>-<span class="hlt">O</span>-methyl-arabinose, <span class="hlt">3</span>-<span class="hlt">O</span>-methyl-galactose and galactose in molar ratio of 1.0:6.<span class="hlt">3</span>:0.9 respectively. Based on per-deuteromethylation, NMR and ESI-MS analyses, TW1 was composed of 1,4-linked <span class="hlt">3</span>-<span class="hlt">O</span>-methyl-β-d-galactopyranose and β-d-galactopyranose backbone with branching at <span class="hlt">O</span>-6 of <span class="hlt">3</span>-<span class="hlt">O</span>-methyl-β-d-galactosyl residues by 1,5-linked <span class="hlt">3</span>-<span class="hlt">O</span>-methyl-α-l-arabinofuranoside chains. <span class="hlt">3</span>-<span class="hlt">O</span>-methyl-arabinose and <span class="hlt">3</span>-<span class="hlt">O</span>-methyl-galactose have first ever been reported in any polysaccharide and Tinospora genus, respectively. Copyright © 2018 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AMT.....9..423M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AMT.....9..423M"><span>A broadband cavity enhanced absorption spectrometer for aircraft measurements of glyoxal, methylglyoxal, nitrous acid, nitrogen dioxide, and water vapor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Min, K.-E.; Washenfelder, R. A.; Dubé, W. P.; Langford, A. O.; Edwards, P. M.; Zarzana, K. J.; Stutz, J.; Lu, K.; Rohrer, F.; Zhang, Y.; Brown, S. S.</p> <p>2016-02-01</p> <p>We describe a two-channel broadband cavity enhanced absorption spectrometer (BBCEAS) for aircraft measurements of glyoxal (CHOCHO), methylglyoxal (CH<span class="hlt">3</span>COCHO), nitrous acid (<span class="hlt">HONO</span>), nitrogen dioxide (NO2), and water (H2<span class="hlt">O</span>). The instrument spans 361-389 and 438-468 nm, using two light-emitting diodes (LEDs) and a single grating spectrometer with a charge-coupled device (CCD) detector. Robust performance is achieved using a custom optical mounting system, high-power LEDs with electronic on/off modulation, high-reflectivity cavity mirrors, and materials that minimize analyte surface losses. We have successfully deployed this instrument during two aircraft and two ground-based field campaigns to date. The demonstrated precision (2σ) for retrievals of CHOCHO, <span class="hlt">HONO</span> and NO2 are 34, 350, and 80 parts per trillion (pptv) in 5 s. The accuracy is 5.8, 9.0, and 5.0 %, limited mainly by the available absorption cross sections.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AMTD....811209M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AMTD....811209M"><span>A broadband cavity enhanced absorption spectrometer for aircraft measurements of glyoxal, methylglyoxal, nitrous acid, nitrogen dioxide, and water vapor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Min, K.-E.; Washenfelder, R. A.; Dubé, W. P.; Langford, A. O.; Edwards, P. M.; Zarzana, K. J.; Stutz, J.; Lu, K.; Rohrer, F.; Zhang, Y.; Brown, S. S.</p> <p>2015-10-01</p> <p>We describe a two-channel broadband cavity enhanced absorption spectrometer (BBCEAS) for aircraft measurements of glyoxal (CHOCHO), methylglyoxal (CH<span class="hlt">3</span>COCHO), nitrous acid (<span class="hlt">HONO</span>), nitrogen dioxide (NO2), and water (H2<span class="hlt">O</span>). The instrument spans 361-389 and 438-468 nm, using two light emitting diodes (LEDs) and a grating spectrometer with a charge-coupled device (CCD) detector. Robust performance is achieved using a custom optical mounting system, high power LEDs with electronic on/off modulation, state-of-the-art cavity mirrors, and materials that minimize analyte surface losses. We have successfully deployed this instrument during two aircraft and two ground-based field campaigns to date. The demonstrated precision (2σ) for retrievals of CHOCHO, <span class="hlt">HONO</span> and NO2 are 34, 350 and 80 pptv in 5 s. The accuracy is 5.8, 9.0 and 5.0 % limited mainly by the available absorption cross sections.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040111416&hterms=space+mapping&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dspace%2Bmapping','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040111416&hterms=space+mapping&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dspace%2Bmapping"><span>Mapping Isoprene Emissions over North America using Formaldehyde Column Observations from Space</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Palmer, Paul I.; Jacob, Daniel J.; Fiore, Arlene M.; Martin, Randall V.; Chance, Kelly; Kurosu, Thomas P.</p> <p>2004-01-01</p> <p>I] We present a methodology for deriving emissions of volatile organic compounds (VOC) using space-based column observations of formaldehyde (<span class="hlt">HCHO</span>) and apply it to data from the Global Ozone Monitoring Experiment (GOME) satellite instrument over North America during July 1996. The <span class="hlt">HCHO</span> column is related to local VOC emissions, with a spatial smearing that increases with the VOC lifetime. lsoprene is the dominant <span class="hlt">HCHO</span> precursor over North America in summer, and its lifetime (approx. = 1 hour) is sufficiently short that the smearing can be neglected. We use the Goddard Earth Observing System global <span class="hlt">3</span>-D model of tropospheric chemistry (GEOS-CHEM) to derive the relationship between isoprene emissions and <span class="hlt">HCHO</span> columns over North America and use these relationships to convert the GOME <span class="hlt">HCHO</span> columns to isoprene emissions. We also use the GEOS-CHEM model as an intermediary to validate the GOME <span class="hlt">HCHO</span> column measurements by comparison with in situ observations. The GEOS-CHEM model including the Global Emissions Inventory Activity (GEIA) isoprene emission inventory provides a good simulation of both the GOME data (r(sup 2) = 0.69, n = 756, bias = +l1 %) and the in situ summertime <span class="hlt">HCHO</span> measurements over North America (r(sup 2) = 0.47, n = 10, bias = -<span class="hlt">3</span>%). The GOME observations show high values over regions of known high isoprene emissions and a day-to-day variability that is consistent with the temperature dependence of isoprene emission. Isoprene emissions inferred from the GOME data are 20% less than GEIA on average over North America and twice those from the U S . EPA Biogenic Emissions Inventory System (BEIS2) inventory. The GOME isoprene inventory when implemented in the GEOS-CHEM model provides a better simulation of the <span class="hlt">HCHO</span> in situ measurements thaneitherGEIAorBEIS2 (r(sup 2) = 0.71,n= 10, bias = -10 %).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A53H..02L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A53H..02L"><span>Formation Of Nitrous Acid In Various Environments And Its Contribution To Hydroxyl Radical Budget</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, X.; Lu, K.; Liu, Y.; Rohrer, F.; Häseler, R.; Bohn, B.; Fuchs, H.; Hofzumahaus, A.; Wahner, A.; Kiendler-Scharr, A.; Zeng, L.; Zhang, Y.</p> <p>2017-12-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) is an important trace gas in the atmosphere due to its contribution to the cycles of nitrogen oxides (NOx) and hydrogen oxides (HOx). Yet the formation mechanism of <span class="hlt">HONO</span> during daytime remains large uncertainty. In the past ten years, we performed ground <span class="hlt">HONO</span> measurements using the LOPAP technique in different seasons in two major mega-city areas, i.e., North China Plain and Pearl River Delta. Spatial distribution of <span class="hlt">HONO</span> over different regions in Europe was also observed by the same technique on-board the Zeppelin NT airship in 2012 and 2013. For all field observations, parameters (OH, HO2, NOx, photolysis frequencies, aerosols, etc.) influencing the <span class="hlt">HONO</span> budget were measured simultaneous, which allows us to investigate the source of <span class="hlt">HONO</span> and its contribution to the hydroxyl radical budget. In this presentation, we summarize the general features of <span class="hlt">HONO</span> in the planetary boundary layer in terms of its daytime concentration, spatial distribution, and source strength, and contribution to the primary OH production. Various proposed <span class="hlt">HONO</span> formation mechanisms have been tested in order to explain the widely exited missing daytime <span class="hlt">HONO</span> production of 100 - 200 ppt h-1 . We will show that there is not a stand alone mechanism can be used for explaining the <span class="hlt">HONO</span> formation in various environments. Design of field experiments which effectively separate different physical and chemical processes would be helpful to pin down the major daytime <span class="hlt">HONO</span> source and to understand its influence on the hydroxyl radical budget.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23368496','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23368496"><span>Nature of weak magnetism in SrTi<span class="hlt">O</span><span class="hlt">3</span>/LaAl<span class="hlt">O</span><span class="hlt">3</span> multilayers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Salman, Z; Ofer, O; Radovic, M; Hao, H; Ben Shalom, M; Chow, K H; Dagan, Y; Hossain, M D; Levy, C D P; Macfarlane, W A; Morris, G M; Patthey, L; Pearson, M R; Saadaoui, H; Schmitt, T; Wang, D; Kiefl, R F</p> <p>2012-12-21</p> <p>We report the observation of weak magnetism in superlattices of LaAl<span class="hlt">O</span>(<span class="hlt">3</span>)/SrTi<span class="hlt">O</span>(<span class="hlt">3</span>) using β-detected nuclear magnetic resonance. The spin lattice relaxation rate of ^{8}Li in superlattices with a spacer layers of 8 and 6 unit cells of LaAl<span class="hlt">O</span>(<span class="hlt">3</span>) exhibits a strong peak near ~35 K, whereas no such peak is observed in a superlattice with spacer layer thickness of <span class="hlt">3</span> unit cells. We attribute the observed temperature dependence to slowing down of weakly coupled electronic moments at the LaAl<span class="hlt">O</span>(<span class="hlt">3</span>)/SrTi<span class="hlt">O</span>(<span class="hlt">3</span>) interface. These results show that the magnetism at the interface depends strongly on the thickness of the spacer layer, and that a minimal thickness of ~4-6 unit cells is required for the appearance of magnetism. A simple model is used to determine that the observed relaxation is due to small fluctuating moments (~0.002μ(B)) in the two samples with a larger LaAl<span class="hlt">O</span>(<span class="hlt">3</span>) spacer thickness.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17407271','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17407271"><span>Quenching of I(2P1/2) by <span class="hlt">O</span><span class="hlt">3</span> and <span class="hlt">O</span>(<span class="hlt">3</span>P).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Azyazov, Valeriy N; Antonov, Ivan O; Heaven, Michael C</p> <p>2007-04-26</p> <p>Oxygen-iodine lasers that utilize electrical or microwave discharges to produce singlet oxygen are currently being developed. The discharge generators differ from conventional chemical singlet oxygen generators in that they produce significant amounts of atomic oxygen. Post-discharge chemistry includes channels that lead to the formation of ozone. Consequently, removal of I(2P1/2) by <span class="hlt">O</span> atoms and <span class="hlt">O</span><span class="hlt">3</span> may impact the efficiency of discharge driven iodine lasers. In the present study, we have measured the rate constants for quenching of I(2P1/2) by <span class="hlt">O</span>(<span class="hlt">3</span>P) atoms and <span class="hlt">O</span><span class="hlt">3</span> using pulsed laser photolysis techniques. The rate constant for quenching by <span class="hlt">O</span><span class="hlt">3</span>, (1.8 +/- 0.4) x 10(-12) cm<span class="hlt">3</span> s-1, was found to be a factor of 5 smaller than the literature value. The rate constant for quenching by <span class="hlt">O</span>(<span class="hlt">3</span>P) was (1.2 +/- 0.2) x 10(-11) cm<span class="hlt">3</span> s-1.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1953i0074H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1953i0074H"><span>Effect of RE (Nd<span class="hlt">3</span>+, Sm<span class="hlt">3</span>+) oxide on structural, optical properties of Na2<span class="hlt">O</span>-Li2<span class="hlt">O-ZnO</span>-B2<span class="hlt">O</span><span class="hlt">3</span> glass system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hivrekar, Mahesh M.; Bhoyar, D. N.; Mande, V. K.; Dhole, V. V.; Solunke, M. B.; Jadhav, K. M.</p> <p>2018-05-01</p> <p>Zinc borate glass activated with rare earth oxide (Nd2<span class="hlt">O</span><span class="hlt">3</span>, Sm2<span class="hlt">O</span><span class="hlt">3</span>) of Na2<span class="hlt">O</span>-Li2<span class="hlt">O-ZnO</span>-B2<span class="hlt">O</span><span class="hlt">3</span> quaternary system has been prepared successfully by melt quenching method. The nucleation and growth of RE oxide were controlled temperature range 950-1000° C and rapid cooling at room temperature. The physical, structural and optical properties were characterized by using X-ray diffraction (XRD), SEM, Ultraviolet-visible spectroscopy (UV-Vis). XRD and SEM studies confirmed the amorphous nature, surface morphology of prepared zinc borate glass. The physical parameters like density, molar volume, molar mass of Nd<span class="hlt">3</span>+, Sm<span class="hlt">3</span>+ doped borate glass are summarized in the present article. The optical absorption spectra along with tauc's plot are presented. The optical energy band gap increases due to the addition of rare earth oxide confirming the role of network modifier.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007ApPhL..91w2912L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007ApPhL..91w2912L"><span>Properties of highly (100) oriented Pb(Mg1/<span class="hlt">3</span>,Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-PbTi<span class="hlt">O</span><span class="hlt">3</span> films on LaNi<span class="hlt">O</span><span class="hlt">3</span> bottom electrodes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Y. W.; Hu, Z. G.; Yue, F. Y.; Yang, G. Y.; Shi, W. Z.; Meng, X. J.; Sun, J. L.; Chu, J. H.</p> <p>2007-12-01</p> <p>The 70%Pb(Mg1/<span class="hlt">3</span>,Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-30%PbTi<span class="hlt">O</span><span class="hlt">3</span> (PMNT) films have been fabricated on LaNi<span class="hlt">O</span><span class="hlt">3</span> (LNO) coated silicon substrate. The conductive LNO films act as a seed layer for the growth of PMNT films, which depresses the formation of pyrochlore phase and induces the high (100) preferred orientation of perovskite PMNT films. Compared with the PMNT films grown on platinum bottom electrode, the ferroelectric properties of PMNT films grown on LNO are enhanced. The frequency dependence of complex permittivity from PMNT films on LNO is the conjunct result of polarization relaxation and movement of oxygen vacancy, which can be fitted by the function containing Debye and universal dielectric response models, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014MatSP..32...80M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014MatSP..32...80M"><span>Optical properties of the Al2<span class="hlt">O</span><span class="hlt">3</span>/Si<span class="hlt">O</span>2 and Al2<span class="hlt">O</span><span class="hlt">3</span>/Hf<span class="hlt">O</span>2/Si<span class="hlt">O</span>2 antireflective coatings</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marszałek, Konstanty; Winkowski, Paweł; Jaglarz, Janusz</p> <p>2014-01-01</p> <p>Investigations of bilayer and trilayer Al2<span class="hlt">O</span><span class="hlt">3</span>/Si<span class="hlt">O</span>2 and Al2<span class="hlt">O</span><span class="hlt">3</span>/Hf<span class="hlt">O</span>2/Si<span class="hlt">O</span>2 antireflective coatings are presented in this paper. The oxide films were deposited on a heated quartz glass by e-gun evaporation in a vacuum of 5 × 10-<span class="hlt">3</span> [Pa] in the presence of oxygen. Depositions were performed at three different temperatures of the substrates: 100 °C, 200 °C and 300 °C. The coatings were deposited onto optical quartz glass (Corning HPFS). The thickness and deposition rate were controlled with Inficon XTC/2 thickness measuring system. Deposition rate was equal to 0.6 nm/s for Al2<span class="hlt">O</span><span class="hlt">3</span>, 0.6 nm - 0.8 nm/s for Hf<span class="hlt">O</span>2 and 0.6 nm/s for Si<span class="hlt">O</span>2. Simulations leading to optimization of the thin film thickness and the experimental results of optical measurements, which were carried out during and after the deposition process, have been presented. The optical thickness values, obtained from the measurements performed during the deposition process were as follows: 78 nm/78 nm for Al2<span class="hlt">O</span><span class="hlt">3</span>/Si<span class="hlt">O</span>2 and 78 nm/156 nm/78 nm for Al2<span class="hlt">O</span><span class="hlt">3</span>/Hf<span class="hlt">O</span>2/Si<span class="hlt">O</span>2. The results were then checked by ellipsometric technique. Reflectance of the films depended on the substrate temperature during the deposition process. Starting from 240 nm to the beginning of visible region, the average reflectance of the trilayer system was below 1 % and for the bilayer, minima of the reflectance were equal to 1.6 %, 1.15 % and 0.8 % for deposition temperatures of 100 °C, 200 °C and 300 °C, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhRvL.109s7202H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhRvL.109s7202H"><span>Interfacial Ferromagnetism and Exchange Bias in CaRu<span class="hlt">O</span><span class="hlt">3</span>/CaMn<span class="hlt">O</span><span class="hlt">3</span> Superlattices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>He, C.; Grutter, A. J.; Gu, M.; Browning, N. D.; Takamura, Y.; Kirby, B. J.; Borchers, J. A.; Kim, J. W.; Fitzsimmons, M. R.; Zhai, X.; Mehta, V. V.; Wong, F. J.; Suzuki, Y.</p> <p>2012-11-01</p> <p>We have found ferromagnetism in epitaxially grown superlattices of CaRu<span class="hlt">O</span><span class="hlt">3</span>/CaMn<span class="hlt">O</span><span class="hlt">3</span> that arises in one unit cell at the interface. Scanning transmission electron microscopy and electron energy loss spectroscopy indicate that the difference in magnitude of the Mn valence states between the center of the CaMn<span class="hlt">O</span><span class="hlt">3</span> layer and the interface region is consistent with double exchange interaction among the Mn ions at the interface. Polarized neutron reflectivity and the CaMn<span class="hlt">O</span><span class="hlt">3</span> thickness dependence of the exchange bias field together indicate that the interfacial ferromagnetism is only limited to one unit cell of CaMn<span class="hlt">O</span><span class="hlt">3</span> at each interface. The interfacial moment alternates between the 1μB/interface Mn ion for even CaMn<span class="hlt">O</span><span class="hlt">3</span> layers and the 0.5μB/interface Mn ion for odd CaMn<span class="hlt">O</span><span class="hlt">3</span> layers. This modulation, combined with the exchange bias, suggests the presence of a modulating interlayer coupling between neighboring ferromagnetic interfaces via the antiferromagnetic CaMn<span class="hlt">O</span><span class="hlt">3</span> layers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1832g0005G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1832g0005G"><span>Thermal characteristics, Raman spectra, optical and structural properties of Ti<span class="hlt">O</span>2-Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Te<span class="hlt">O</span>2 glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gupta, Nupur; Khanna, Atul; Gonzàlez, Fernando; Iordanova, Reni</p> <p>2017-05-01</p> <p>Tellurite and borotellurite glasses containing Bi2<span class="hlt">O</span><span class="hlt">3</span> and Ti<span class="hlt">O</span>2 were prepared and structure-property correlations were carried out by density measurements, X-ray Diffraction (XRD), Differential Scanning Calorimetry (DSC), Raman and UV-visible spectroscopy. Titanium tellurite glasses require high melt-cooling rates and were fabricated by splat quenching. On adding B2<span class="hlt">O</span><span class="hlt">3</span>, the glass forming ability (GFA) enhances, and glasses could be synthesized at lower quenching rates. The density of glasses shows a direct correlation with molecular mass of the constituents. UV-visible studies were used to determine the optical band gap and refractive index. Raman studies found that the co-ordination number of tellurium ions with oxygen (NTe-<span class="hlt">O</span>) decreases with the increase in B2<span class="hlt">O</span><span class="hlt">3</span> as well as Bi2<span class="hlt">O</span><span class="hlt">3</span> content while, Ti<span class="hlt">O</span>2 produce only a small decrease in NTe-<span class="hlt">O</span>, which explains the lower GFA of titanium tellurite glasses that do not contain Bi2<span class="hlt">O</span><span class="hlt">3</span> and B2<span class="hlt">O</span><span class="hlt">3</span>. DSC studies show that the glass transition temperature (Tg) increases with B2<span class="hlt">O</span><span class="hlt">3</span> and Ti<span class="hlt">O</span>2 concentrations and that Tg correlates well with bond enthalpy of the metal oxides.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.A53F..06G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.A53F..06G"><span>Light-induced heterogeneous reactions of NO2 on indoor surfaces: How they affect the balance of nitrous acid</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gomez Alvarez, E.; Soergel, M.; Bassil, S.; Zetzsch, C.; Gligorovski, S.; Wortham, H.</p> <p>2011-12-01</p> <p>Nitrous acid (<span class="hlt">HONO</span>) is an important indoor pollutant. The adverse health effects due to the formation of nitrosamines are well known. <span class="hlt">HONO</span> acts as a nitrosating agent after wall reactions of <span class="hlt">HONO</span> with nicotine [Sleiman et al., 2010]. Indoor air can be surprisingly rich in <span class="hlt">HONO</span> (homes with fireplaces, stoves, gas heating and cooking) and also surfaces are abundant. High <span class="hlt">HONO</span> concentrations have been measured in indoor environments, from the direct emissions and heterogeneous reactions of NO2 in darkness. However, the measured <span class="hlt">HONO</span> concentrations do not correspond to the <span class="hlt">HONO</span> levels determined by the models [Carslaw, 2007]. We have tested in a flow tube reactor on-line coupled to a NOx analyzer and a sensitive Long Path Absorption Photometry instrument, the behaviour of various indoor surfaces towards NO2 under simulated solar light irradiation (λ= 300-700 nm). Our study has allowed us to obtain a deeper knowledge on the mechanisms of heterogeneous formation of <span class="hlt">HONO</span>, quantifying the dependence of <span class="hlt">HONO</span> formation on behalf of NO2 concentration and relative humidity and the enhancement of <span class="hlt">HONO</span> formation in the presence of light. Pyrex, acidic detergent, alkaline detergent, paint and lacquer were tested on behalf of their heterogeneous reactivity towards NO2 in the absence and in presence of light. The results obtained demonstrated that indoor surfaces are photo-chemically active under atmospherically relevant conditions. The strongly alkaline surfaces (such as certain types of detergent) show a strong long-term uptake capacity. However, other surfaces such as detergents with a more acidic character released <span class="hlt">HONO</span>. In some cases such as paint and varnish, a strong <span class="hlt">HONO</span> release with light was detected, which was significantly higher than that obtained over clean glass surfaces. Certain organics present on their composition could exert a photo-sensitizing effect that may explain their increased reactivity. Unfortunately, the final balance points towards an important net</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004JSSCh.177.1501C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004JSSCh.177.1501C"><span>A new ferroelectric solid solution system of LaCr<span class="hlt">O</span> <span class="hlt">3</span>-BiCr<span class="hlt">O</span> <span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, J. I. L.; Kumar, M. Mahesh; Ye, Z.-G.</p> <p>2004-04-01</p> <p>A new perovskite solid solution system of (1- x)LaCr<span class="hlt">O</span> <span class="hlt">3</span>- xBiCr<span class="hlt">O</span> <span class="hlt">3</span> has been prepared by conventional solid-state reaction and sintering processes at 1200°C in a sealed Pt tube or a Bi 2<span class="hlt">O</span> <span class="hlt">3</span>-rich environment. A clean orthorhombic phase of LaCr<span class="hlt">O</span> <span class="hlt">3</span>-type structure is established at room temperature for compositions with 0⩽ x⩽0.35. The relative density, lattice parameters, sintering mechanism, microstructure and ferroelectricity of the compounds are investigated. The substitution of Bi 2<span class="hlt">O</span> <span class="hlt">3</span> for La 2<span class="hlt">O</span> <span class="hlt">3</span> is found to decrease the unit cell volume and increase the grain size of the ceramics. The relative density of the ceramics sintered at 1200°C is significantly improved from 40% for LaCr<span class="hlt">O</span> <span class="hlt">3</span> up to about 90% for La 0.65Bi 0.35Cr<span class="hlt">O</span> <span class="hlt">3</span> through a liquid phase sintering mechanism. The ferroelectricity is revealed in La 1- xBi xCr<span class="hlt">O</span> <span class="hlt">3</span> with 0.1⩽ x⩽0.35 by dielectric hysteresis loops displayed at 77 K. The remnant polarization is found to increase with increasing Bi <span class="hlt">3</span>+ content. The origin of the ferroelectricity is attributed to the structural distortion induced by the stereochemically active Bi <span class="hlt">3</span>+ ion on the A site.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17249819','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17249819"><span>Self-assembly and hierarchical organization of Ga2<span class="hlt">O</span><span class="hlt">3</span>/In2<span class="hlt">O</span><span class="hlt">3</span> nanostructures.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Liang; Su, Yong; Li, Sen; Chen, Yiqing; Zhou, Qingtao; Yin, Song; Feng, Yi</p> <p>2007-02-01</p> <p>We report on the realization of novel <span class="hlt">3</span>-D hierarchical heterostructures with 6-and 4-fold symmetries by a transport and condensation technique. It was found that the major core nanowires or nanobelts are single-crystalline In2<span class="hlt">O</span><span class="hlt">3</span>, and the secondary nanorods are single-crystalline monoclinic beta-Ga2<span class="hlt">O</span><span class="hlt">3</span> and grow either perpendicular on or slanted to all the facets of the core In2<span class="hlt">O</span><span class="hlt">3</span> nanobelts. Depending on the diameter of the core In2<span class="hlt">O</span><span class="hlt">3</span> nanostructures, the secondary Ga2<span class="hlt">O</span><span class="hlt">3</span> nanorods grow either as a single row or multiple rows. The one-step growth of the unique Ga2<span class="hlt">O</span><span class="hlt">3</span>/In2<span class="hlt">O</span><span class="hlt">3</span> heteronanostructures is a spontaneous and self-organized process. The simultaneous control of nanocrystal size and shape together with the possibility of growing heterostructures on certain nanocrystal facets opens up novel routes to the synthesis of more sophisticated heterostructures as building blocks for opto- and nanoelectronics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JEMat..44.4846Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JEMat..44.4846Y"><span>Effects of Pb<span class="hlt">O</span>-B2<span class="hlt">O</span><span class="hlt">3</span> Glass Doping on the Sintering Temperature and Piezoelectric Properties of 0.35Pb (Ni1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-0.65Pb(Zr0.41Ti0.59)<span class="hlt">O</span><span class="hlt">3</span> Ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yi, Jinqiao; Shen, Meng; Liu, Sisi; Jiang, Shenglin</p> <p>2015-12-01</p> <p>0.35Pb(Ni1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-0.65Pb(Zr0.41Ti0.59)<span class="hlt">O</span><span class="hlt">3</span> (PNN-PZT) ceramics doped with 0.5Pb<span class="hlt">O</span>-0.5B2<span class="hlt">O</span><span class="hlt">3</span> glass have been synthesized by the conventional solid-state sintering technique. The effects of 0.5Pb<span class="hlt">O</span>-0.5B2<span class="hlt">O</span><span class="hlt">3</span> glass on the sintering temperature and piezoelectric properties of PNN-PZT ceramics were studied. The results indicated that the sintering temperature of PNN-PZT was significantly reduced due to the incorporation of 0.5Pb<span class="hlt">O</span>-0.5B2<span class="hlt">O</span><span class="hlt">3</span> glass dopant. When the content of 0.5Pb<span class="hlt">O</span>-0.5B2<span class="hlt">O</span><span class="hlt">3</span> glass was 0.5 wt.%, the sintering temperature of PNN-PZT was observed to reduce from above 1200°C to 920°C while the samples maintained high density (7.91 g/cm<span class="hlt">3</span>), excellent piezoelectric constant ( d 33 = 479 pC/N), large electromechanical coupling coefficient ( K p = 0.55), and relatively low electromechanical quality factor ( Q m = 79). Moreover, large dielectric constant ( ɛ 33 T / ɛ 0 = 2904) and low dielectric loss (tan δ = 0.0166) were obtained in this work.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26010564','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26010564"><span>Optical spectroscopy of Sm(<span class="hlt">3</span>+) doped Na2<span class="hlt">O-ZnO</span>-La2<span class="hlt">O</span><span class="hlt">3</span>-Te<span class="hlt">O</span>2 glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sobczyk, Marcin</p> <p>2015-10-05</p> <p>Telluride glasses with the composition xSm2<span class="hlt">O</span><span class="hlt">3</span>-(7-x)La2<span class="hlt">O</span><span class="hlt">3-3</span>Na2<span class="hlt">O</span>-25Zn<span class="hlt">O</span>-65Te<span class="hlt">O</span>2 (where x=0.1, 1, 2, 5 and 7 mol%) were obtained by the melt quenching technique. Electronic absorption and fluorescence spectra as well as fluorescence dynamics of the Sm(<span class="hlt">3</span>+)-doped title glasses are presented and analysed in detail. A Judd-Ofelt intensity analysis of the absorption spectrum at 300 K has been applied for determination of Ωλ parameters (Ω2=<span class="hlt">3</span>.10, Ω4=<span class="hlt">3</span>.80, Ω6=1.61×10(-20) cm(2)) which in turn have been used for calculations of the radiative transition probabilities (AT), the natural (radiative) lifetimes (τR) of the (4)G5/2 level of Sm(<span class="hlt">3</span>+), the fluorescence branching ratios (β) and the emission cross-sections (σem). The τR value of the (4)G5/2 level amount to 1546 μs and is slightly higher than the measured decay time of 1306 μs. With the increasing of Sm2<span class="hlt">O</span><span class="hlt">3</span> concentration from 0.1 to 7.0 mol% the experimental lifetime of the fluorescent level decreases from 1306 to 41 μs. An analysis of the non-radiative decay was based on the cross-relaxation mechanisms. The optical achieved results indicate that the investigated glasses are potentially applicable as an orange and/or red laser host. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10563E..43W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10563E..43W"><span>Proposed concept and preliminary design for the sentinel-5 UVNs spectrometer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Windpassinger, R.; Schubert, J.; Kampf, D.</p> <p>2017-11-01</p> <p>Sentinel-5 is an atmospheric monitoring mission within the European Copernicus programme, formerly GMES (Global Monitoring for Environment and Security). Its main objective is trace-gas and aerosol optical depth measurements for air quality and climate monitoring and forecast with daily global coverage. Constituents of interest are <span class="hlt">O</span><span class="hlt">3</span>, SO2, <span class="hlt">HCHO</span> (formaldehyde), Br<span class="hlt">O</span>, NO2, CHCHO (glyoxal), <span class="hlt">O</span>2, CH4 (methane), and CO. Sentinel-5 will complement the Sentinel-4 GEO data over Europe. Both Sentinel-4 and -5 are intended to start operation in 2020.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4907701','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4907701"><span>Phase Equilibria and Crystal Chemistry in Portions of the System Sr<span class="hlt">O-CaO</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-Cu<span class="hlt">O</span>, Part IV— The System Ca<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-Cu<span class="hlt">O</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Burton, B. P.; Rawn, C. J.; Roth, R. S.; Hwang, N. M.</p> <p>1993-01-01</p> <p>New data are presented on the phase equilibria and crystal chemistry of the binary systems Ca<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span> and Ca<span class="hlt">O-CuO</span> and the ternary Ca<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-Cu<span class="hlt">O</span>. Symmetry data and unit cell dimensions based on single crystal and powder x-ray diffraction measurements are reported for several of the binary Ca<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span> phases, including corrected compositions for Ca4Bi6<span class="hlt">O</span>13 and Ca2Bi2<span class="hlt">O</span>5. The ternary system contains no new ternary phases which can be formed in air at ~700–900 °C. PMID:28053484</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15863070','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15863070"><span>Investigation of thermal stability and spectroscopic properties in Er<span class="hlt">3</span>+/Yb<span class="hlt">3</span>+-codoped Te<span class="hlt">O</span>2-Li2<span class="hlt">O</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Ge<span class="hlt">O</span>2 glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nie, Qiu-Hua; Gao, Yuan; Xu, Tie-Feng; Shen, Xiang</p> <p>2005-06-01</p> <p>The new Er<span class="hlt">3</span>+/Yb<span class="hlt">3</span>+ co-doped 70Te<span class="hlt">O</span>2-5Li2<span class="hlt">O</span>-(25-x)B2<span class="hlt">O</span><span class="hlt">3</span>-xGe<span class="hlt">O</span>2 (x = 0, 5, 10, 15 fand 20 mol.%) glasses were prepared. The thermal stability, absorption spectra, emission spectra and lifetime of the 4I(13/2) level of Er<span class="hlt">3</span>+ ions were measured and studied. The FT-IR spectra were carried out in order to investigate the structure of local arrangements in glasses. It is found that the thermal stability, absorption cross-section of Yb<span class="hlt">3</span>+, emission intensity and lifetime of the 4I(13/2) level of Er<span class="hlt">3</span>+ increase with increasing Ge<span class="hlt">O</span>2 content in the glass composition, while the fluorescence width at half maximum (FWHM) at 1.5 um of Er<span class="hlt">3</span>+ is about 70 nm. The obtained data suggest that this system glass can be used as a candidate host material for potential broadband optical amplifiers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1265482-thermodynamic-properties-fe2o3-fe3o4-nanoparticles','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1265482-thermodynamic-properties-fe2o3-fe3o4-nanoparticles"><span>Thermodynamic Properties of α-Fe 2<span class="hlt">O</span> <span class="hlt">3</span> and Fe <span class="hlt">3</span><span class="hlt">O</span> 4 Nanoparticles</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Spencer, Elinor C.; Ross, Nancy L.; Olsen, Rebecca E.; ...</p> <p>2015-04-21</p> <p>Here we comprehansively assessed the thermodynamic properties of hydrated α-Fe 2<span class="hlt">O</span> <span class="hlt">3</span> (hematite) and Fe <span class="hlt">3</span><span class="hlt">O</span> 4 (magnetite) nanoparticles. In addition to 9 nm Fe <span class="hlt">3</span><span class="hlt">O</span> 4, three α-e 2<span class="hlt">O</span> <span class="hlt">3</span>nanoparticles samples of different sizes (11, 14, and 25 nm) and bulk α-e 2<span class="hlt">O</span> <span class="hlt">3</span> have been evaluated by inelastic neutron scattering methods. The contribution of the two-level magnetic spin flip transition to the heat capacity of the α-e 2<span class="hlt">O</span> <span class="hlt">3</span> particles has been determined. The isochoric heat capacity of the water confined on the surface of these two types of iron oxide particles have been calculated from their INSmore » spectra, and is affected by the chemical composition of the underlying particle. Furthermore, the heat capacity and dynamics of the particle hydration layers appear to be influenced by a complex array of factors including particle size, water coverage, and possibly the magnetic state of the particle itself.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27840059','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27840059"><span>Preparation of Er<span class="hlt">3</span>+:Y<span class="hlt">3</span>Al5<span class="hlt">O</span>12/WO<span class="hlt">3</span>-KNb<span class="hlt">O</span><span class="hlt">3</span> composite and application in treatment of methamphetamine under ultrasonic irradiation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Hongbo; Huang, Yingying; Li, Guanshu; Wang, Guowei; Fang, Dawei; Song, Youtao; Wang, Jun</p> <p>2017-03-01</p> <p>Er <span class="hlt">3</span>+ :Y <span class="hlt">3</span> Al 5 <span class="hlt">O</span> 12 /WO <span class="hlt">3</span> -KNb<span class="hlt">O</span> <span class="hlt">3</span> composite powder as an effective sonocatalyst was prepared via collosol-gelling-hydrothermal and high-temperature calcination methods. The textures of materials were observed by X-ray diffractometer (XRD), scanning electron microscopy (SEM) and X-ray photoelectron spectroscopy (XPS). In order to estimate the sonocatalytic activity of Er <span class="hlt">3</span>+ :Y <span class="hlt">3</span> Al 5 <span class="hlt">O</span> 12 /WO <span class="hlt">3</span> -KNb<span class="hlt">O</span> <span class="hlt">3</span> composite powder, the sonocatalytic degradation of methamphetamine (MAPA) was performed. Furthermore, the influences of mass ratio of WO <span class="hlt">3</span> and KNb<span class="hlt">O</span> <span class="hlt">3</span> , ultrasonic irradiation time, catalyst addition amount, initial methamphetamine (MAPA) concentration and used times on the sonocatalytic degradation of methamphetamine (MAPA) caused by Er <span class="hlt">3</span>+ :Y <span class="hlt">3</span> Al 5 <span class="hlt">O</span> 12 /WO <span class="hlt">3</span> -KNb<span class="hlt">O</span> <span class="hlt">3</span> composite powder were investigated by using gas chromatography. Under optimal conditions of 1.00g/L Er <span class="hlt">3</span>+ :Y <span class="hlt">3</span> Al 5 <span class="hlt">O</span> 12 /WO <span class="hlt">3</span> -KNb<span class="hlt">O</span> <span class="hlt">3</span> addition amount and 10.00mg/L methamphetamine (MAPA) initial concentration, 68% of methamphetamine (MAPA) could be removed after 150min ultrasonic irradiation. The experimental results showed that the Er <span class="hlt">3</span>+ :Y <span class="hlt">3</span> Al 5 <span class="hlt">O</span> 12 /WO <span class="hlt">3</span> -KNb<span class="hlt">O</span> <span class="hlt">3</span> as sonocatalyst displayed an excellent sonocatalytic activity in degradation of methamphetamine (MAPA) under ultrasonic irradiation. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21809872','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21809872"><span>Measurements of nitrous acid in commercial aircraft exhaust at the Alternative Aviation Fuel Experiment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lee, Ben H; Santoni, Gregory W; Wood, Ezra C; Herndon, Scott C; Miake-Lye, Richard C; Zahniser, Mark S; Wofsy, Steven C; Munger, J William</p> <p>2011-09-15</p> <p>The Alternative Aviation Fuel Experiment (AAFEX), conducted in January of 2009 in Palmdale, California, quantified aerosol and gaseous emissions from a DC-8 aircraft equipped with CFM56-2C1 engines using both traditional and synthetic fuels. This study examines the emissions of nitrous acid (<span class="hlt">HONO</span>) and nitrogen oxides (NO(x) = NO + NO(2)) measured 145 m behind the grounded aircraft. The fuel-based emission index (EI) for <span class="hlt">HONO</span> increases approximately 6-fold from idle to takeoff conditions but plateaus between 65 and 100% of maximum rated engine thrust, while the EI for NO(x) increases continuously. At high engine power, NO(x) EI is greater when combusting traditional (JP-8) rather than Fischer-Tropsch fuels, while <span class="hlt">HONO</span> exhibits the opposite trend. Additionally, hydrogen peroxide (H(2)<span class="hlt">O</span>(2)) was identified in exhaust plumes emitted only during engine idle. Chemical reactions responsible for emissions and comparison to previous measurement studies are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22814313','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22814313"><span>Porous nanocubic Mn<span class="hlt">3</span><span class="hlt">O</span>4-Co<span class="hlt">3</span><span class="hlt">O</span>4 composites and their application as electrochemical supercapacitors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pang, Huan; Deng, Jiawei; Du, Jimin; Li, Sujuan; Li, Juan; Ma, Yahui; Zhang, Jiangshan; Chen, Jing</p> <p>2012-09-14</p> <p>A simple approach has been developed to fabricate ideal supercapacitors based on porous Mn(<span class="hlt">3</span>)<span class="hlt">O</span>(4)-Co(<span class="hlt">3</span>)<span class="hlt">O</span>(4) nanocubic composite electrodes. We can easily obtain porous corner-truncated nanocubic Mn(<span class="hlt">3</span>)<span class="hlt">O</span>(4)-Co(<span class="hlt">3</span>)<span class="hlt">O</span>(4) composite nanomaterials without any subsequent complicated workup procedure for the removal of a hard template, seed or by using a soft template. In such a composite, the porous Mn(<span class="hlt">3</span>)<span class="hlt">O</span>(4)-Co(<span class="hlt">3</span>)<span class="hlt">O</span>(4) enables a fast and reversible redox reaction to improve the specific capacitance. The porous nanocubic Mn(<span class="hlt">3</span>)<span class="hlt">O</span>(4)-Co(<span class="hlt">3</span>)<span class="hlt">O</span>(4) composite electrode can effectively transport electrolytes and shorten the ion diffusion path, which offers excellent electrochemical performance. These results suggest that such porous Mn(<span class="hlt">3</span>)<span class="hlt">O</span>(4)-Co(<span class="hlt">3</span>)<span class="hlt">O</span>(4) composite nanocubes are very promising for next generation high-performance supercapacitors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018APExp..11f5501O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018APExp..11f5501O"><span>α-Al2<span class="hlt">O</span><span class="hlt">3</span>/Ga2<span class="hlt">O</span><span class="hlt">3</span> superlattices coherently grown on r-plane sapphire</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Oshima, Takayoshi; Kato, Yuji; Imura, Masataka; Nakayama, Yoshiko; Takeguchi, Masaki</p> <p>2018-06-01</p> <p>Ten-period binary α-Al2<span class="hlt">O</span><span class="hlt">3</span>/Ga2<span class="hlt">O</span><span class="hlt">3</span> superlattices were fabricated on r-plane sapphire substrates by molecular beam epitaxy. By systematic variation of α-Ga2<span class="hlt">O</span><span class="hlt">3</span> thickness and evaluation through X-ray reflectivity and diffraction measurements and scanning transmission electron microscopy, we verified that the superlattice with α-Ga2<span class="hlt">O</span><span class="hlt">3</span> thickness up to ∼1 nm had coherent interfaces without misfit dislocation in spite of the large lattice mismatches. This successful fabrication of coherent α-Al2<span class="hlt">O</span><span class="hlt">3</span>/Ga2<span class="hlt">O</span><span class="hlt">3</span> superlattices will encourage further development of α-(Al x Ga1‑ x )2<span class="hlt">O</span><span class="hlt">3</span>-based heterostructures including superlattices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1347292-orbital-configuration-catio3-films-ndgao3','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1347292-orbital-configuration-catio3-films-ndgao3"><span>Orbital configuration in CaTi<span class="hlt">O</span> <span class="hlt">3</span> films on NdGa<span class="hlt">O</span> <span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Cao, Yanwei; Park, Se Young; Liu, Xiaoran; ...</p> <p>2016-10-13</p> <p>Despite its use as a constituent layer for realization of a polar metal and interfacial conductivity, the microscopic study of electronic structure of CaTi<span class="hlt">O</span> <span class="hlt">3</span> is still very limited. Here, we epitaxially stabilized CaTi<span class="hlt">O</span> <span class="hlt">3</span> films on NdGa<span class="hlt">O</span> <span class="hlt">3</span> (110) substrates in a layer-by-layer way by pulsed laser deposition. The structural and electronic properties of the films were characterized by reflection-high-energy-electron-diffraction, X-ray diffraction, and element-specific resonant X-ray absorption spectroscopy. To reveal the orbital polarization and the crystal field splitting of the titanium <span class="hlt">3</span>d state, X-ray linear dichroism was carried out on CaTi<span class="hlt">O</span> <span class="hlt">3</span> films, demonstrating the orbital configuration of dmore » xz/d yz < d xy < d <span class="hlt">3</span>z2-r2 < d x2-y2. To further explore the origin of this configuration, we performed the first-principles density function theory calculations, which linked the orbital occupation to the on-site energy of Ti <span class="hlt">3</span>d orbitals. Finally, these findings can be important for understanding and designing exotic quantum states in heterostructures based on CaTi<span class="hlt">O</span> <span class="hlt">3</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010LanB..48B.1172C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010LanB..48B.1172C"><span>Nqrs Data for H6I<span class="hlt">3</span>In<span class="hlt">O</span>12 [I<span class="hlt">3</span>In<span class="hlt">O</span>9·<span class="hlt">3</span>(H2<span class="hlt">O</span>)] (Subst. No. 2289)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chihara, H.; Nakamura, N.</p> <p></p> <p>This document is part of Subvolume B 'Substances Containing C10H16 … Zn' of Volume 48 'Nuclear Quadrupole Resonance Spectroscopy Data' of Landolt-Börnstein - Group III 'Condensed Matter'. It contains an extract of Section '<span class="hlt">3</span>.2 Data tables' of the Chapter '<span class="hlt">3</span> Nuclear quadrupole resonance data' providing the NQRS data for H6I<span class="hlt">3</span>In<span class="hlt">O</span>12 [I<span class="hlt">3</span>In<span class="hlt">O</span>9·<span class="hlt">3</span>(H2<span class="hlt">O</span>)] (Subst. No. 2289)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006SPIE.6101..531A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006SPIE.6101..531A"><span>Quenching of I(2P 1/2) by <span class="hlt">O</span> <span class="hlt">3</span> and <span class="hlt">O</span>( <span class="hlt">3</span>P)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Azyazov, V. N.; Antonov, I. O.; Ruffner, S.; Heaven, M. C.</p> <p>2006-02-01</p> <p>Oxygen-iodine lasers that utilize electrical or microwave discharges to produce singlet oxygen are currently being developed. The discharge generators differ from conventional chemical singlet oxygen generators in that they produce significant amounts of atomic oxygen. Post-discharge chemistry includes channels that lead to the formation of ozone. Consequently, removal of I(2P 1/2) by <span class="hlt">O</span> atoms and <span class="hlt">O</span> <span class="hlt">3</span> may impact the efficiency of discharge driven iodine lasers. In the present study we have measured the rate constants for quenching of I(2P 1/2) by <span class="hlt">O</span>( <span class="hlt">3</span>P) atoms and <span class="hlt">O</span> <span class="hlt">3</span> using pulsed laser photolysis techniques. The rate constant for quenching by <span class="hlt">O</span> <span class="hlt">3</span>, 1.8x10 -12 cm <span class="hlt">3</span> s -1, was found to be a factor of five smaller than the literature value. The rate constant for quenching by <span class="hlt">O</span>( <span class="hlt">3</span>P) was 1.2x10 -11 cm <span class="hlt">3</span> s -1. This was six times larger than a previously reported upper bound, but consistent with estimates obtained by modeling the kinetics of discharge-driven laser systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhD...51k4004A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhD...51k4004A"><span>Enhanced magnetoelectric response in 2-2 bilayer 0.50Pb(Ni1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-0.35PbTi<span class="hlt">O</span><span class="hlt">3</span>-0.15PbZr<span class="hlt">O</span><span class="hlt">3</span>/NiFe2<span class="hlt">O</span>4 thin films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ade, Ramesh; Sambasiva, V.; Kolte, Jayant; Karthik, T.; Kulkarni, Ajit R.; Venkataramani, N.</p> <p>2018-03-01</p> <p>In this work, room temperature magnetoelectric (ME) properties of 0.50Pb(Ni1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>-0.35PbTi<span class="hlt">O</span><span class="hlt">3</span>-0.15PbZr<span class="hlt">O</span><span class="hlt">3</span> (PNNZT)/NiFe2<span class="hlt">O</span>4 (NFO) 2-2 bilayer thin films grown on Pt/Ti/Si<span class="hlt">O</span>2/Si substrate, using pulsed laser deposition technique, are reported. Structural studies confirm single phase PNNZT/NFO 2-2 bilayer structure formation. PNNZT/NFO 2-2 bilayer thin film shows a maximum ME voltage coefficient (α E ) of ~0.70 V cm-1. Oe-1 at a frequency of 1 kHz. The present study reveals that PNNZT/NFO bilayer thin film can be a potential candidate for technological applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JTST...26.1076C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JTST...26.1076C"><span>Air Plasma-Sprayed La2Zr2<span class="hlt">O</span>7-SrZr<span class="hlt">O</span><span class="hlt">3</span> Composite Thermal Barrier Coating Subjected to Ca<span class="hlt">O-MgO</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 (CMAS)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cai, Lili; Ma, Wen; Ma, Bole; Guo, Feng; Chen, Weidong; Dong, Hongying; Shuang, Yingchai</p> <p>2017-08-01</p> <p>La2Zr2<span class="hlt">O</span>7-SrZr<span class="hlt">O</span><span class="hlt">3</span> composite thermal barrier coatings (TBCs) were prepared by air plasma spray (APS). The La2Zr2<span class="hlt">O</span>7-SrZr<span class="hlt">O</span><span class="hlt">3</span> composite TBCs covered with calcium-magnesium-aluminum-silicate (CMAS) powder, as well as the powder mixture of CMAS and spray-dried La2Zr2<span class="hlt">O</span>7-SrZr<span class="hlt">O</span><span class="hlt">3</span> composite powder, were heat-treated at 1250 °C in air for 1, 4, 8, and 12 h. The phase constituents and microstructures of the reaction products were characterized by x-ray diffraction, scanning electron microscopy, and energy-dispersive spectroscopy. Experimental results showed that the La2Zr2<span class="hlt">O</span>7-SrZr<span class="hlt">O</span><span class="hlt">3</span> composite TBCs had higher CMAS resistance than 8YSZ coating. A dense new layer developed between CMAS and La2Zr2<span class="hlt">O</span>7-SrZr<span class="hlt">O</span><span class="hlt">3</span> composite TBCs during interaction, and this new layer consisted mostly of apatite (Ca2La8(Si<span class="hlt">O</span>4)6<span class="hlt">O</span>2) and c-Zr<span class="hlt">O</span>2. The newly developed layer effectively protected the La2Zr2<span class="hlt">O</span>7-SrZr<span class="hlt">O</span><span class="hlt">3</span> composite TBCs from further CMAS attack.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5397043-removal-atmospheric-oxidants-annular-denuders','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5397043-removal-atmospheric-oxidants-annular-denuders"><span>Removal of atmospheric oxidants with annular denuders</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Williams, E.L. II; Grosjean, D.</p> <p>1990-06-01</p> <p>Annular denuders have been tested for their ability to remove atmospheric photochemical oxidants including ozone (150-170 ppb), nitrogen dioxide (40-200 ppb), nitric acid (35 ppb), and peroxyacetyl nitrate (PAN, 6-25 ppb). Formaldehyde (80-140 ppb) was also tested as a copollutant. Tests were carried out at low rates of 1,2, and 20 L/min, with oxidants tested singly or as part of photochemical mixtures in purified air. Efficient collection was obtained with annular denuders coated with potassium iodide (<span class="hlt">O</span>{sub <span class="hlt">3</span>}), phenoxamine (<span class="hlt">O</span>{sub <span class="hlt">3</span>}), sodium carbonate (HNO{sub <span class="hlt">3</span>}), potassium hydroxide (PAN), and 2,4-dinitrophenylhydrazine (<span class="hlt">HCHO</span>).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28387121','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28387121"><span>Al2<span class="hlt">O</span><span class="hlt">3</span> Passivation Effect in Hf<span class="hlt">O</span>2·Al2<span class="hlt">O</span><span class="hlt">3</span> Laminate Structures Grown on InP Substrates.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kang, Hang-Kyu; Kang, Yu-Seon; Kim, Dae-Kyoung; Baik, Min; Song, Jin-Dong; An, Youngseo; Kim, Hyoungsub; Cho, Mann-Ho</p> <p>2017-05-24</p> <p>The passivation effect of an Al 2 <span class="hlt">O</span> <span class="hlt">3</span> layer on the electrical properties was investigated in Hf<span class="hlt">O</span> 2 -Al 2 <span class="hlt">O</span> <span class="hlt">3</span> laminate structures grown on indium phosphide (InP) substrate by atomic-layer deposition. The chemical state obtained using high-resolution X-ray photoelectron spectroscopy showed that interfacial reactions were dependent on the presence of the Al 2 <span class="hlt">O</span> <span class="hlt">3</span> passivation layer and its sequence in the Hf<span class="hlt">O</span> 2 -Al 2 <span class="hlt">O</span> <span class="hlt">3</span> laminate structures. Because of the interfacial reaction, the Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Hf<span class="hlt">O</span> 2 /Al 2 <span class="hlt">O</span> <span class="hlt">3</span> structure showed the best electrical characteristics. The top Al 2 <span class="hlt">O</span> <span class="hlt">3</span> layer suppressed the interdiffusion of oxidizing species into the Hf<span class="hlt">O</span> 2 films, whereas the bottom Al 2 <span class="hlt">O</span> <span class="hlt">3</span> layer blocked the outdiffusion of In and P atoms. As a result, the formation of In-<span class="hlt">O</span> bonds was more effectively suppressed in the Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /Hf<span class="hlt">O</span> 2 /Al 2 <span class="hlt">O</span> <span class="hlt">3</span> /InP structure than that in the Hf<span class="hlt">O</span> 2 -on-InP system. Moreover, conductance data revealed that the Al 2 <span class="hlt">O</span> <span class="hlt">3</span> layer on InP reduces the midgap traps to 2.6 × 10 12 eV -1 cm -2 (compared to that of Hf<span class="hlt">O</span> 2 /InP, that is, 5.4 × 10 12 eV -1 cm -2 ). The suppression of gap states caused by the outdiffusion of In atoms significantly controls the degradation of capacitors caused by leakage current through the stacked oxide layers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28256752','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28256752"><span>Interface Control of Ferroelectricity in an SrRu<span class="hlt">O</span><span class="hlt">3</span> /BaTi<span class="hlt">O</span><span class="hlt">3</span> /SrRu<span class="hlt">O</span><span class="hlt">3</span> Capacitor and its Critical Thickness.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shin, Yeong Jae; Kim, Yoonkoo; Kang, Sung-Jin; Nahm, Ho-Hyun; Murugavel, Pattukkannu; Kim, Jeong Rae; Cho, Myung Rae; Wang, Lingfei; Yang, Sang Mo; Yoon, Jong-Gul; Chung, Jin-Seok; Kim, Miyoung; Zhou, Hua; Chang, Seo Hyoung; Noh, Tae Won</p> <p>2017-05-01</p> <p>The atomic-scale synthesis of artificial oxide heterostructures offers new opportunities to create novel states that do not occur in nature. The main challenge related to synthesizing these structures is obtaining atomically sharp interfaces with designed termination sequences. In this study, it is demonstrated that the oxygen pressure (PO2) during growth plays an important role in controlling the interfacial terminations of SrRu<span class="hlt">O</span> <span class="hlt">3</span> /BaTi<span class="hlt">O</span> <span class="hlt">3</span> /SrRu<span class="hlt">O</span> <span class="hlt">3</span> (SRO/BTO/SRO) ferroelectric (FE) capacitors. The SRO/BTO/SRO heterostructures are grown by a pulsed laser deposition method. The top SRO/BTO interface, grown at high PO2 (around 150 mTorr), usually exhibits a mixture of Ru<span class="hlt">O</span> 2 -Ba<span class="hlt">O</span> and Sr<span class="hlt">O-TiO</span> 2 terminations. By reducing PO2, the authors obtain atomically sharp SRO/BTO top interfaces with uniform Sr<span class="hlt">O-TiO</span> 2 termination. Using capacitor devices with symmetric and uniform interfacial termination, it is demonstrated for the first time that the FE critical thickness can reach the theoretical limit of <span class="hlt">3</span>.5 unit cells. © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..MAR.M1004W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..MAR.M1004W"><span>Magnetic Orders of LaTi<span class="hlt">O</span><span class="hlt">3</span> and YTi<span class="hlt">O</span><span class="hlt">3</span> Under Epitaxial Strain: a First-Principles study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Weng, Yakui; Huang, Xin; Tang, Yankun; Dong, Shuai</p> <p>2014-03-01</p> <p>Perovskite RTi<span class="hlt">O</span><span class="hlt">3</span> family is a typical Mott-insulator with localized <span class="hlt">3</span>d electrons. In this work, the epitaxial strain effects on the ground magnetic orders of LaTi<span class="hlt">O</span><span class="hlt">3</span> and YTi<span class="hlt">O</span><span class="hlt">3</span> films have been studied using the first-principles density-functional theory. For the YTi<span class="hlt">O</span><span class="hlt">3</span> films, A-type antiferromagnetic order emerges against the original ferromagnetic order under the in-plane compressive strain by LaAl<span class="hlt">O</span><span class="hlt">3</span> (001) substrate, although the A-type antiferromagnetic order does not exist in any RTi<span class="hlt">O</span><span class="hlt">3</span> bulks. Then, for the LaTi<span class="hlt">O</span><span class="hlt">3</span> films under the compressive strain, e.g. LaTi<span class="hlt">O</span><span class="hlt">3</span> films grown on LaAl<span class="hlt">O</span><span class="hlt">3</span>, LaGa<span class="hlt">O</span><span class="hlt">3</span>, and SrTi<span class="hlt">O</span><span class="hlt">3</span> substrates, undergo a phase transition from the original G-type antiferromagnetism to A-type antiferromagnetism. While under the tensile strain, e.g. grown on the BaTi<span class="hlt">O</span><span class="hlt">3</span> and LaSc<span class="hlt">O</span><span class="hlt">3</span> substrate, LaTi<span class="hlt">O</span><span class="hlt">3</span> films show a tendency to transit to the C-type antiferromagnetism. Furthermore, our calculations find that the magnetic transitions under epitaxial strain do not change the insulating fact of LaTi<span class="hlt">O</span><span class="hlt">3</span> and YTi<span class="hlt">O</span><span class="hlt">3</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JAP...114r3909H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JAP...114r3909H"><span>Voltage-controlled ferromagnetism and magnetoresistance in LaCo<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hu, Chengqing; Park, Keun Woo; Posadas, Agham; Jordan-Sweet, Jean L.; Demkov, Alexander A.; Yu, Edward T.</p> <p>2013-11-01</p> <p>A LaCo<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> heterostructure grown on Si (001) is shown to provide electrically switchable ferromagnetism, a large, electrically tunable magnetoresistance, and a vehicle for achieving and probing electrical control over ferromagnetic behavior at submicron dimensions. Fabrication of devices in a field-effect transistor geometry enables application of a gate bias voltage that modulates strain in the heterostructure via the converse piezoelectric effect in SrTi<span class="hlt">O</span><span class="hlt">3</span>, leading to an artificial inverse magnetoelectric effect arising from the dependence of ferromagnetism in the LaCo<span class="hlt">O</span><span class="hlt">3</span> layer on strain. Below the Curie temperature of the LaCo<span class="hlt">O</span><span class="hlt">3</span> layer, this effect leads to modulation of resistance in LaCo<span class="hlt">O</span><span class="hlt">3</span> as large as 100%, and magnetoresistance as high as 80%, both of which arise from carrier scattering at ferromagnetic-nonmagnetic interfaces in LaCo<span class="hlt">O</span><span class="hlt">3</span>. Finite-element numerical modeling of electric field distributions is used to explain the dependence of carrier transport behavior on gate contact geometry, and a Valet-Fert transport model enables determination of spin polarization in the LaCo<span class="hlt">O</span><span class="hlt">3</span> layer. Piezoresponse force microscopy is used to confirm the existence of piezoelectric response in SrTi<span class="hlt">O</span><span class="hlt">3</span> grown on Si (001). It is also shown that this structure offers the possibility of achieving exclusive-NOR logic functionality within a single device.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JPhD...45R5305Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JPhD...45R5305Y"><span>Positron annihilation studies on the behaviour of vacancies in LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yuan, Guoliang; Li, Chen; Yin, Jiang; Liu, Zhiguo; Wu, Di; Uedono, Akira</p> <p>2012-11-01</p> <p>The formation and diffusion of vacancies are studied in LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> heterostructures. Oxygen vacancies (VOS) appear easily in the SrTi<span class="hlt">O</span><span class="hlt">3</span> substrate during LaAl<span class="hlt">O</span><span class="hlt">3</span> film growth at 700 °C and 10-4 Pa oxygen pressure rather than at 10-<span class="hlt">3</span>-10-1 Pa, thus the latter two-dimensional electron gas should come from the polarity discontinuity at the (La<span class="hlt">O)+/(TiO</span>2)0 interface. For SrTi<span class="hlt">O</span><span class="hlt">3</span>-δ/LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span>, high-density VOS of the SrTi<span class="hlt">O</span><span class="hlt">3</span>-δ film can pass through the LaAl<span class="hlt">O</span><span class="hlt">3</span> film and then diffuse to 1.7 µm depth in the SrTi<span class="hlt">O</span><span class="hlt">3</span> substrate, suggesting that LaAl<span class="hlt">O</span><span class="hlt">3</span> has VOS at its middle-deep energy levels within the band gap. Moreover, high-density VOS may combine with a strontium/titanium vacancy (VSr/Ti) to form VSr/Ti-<span class="hlt">O</span> complexes in the SrTi<span class="hlt">O</span><span class="hlt">3</span> substrate at 700 °C.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPA....8e5820M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPA....8e5820M"><span>Magneto-optical properties of BaTi<span class="hlt">O</span><span class="hlt">3</span>/La0.76Sr0.24Mn<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Moog, M.; Singamaneni, S. R.; Prater, J. T.; Biegalski, M. D.; Tsui, F.</p> <p>2018-05-01</p> <p>The magnetic properties of epitaxial BaTi<span class="hlt">O</span><span class="hlt">3</span>/La0.76Sr0.24Mn<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span> (BTO/LSMO/BTO) heterostructures have been studied using magneto-optic Kerr effect (MOKE) technique. Both longitudinal and polar MOKE were probed as a function of magnetic field and temperature (in the range between 80 and 320 K) for epitaxial films of BTO/LSMO/BTO and LSMO grown on Ti<span class="hlt">O</span>2-terminated SrTi<span class="hlt">O</span><span class="hlt">3</span> (001) substrates by pulsed laser deposition technique. The LSMO film without the BTO layers exhibits nearly square field-dependent MOKE hysteresis loops with low saturation fields below a bulk-like Curie temperature (TC) of ˜ 350K. In contrast, the film with the BTO layers exhibits a significantly suppressed TC of 155 K, accompanied by significantly enhanced coercive fields and perpendicular magnetic anisotropy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MRE.....5a5913U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MRE.....5a5913U"><span>Fe doped BaTi<span class="hlt">O</span><span class="hlt">3</span> sensitized by Fe<span class="hlt">3</span><span class="hlt">O</span>4 nanoparticles for improved photoelectrochemical response</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Upadhyay, Rishibrind Kumar; Sharma, Dipika</p> <p>2018-01-01</p> <p>Nanostructured powders of pristine Fe<span class="hlt">3</span><span class="hlt">O</span>4, BaTi<span class="hlt">O</span><span class="hlt">3</span>, and Fe-BaTi<span class="hlt">O</span><span class="hlt">3</span> were synthesized using hydrothermal method and BaTi<span class="hlt">O</span><span class="hlt">3</span>/Fe<span class="hlt">3</span><span class="hlt">O</span>4 and Fe-BaTi<span class="hlt">O</span><span class="hlt">3</span>/Fe<span class="hlt">3</span><span class="hlt">O</span>4 composite sample were also prepared by mixing the appropriate amount of pristine powders. All samples were characterized using x-ray diffraction, SEM and UV-vis spectrometry. Photoelectrochemical properties were investigated in a three-electrode cell system. Maximum photocurrent density of 2.1 mA cm-2 at 0.95 V/SCE was observed for Fe-BaTi<span class="hlt">O</span><span class="hlt">3</span>/Fe<span class="hlt">3</span><span class="hlt">O</span>4 composite sample. Increased photocurrent density offered by composite may be attributed to improved conductivity and better separation of the photogenerated charge carriers at interface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018yCat..74542292O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018yCat..74542292O"><span>VizieR Online Data Catalog: H<span class="hlt">3</span><span class="hlt">O</span>+ and D<span class="hlt">3</span><span class="hlt">O</span>+ rota</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Owens, A.; Yurchenko, S. N.; Polyansky, O. L.; Ovsyannikov, R. I.; Thiel, W.; Spirko, V.</p> <p>2018-01-01</p> <p>Given the astronomical relevance of H<span class="hlt">3</span><span class="hlt">O</span>+, and a good set of accurately measured experimental data (Uy, White & Oka 1997JMoSp.183..240U; Araki, Ozeki & Saito 1999, Mol. Phys., 97, 177); Tang & Oka 1999JMoSp.196..120T ; Furuya & Saito 2005A&A...441.1039F; Yu et al. 2009ApJS..180..119Y; Yu & Pearson 2014ApJ...786..133Y), we find it worthwhile to carry out a comprehensive study of hydronium, H316<span class="hlt">O</span>+ (also referred to as H<span class="hlt">3</span><span class="hlt">O</span>+), and its two symmetric top isotopologues, H318<span class="hlt">O</span>+ and D316<span class="hlt">O</span>+. To do this we employ a highly accurate variational approach, which was recently applied to ammonia (Owens et al. 2015MNRAS.450.3191<span class="hlt">O</span>). Like NH<span class="hlt">3</span> (Jansen, Bethlem & Ubachs 2014JChPh.140a0901J; Spirko 2014, J. Phys. Chem. Lett., 5, 919; Owens et al. 2015MNRAS.450.3191<span class="hlt">O</span>), there is a possibility to find transitions with strongly anomalous sensitivities caused by the Δk=+/-<span class="hlt">3</span> interactions (see Papousek et al. 1986JMoSt.141..361P), which have not yet been considered. (11 data files).</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900034120&hterms=Glasses+SiO2&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DGlasses%2BSiO2','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900034120&hterms=Glasses+SiO2&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DGlasses%2BSiO2"><span>Crystallization kinetics of Ba<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 glasses</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bansal, Narottam P.; Hyatt, Mark J.</p> <p>1989-01-01</p> <p>Barium aluminosilicate glasses are being investigated as matrix materials in high-temperature ceramic composites for structural applications. Kinetics of crystallization of two refractory glass compositions in the barium aluminosilicate system were studied by differential thermal analysis (DTA), X-ray diffraction (XRD), and scanning electron microscopy (SEM). From variable heating rate DTA, the crystallization activation energies for glass compositions (wt percent) 10Ba<span class="hlt">O</span>-38Al2<span class="hlt">O</span><span class="hlt">3</span>-51Si<span class="hlt">O</span>2-1Mo<span class="hlt">O</span><span class="hlt">3</span> (glass A) and 39Ba<span class="hlt">O</span>-25Al2<span class="hlt">O</span><span class="hlt">3</span>-35Si<span class="hlt">O</span>2-1Mo<span class="hlt">O</span><span class="hlt">3</span> (glass B) were determined to be 553 and 558 kJ/mol, respectively. On thermal treatment, the crystalline phases in glasses A and B were identified as mullite (<span class="hlt">3</span>Al2<span class="hlt">O</span><span class="hlt">3</span>-2Si<span class="hlt">O</span>2) and hexacelsian (Ba<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-2Si<span class="hlt">O</span>2), respectively. Hexacelsian is a high-temperature polymorph which is metastable below 1590 C. It undergoes structural transformation into the orthorhombic form at approximately 300 C accompanied by a large volume change which is undesirable for structural applications. A process needs to be developed where stable monoclinic celsian, rather than hexacelsian, precipitates out as the crystal phase in glass B.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890010881','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890010881"><span>Crystallization kinetics of Ba<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 glasses</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bansal, Narottam P.; Hyatt, Mark J.</p> <p>1988-01-01</p> <p>Barium aluminosilicate glasses are being investigated as matrix materials in high-temperature ceramic composites for structural applications. Kinetics of crystallization of two refractory glass compositions in the barium aluminosilicate system were studied by differential thermal analysis (DTA), X-ray diffraction (XRD), and scanning electron microscopy (SEM). From variable heating rate DTA, the crystallization activation energies for glass compositions (wt percent) 10Ba<span class="hlt">O</span>-38Al2<span class="hlt">O</span><span class="hlt">3</span>-51Si<span class="hlt">O</span>2-1Mo<span class="hlt">O</span><span class="hlt">3</span> (glass A) and 39Ba<span class="hlt">O</span>-25Al2<span class="hlt">O</span><span class="hlt">3</span>-35Si<span class="hlt">O</span>2-1Mo<span class="hlt">O</span><span class="hlt">3</span> (glass B) were determined to be 553 and 558 kJ/mol, respectively. On thermal treatment, the crystalline phases in glasses A and B were identified as mullite (<span class="hlt">3</span>Al2<span class="hlt">O</span><span class="hlt">3</span>-2Si<span class="hlt">O</span>2) and hexacelsian (Ba<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-2Si<span class="hlt">O</span>2), respectively. Hexacelsian is a high-temperature polymorph which is metastable below 1590 C. It undergoes structural transformation into the orthorhombic form at approximately 300 C accompanied by a large volume change which is undesirable for structural applications. A process needs to be developed where stable monoclinic celsian, rather than hexacelsian, precipitates out as the crystal phase in glass B.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018OptMa..80..216C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018OptMa..80..216C"><span>Structure, spectra and thermal, mechanical, Faraday rotation properties of novel diamagnetic Se<span class="hlt">O</span>2-Pb<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span> glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Qiuling; Su, Kai; Li, Yantao; Zhao, Zhiwei</p> <p>2018-06-01</p> <p>Faraday rotation diamagnetic glass has attracted research attentions in photonics, sensing and magneto optical devices due to their high refractive index, wide transmittance in UV and Fourier transform infrared (FT-IR) range and temperature independent Faraday rotation. Selenite modified heavy metal oxides glasses with composition of xSe<span class="hlt">O</span>2-(10-x) B2<span class="hlt">O</span><span class="hlt">3</span>-45Pb<span class="hlt">O</span>-45Bi2<span class="hlt">O</span><span class="hlt">3</span> (x = 0, 1, 5 and 10mol%) and 15%Se<span class="hlt">O</span>2-40%Pb<span class="hlt">O</span>-45%Bi2<span class="hlt">O</span><span class="hlt">3</span> have been fabricated by melt-quenching method in present study. The influence of Se<span class="hlt">O</span>2 on glass forming ability, thermal, mechanical properties and Faraday rotation were evaluated through X-ray Diffraction (XRD), Fourier transforms infrared spectra (FT-IR), Raman, X-ray photoelectron spectroscopy (XPS), differential scanning calorimetry (DSC), Vicker's hardness and Verdet constant measurements. XRD spectra reveal that the good vitrification was achieved for glass with Se<span class="hlt">O</span>2 amounts ≤10% even without B2<span class="hlt">O</span><span class="hlt">3</span>. FT-IR, Raman and XPS spectra ascertain the existence of characteristic vibration of Se<span class="hlt">O</span>4, Se<span class="hlt">O</span><span class="hlt">3</span>, Pb<span class="hlt">O</span>4, Bi<span class="hlt">O</span><span class="hlt">3</span> and BO<span class="hlt">3</span> units. The incorporation of Se<span class="hlt">O</span>2 increases the connectivity of glassy network by increasing the Tg, thermal stability and mechanical hardness. The small band gap, high polarizable Se4+ ions and isolated Se<span class="hlt">O</span><span class="hlt">3</span> units contribute to Faraday rotation improvement.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApPhL.112x2101P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApPhL.112x2101P"><span>Structural and electronic properties of Ga2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> alloys</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Peelaers, Hartwin; Varley, Joel B.; Speck, James S.; Van de Walle, Chris G.</p> <p>2018-06-01</p> <p>Ga2<span class="hlt">O</span><span class="hlt">3</span> is emerging as an important electronic material. Alloying with Al2<span class="hlt">O</span><span class="hlt">3</span> is a viable method to achieve carrier confinement, to increase the bandgap, or to modify the lattice parameters. However, the two materials have very different ground-state crystal structures (monoclinic β-gallia for Ga2<span class="hlt">O</span><span class="hlt">3</span> and corundum for Al2<span class="hlt">O</span><span class="hlt">3</span>). Here, we use hybrid density functional theory calculations to assess the alloy stabilities and electronic properties of the alloys. We find that the monoclinic phase is the preferred structure for up to 71% Al incorporation, in close agreement with experimental phase diagrams, and that the ordered monoclinic AlGa<span class="hlt">O</span><span class="hlt">3</span> alloy is exceptionally stable. We also discuss bandgap bowing, lattice constants, and band offsets that can guide future synthesis and device design efforts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PEPS....4...34N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PEPS....4...34N"><span>Melting phase relations in the MgSi<span class="hlt">O</span><span class="hlt">3</span>-CaSi<span class="hlt">O</span><span class="hlt">3</span> system at 24 GPa</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nomura, Ryuichi; Zhou, Youmo; Irifune, Tetsuo</p> <p>2017-12-01</p> <p>The Earth's lower mantle is composed of bridgmanite, ferropericlase, and CaSi<span class="hlt">O</span><span class="hlt">3</span>-rich perovskite. The melting phase relations between each component are key to understanding the melting of the Earth's lower mantle and the crystallization of the deep magma ocean. In this study, melting phase relations in the MgSi<span class="hlt">O</span><span class="hlt">3</span>-CaSi<span class="hlt">O</span><span class="hlt">3</span> system were investigated at 24 GPa using a multi-anvil apparatus. The eutectic composition is (Mg,Ca)Si<span class="hlt">O</span><span class="hlt">3</span> with 81-86 mol% MgSi<span class="hlt">O</span><span class="hlt">3</span>. The solidus temperature is 2600-2620 K. The solubility of CaSi<span class="hlt">O</span><span class="hlt">3</span> component into bridgmanite increases with temperature, reaching a maximum of <span class="hlt">3</span>-6 mol% at the solidus, and then decreases with temperature. The same trend was observed for the solubility of MgSi<span class="hlt">O</span><span class="hlt">3</span> component into CaSi<span class="hlt">O</span><span class="hlt">3</span>-rich perovskite, with a maximum of 14-16 mol% at the solidus. The asymmetric regular solutions between bridgmanite and CaSi<span class="hlt">O</span><span class="hlt">3</span>-rich perovskite and between MgSi<span class="hlt">O</span><span class="hlt">3</span> and CaSi<span class="hlt">O</span><span class="hlt">3</span> liquid components well reproduce the melting phase relations constrained experimentally. [Figure not available: see fulltext.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.858a2009D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.858a2009D"><span>Effect of composition on properties of In2<span class="hlt">O</span><span class="hlt">3</span>-Ga2<span class="hlt">O</span><span class="hlt">3</span> thin films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Demin, I. E.; Kozlov, A. G.</p> <p>2017-06-01</p> <p>The In2<span class="hlt">O</span><span class="hlt">3</span>-Ga2<span class="hlt">O</span><span class="hlt">3</span> mixed oxide polycrystalline thin films with various ratios of components were obtained by pulsed laser deposition. The effect of films composition on surface morphology, electrophysical and gas sensing properties and energies of adsorption and desorption of combustible gases was studied. The films with50%In2<span class="hlt">O</span><span class="hlt">3</span>-50%Ga2<span class="hlt">O</span><span class="hlt">3</span> composition showed maximum gas response (˜25 times) combined with minimum optimal working temperature (˜530 °C) as compared with the other films. The optical transmittance of the films in visible range was investigated. For 50%In2<span class="hlt">O</span><span class="hlt">3</span>-50%Ga2<span class="hlt">O</span><span class="hlt">3</span> films, the transmittance is higher in comparison with the other films. The explanation of the dependency of films behaviors on their composition was presented.The In2<span class="hlt">O</span><span class="hlt">3</span>-Ga2<span class="hlt">O</span><span class="hlt">3</span> films were assumed to have perspectives as gas sensing material for semiconducting gas sensors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.935a2029D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.935a2029D"><span>Interfacial coupling in multiferroic BiFe<span class="hlt">O</span><span class="hlt">3</span> and ferromagnetic La2/<span class="hlt">3</span>Sr1/<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> thin films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dominguez, C.; E Ordoñez, J.; E Gomez, M.</p> <p>2017-12-01</p> <p>Antiferromagnetic/Ferromagnetic coupling mechanics have been studying by growing successfully BiFe<span class="hlt">O</span><span class="hlt">3</span>/La2/<span class="hlt">3</span>Sr1/<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> bilayers on SrTi<span class="hlt">O</span><span class="hlt">3</span> single crystals by using rf and dc sputtering technique at pure oxygen pressures. We have investigated the magnetic behaviour of this samples, field cooling loops evidence interfacial coupling effect when antiferromagnetic ferroelectric BiFe<span class="hlt">O</span><span class="hlt">3</span> is placed in contact with ferromagnetic La2/<span class="hlt">3</span>Sr1/<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> indicate by the shift of the magnetization loop to negative values of the applied magnetic field. Our samples exhibited an exchange bias of 76Oe at 5K after field cooling the sample under 5000Oe. Temperature dependence of the exchange bias field showed exponential decay. The BFO/LSMO bilayer exhibits excellent ferroelectric behaviour (Ps=65μC/cm2 at 4V and 100Hz). Coexistence of ferroelectric and ferromagnetic properties in the BFO/LSMO bilayer make it a promising candidate system for applications where the magnetoelectric behaviour is required.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20418048','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20418048"><span>Is the Health of the Nation Outcome Scales appropriate for the assessment of symptom severity in patients with substance-related disorders?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Andreas, Sylke; Harries-Hedder, Karin; Schwenk, Wolfgang; Hausberg, Maria; Koch, Uwe; Schulz, Holger</p> <p>2010-07-01</p> <p>The Health of the Nation Outcome Scales (<span class="hlt">HoNOS</span>) is an internationally established clinician-rated instrument. The aim of the study was to assess the psychometric properties in inpatients with substance-related disorders. The <span class="hlt">HoNOS</span> was applied in a multicenter, consecutive sample of 417 inpatients. Interrater reliability coefficients, confirmatory factor analysis, and regression tree analyses were calculated to assess the reliability and validity of the <span class="hlt">HoNOS</span>. The factor validity of the <span class="hlt">HoNOS</span> and its total score could not be confirmed. After training, all items of the <span class="hlt">HoNOS</span> revealed sufficient values of interrater reliabilities. As the results of the regression tree analyses showed, the single items of the <span class="hlt">HoNOS</span> were one of the most important predictor of service utilization. The <span class="hlt">HoNOS</span> can be recommended for obtaining detailed ratings of the problems of inpatients with substance-related disorders as a clinical application in routine mental health care at present. Further studies should include comparisons of <span class="hlt">HoNOS</span> and Addiction Severity Index. Copyright 2010 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..MARA13005P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..MARA13005P"><span>High-carrier-density phase in LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> superlattices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Park, Se Young; Rabe, Karin; Millis, Andrew</p> <p>2015-03-01</p> <p>We investigate superlattices composed of alternating layers of Mott insulating LaTi<span class="hlt">O</span><span class="hlt">3</span> and band insulating SrTi<span class="hlt">O</span><span class="hlt">3</span> from first principles, using the density functional theory plus U (DFT+U) method. For values of U above a critical threshold, we find that melting of the Mott-insulating phase can extend from the interface into the LaTi<span class="hlt">O</span><span class="hlt">3</span> layer, resulting in a sheet carrier density exceeding the density of 0.5 electrons per in-plane unit cell found in previous studies. The critical U for the melting transition is larger than the critical Coulomb correlation required for the insulating LaTi<span class="hlt">O</span><span class="hlt">3</span>, suggesting the existence of a high sheet carrier density phase in LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> superlattices. The effects of in-plane strain and varying layer thickness on the melting transition are discussed. For insulating superlattices, we study the strain and thickness dependence of the polarization and its relation to near-interface local atomic distortions. Support: DOE ER 046169, ONR N00014-11-0666.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3522926','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3522926"><span>Study on Decomposition of Indoor Air Contaminants by Pulsed Atmospheric Microplasma</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Shimizu, Kazuo; Kuwabara, Tomoya; Blajan, Marius</p> <p>2012-01-01</p> <p>Decomposition of formaldehyde (<span class="hlt">HCHO</span>) by a microplasma reactor in order to improve Indoor Air Quality (IAQ) was achieved. <span class="hlt">HCHO</span> was removed from air using one pass through reactor treatment (5 L/min). From an initial concentration of <span class="hlt">HCHO</span> of 0.7 ppm about 96% was removed in one pass treatment using a discharge power of 0.<span class="hlt">3</span> W provided by a high voltage amplifier and a Marx Generator with MOSFET switches as pulsed power supplies. Moreover microplasma driven by the Marx Generator did not generate NOx as detected by a chemiluminescence NOx analyzer. In the case of large volume treatment the removal ratio of <span class="hlt">HCHO</span> (initial concentration: 0.5 ppm) after 60 minutes was 51% at 1.2 kV when using HV amplifier considering also a 41% natural decay ratio of <span class="hlt">HCHO</span>. The removal ratio was 54% at 1.2 kV when a Marx Generator energized the electrodes with a 44% natural decay ratio after 60 minutes of treatment. PMID:23202173</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARA37004X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARA37004X"><span>Carrier Density at LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> Interfaces: Evidence of Electronic Reconstruction.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xi, Xiaoxing</p> <p></p> <p>The origin of the 2D electron gas at the LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> interface has been a controversial subject ever since its discovery. A serious inconsistency with the most accepted mechanism, an electronic reconstruction in response to a polar discontinuity at the interface, is that the carrier densities reported experimentally are invariably lower than the expected value except under conditions where reduction of SrTi<span class="hlt">O</span><span class="hlt">3</span> substrate is suspected. We have grown LaAl<span class="hlt">O</span><span class="hlt">3</span> films of different stoichiometry on Ti<span class="hlt">O</span>2-terminated SrTi<span class="hlt">O</span><span class="hlt">3</span> substrates using atomic layer-by-layer laser molecular beam epitaxy (ALL-Laser MBE), in which La2<span class="hlt">O</span><span class="hlt">3</span> and Al2<span class="hlt">O</span><span class="hlt">3</span> targets were sequentially ablated in 37 mTorr oxygen. The high oxygen pressure during growth prevents the possible oxygen reduction in SrTi<span class="hlt">O</span><span class="hlt">3</span>, ensures that the LaAl<span class="hlt">O</span><span class="hlt">3</span> films are sufficiently oxygenated, and suppresses the La-Sr intermixing due to the bombardment effect. X-ray linear dichroism (XLD) and x-ray magnetic circular dichroism (XMCD) measurements show characteristics of oxygenated samples. In the electronic reconstruction picture, instead of the charge transfer of half of an electron in the case of a sufficiently thick stoichiometric LaAl<span class="hlt">O</span><span class="hlt">3</span>, a LaAl<span class="hlt">O</span><span class="hlt">3</span> film thickness dependence is expected as well as a linear dependence on stoichiometry. Our experimental results on carrier densities in 10 nm-thick LaAl1 +y<span class="hlt">O</span><span class="hlt">3</span>(1 +0.5y) films agree quantitatively with the theoretical expectations, lending a strong support for the electronic reconstruction mechanism. This material is based upon work supported by the U.S. Department of Energy, Office of Science, under Grant No. DE-SC0004764.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1438181','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1438181"><span>Modulation of superconducting transition temperature in LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> by SrTi<span class="hlt">O</span><span class="hlt">3</span> structural domains</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Noad, Hilary; Moler, Kathryn</p> <p>2018-01-01</p> <p>The tetragonal domain structure in SrTi<span class="hlt">O</span><span class="hlt">3</span> (STO) is known to modulate the normal-state carrier density in LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> (LAO/STO) heterostructures, among other electronic properties, but the effect of STO domains on the superconductivity in LAO/STO has not been fully explored. Using a scanning SQUID susceptometer microscope to map the superconducting response as a function of temperature in LAO/STO, we find that the superconducting transition temperature is spatially inhomogeneous and modulated in a pattern that is characteristic of structural domains in the STO.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011APS..MARH11005P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011APS..MARH11005P"><span>Thermoelectric Properties and Microstructure of Ca<span class="hlt">3</span> Co 4 <span class="hlt">O</span> 9 thin films on SrTi<span class="hlt">O</span><span class="hlt">3</span> and Al2 <span class="hlt">O</span> <span class="hlt">3</span> Substrates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Paulauskas, T.; Qiao, Q.; Gulec, A.; Klie, R. F.; Ozdemir, M.; Boyraz, C.; Mazumdar, D.; Gupta, A.</p> <p>2011-03-01</p> <p>Ca <span class="hlt">3</span> Co 4 <span class="hlt">O</span>9 (CCO), a misfit layered structure exhibiting large Seebeck coefficient at temperatures up to 1000K has attracted increasing attention as a novel high-temperature thermoelectric material. In this work, we investigate CCO thin films grown on SrTi <span class="hlt">O</span><span class="hlt">3</span> (001) and Al 2 <span class="hlt">O</span><span class="hlt">3</span> (0001) using pulsed laser deposition. Quality of the thin films was examined using high-resolution transmission electron microscopy and thermoelectric transport measurements. HRTEM images show incommensurate stacks of Cd I2 -type Co <span class="hlt">O</span>2 layer alternating with rock-salt-type Ca 2 Co <span class="hlt">O</span><span class="hlt">3</span> layer along the c-axis. Perovskite buffer layer about 10nm thick was found present between CCO and SrTi <span class="hlt">O</span><span class="hlt">3</span> accompanied by higher density of stacking faults. The CCO grown on Al 2 <span class="hlt">O</span><span class="hlt">3</span> exhibited numerous misoriented grains and presence of Ca x Co <span class="hlt">O</span>2 phase. Seebeck coefficient measurements yield an improvement for both samples compared to the bulk value. We suggest that thermoelectric properties of CCO increase due to additional phonon scattering at the stacking faults as well as at the film surfaces/interfaces. This research was supported by the US Army Research Office (W911NF-10-1-0147) and the Sivananthan Undergraduate Research Fellowship.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011MMTB...42...50Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011MMTB...42...50Z"><span>Effects of Al2<span class="hlt">O</span><span class="hlt">3</span> and Ca<span class="hlt">O/SiO</span>2 Ratio on Phase Equilbria in the Zn<span class="hlt">O-"FeO</span>"-Al2<span class="hlt">O</span><span class="hlt">3</span>-Ca<span class="hlt">O-SiO</span>2 System in Equilibrium with Metallic Iron</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Baojun; Hayes, Peter C.; Jak, Evgueni</p> <p>2011-02-01</p> <p>The phase equilibria and liquidus temperatures in the Zn<span class="hlt">O-"FeO</span>"-Al2<span class="hlt">O</span><span class="hlt">3</span>-Ca<span class="hlt">O-SiO</span>2 system in equilibrium with metallic iron have been determined experimentally in the temperature range 1383 K to 1573 K (1150 °C to 1300 °C). The experimental conditions were selected to characterize lead blast furnace and imperial smelting furnace slags. The results are presented in a form of pseudoternary sections Zn<span class="hlt">O-"FeO</span>"-(Al2<span class="hlt">O</span><span class="hlt">3</span> + Ca<span class="hlt">O</span> + Si<span class="hlt">O</span>2) with fixed Ca<span class="hlt">O/SiO</span>2 and (Ca<span class="hlt">O</span> + Si<span class="hlt">O</span>2)/Al2<span class="hlt">O</span><span class="hlt">3</span> ratios. It was found that wustite and spinel are the major primary phases in the composition range investigated. Effects of Al2<span class="hlt">O</span><span class="hlt">3</span> concentration as well as the Ca<span class="hlt">O/SiO</span>2 ratio on the primary phase field, the liquidus temperature, and the partitioning of Zn<span class="hlt">O</span> between liquid and solid phases have been discussed for zinc-containing slags.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MS%26E..275a2032L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MS%26E..275a2032L"><span>Microstructure, Thermal, Mechanical, and Dielectric Properties of Ba<span class="hlt">O-CaO</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 Glass-Ceramics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Bo; Bian, Haibo; Fang, Yi</p> <p>2017-12-01</p> <p>Ba<span class="hlt">O-CaO</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 (BCABS) glass-ceramics were prepared via the method of controlled crystallization. The effect of Ca<span class="hlt">O</span> modification on the microstructure, phase evolution, as well as thermal, mechanical, and dielectric properties was investigated. XRD identified that quartz is the major crystal phase; cristobalite and bazirite are the minor crystal phases. Moreover, the increase of Ca<span class="hlt">O</span> could inhibit the phase transformation from quartz to cristobalite, but excessive Ca<span class="hlt">O</span> would increase the porosity of the ceramics. Additionally, with increasing the amount of Ca<span class="hlt">O</span>, the thermal expansion curve tends to be linear, and subsequently the CTE value decreases gradually, which is attributed to the decrease of cristobalite with high CTE and the formation of CaSi<span class="hlt">O</span><span class="hlt">3</span> with low CTE. The results indicated that a moderate amount of Ca<span class="hlt">O</span> helps attaining excellent mechanical, thermal, and dielectric properties, that is, the specimen with 9 wt% Ca<span class="hlt">O</span> sintered at 950 °C has a high CTE value (11.5 × 10-6/°C), a high flexural strength (165.7 MPa), and good dielectric properties (ɛr = 6.2, tanδ = 1.8 × 10-4, ρ = 4.6 × 1011 Ω•cm).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JOM...tmp..264H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JOM...tmp..264H"><span>Fabrication by Electrophoretic Deposition of Nano-Fe<span class="hlt">3</span><span class="hlt">O</span>4 and Fe<span class="hlt">3</span><span class="hlt">O</span>4@Si<span class="hlt">O</span>2 <span class="hlt">3</span>D Structure on Carbon Fibers as Supercapacitor Materials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hajalilou, Abdollah; Abouzari-Lotf, Ebrahim; Etemadifar, Reza; Abbasi-Chianeh, Vahid; Kianvash, Abbas</p> <p>2018-05-01</p> <p>Core-shell nanostructured magnetic Fe<span class="hlt">3</span><span class="hlt">O</span>4@Si<span class="hlt">O</span>2 with particle size ranging from <span class="hlt">3</span> nm to 40 nm has been synthesized via a facile precipitation method. Tetraethyl orthosilicate was employed as surfactant to prepare core-shell structures from Fe<span class="hlt">3</span><span class="hlt">O</span>4 nanoparticles synthesized from pomegranate peel extract using a green method. X-ray diffraction analysis, Fourier-transform infrared and ultraviolet-visible (UV-Vis) spectroscopies, transmission electron microscopy, and scanning electron microscopy with energy-dispersive spectroscopy were employed to characterize the samples. The prepared Fe<span class="hlt">3</span><span class="hlt">O</span>4 nanoparticles were approximately 12 nm in size, and the thickness of the Si<span class="hlt">O</span>2 shell was 4 nm. Evaluation of the magnetic properties indicated lower saturation magnetization for Fe<span class="hlt">3</span><span class="hlt">O</span>4@Si<span class="hlt">O</span>2 powder ( 11.26 emu/g) compared with Fe<span class="hlt">3</span><span class="hlt">O</span>4 powder ( 13.30 emu/g), supporting successful wrapping of the Fe<span class="hlt">3</span><span class="hlt">O</span>4 nanoparticles by Si<span class="hlt">O</span>2. As-prepared powders were deposited on carbon fibers (CFs) using electrophoretic deposition and their electrochemical behavior investigated. The rectangular-shaped cyclic voltagrams of Fe<span class="hlt">3</span><span class="hlt">O</span>4@CF and Fe<span class="hlt">3</span><span class="hlt">O</span>4@C@CF samples indicated electrochemical double-layer capacitor (EDLC) behavior. The higher specific capacitance of 477 F/g for Fe<span class="hlt">3</span><span class="hlt">O</span>4@C@CF (at scan rate of 0.05 V/s in the potential range of - 1.13 to 0.45 V) compared with 205 F/g for Fe<span class="hlt">3</span><span class="hlt">O</span>4@CF (at the same scan rate in the potential range of - 1.04 to 0.24 V) makes the former a superior candidate for use in energy storage applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28129588','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28129588"><span>Study on spectroscopic properties and effects of tungsten ions in 2Bi2<span class="hlt">O</span><span class="hlt">3-3</span>Ge<span class="hlt">O</span>2/Si<span class="hlt">O</span>2 glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yu, Pingsheng; Su, Liangbi; Cheng, Junhua; Zhang, Xia; Xu, Jun</p> <p>2017-04-01</p> <p>The 2Bi 2 <span class="hlt">O</span> <span class="hlt">3</span> -<span class="hlt">3</span>Ge<span class="hlt">O</span> 2 /Si<span class="hlt">O</span> 2 glass samples have been prepared by the conventional melt quenching technique. XRD patterns, absorption spectra, excitation-emission spectra and Raman measurements were utilized to characterize the synthesized glasses. When substitute Si<span class="hlt">O</span> 2 for Ge<span class="hlt">O</span> 2 , the 0.4Bi 2 <span class="hlt">O</span> <span class="hlt">3</span> -(0.4-0.1)Ge<span class="hlt">O</span> 2 -(0.2-0.5)Si<span class="hlt">O</span> 2 glasses exhibit strong emission centered at about 475nm (under 300nm excitation), and the decay constants are within the scope of 20-40ns. W doping into 2Bi 2 <span class="hlt">O</span> <span class="hlt">3</span> -<span class="hlt">3</span>Si<span class="hlt">O</span> 2 glass could increase the emission intensity of 470nm, and the W-doped 2Bi 2 <span class="hlt">O</span> <span class="hlt">3</span> -<span class="hlt">3</span>Si<span class="hlt">O</span> 2 glass has shown another emission at about 433nm with much shorter decay time (near 10ns). The 2Bi 2 <span class="hlt">O</span> <span class="hlt">3</span> -<span class="hlt">3</span>Ge<span class="hlt">O</span> 2 /Si<span class="hlt">O</span> 2 glass system could be the possible candidate for scintillator in high energy physics applications. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990078735&hterms=1575&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3D%2526%25231575','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990078735&hterms=1575&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3D%2526%25231575"><span>Rate Coefficient Measurements of the Reaction CH<span class="hlt">3</span> + <span class="hlt">O</span>2 = CH<span class="hlt">3</span><span class="hlt">O</span> + <span class="hlt">O</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hwang, S. M.; Ryu, Si-Ok; DeWitt, K. J.; Rabinowitz, M. J.</p> <p>1999-01-01</p> <p>Rate coefficients for the reaction CH<span class="hlt">3</span> + <span class="hlt">O</span>2 = CH<span class="hlt">3</span><span class="hlt">O</span> + <span class="hlt">O</span> were measured behind reflected shock waves in a series of lean CH4-<span class="hlt">O</span>2-Ar mixtures using hydroxyl and methyl radical diagnostics. The rate coefficients are well represented by an Arrhenius expression given as k = (1.60(sup +0.67, sub -0.47 ) x 10(exp 13) e(-15813 +/- 587 K/T)/cubic cm.mol.s. This expression, which is valid in the temperature range 1575-1822 K, supports the downward trend in the rate coefficients that has been reported in recent determinations. All measurements to date, including the present study, have been to some extent affected by secondary reactions. The complications due to secondary reactions, choice of thermochemical data, and shock-boundary layer interactions that affect the determination of the rate coefficients are examined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000004898&hterms=1575&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3D%2526%25231575','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000004898&hterms=1575&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3D%2526%25231575"><span>Rate Coefficient Measurements of the Reaction CH<span class="hlt">3</span>+<span class="hlt">O</span>2+CH<span class="hlt">3</span><span class="hlt">O+O</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hwang, S. M.; Ryu, Si-Ok; DeWitt, K. J.; Rabinowitz, M. J.</p> <p>1999-01-01</p> <p>Rate coefficients for the reaction CH<span class="hlt">3</span> + <span class="hlt">O</span>2 = CH<span class="hlt">3</span><span class="hlt">O</span> + <span class="hlt">O</span> were measured behind reflected shock waves in a series of lean CH4-<span class="hlt">O</span>2-Ar mixtures using hydroxyl and methyl radical diagnostics. The rate coefficients are well represented by an Arrhenius expression given as k = (1.60(sup +0.67, -0.47)) X 10(exp 13) exp(- 15813 +/- 587 K/T)cc/mol s. This expression, which is valid in the temperature range 1575-1822 K, supports the downward trend in the rate coefficients that has been reported in recent determinations. All measurements to date, including the present study, have been to some extent affected by secondary reactions. The complications due to secondary reactions, choice of thermochemical data, and shock-boundary layer interactions that affect the determination of the rate coefficients are examined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27344231','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27344231"><span>Evaluation to the effect of B2<span class="hlt">O</span><span class="hlt">3</span>-La2<span class="hlt">O</span><span class="hlt">3</span>-Sr<span class="hlt">O</span>-Na2<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> bonding agent on Ti6Al4V-porcelain bonding.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, C Q; Wu, S Q; Lu, Y J; Gan, Y L; Guo, S; Lin, J J; Huang, T T; Lin, J X</p> <p>2016-10-01</p> <p>Low-fusing bonding agents have been widely applied in Ti-ceramics restorations. As an important category, borate bonding agents have great potentials in increasing Ti-porcelain bonding. The purpose of this study is to evaluate the effect of borate bonding agent with addition of Na2<span class="hlt">O</span> and Al2<span class="hlt">O</span><span class="hlt">3</span> on Ti6Al4V-porcelain bonding. The thermal properties of borate bonding agent, such as glass transition temperature (Tg) and crystallization peak temperature (Tp), were investigated to establish the sintering process. And the coefficient of thermal expansion (CTE) was to evaluate the matching effect of porcelain to Ti6Al4V. The bond strength was analyzed by the three point bending test. The microscopic morphology of the borate bonding agent surface after sintering, the interface of Ti-borate bonding agent-porcelain, and the fracture mode after porcelains fracture, were studied to assess the influence of borate bonding agent on Ti6Al4V-ceramics. With the addition of Na2<span class="hlt">O</span> and Al2<span class="hlt">O</span><span class="hlt">3</span>, the porcelain residues were observed increased indication on the Ti6Al4V surface after porcelain fracture and the bond strength was acquired the maximum (49.45MPa) in the bonding agent composition of 75.70B2<span class="hlt">O</span><span class="hlt">3</span>-5.92La2<span class="hlt">O</span><span class="hlt">3</span>-11.84Sr<span class="hlt">O</span>-4.67Na2<span class="hlt">O</span>-1.87Al2<span class="hlt">O</span><span class="hlt">3</span>. Those results suggest that borate bonding agent is an effective way to improve the Ti6Al4V-ceramics bond strength. And the addition of Na2<span class="hlt">O</span> and Al2<span class="hlt">O</span><span class="hlt">3</span> strengthen this effect. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1243173-iso-oriented-monolayer-moo-films-epitaxially-grown-srtio','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1243173-iso-oriented-monolayer-moo-films-epitaxially-grown-srtio"><span>Iso-oriented monolayer α-Mo<span class="hlt">O</span> <span class="hlt">3</span> (010) films epitaxially grown on SrTi<span class="hlt">O</span> <span class="hlt">3</span> (001)</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Du, Yingge; Li, Guoqiang; Peterson, Erik W.</p> <p></p> <p>The ability to synthesis well-ordered two-dimensional materials under ultra-high vacuum and directly characterize them by other techniques in-situ can greatly advance our current understanding on their physical and chemical properties. In this paper, we demonstrate that iso-oriented α-Mo<span class="hlt">O</span><span class="hlt">3</span> films with as low as single monolayer thickness can be reproducibly grown on SrTi<span class="hlt">O</span><span class="hlt">3</span>(001) substrates by molecular beam epitaxy ( (010)Mo<span class="hlt">O</span><span class="hlt">3</span> || (001)STO, [100]Mo<span class="hlt">O</span><span class="hlt">3</span> || [100]STO or [010]STO) through a self-limiting process. While one in-plane lattice parameter of the Mo<span class="hlt">O</span><span class="hlt">3</span> is very close to that of the SrTi<span class="hlt">O</span><span class="hlt">3</span> (aMo<span class="hlt">O</span><span class="hlt">3</span> = <span class="hlt">3</span>.96 Å, aSTO = <span class="hlt">3</span>.905 Å), the lattice mismatch along other directionmore » is large (~5%, cMo<span class="hlt">O</span><span class="hlt">3</span> = <span class="hlt">3</span>.70 Å), which leads to relaxation as clearly observed from the splitting of streaks in reflection high-energy electron diffraction (RHEED) patterns. A narrow range in the growth temperature is found to be optimal for the growth of monolayer α-Mo<span class="hlt">O</span><span class="hlt">3</span> films. Increasing deposition time will not lead to further increase in thickness, which is explained by a balance between deposition and thermal desorption due to the weak van der Waals force between α-Mo<span class="hlt">O</span><span class="hlt">3</span> layers. Lowering growth temperature after the initial iso-oriented α-Mo<span class="hlt">O</span><span class="hlt">3</span> monolayer leads to thicker α-Mo<span class="hlt">O</span><span class="hlt">3</span>(010) films with excellent crystallinity.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015STAdM..16b6003B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015STAdM..16b6003B"><span>Magnetic properties of solid solutions between BiCr<span class="hlt">O</span><span class="hlt">3</span> and BiGa<span class="hlt">O</span><span class="hlt">3</span> with perovskite structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Belik, Alexei A.</p> <p>2015-04-01</p> <p>Magnetic properties of BiCr1-xGax<span class="hlt">O</span><span class="hlt">3</span> perovskite-type solid solutions are reported, and a magnetic phase diagram is established. As-synthesized BiCr<span class="hlt">O</span><span class="hlt">3</span> and BiCr0.9Ga0.1<span class="hlt">O</span><span class="hlt">3</span> crystallize in a monoclinic (m) C2/c structure. The Néel temperature (TN) decreases from 111 K in BiCr<span class="hlt">O</span><span class="hlt">3</span> to 98 K in BiCr0.9Ga0.1<span class="hlt">O</span><span class="hlt">3</span>, and spin-reorientation transition temperature increases from 72 K in BiCr<span class="hlt">O</span><span class="hlt">3</span> to 83 K in BiCr0.9Ga0.1<span class="hlt">O</span><span class="hlt">3</span>. <span class="hlt">o</span>-BiCr0.9Ga0.1<span class="hlt">O</span><span class="hlt">3</span> with a PbZr<span class="hlt">O</span><span class="hlt">3</span>-type orthorhombic structure is obtained by heating m-BiCr0.9Ga0.1<span class="hlt">O</span><span class="hlt">3</span> up to 573 K in air; it shows similar magnetic properties with those of m-BiCr0.9Ga0.1<span class="hlt">O</span><span class="hlt">3</span>. TN of BiCr0.8Ga0.2<span class="hlt">O</span><span class="hlt">3</span> is 81 K, and TN of BiCr0.7Ga0.<span class="hlt">3</span><span class="hlt">O</span><span class="hlt">3</span> is 63 K. Samples with x = 0.4, 0.5, 0.6 and 0.7 crystallize in a polar R<span class="hlt">3</span>c structure. Long-range antiferromagnetic order with weak ferromagnetism is observed below TN = 56 K in BiCr0.6Ga0.4<span class="hlt">O</span><span class="hlt">3</span>, TN = 36 K in BiCr0.5Ga0.5<span class="hlt">O</span><span class="hlt">3</span> and TN = 18 K in BiCr0.4Ga0.6<span class="hlt">O</span><span class="hlt">3</span>. BiCr0.<span class="hlt">3</span>Ga0.7<span class="hlt">O</span><span class="hlt">3</span> shows a paramagnetic behaviour because the Cr concentration is below the percolation threshold of 31%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1376467-diffusion-quantum-monte-carlo-calculations-srfeo3-lafeo3','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1376467-diffusion-quantum-monte-carlo-calculations-srfeo3-lafeo3"><span>Diffusion quantum Monte Carlo calculations of SrFe<span class="hlt">O</span> <span class="hlt">3</span> and LaFe<span class="hlt">O</span> <span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Santana, Juan A.; Krogel, Jaron T.; Kent, Paul R. C.; ...</p> <p>2017-07-18</p> <p>The equations of state, formation energy, and migration energy barrier of the oxygen vacancy in SrFe<span class="hlt">O</span> <span class="hlt">3</span> and LaFe<span class="hlt">O</span> <span class="hlt">3</span> were calculated in this paper with the diffusion quantum Monte Carlo (DMC) method. Calculations were also performed with various Density Functional Theory (DFT) approximations for comparison. DMC reproduces the measured cohesive energies of these materials with errors below 0.23(5) eV and the structural properties within 1% of the experimental values. The DMC formation energies of the oxygen vacancy in SrFe<span class="hlt">O</span> <span class="hlt">3</span> and LaFe<span class="hlt">O</span> <span class="hlt">3</span> under oxygen-rich conditions are 1.<span class="hlt">3</span>(1) and 6.24(7) eV, respectively. Similar calculations with semi-local DFT approximations formore » LaFe<span class="hlt">O</span> <span class="hlt">3</span> yielded vacancy formation energies 1.5 eV lower. Comparison of charge density evaluated with DMC and DFT approximations shows that DFT tends to overdelocalize the electrons in defected SrFe<span class="hlt">O</span> <span class="hlt">3</span> and LaFe<span class="hlt">O</span> <span class="hlt">3</span>. Finally, calculations with DMC and local density approximation yield similar vacancy migration energy barriers, indicating that steric/electrostatic effects mainly determine migration barriers in these materials.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARY43007P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARY43007P"><span>Low energy spin dynamics of rare-earth orthoferrites YFe<span class="hlt">O</span><span class="hlt">3</span> and LaFe<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Park, Kisoo; Sim, Hasung; Leiner, Jonathan; Yoshida, Yoshiyuki; Eisaki, Hiroshi; Yano, Shinichiro; Gardner, Jason; Park, Je-Geun</p> <p></p> <p>YFe<span class="hlt">O</span><span class="hlt">3</span> and LaFe<span class="hlt">O</span><span class="hlt">3</span>\\ are members of the rare-earth orthoferrites (RFe<span class="hlt">O</span><span class="hlt">3</span>) family with Pbnm space group. With the strong superexchange interaction between Fe<span class="hlt">3</span> + ions, both compounds exhibit the room temperature antiferromagnetic order (TN >600 K) with a slight spin canting. Here we report low-energy magnetic excitation of YFe<span class="hlt">O</span><span class="hlt">3</span> and LaFe<span class="hlt">O</span><span class="hlt">3</span> using inelastic neutron scattering measurements, showing evidence of magnon mode splitting and a spin anisotropy gap at the zone center. Spin wave calculations with the spin Hamiltonian including both Dzyaloshinsky-Moriya interaction and single-ion anisotropy accounts for the observed features well. Our results offer insight into the underlying physics of other RFe<span class="hlt">O</span><span class="hlt">3</span>\\ with magnetic rare-earth ions or related Fe<span class="hlt">3</span>+-based multiferroic perovskites such as BiFe<span class="hlt">O</span><span class="hlt">3</span>. The work at the IBS CCES (South Korea) was supported by the research program of the Institute for Basic Science (IBS-R009-G1).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JEMat..47..766L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JEMat..47..766L"><span>Influence of Y2<span class="hlt">O</span><span class="hlt">3</span> Addition on Crystallization, Thermal, Mechanical, and Electrical Properties of Ba<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2 Glass-Ceramic for Ceramic Ball Grid Array Package</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Bo; Li, Wei; Zheng, Jingguo</p> <p>2018-01-01</p> <p>Y2<span class="hlt">O</span><span class="hlt">3</span> addition has a significant influence on the crystallization, thermal, mechanical, and electrical properties of Ba<span class="hlt">O</span> -Al2<span class="hlt">O</span><span class="hlt">3</span> -B2<span class="hlt">O</span><span class="hlt">3</span> -Si<span class="hlt">O</span>2 (BABS) glass-ceramics. Semi-quantitative calculation based on x-ray diffraction demonstrated that with increasing Y2<span class="hlt">O</span><span class="hlt">3</span> content, both the crystallinity and the phase content of cristobalite gradually decreased. It is effective for the additive Y2<span class="hlt">O</span><span class="hlt">3</span> to inhibit the formation of cristobalite phase with a large coefficient of thermal expansion value. The flexural strength and the Young's modulus, thus, are remarkably increased from 140 MPa to 200 MPa and 56.5 GPa to 63.7 GPa, respectively. Also, the sintering kinetics of BABS glass-ceramics with various Y2<span class="hlt">O</span><span class="hlt">3</span> were investigated using the isothermal sintering shrinkage curve at different sintering temperatures. The sintering activation energy Q sharply decreased from 99.8 kJ/mol to 81.5 kJ/mol when 0.2% Y2<span class="hlt">O</span><span class="hlt">3</span> was added, which indicated that a small amount of Y2<span class="hlt">O</span><span class="hlt">3</span> could effectively promote the sintering procedure of BABS glass-ceramics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20075908-infrared-tpd-studies-nitrates-adsorbed-tb-sub-sub-la-sub-sub-bao-mgo-gamma-al-sub-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20075908-infrared-tpd-studies-nitrates-adsorbed-tb-sub-sub-la-sub-sub-bao-mgo-gamma-al-sub-sub"><span>Infrared and TPD studies of nitrates adsorbed on Tb{sub 4}<span class="hlt">O</span>{sub 7}, La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}, Ba<span class="hlt">O</span>, and Mg<span class="hlt">O</span>/{gamma}-Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Chi, Y.; Chuang, S.S.C.</p> <p>2000-05-18</p> <p>NO and <span class="hlt">O</span>{sub 2} coadsorption on {gamma}-Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}-supported Tb{sub 4}<span class="hlt">O</span>{sub 7}, La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}, Ba<span class="hlt">O</span>, and Mg<span class="hlt">O</span> has been investigated by in situ infrared spectroscopy coupled with temperature-programmed decomposition and desorption. Ba<span class="hlt">O</span>/{gamma}-Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} and Mg<span class="hlt">O</span>/{gamma}-Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} possess a higher NO{sub x} storage capability than Tb{sub 4}<span class="hlt">O</span>{sub 7}/{gamma}-Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} and La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}/{gamma}-Al{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}. NO/<span class="hlt">O</span>{sub 2} coadsorbed on Tb{sub 4}<span class="hlt">O</span>{sub 7}, La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}, and Ba<span class="hlt">O</span> in the form of bridging bidentate, chelating bidentate, and monodentate nitrates, and on Mg<span class="hlt">O</span> in the form of bridging bidentate and monodentate nitrates via the reaction of adsorbed NO withmore » adsorbed oxygen at 298 K. NO/<span class="hlt">O</span>{sub 2} coadsorbed as a chelating bidentate nitrate on Tb{sub 4}<span class="hlt">O</span>{sub 7} and La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}, and as a distinctive bridging bidentate nitrate on Ba<span class="hlt">O</span> and Mg<span class="hlt">O</span> via the reaction of adsorbed NO with surface lattice oxygen at 523 K. These various forms of adsorbed nitrate differ in structure and reactivity from Tb(NO{sub <span class="hlt">3</span>}){sub <span class="hlt">3</span>}, La(NO{sub <span class="hlt">3</span>}){sub <span class="hlt">3</span>}, Ba(NO{sub <span class="hlt">3</span>}){sub 2}, and Mg(NO{sub <span class="hlt">3</span>}){sub 2}, the precursor used to prepare metal oxides for NO/<span class="hlt">O</span>{sub 2} coadsorption. Temperature-programmed desorption (TPD) of chelating bidentate nitrate on Tb{sub 4}<span class="hlt">O</span>{sub 7}, La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}, and Ba<span class="hlt">O</span> produced primarily NO and <span class="hlt">O</span>{sub 2}, with maxima at 640 and 670 K, respectively. TPD of bridging bidentate nitrate and monodentate nitrate on Tb{sub 4}<span class="hlt">O</span>{sub 7}, La{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}, and Ba<span class="hlt">O</span> produced NO and <span class="hlt">O</span>{sub 2} as major products and N{sub 2} and N{sub 2}<span class="hlt">O</span> as minor products, at 320--500 K. Decomposition of bridging bidentate on Mg<span class="hlt">O</span> produced NO as a major product and N{sub 2}<span class="hlt">O</span> as a minor product at a peak temperature of 690 K. Peak temperatures for Tb(NO{sub <span class="hlt">3</span>}){sub <span class="hlt">3</span>}, La(NO{sub <span class="hlt">3</span>}){sub <span class="hlt">3</span>}, Ba(NO{sub <span class="hlt">3</span>}){sub 2}, and Mg(NO{sub <span class="hlt">3</span>}){sub 2} decomposition occurred between those for bridging and chelating nitrates. The difference in stability between</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhDT........56K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhDT........56K"><span>Modulation-Doped SrTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi1-xZrx<span class="hlt">O</span><span class="hlt">3</span> Heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kajdos, Adam Paul</p> <p></p> <p>Two-dimensional electron gases (2DEGs) in SrTi<span class="hlt">O</span><span class="hlt">3</span> have attracted considerable attention for exhibiting a variety of interesting physical phenomena, such as superconductivity and magnetism. So far, most of the literature has focused on interfaces between nonpolar SrTi<span class="hlt">O</span><span class="hlt">3</span> and polar perovskite oxides (e.g. LaAl<span class="hlt">O</span><span class="hlt">3</span> or rare-earth titanates), where high carrier density 2DEGs (˜<span class="hlt">3</span> x 1014 cm-2) are generated by polar discontinuity. Modulation doping is an alternative approach to generating a 2DEG that has been explored extensively in III-V semiconductors but has not heretofore been explored in complex oxides. This approach involves interfacing an undoped semiconductor with a doped semiconductor whose conduction band edge lies at a higher energy, which results in electrons diffusing into the undoped semiconductor transport channel, where scattering from ionized dopants is minimized. Realizing a high-mobility modulation-doped structure with a SrTi<span class="hlt">O</span><span class="hlt">3</span> transport channel therefore requires both the optimization of the transport channel by minimizing native defects as well as the development of a perovskite oxide which has a suitable band offset with SrTi<span class="hlt">O</span><span class="hlt">3</span> and can be electron-doped. The growth of high electron mobility SrTi<span class="hlt">O</span><span class="hlt">3</span> as a suitable transport channel material was previously demonstrated using the hybrid molecular beam epitaxy (MBE) approach, where Sr is delivered via a solid source and Ti is delivered using a metal-organic precursor, titanium (IV) tetra-isopropoxide (TTIP). Expanding on this, in-situ reflection high-energy electron diffraction (RHEED) is used to track the surface and resulting film cation stoichiometry of homoepitaxial SrTi<span class="hlt">O</span><span class="hlt">3</span> (001) thin films grown by hybrid MBE. It is shown that films with lattice parameters identical to bulk single-crystal substrates within the detection limit of high-resolution X-ray diffraction (XRD) measurements exhibit an evolution in surface reconstruction with increasing TTIP beam-equivalent pressure. The change in the observed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AtmEn.165...57F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AtmEn.165...57F"><span>Artificial <span class="hlt">O</span><span class="hlt">3</span> formation during fireworks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fiedrich, M.; Kurtenbach, R.; Wiesen, P.; Kleffmann, J.</p> <p>2017-09-01</p> <p>In several previous studies emission of ozone (<span class="hlt">O</span><span class="hlt">3</span>) during fireworks has been reported, which was attributed to either photolysis of molecular oxygen (<span class="hlt">O</span>2) or nitrogen dioxide (NO2) by short/near UV radiation emitted during the high-temperature combustion of fireworks. In contrast, in the present study no <span class="hlt">O</span><span class="hlt">3</span> formation was observed using a selective <span class="hlt">O</span><span class="hlt">3</span>-LOPAP instrument during the combustion of pyrotechnical material in the laboratory, while a standard <span class="hlt">O</span><span class="hlt">3</span> monitor using UV absorption showed extremely high <span class="hlt">O</span><span class="hlt">3</span> signals. The artificial <span class="hlt">O</span><span class="hlt">3</span> response of the standard <span class="hlt">O</span><span class="hlt">3</span> monitor was caused by known interferences associated with high levels of co-emitted VOCs and could also be confirmed in field measurements during New Year's Eve in the city of Wuppertal, Germany. The present results help to explain unreasonably high ozone levels documented during ambient fireworks, which are in contradiction to the fast titration of <span class="hlt">O</span><span class="hlt">3</span> by nitrogen monoxide (NO) in the night-time atmosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97c5438R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97c5438R"><span>Ultrafast modification of the polarity at LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> interfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rubano, A.; Günter, T.; Fiebig, M.; Granozio, F. Miletto; Marrucci, L.; Paparo, D.</p> <p>2018-01-01</p> <p>Oxide growth with semiconductorlike accuracy has led to atomically precise thin films and interfaces that exhibit a plethora of phases and functionalities not found in the oxide bulk material. This has yielded spectacular discoveries such as the conducting, magnetic, and even superconducting LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> interfaces separating two prototypical insulating perovskite materials. All these investigations, however, consider the static state at the interface, although studies on fast oxide interface dynamics would introduce a powerful degree of freedom to understanding the nature of the LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> interface state. Here, we show that the polarization state at the LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> interface can be optically enhanced or attenuated within picoseconds. Our observations are explained by a model based on charge propagation effects in the interfacial vicinity and transient polarization buildup at the interface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016FrP.....4...41P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016FrP.....4...41P"><span>Formation of self-organized Mn<span class="hlt">3</span><span class="hlt">O</span>4 nanoinclusions in LaMn<span class="hlt">O</span><span class="hlt">3</span> films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pomar, Alberto; Konstantinović, Zorica; Bagués, Nuria; Roqueta, Jaume; López-Mir, Laura; Balcells, Lluis; Frontera, Carlos; Mestres, Narcis; Gutiérrez-Llorente, Araceli; Šćepanović, Maja; Lazarević, Nenad; Popović, Zoran; Sandiumenge, Felip; Martínez, Benjamín; Santiso, José</p> <p>2016-09-01</p> <p>We present a single-step route to generate ordered nanocomposite thin films of secondary phase inclusions (Mn<span class="hlt">3</span><span class="hlt">O</span>4) in a pristine perovskite matrix (LaMn<span class="hlt">O</span><span class="hlt">3</span>) by taking advantage of the complex phase diagram of manganese oxides. We observed that in samples grown under vacuum growth conditions from a single LaMn<span class="hlt">O</span><span class="hlt">3</span> stoichiometric target by Pulsed Laser Deposition, the most favourable mechanism to accommodate Mn2+ cations is the spontaneous segregation of self-assembled wedge-like Mn<span class="hlt">3</span><span class="hlt">O</span>4 ferrimagnetic inclusions inside a LaMn<span class="hlt">O</span><span class="hlt">3</span> matrix that still preserves its orthorhombic structure and its antiferromagnetic bulk-like behaviour. A detailed analysis on the formation of the self-assembled nanocomposite films evidences that Mn<span class="hlt">3</span><span class="hlt">O</span>4 inclusions exhibit an epitaxial relationship with the surrounding matrix that it may be explained in terms of a distorted cubic spinel with slight ( 9º) c-axis tilting. Furthermore, a Ruddlesden-Popper La2Mn<span class="hlt">O</span>4 phase, helping to the stoichiometry balance, has been identified close to the interface with the substrate. We show that ferrimagnetic Mn<span class="hlt">3</span><span class="hlt">O</span>4 columns influence the magnetic and transport properties of the nanocomposite by increasing its coercive field and by creating local areas with enhanced conductivity in the vicinity of the inclusions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADD018987','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADD018987"><span>Method of Preparing Monoclinic Ba<span class="hlt">O</span>.Al2<span class="hlt">O</span><span class="hlt">3</span>.2Si<span class="hlt">O</span>2</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p></p> <p>Monoclinic celsian (Ba<span class="hlt">O</span>.Al2<span class="hlt">O</span><span class="hlt">3</span>.2Si<span class="hlt">O</span>2) is produced by heating a stoichiometric, powder mixture of BaCO<span class="hlt">3</span> (or BaC2<span class="hlt">O</span>4), Al2<span class="hlt">O</span><span class="hlt">3</span>, and Si<span class="hlt">O</span>2 (preferably Si<span class="hlt">O</span>2 gel) with monoclinic celsian seeds at from 1250 deg C to 1500 deg C.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA593688','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA593688"><span>Interfacial Ferromagnetism and Exchange Bias in CaRu<span class="hlt">O</span><span class="hlt">3</span>/CaMn<span class="hlt">O</span><span class="hlt">3</span> Superlattices</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2012-11-07</p> <p>microscopy and electron energy loss spectroscopy indicate that the difference in magnitude of the Mn valence states between the center of the CaMn<span class="hlt">O</span><span class="hlt">3</span> layer...CaMn<span class="hlt">O</span><span class="hlt">3</span> thickness dependence of the exchange bias field together indicate that the interfacial 1. REPORT DATE (DD-MM-YYYY) 4. TITLE AND SUBTITLE 13...superlattices of CaRu<span class="hlt">O</span><span class="hlt">3</span>/CaMn<span class="hlt">O</span><span class="hlt">3</span> that arises in one unit cell at the interface. Scanning transmission electron microscopy and electron energy loss</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28254711','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28254711"><span>Kinetic removal of haloacetonitrile precursors by photo-based advanced oxidation processes (UV/H2<span class="hlt">O</span>2, UV/<span class="hlt">O</span><span class="hlt">3</span>, and UV/H2<span class="hlt">O</span>2/<span class="hlt">O</span><span class="hlt">3</span>).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Srithep, Sirinthip; Phattarapattamawong, Songkeart</p> <p>2017-06-01</p> <p>The objective of the study is to evaluate the performance of conventional treatment process (i.e., coagulation, flocculation, sedimentation and sand filtration) on the removals of haloacetonitrile (HAN) precursors. In addition, the removals of HAN precursors by photo-based advanced oxidation processes (Photo-AOPs) (i.e., UV/H 2 <span class="hlt">O</span> 2 , UV/<span class="hlt">O</span> <span class="hlt">3</span> , and UV/H 2 <span class="hlt">O</span> 2 /<span class="hlt">O</span> <span class="hlt">3</span> ) are investigated. The conventional treatment process was ineffective to remove HAN precursors. Among Photo-AOPs, the UV/H 2 <span class="hlt">O</span> 2 /<span class="hlt">O</span> <span class="hlt">3</span> was the most effective process for removing HAN precursors, followed by UV/H 2 <span class="hlt">O</span> 2 , and UV/<span class="hlt">O</span> <span class="hlt">3</span> , respectively. For 20min contact time, the UV/H 2 <span class="hlt">O</span> 2 /<span class="hlt">O</span> <span class="hlt">3</span> , UV/H 2 <span class="hlt">O</span> 2 , and UV/<span class="hlt">O</span> <span class="hlt">3</span> suppressed the HAN formations by 54, 42, and 27% reduction. Increasing ozone doses from 1 to 5 mgL -1 in UV/<span class="hlt">O</span> <span class="hlt">3</span> systems slightly improved the removals of HAN precursors. Changes in pH (6-8) were unaffected most of processes (i.e., UV, UV/H 2 <span class="hlt">O</span> 2 , and UV/H 2 <span class="hlt">O</span> 2 /<span class="hlt">O</span> <span class="hlt">3</span> ), except for the UV/<span class="hlt">O</span> <span class="hlt">3</span> system that its efficiency was low in the weak acid condition. The pseudo first-order kinetic constant for removals of dichloroacetonitrile precursors (k' DCANFP ) by the UV/H 2 <span class="hlt">O</span> 2 /<span class="hlt">O</span> <span class="hlt">3</span> , UV/H 2 <span class="hlt">O</span> 2 and standalone UV systems were 1.4-2.8 orders magnitude higher than the UV/<span class="hlt">O</span> <span class="hlt">3</span> process. The kinetic degradation of dissolved organic nitrogen (DON) tended to be higher than the k' DCANFP value. This study firstly differentiates the kinetic degradation between DON and HAN precursors. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010SSSci..12.1756T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010SSSci..12.1756T"><span>Effects of nano-YAG (Y <span class="hlt">3</span>Al 5<span class="hlt">O</span> 12) crystallization on the structure and photoluminescence properties of Nd <span class="hlt">3</span>+-doped K 2<span class="hlt">O-SiO</span> 2-Y 2<span class="hlt">O</span> <span class="hlt">3</span>-Al 2<span class="hlt">O</span> <span class="hlt">3</span> glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tarafder, Anal; Molla, Atiar Rahaman; Karmakar, Basudeb</p> <p>2010-10-01</p> <p>Nd <span class="hlt">3</span>+-doped precursor glass in the K 2<span class="hlt">O-SiO</span> 2-Y 2<span class="hlt">O</span> <span class="hlt">3</span>-Al 2<span class="hlt">O</span> <span class="hlt">3</span> (KSYA) system was prepared by the melt-quench technique. The transparent Y <span class="hlt">3</span>Al 5<span class="hlt">O</span> 12 (YAG) glass-ceramics were derived from this glass by a controlled crystallization process at 750 °C for 5-100 h. The formation of YAG crystal phase, size and morphology with progress of heat-treatment was examined by X-ray diffraction (XRD), field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM) and Fourier transformed infrared reflectance spectroscopy (FT-IRRS). The crystallite sizes obtained from XRD are found to increase with heat-treatment time and vary in the range 25-40 nm. The measured photoluminescence spectra have exhibited emission transitions of 4F <span class="hlt">3</span>/2 → 4I J ( J = 9/2, 11/2 and 13/2) from Nd <span class="hlt">3</span>+ ions upon excitation at 829 nm. It is observed that the photoluminescence intensity and excited state lifetime of Nd <span class="hlt">3</span>+ ions decrease with increase in heat-treatment time. The present study indicates that the incorporation of Nd <span class="hlt">3</span>+ ions into YAG crystal lattice enhance the fluorescence performance of the glass-ceramic nanocomposites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008ApPhL..92f2508Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008ApPhL..92f2508Y"><span>Optical magnetoelectric effect at CaRu<span class="hlt">O</span><span class="hlt">3</span>-CaMn<span class="hlt">O</span><span class="hlt">3</span> interfaces as a polar ferromagnet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yamada, Hiroyuki; Sato, H.; Akoh, H.; Kida, N.; Arima, T.; Kawasaki, M.; Tokura, Y.</p> <p>2008-02-01</p> <p>A correlated electron interface between paramagnetic CaRu<span class="hlt">O</span><span class="hlt">3</span> and antiferromagnetic CaMn<span class="hlt">O</span><span class="hlt">3</span> has been characterized with optical magnetoelectric (OME) effect as an interface-selective probe for spin and charge states. To detect the OME effect, i.e., nonreciprocal directional dichroism for visible or near-infrared light, we have constructed a "tricolor" superlattice with artificially broken inversion symmetry by stacking CaRu<span class="hlt">O</span><span class="hlt">3</span>, CaMn<span class="hlt">O</span><span class="hlt">3</span>, and CaTi<span class="hlt">O</span><span class="hlt">3</span>, and patterned a grating structure with 4μm period on the superlattice. The observed intensity modulation (0.<span class="hlt">3</span>% at 50K) in the Bragg diffraction verifies a charge transfer and concomitant ferromagnetism at the CaRu<span class="hlt">O</span><span class="hlt">3</span>-CaMn<span class="hlt">O</span><span class="hlt">3</span> interface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PhRvB..84d5108Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PhRvB..84d5108Y"><span>Far-infrared and dc magnetotransport of CaMn<span class="hlt">O</span><span class="hlt">3</span>-CaRu<span class="hlt">O</span><span class="hlt">3</span> superlattices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yordanov, P.; Boris, A. V.; Freeland, J. W.; Kavich, J. J.; Chakhalian, J.; Lee, H. N.; Keimer, B.</p> <p>2011-07-01</p> <p>We report temperature- and magnetic-field-dependent measurements of the dc resistivity and the far-infrared reflectivity (FIR) (photon energies ℏω=50-700 cm-1) of superlattices comprising ten consecutive unit cells of the antiferromagnetic insulator CaMn<span class="hlt">O</span><span class="hlt">3</span>, and four to ten unit cells of the correlated paramagnetic metal CaRu<span class="hlt">O</span><span class="hlt">3</span>. Below the Néel temperature of CaMn<span class="hlt">O</span><span class="hlt">3</span>, the dc resistivity exhibits a logarithmic divergence upon cooling, which is associated with a large negative, isotropic magnetoresistance. The ω→0 extrapolation of the resistivity extracted from the FIR reflectivity, on the other hand, shows a much weaker temperature and field dependence. We attribute this behavior to scattering of itinerant charge carriers in CaRu<span class="hlt">O</span><span class="hlt">3</span> from sparse, spatially isolated magnetic defects at the CaMn<span class="hlt">O</span><span class="hlt">3</span>-CaRu<span class="hlt">O</span><span class="hlt">3</span> interfaces. This field-tunable “transport bottleneck” effect may prove useful for functional metal-oxide devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11874353','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11874353"><span>Hydrothermal syntheses, structures, and properties of the new uranyl selenites Ag(2)(UO(2))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))(2), M[(UO(2))(HSe<span class="hlt">O</span>(<span class="hlt">3</span>))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))] (M = K, Rb, Cs, Tl), and Pb(UO(2))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))(2).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Almond, Philip M; Albrecht-Schmitt, Thomas E</p> <p>2002-03-11</p> <p>The transition metal, alkali metal, and main group uranyl selenites, Ag(2)(UO(2))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))(2) (1), K[(UO(2))(HSe<span class="hlt">O</span>(<span class="hlt">3</span>))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))] (2), Rb[(UO(2))(HSe<span class="hlt">O</span>(<span class="hlt">3</span>))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))] (<span class="hlt">3</span>), Cs[(UO(2))(HSe<span class="hlt">O</span>(<span class="hlt">3</span>))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))] (4), Tl[(UO(2))(HSe<span class="hlt">O</span>(<span class="hlt">3</span>))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))] (5), and Pb(UO(2))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))(2) (6), have been prepared from the hydrothermal reactions of AgNO(<span class="hlt">3</span>), KCl, RbCl, CsCl, TlCl, or Pb(NO(<span class="hlt">3</span>))(2) with UO(<span class="hlt">3</span>) and Se<span class="hlt">O</span>(2) at 180 degrees C for <span class="hlt">3</span> d. The structures of 1-5 contain similar [(UO(2))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))(2)](2-) sheets constructed from pentagonal bipyramidal UO(7) units that are joined by bridging Se<span class="hlt">O</span>(<span class="hlt">3</span>)(2-) anions. In 1, the selenite oxo ligands that are not utilized within the layers coordinate the Ag(+) cations to create a three-dimensional network structure. In 2-5, half of the selenite ligands are monoprotonated to yield a layer composition of [(UO(2))(HSe<span class="hlt">O</span>(<span class="hlt">3</span>))(Se<span class="hlt">O</span>(<span class="hlt">3</span>))](1-), and coordination of the K(+), Rb(+), Cs(+), and Tl(+) cations occurs through long ionic contacts. The structure of 6 contains a uranyl selenite layered substructure that differs substantially from those in 1-5 because the selenite anions adopt both bridging and chelating binding modes to the uranyl centers. Furthermore, the Pb(2+) cations form strong covalent bonds with these anions creating a three-dimensional framework. These cations occur as distorted square pyramidal Pb<span class="hlt">O</span>(5) units with stereochemically active lone pairs of electrons. These polyhedra align along the c-axis to create a polar structure. Second-harmonic generation (SHG) measurements revealed a response of 5x alpha-quartz for 6. The diffuse reflectance spectrum of 6 shows optical transitions at 330 and 440 nm. The trailing off of the 440 nm transition to longer wavelengths is responsible for the orange coloration of 6.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SuMi...97..116W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SuMi...97..116W"><span>Study of the electronic structure and half-metallicity of CaMn<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span> superlattice</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Kai; Jiang, Wei; Chen, Jun-Nan; Huang, Jian-Qi</p> <p>2016-09-01</p> <p>In this paper, the electronic structure, magnetic properties and half-metallicity of the CaMn<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span> superlattice are investigated by employing the first-principle calculation based on density functional theory within the GGA or GGA + U exchange-correlation functional. The CaMn<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span> superlattice is constructed by the cubic CaMn<span class="hlt">O</span><span class="hlt">3</span> and the tetragonal ferroelectric BaTi<span class="hlt">O</span><span class="hlt">3</span> growing alternately along (0 0 1) direction. The cubic CaMn<span class="hlt">O</span><span class="hlt">3</span> presents a robust half-metallicity and a metastable ferromagnetic phase. Its magnetic moment is an integral number of <span class="hlt">3</span>.000 μB per unit cell. However, the CaMn<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span> superlattice has a stable ferromagnetic phase, for which the magnetic moment is 12.000 μB per unit cell. It also retains the robust half-metallicity which mainly results from the strong hybridization between Mn and <span class="hlt">O</span> atoms. The results show that the constructed CaMn<span class="hlt">O</span><span class="hlt">3</span>/BaTi<span class="hlt">O</span><span class="hlt">3</span> superlattice exhibits superior magnetoelectric properties. It may provide a theoretical reference for the design and preparation of new multiferroic materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.7982M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.7982M"><span>Simulation of the diurnal variations of the isotope anomaly (?17<span class="hlt">O</span>) of reactive trace gases (NOx, HOx) and implications for the ?17<span class="hlt">O</span> of nitrate.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Morin, Samuel; Sander, Rolf; Savarino, Joël.</p> <p>2010-05-01</p> <p>The isotope anomaly of secondary atmospheric species such as nitrate (NO<span class="hlt">3</span>-) has potential to provide useful constrains on their formation pathways. Indeed, the ?17<span class="hlt">O</span> of their precursors (NOx, HOx etc.) differs and depends on their interactions with ozone, which is the main source of non-zero ?17<span class="hlt">O</span> in the atmosphere. Interpreting variations of ?17<span class="hlt">O</span> in nitrate requires an in-depth understanding of the ?17<span class="hlt">O</span> of its precursors taking into account non-linear chemical regimes operating under various environmental settings. In addition, the role of isotope exchange reactions must be carefully accounted for. To investigate the relevance of various assessments of the isotopic signature of nitrate production pathways that have recently been proposed in the literature, an atmospheric chemistry box model (MECCA, Sander et al., 2005, ACP)) was used to explicitly compute the diurnal variations of the isotope anomaly of NOx, HOx under several conditions prevailing in the marine boundary layer. ?17<span class="hlt">O</span> was propagated from ozone to other species (NO, NO2, OH, HO2, RO2, NO<span class="hlt">3</span>, N2<span class="hlt">O</span>5, <span class="hlt">HONO</span>, HNO<span class="hlt">3</span>, HNO4, H2<span class="hlt">O</span>2) according to the classical mass-balance equation applied at each time step of the model (30 seconds typically). The model confirms that diurnal variations in ?17<span class="hlt">O</span> of NOx are well predicted by the photochemical steady-state relationship introduced by Michalski et al. (2003, GRL) during the day, but that at night a different approach must be employed (e.g. « fossilization » of the ?17<span class="hlt">O</span> of NOx as soon as the photochemical lifetime of NOx drops below ca. 5 minutes). The model also allows to evaluate the impact on ?17<span class="hlt">O</span> of NOx and nitrate of the frequently made simplifying assumption that ?17<span class="hlt">O</span>(HOx)=0 permil, with and without mass-independent fractionation during the H+<span class="hlt">O</span>2-HO2 reaction. Recommendations for the modeling of ?17<span class="hlt">O</span> of nitrate will be given, based on the extensive model work carried out. In addition, the link between diurnal variations of the ?17<span class="hlt">O</span> of nitrate precursors and seasonal</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.5160Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.5160Y"><span>Simultaneous retrieval of daytime <span class="hlt">O</span>(<span class="hlt">3</span>P) and <span class="hlt">O</span><span class="hlt">3</span> concentrations in the altitude interval 80 - 100 km.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yankovsky, Valentine; Manuilova, Rada; Koval, Andrey</p> <p>2017-04-01</p> <p>We propose methods of simultaneously independent retrievals of the key components of Mesosphere and Lower Thermosphere (MLT) [<span class="hlt">O</span><span class="hlt">3</span>] and [<span class="hlt">O</span>(<span class="hlt">3</span>P)]. The altitude profile of ozone concentration, [<span class="hlt">O</span><span class="hlt">3</span>], can be measured by direct method of the measurement of absorbing radiation from the Sun or the stars in the UV range of the spectrum. However, this method is most often realized in twilight. Retrieval of daytime [<span class="hlt">O</span><span class="hlt">3</span>] depends on a prior information about the <span class="hlt">O</span>(<span class="hlt">3</span>P) altitude profile. Vice versa, atomic oxygen concentration, [<span class="hlt">O</span>(<span class="hlt">3</span>P)], is usually retrieved from the measured values of [<span class="hlt">O</span><span class="hlt">3</span>]. The problem of independent and simultaneous retrieval of [<span class="hlt">O</span><span class="hlt">3</span>] and [<span class="hlt">O</span>(<span class="hlt">3</span>P)] can be solved by using individual proxy for each of the target component. Using a sensitivity study and uncertainty analysis of the contemporary model of <span class="hlt">O</span><span class="hlt">3</span> and <span class="hlt">O</span>2 photolysis in the MLT, YM2011, we determined that populations of three excited electronic-vibrational levels <span class="hlt">O</span>2(b1, v = 0, 1, 2) and of metastable <span class="hlt">O</span>(1D) atom depend on [<span class="hlt">O</span>(<span class="hlt">3</span>P)] and [<span class="hlt">O</span><span class="hlt">3</span>] concentrations. For [<span class="hlt">O</span>(<span class="hlt">3</span>P)] retrieval the following transitions should be used: <span class="hlt">O</span>2(b1, v') -> <span class="hlt">O</span>2(X<span class="hlt">3</span>, v") which produce emissions: (a) at 780.4 nm in the band (v' = 2, v" = 2) and at 697.0 nm in the band (2, 1) with the uncertainty of retrieval smaller than 30% for the whole altitude range 80 - 100 km; (b) at 771.0 nm in the band (1, 1), 688.4 nm in the band (1, 0) and at 874.4 nm in the band (1, 2) with the uncertainty of retrieval about 30% above 90 km. For [<span class="hlt">O</span><span class="hlt">3</span>] retrieval the following transitions should be used: <span class="hlt">O</span>2(b1, v') -> <span class="hlt">O</span>2(X<span class="hlt">3</span>, v") which produce emissions: (c) at 762.1 nm in the band (0, 0) and at 864.7 nm in the band (0, 1) with the uncertainty of retrieval 20 - 25% for the altitude range 80 - 85 km and smaller than 20% in the interval 85 - 95 km; (d) in the line of <span class="hlt">O</span>(1D) 630.0 nm with the uncertainty of retrieval 10 - 15% in the interval 80 - 95 km. Above 95 km the uncertainty of [<span class="hlt">O</span><span class="hlt">3</span>] retrieval grows and reaches up to 80% at 100 km for all suggested proxies. For</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MARV30010V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MARV30010V"><span>Quantum Oscillations at LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> Interfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Veit, Michael; Suzuki, Yuri</p> <p></p> <p>Emergent metallic behavior at the interface of the Mott insulator LaTi<span class="hlt">O</span><span class="hlt">3</span> and the band insulator SrTi<span class="hlt">O</span><span class="hlt">3</span> was observed for the first time more than a decade ago. Since then the metallicity has been explained in terms of charge redistribution at the interface combined with lattice relaxation. However to date, Shubnikov de Haas oscillations have not been reported in this two dimensional metallic system. For ultrathin (<span class="hlt">3</span>-4 unit cells) LaTi<span class="hlt">O</span><span class="hlt">3</span> thin films on SrTi<span class="hlt">O</span><span class="hlt">3</span>, we report the observation of Shubnikov-de Haas oscillations whose frequency corresponds to a small Fermi pocket. Surprisingly the oscillation are only observed between 1 and 4 T. Above this range, the quantum limit is reached for this pocket so no more oscillations are observed. A Berry's phase of π is also detected in these oscillations. Additionally a strong in-plane anisotropic magnetoresistance was measured in the heterostructures which, along with the Berry's phase, is attributed to a giant Rashba coupling at the interface. This work is funded by a National Security Science Engineering Faculty Fellowship of the Department of Defense under N00014-15-1-0045.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27506152','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27506152"><span>Optical characterization of Tm(<span class="hlt">3</span>+) doped Bi2<span class="hlt">O</span><span class="hlt">3</span>-Ge<span class="hlt">O</span>2-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses in absence and presence of BaF2.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Han, Kexuan; Zhang, Peng; Wang, Shunbin; Guo, Yanyan; Zhou, Dechun; Yu, Fengxia</p> <p>2016-08-10</p> <p>In this paper, Two new Bi2<span class="hlt">O</span><span class="hlt">3</span>-Ge<span class="hlt">O</span>2-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses (one presence of BaF2) doped with 1mol% Tm2<span class="hlt">O</span><span class="hlt">3</span> were prepared by melt-quenching technique. Differential thermal analysis (DTA), the absorption, Raman, IR spectra and fluorescence spectra were measured. The Judd-Ofelt intensity parameters, emission cross section, absorption cross section, and gain coefficient of Tm(<span class="hlt">3</span>+) ions were comparatively investigated. After the BaF2 introduced, the glass showed a better thermal stability, lower phonon energy and weaker OH(-) absorption coefficient, meanwhile, a larger ~1.8 μm emission cross section σem (7.56 × 10(-21) cm(2)) and a longer fluorescence lifetime τmea (2.25 ms) corresponding to the Tm(<span class="hlt">3</span>+): (4)F<span class="hlt">3</span> → (<span class="hlt">3</span>)H6 transition were obtained, which is due to the addition of fluoride in glass could reduce the quenching rate of hydroxyls and raise the cross-relaxation ((<span class="hlt">3</span>)H6 + (<span class="hlt">3</span>)H4 → (<span class="hlt">3</span>)F4 + (<span class="hlt">3</span>)F4) rate. Our results suggest that the Tm(<span class="hlt">3</span>+) doped Bi2<span class="hlt">O</span><span class="hlt">3</span>-Ge<span class="hlt">O</span>2-Ga2<span class="hlt">O</span><span class="hlt">3</span> glass with BaF2 might be potential to the application in efficient ~1.8 μm lasers system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4979014','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4979014"><span>Optical characterization of Tm<span class="hlt">3</span>+ doped Bi2<span class="hlt">O</span><span class="hlt">3</span>-Ge<span class="hlt">O</span>2-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses in absence and presence of BaF2</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Han, Kexuan; Zhang, Peng; Wang, Shunbin; Guo, Yanyan; Zhou, Dechun; Yu, Fengxia</p> <p>2016-01-01</p> <p>In this paper, Two new Bi2<span class="hlt">O</span><span class="hlt">3</span>-Ge<span class="hlt">O</span>2-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses (one presence of BaF2) doped with 1mol% Tm2<span class="hlt">O</span><span class="hlt">3</span> were prepared by melt-quenching technique. Differential thermal analysis (DTA), the absorption, Raman, IR spectra and fluorescence spectra were measured. The Judd–Ofelt intensity parameters, emission cross section, absorption cross section, and gain coefficient of Tm<span class="hlt">3</span>+ ions were comparatively investigated. After the BaF2 introduced, the glass showed a better thermal stability, lower phonon energy and weaker OH− absorption coefficient, meanwhile, a larger ~1.8 μm emission cross section σem (7.56 × 10−21 cm2) and a longer fluorescence lifetime τmea (2.25 ms) corresponding to the Tm<span class="hlt">3</span>+: 4F<span class="hlt">3</span> → <span class="hlt">3</span>H6 transition were obtained, which is due to the addition of fluoride in glass could reduce the quenching rate of hydroxyls and raise the cross-relaxation (<span class="hlt">3</span>H6 + <span class="hlt">3</span>H4 → <span class="hlt">3</span>F4 + <span class="hlt">3</span>F4) rate. Our results suggest that the Tm<span class="hlt">3</span>+ doped Bi2<span class="hlt">O</span><span class="hlt">3</span>-Ge<span class="hlt">O</span>2-Ga2<span class="hlt">O</span><span class="hlt">3</span> glass with BaF2 might be potential to the application in efficient ~1.8 μm lasers system. PMID:27506152</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4555181','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4555181"><span>Trends in (LaMn<span class="hlt">O</span><span class="hlt">3</span>)n/(SrTi<span class="hlt">O</span><span class="hlt">3</span>)m superlattices with varying layer thicknesses</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jilili, J.; Cossu, F.; Schwingenschlögl, U.</p> <p>2015-01-01</p> <p>We investigate the thickness dependence of the structural, electronic, and magnetic properties of (LaMn<span class="hlt">O</span><span class="hlt">3</span>)n/(SrTi<span class="hlt">O</span><span class="hlt">3</span>)m (n, m = 2, 4, 6, 8) superlattices using density functional theory. The electronic structure turns out to be highly sensitive to the onsite Coulomb interaction. In contrast to bulk SrTi<span class="hlt">O</span><span class="hlt">3</span>, strongly distorted <span class="hlt">O</span> octahedra are observed in the SrTi<span class="hlt">O</span><span class="hlt">3</span> layers with a systematic off centering of the Ti atoms. The systems favour ferromagnetic spin ordering rather than the antiferromagnetic spin ordering of bulk LaMn<span class="hlt">O</span><span class="hlt">3</span> and all show half-metallicity, while a systematic reduction of the minority spin band gaps as a function of the LaMn<span class="hlt">O</span><span class="hlt">3</span> and SrTi<span class="hlt">O</span><span class="hlt">3</span> layer thicknesses originates from modifications of the Ti dxy states. PMID:26323361</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012APS..MAR.T9001F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012APS..MAR.T9001F"><span>Upper limit to magnetism in LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fitzsimmons, Michael</p> <p>2012-02-01</p> <p>In 2004 Ohtomo and Hwang reported unusually high conductivity in LaAl<span class="hlt">O</span><span class="hlt">3</span> and SrTi<span class="hlt">O</span><span class="hlt">3</span> bilayer samples. Since then, metallic conduction, superconductivity, magnetism, and coexistence of superconductivity and ferromagnetism have been attributed to LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> interfaces. Very recently, two studies have reported large magnetic moments attributed to interfaces from measurement techniques that are unable to distinguish between interfacial and bulk magnetism. Consequently, it is imperative to perform magnetic measurements that by being intrinsically sensitive to interface magnetism are impervious to experimental artifacts suffered by bulk measurements. Using polarized neutron reflectometry, we measured the neutron spin dependent reflectivity from four LaAl<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> superlattices. Our results indicate the upper limit for the magnetization averaged over the lateral dimensions of the sample induced by an 11 T magnetic field at 1.7 K is less than 2 G. SQUID magnetometry of the neutron superlattice samples sporadically finds an enhanced moment (consistent with past reports), possibly due to experimental artifacts. These observations set important restrictions on theories which imply a strongly enhanced magnetism at the interface between LaAl<span class="hlt">O</span><span class="hlt">3</span> and SrTi<span class="hlt">O</span><span class="hlt">3</span>. Work performed in collaboration with N.W. Hengartner, S. Singh, M. Zhernenkov (LANL), F.Y. Bruno, J. Santamaria (Universidad Complutense de Madrid), A. Brinkman, M.J.A. Huijben, H. Molegraaf (MESA+ Institute for Nanotechnology), J. de la Venta and Ivan K. Schuller (UCSD). [4pt] Work supported by the Office of Basic Energy Science, U.S. Department of Energy, BES-DMS and DMR under grant DE FG03-87ER-45332. Work at UCM is supported by Consolider Ingenio CSD2009-00013 (IMAGINE), CAM S2009-MAT 1756 (PHAMA) and work at Twente is supported by the Foundation for Fundamental Research on Matter (FOM).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MRE.....5c5204L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MRE.....5c5204L"><span>EPR and FTIR spectroscopic studies of MO-Al2<span class="hlt">O</span><span class="hlt">3</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Mn<span class="hlt">O</span>2(M = Pb, Zn and Cd) glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lalitha Phani, A. V.; Sekhar, K. Chandra; Chakradhar, R. P. S.; Narasimha Chary, M.; Shareefuddin, Md</p> <p>2018-03-01</p> <p>Glasses of the system (30-x)MO-xAl2<span class="hlt">O</span><span class="hlt">3</span>-15Bi2<span class="hlt">O</span><span class="hlt">3</span>-54.5B2<span class="hlt">O</span><span class="hlt">3</span>-0.5Mn<span class="hlt">O</span>2 [M = Pb, Zn & Cd] (x = 0, 5, 10 & 15 mol%) were prepared by the normal melt quenching method. The amorphous nature of the prepared glasses was confirmed by the XRD studies. The EPR and FTIR studies were carried out at room temperature (RT). The EPR spectra exhibited three resonance signals at g ≈ 2.0 with a hyperfine structure, an absorption around g = 4.<span class="hlt">3</span> and a distinct shoulder at g = <span class="hlt">3.3</span>. Deconvoluted spectra were drawn for g ≈ 2.0 to resolve the six hyperfine lines. The electron paramagnetic resonance signal at g ≈ 2.0 indicates that the Mn2+ ions are in nearly perfectly octahedral symmetry. The low field signals at g = <span class="hlt">3.3</span> and g = 4.<span class="hlt">3</span> are attributed to the Mn2+ ion which are in distorted rhombic symmetries. The hyperfine (HF) splitting constant (A) values suggested that the bonding between Mn2+ ions and its ligands is ionic in nature. The presence of BO<span class="hlt">3</span> and BO4 borate units, metal oxide cation units, Mn2+ and Bi-<span class="hlt">O</span> bond vibrations in Bi<span class="hlt">O</span><span class="hlt">3</span> units were noticed from the FTIR spectra.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009PhRvB..80l5115Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009PhRvB..80l5115Y"><span>Electronic and magnetic properties of RMn<span class="hlt">O</span><span class="hlt">3</span>/AMn<span class="hlt">O</span><span class="hlt">3</span> heterostructures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Rong; Yunoki, Seiji; Dong, Shuai; Dagotto, Elbio</p> <p>2009-09-01</p> <p>The ground-state properties of RMn<span class="hlt">O</span><span class="hlt">3</span>/AMn<span class="hlt">O</span><span class="hlt">3</span> (RMO/AMO) heterostructures (with R=La,Pr,… , a trivalent rare-earth cation and A=Sr,Ca,… , a divalent alkaline cation) are studied using a two-orbital double-exchange model including the superexchange coupling and Jahn-Teller lattice distortions. To describe the charge transfer across the interface, the long-range Coulomb interaction is taken into account at the mean-field level, by self-consistently solving the Poisson’s equation. The calculations are carried out numerically on finite clusters. We find that the state stabilized near the interface of the heterostructure is similar to the state of the bulk compound (R,A)MO at electronic density close to 0.5. For instance, a charge and orbitally ordered CE state is found at the interface if the corresponding bulk (R,A)MO material is a narrow-to-intermediate bandwidth manganite. But instead the interface regime accommodates an A-type antiferromagnetic state with a uniform x2-y2 orbital order, if the bulk (R,A)MO corresponds to a wide bandwidth manganite. We argue that these results explain some of the properties of long-period (RMO)m/(AMO)n superlattices, such as (PrMn<span class="hlt">O</span><span class="hlt">3</span>)m/(CaMn<span class="hlt">O</span><span class="hlt">3</span>)n and (LaMn<span class="hlt">O</span><span class="hlt">3</span>)m/(SrMn<span class="hlt">O</span><span class="hlt">3</span>)n . We also remark that the intermediate states in between the actual interface and the bulklike regimes of the heterostructure are dependent on the bandwidth and the screening of the Coulomb interaction. In these regions of the heterostructures, states are found that do not have an analog in experimentally known bulk phase diagrams. These new states of the heterostructures provide a natural interpolation between magnetically ordered states that are stable in the bulk at different electronic densities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5111568','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5111568"><span>Properties of epitaxial, (001)- and (110)-oriented (PbMg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span><span class="hlt">O</span><span class="hlt">3)2/3</span>-(PbTi<span class="hlt">O</span><span class="hlt">3)1/3</span> films on silicon described by polarization rotation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Boota, Muhammad; Houwman, Evert P.; Dekkers, Matthijn; Nguyen, Minh D.; Vergeer, Kurt H.; Lanzara, Giulia; Koster, Gertjan; Rijnders, Guus</p> <p>2016-01-01</p> <p>Abstract Epitaxial (PbMg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span><span class="hlt">O</span><span class="hlt">3)2/3</span>-(PbTi<span class="hlt">O</span><span class="hlt">3)1/3</span> (PMN-PT) films with different out-of-plane orientations were prepared using a Ce<span class="hlt">O</span>2/yttria stabilized Zr<span class="hlt">O</span>2 bilayer buffer and symmetric SrRu<span class="hlt">O</span><span class="hlt">3</span> electrodes on silicon substrates by pulsed laser deposition. The orientation of the SrRu<span class="hlt">O</span><span class="hlt">3</span> bottom electrode, either (110) or (001), was controlled by the deposition conditions and the subsequent PMN-PT layer followed the orientation of the bottom electrode. The ferroelectric, dielectric and piezoelectric properties of the (SrRu<span class="hlt">O</span><span class="hlt">3</span>/PMN-PT/SrRu<span class="hlt">O</span><span class="hlt">3</span>) ferroelectric capacitors exhibit orientation dependence. The properties of the films are explained in terms of a model based on polarization rotation. At low applied fields domain switching dominates the polarization change. The model indicates that polarization rotation is easier in the (110) film, which is ascribed to a smaller effect of the clamping on the shearing of the pseudo-cubic unit cell compared to the (001) case. PMID:27877857</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.853a2045D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.853a2045D"><span>The effect of Cr2<span class="hlt">O</span><span class="hlt">3</span> doping on structures and dielectric constants of Si<span class="hlt">O</span>2-Bi2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span>-Na2CO<span class="hlt">3</span> glass based on silica gel of natural sand</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Diantoro, M.; Zaini, M. B.; Muniroh, Z.; Nasikhudin; Hidayat, A.</p> <p>2017-05-01</p> <p>One of the abundant natural resources along the coastal lines of Indonesia is silica sand. One of the beaches which has a lot of silica content is Bancar-Tuban beach. Silica can be used as a raw material of glass that has multiple properties in optic, dielectric, and other physical properties by introducing specific dopants. Some oxides have been used as dopant e.g. Al2<span class="hlt">O</span><span class="hlt">3</span>, Fe<span class="hlt">3</span><span class="hlt">O</span>4, and Ni<span class="hlt">O</span>. However, there has not been any comprehensive study discussing the multiple properties of natural silica-sand-based glass with Cr2<span class="hlt">O</span><span class="hlt">3</span> dopant so far. A series of samples have been prepared, which mean two solid steps to state melting technique. Cr2<span class="hlt">O</span><span class="hlt">3</span> was selected as a dopant due to its potential to control its color and to increase the dielectric constant of the glass. The synthesis of silica (Si<span class="hlt">O</span>2) sand from BancarTuban beach was conducted through the sol-gel process. The composition varied as the addition of Cr2<span class="hlt">O</span><span class="hlt">3</span>on 50Si<span class="hlt">O</span>2-25B2<span class="hlt">O</span><span class="hlt">3</span>-(6.5-x) Bi2<span class="hlt">O</span><span class="hlt">3</span>-18.5Na2CO<span class="hlt">3</span>-xCr2<span class="hlt">O</span><span class="hlt">3</span> (x = 0, 0.02, 0.04, 0.06 and 0.08mol), later called SBBN glass. The samples’ characterizations of the structure and morphology were conducted through the use of XRD, and SEM-EDX. The measurements were done by using a DC capacitance meter in order to investigate the dielectric properties of the sample, under the influence of light. It is shown that addition of Cr2<span class="hlt">O</span><span class="hlt">3</span> did not alter the crystal structure but changed the structure of the functional bond formation. It is also revealed that the dielectric constant increased along with the increasing of Cr2<span class="hlt">O</span><span class="hlt">3</span>. An interesting result was that the dielectric constant of the glass was quantized decreasingly as the increase of light.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1360035-strain-dependence-antiferromagnetic-interface-coupling-la0-srruo3-superlattices','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1360035-strain-dependence-antiferromagnetic-interface-coupling-la0-srruo3-superlattices"><span>Strain dependence of antiferromagnetic interface coupling in La 0.7Sr 0.<span class="hlt">3</span>Mn<span class="hlt">O</span> <span class="hlt">3</span>/SrRu<span class="hlt">O</span> <span class="hlt">3</span> superlattices</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Das, Sujit; Herklotz, Andreas; Pippel, Eckhard; ...</p> <p>2015-04-06</p> <p>We have investigated the magnetic response of La 0.7Sr 0.<span class="hlt">3</span>Mn<span class="hlt">O</span> <span class="hlt">3</span>/SrRu<span class="hlt">O</span> <span class="hlt">3</span> superlattices to biaxial in-plane strain applied in situ. Superlattices grown on piezoelectric substrates of 0.72PbMg 1/<span class="hlt">3</span>Nb 2/<span class="hlt">3</span><span class="hlt">O</span> <span class="hlt">3</span>-0.28PbTi<span class="hlt">O</span> <span class="hlt">3</span>(001) (PMN-PT) show strong antiferromagnetic coupling of the two ferromagnetic components. The coupling field of mu H-0(AF) = 1.8 T is found to change by mu(0)Delta H-AF/Delta epsilon similar to -520 mT %(-1) under reversible biaxial strain Delta epsilon at 80 K in a [La 0.7Sr 0.<span class="hlt">3</span>Mn<span class="hlt">O</span> <span class="hlt">3</span>(22 angstrom)/SrRu<span class="hlt">O</span> <span class="hlt">3</span>(55 angstrom)] 15 superlattice. This reveals a significant strain effect on interfacial coupling. The applied in-plane compression enhances the ferromagnetic ordermore » in the manganite layers, which are under as-grown tensile strain, leading to a larger net coupling of SrRu<span class="hlt">O</span> <span class="hlt">3</span> layers at the interface. It is thus difficult to disentangle the contributions from strain-dependent antiferromagnetic Mn-<span class="hlt">O</span>-Ru interface coupling and Mn-<span class="hlt">O</span>-Mn ferromagnetic double exchange near the interface for the strength of the apparent antiferromagnetic coupling. We discuss our results in the framework of available models.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1240697-octonary-resistance-states-la0-batio3-la0-multiferroic-tunnel-junctions','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1240697-octonary-resistance-states-la0-batio3-la0-multiferroic-tunnel-junctions"><span>Octonary resistance states in La 0.7Sr 0.<span class="hlt">3</span>Mn<span class="hlt">O</span> <span class="hlt">3</span>/BaTi<span class="hlt">O</span> <span class="hlt">3</span>/La 0.7Sr 0.<span class="hlt">3</span>Mn<span class="hlt">O</span> <span class="hlt">3</span> multiferroic tunnel junctions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Yue -Wei Yin; Tao, Jing; Huang, Wei -Chuan; ...</p> <p>2015-10-06</p> <p>General drawbacks of current electronic/spintronic devices are high power consumption and low density storage. A multiferroic tunnel junction (MFTJ), employing a ferroelectric barrier layer sandwiched between two ferromagnetic layers, presents four resistance states in a single device and therefore provides an alternative way to achieve high density memories. Here, an MFTJ device with eight nonvolatile resistance states by further integrating the design of noncollinear magnetization alignments between the ferromagnetic layers is demonstrated. Through the angle-resolved tunneling magnetoresistance investigations on La 0.7Sr 0.<span class="hlt">3</span>Mn<span class="hlt">O</span> <span class="hlt">3</span>/BaTi<span class="hlt">O</span> <span class="hlt">3</span>/La 0.7Sr 0.<span class="hlt">3</span>Mn<span class="hlt">O</span> <span class="hlt">3</span> junctions, it is found that, besides collinear parallel/antiparallel magnetic configurations, the MFTJ showsmore » at least two other stable noncollinear (45° and 90°) magnetic configurations. As a result, combining the tunneling electroresistance effect caused by the ferroelectricity reversal of the BaTi<span class="hlt">O</span> <span class="hlt">3</span> barrier, an octonary memory device is obtained, representing potential applications in high density nonvolatile storage in the future.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17026128','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17026128"><span>Lattice relaxation in oxide heterostructures: LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> superlattices.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Okamoto, Satoshi; Millis, Andrew J; Spaldin, Nicola A</p> <p>2006-08-04</p> <p>Local density approximation + Hubbard U and many-body effective Hamiltonian calculations are used to determine the effects of lattice relaxation in LaTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span> superlattices. Large ferroelectric-like distortions of the Ti<span class="hlt">O</span>6 octahedra are found, which substantially affect the Ti d-electron density, bringing the calculated results into good agreement with experimental data. The relaxations also change the many-body physics, leading to a novel symmetry-breaking-induced ordering of the xy orbitals, which does not occur in bulk LaTi<span class="hlt">O</span><span class="hlt">3</span>, or in the hypothetical unrelaxed structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12377045','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12377045"><span>Expanding the remarkable structural diversity of uranyl tellurites: hydrothermal preparation and structures of K[UO(2)Te(2)<span class="hlt">O</span>(5)(OH)], Tl(<span class="hlt">3</span>)[(UO(2))(2)[Te(2)<span class="hlt">O</span>(5)(OH)](Te(2)<span class="hlt">O</span>(6))].2H(2)<span class="hlt">O</span>, beta-Tl(2)[UO(2)(Te<span class="hlt">O</span>(<span class="hlt">3</span>))(2)], and Sr(<span class="hlt">3</span>)[UO(2)(Te<span class="hlt">O</span>(<span class="hlt">3</span>))(2)](Te<span class="hlt">O</span>(<span class="hlt">3</span>))(2).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Almond, Philip M; Albrecht-Schmitt, Thomas E</p> <p>2002-10-21</p> <p>The reactions of UO(2)(C(2)H(<span class="hlt">3</span>)<span class="hlt">O</span>(2))(2).2H(2)<span class="hlt">O</span> with K(2)Te<span class="hlt">O</span>(<span class="hlt">3</span>).H(2)<span class="hlt">O</span>, Na(2)Te<span class="hlt">O</span>(<span class="hlt">3</span>) and TlCl, or Na(2)Te<span class="hlt">O</span>(<span class="hlt">3</span>) and Sr(OH)(2).8H(2)<span class="hlt">O</span> under mild hydrothermal conditions yield K[UO(2)Te(2)<span class="hlt">O</span>(5)(OH)] (1), Tl(<span class="hlt">3</span>)[(UO(2))(2)[Te(2)<span class="hlt">O</span>(5)(OH)](Te(2)<span class="hlt">O</span>(6))].2H(2)<span class="hlt">O</span> (2) and beta-Tl(2)[UO(2)(Te<span class="hlt">O</span>(<span class="hlt">3</span>))(2)] (<span class="hlt">3</span>), or Sr(<span class="hlt">3</span>)[UO(2)(Te<span class="hlt">O</span>(<span class="hlt">3</span>))(2)](Te<span class="hlt">O</span>(<span class="hlt">3</span>))(2) (4), respectively. The structure of 1 consists of tetragonal bipyramidal U(VI) centers that are bound by terminal oxo groups and tellurite anions. These UO(6) units span between one-dimensional chains of corner-sharing, square pyramidal Te<span class="hlt">O</span>(4) polyhedra to create two-dimensional layers. Alternating corner-shared oxygen atoms in the tellurium oxide chains are protonated to create short/long bonding patterns. The one-dimensional chains of corner-sharing Te<span class="hlt">O</span>(4) units found in 1 are also present in 2. However, in 2 there are two distinct chains present, one where alternating corner-shared oxygen atoms are protonated, and one where the chains are unprotonated. The uranyl moieties in 2 are bound by five oxygen atoms from the tellurite chains to create seven-coordinate pentagonal bipyramidal U(VI). The structures of <span class="hlt">3</span> and 4 both contain one-dimensional [UO(2)(Te<span class="hlt">O</span>(<span class="hlt">3</span>))(2)](2-) chains constructed from tetragonal bipyramidal U(VI) centers that are bridged by tellurite anions. The chains differ between <span class="hlt">3</span> and 4 in that all of the pyramidal tellurite anions in <span class="hlt">3</span> have the same orientation, whereas the tellurite anions in 4 have opposite orientations on each side of the chain. In 4, there are also additional isolated Te<span class="hlt">O</span>(<span class="hlt">3</span>)(2-) anions present. Crystallographic data: 1, orthorhombic, space group Cmcm, a = 7.9993(5) A, b = 8.7416(6) A, c = 11.4413(8) A, Z = 4; 2, orthorhombic, space group Pbam, a = 10.0623(8) A, b = 23.024(2) A, c = 7.9389(6) A, Z = 4; <span class="hlt">3</span>, monoclinic, space group P2(1)/n, a = 5.4766(4) A, b = 8.2348(6) A, c = 20.849(<span class="hlt">3</span>) A, beta = 92.329(1) degrees, Z = 4; 4, monoclinic, space group C2/c, a = 20.546(1) A, b = 5.6571(<span class="hlt">3</span>) A, c = 13.0979(8) A, beta</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22610675','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22610675"><span>Structure of (Ga2<span class="hlt">O</span><span class="hlt">3</span>)2(Zn<span class="hlt">O</span>)13 and a unified description of the homologous series (Ga2<span class="hlt">O</span><span class="hlt">3</span>)2(Zn<span class="hlt">O</span>)(2n + 1).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Michiue, Yuichi; Kimizuka, Noboru; Kanke, Yasushi; Mori, Takao</p> <p>2012-06-01</p> <p>The structure of (Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>))(2)(Zn<span class="hlt">O</span>)(13) has been determined by a single-crystal X-ray diffraction technique. In the monoclinic structure of the space group C2/m with cell parameters a = 19.66 (4), b = <span class="hlt">3</span>.2487 (5), c = 27.31 (2) Å, and β = 105.9 (1)°, a unit cell is constructed by combining the halves of the unit cell of Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>)(Zn<span class="hlt">O</span>)(6) and Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>)(Zn<span class="hlt">O</span>)(7) in the homologous series Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>)(Zn<span class="hlt">O</span>)(m). The homologous series (Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>))(2)(Zn<span class="hlt">O</span>)(2n + 1) is derived and a unified description for structures in the series is presented using the (<span class="hlt">3</span>+1)-dimensional superspace formalism. The phases are treated as compositely modulated structures consisting of two subsystems. One is constructed by metal ions and another is by <span class="hlt">O</span> ions. In the (<span class="hlt">3</span> + 1)-dimensional model, displacive modulations of ions are described by the asymmetric zigzag function with large amplitudes, which was replaced by a combination of the sawtooth function in refinements. Similarities and differences between the two homologous series (Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>))(2)(Zn<span class="hlt">O</span>)(2n + 1) and Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>)(Zn<span class="hlt">O</span>)(m) are clarified in (<span class="hlt">3</span> + 1)-dimensional superspace. The validity of the (<span class="hlt">3</span> + 1)-dimensional model is confirmed by the refinements of (Ga(2)<span class="hlt">O</span>(<span class="hlt">3</span>))(2)(Zn<span class="hlt">O</span>)(13), while a few complex phenomena in the real structure are taken into account by modifying the model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AIPC.1756b0007R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AIPC.1756b0007R"><span>Microwave dielectric properties of CaCu<span class="hlt">3</span>Ti4<span class="hlt">O</span>12-Al2<span class="hlt">O</span><span class="hlt">3</span> composite</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rahman, Mohd Fariz Ab; Abu, Mohamad Johari; Karim, Saniah Ab; Zaman, Rosyaini Afindi; Ain, Mohd Fadzil; Ahmad, Zainal Arifin; Mohamed, Julie Juliewatty</p> <p>2016-07-01</p> <p>(1-x)CaCu<span class="hlt">3</span>Ti4<span class="hlt">O</span>12 + (x)Al2<span class="hlt">O</span><span class="hlt">3</span> composite (0 ≤ x ≤0.25) was prepared via conventional solid-state reaction method. The fabrication of sample was started with synthesizing stoichiometric CCTO from CaCO<span class="hlt">3</span>, Cu<span class="hlt">O</span> and Ti<span class="hlt">O</span>2 powders, then wet-mixed in deionized water for 24 h. The process was continued with calcined CCTO powder at 900 °C for 12 h before sintered at 1040 °C for 10 h. Next, the calcined CCTO powder with different amount of Al2<span class="hlt">O</span><span class="hlt">3</span> were mixed for 24 h, then palletized and sintered at 1040 °C for 10. X-ray diffraction analysis on the sintered samples showed that CCTO powder was in a single phase, meanwhile the trace of secondary peaks which belong to CaAl2<span class="hlt">O</span>4 and Corundum (Al2<span class="hlt">O</span><span class="hlt">3</span>) could be observed in the other samples Scanning electron microscopy analysis showed that the grain size of the sample is firstly increased with addition of Al2<span class="hlt">O</span><span class="hlt">3</span> (x = 0.01), then become smaller with the x > 0.01. Microwave dielectric properties showed that the addition of Al2<span class="hlt">O</span><span class="hlt">3</span> (x = 0.01) was remarkably reduced the dielectric loss while slightly increased the dielectric permittivity. However, further addition of Al2<span class="hlt">O</span><span class="hlt">3</span> was reduced both dielectric loss and permittivity at least for an order of magnitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018RaPC..145...26S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018RaPC..145...26S"><span>Investigation on gamma and neutron radiation shielding parameters for Ba<span class="hlt">O</span>/SrO‒Bi2<span class="hlt">O</span><span class="hlt">3</span>‒B2<span class="hlt">O</span><span class="hlt">3</span> glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sayyed, M. I.; Lakshminarayana, G.; Dong, M. G.; Ersundu, M. Çelikbilek; Ersundu, A. E.; Kityk, I. V.</p> <p>2018-04-01</p> <p>In this work, mass attenuation coefficients (μ/ρ), effective atomic number (Zeff), electron density (Ne), mean free path (MFP), and half-value layer (HVL) of 20 Ba<span class="hlt">O</span>/SrO‒(x) Bi2<span class="hlt">O</span><span class="hlt">3</span>‒(80‒x) B2<span class="hlt">O</span><span class="hlt">3</span> glasses (where x=10, 20, 30, 40, 50 and 60 mol%) were calculated using WinXCom program and MCNP5 code. The obtained (μ/ρ) results using both MCNP5 code and WinXCom program were in good agreement. It is found that the addition of Bi2<span class="hlt">O</span><span class="hlt">3</span> leads to increase the Zeff values in both Ba<span class="hlt">O</span>/SrO‒Bi2<span class="hlt">O</span><span class="hlt">3</span>‒B2<span class="hlt">O</span><span class="hlt">3</span> glass systems. However, the Zeff values of the BaO‒Bi2<span class="hlt">O</span><span class="hlt">3</span>‒B2<span class="hlt">O</span><span class="hlt">3</span> glass system are higher than those of the SrO‒Bi2<span class="hlt">O</span><span class="hlt">3</span>‒B2<span class="hlt">O</span><span class="hlt">3</span> glasses. The fast neutrons effective removal cross sections (ΣR) for 20 SrO‒40 Bi2<span class="hlt">O</span><span class="hlt">3</span>‒40 B2<span class="hlt">O</span><span class="hlt">3</span> glass is the highest among all studied glasses. The calculated half-value layer values were compared with different glass systems and it was found that the shielding properties of the selected glasses are comparable or even better than other glass systems such as phosphate glasses.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008APS..MARA31012L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008APS..MARA31012L"><span>Structure and properties of CaMn<span class="hlt">O</span><span class="hlt">3</span>/SrMn<span class="hlt">O</span><span class="hlt">3</span>/BaMn<span class="hlt">O</span><span class="hlt">3</span> superlattices from first principles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Shen; Oh, Seongshik; Rabe, Karin</p> <p>2008-03-01</p> <p>Previous theoretical and experimental studies have shown that three-component, or ``tri-color'' superlattices can exhibit intrinsic electric polarization due to inversion-symmetry breaking in the layer sequence. In ferromagnetic inversion-symmetry-breaking superlattices, controlled symmetry lowering is similarly expected to lead to interesting new and tunable properties. Here, we present results of first-principles density-functional-theory calculations for short-period CaMn<span class="hlt">O</span><span class="hlt">3</span>/SrMn<span class="hlt">O</span><span class="hlt">3</span>/BaMn<span class="hlt">O</span><span class="hlt">3</span> superlattices, using VASP. The ground state structure, magnetic ordering, polarization and dielectric response will be presented. The role of epitaxial strain in the individual layers and the role of layer sequence will be explored. Connections to experimental studies and prospects for future work will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20783285-addition-nh-sub-al-sub-sub-sup','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20783285-addition-nh-sub-al-sub-sub-sup"><span>Addition of NH{sub <span class="hlt">3</span>} to Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub <span class="hlt">3</span>}{sup -}</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wyrwas, Richard B.; Jarrold, Caroline Chick; Das, Ujjal</p> <p>2006-05-28</p> <p>Recent computational studies on the addition of ammonia (NH{sub <span class="hlt">3</span>}) to the Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub <span class="hlt">3</span>}{sup -} cluster anion [A. Guevara-Garcia, A. Martinez, and J. V. Ortiz, J. Chem. Phys. 122, 214309 (2005)] have motivated experimental and additional computational studies, reported here. Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub <span class="hlt">3</span>}{sup -} is observed to react with a single NH{sub <span class="hlt">3</span>} molecule to form the Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub <span class="hlt">3</span>}NH{sub <span class="hlt">3</span>}{sup -} ion in mass spectrometric studies. This is in contrast to similarly performed studies with water, in which the Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 5}H{sub 4}{sup -} product was highly favored. However, the anion PE spectrum of the ammoniated species ismore » very similar to that of Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub 4}H{sub 2}{sup -}. The adiabatic electron affinity of Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub <span class="hlt">3</span>}NH{sub <span class="hlt">3</span>} is determined to be 2.35(5) eV. Based on comparison between the spectra and calculated electron affinities, it appears that NH{sub <span class="hlt">3</span>} adds dissociatively to Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub <span class="hlt">3</span>}{sup -}, suggesting that the time for the Al{sub <span class="hlt">3</span>}<span class="hlt">O</span>{sub <span class="hlt">3</span>}{sup -}{center_dot}NH{sub <span class="hlt">3</span>} complex to either overcome or tunnel through the barrier to proton transfer (which is higher for NH{sub <span class="hlt">3</span>} than for water) is short relative to the time for collisional cooling in the experiment.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016RJPCA..90...65M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016RJPCA..90...65M"><span>Catalytic performance of V2<span class="hlt">O</span>5-Mo<span class="hlt">O</span><span class="hlt">3</span>/γ-Al2<span class="hlt">O</span><span class="hlt">3</span> catalysts for partial oxidation of n-hexane1</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mahmoudian, R.; Khodadadi, Z.; Mahdavi, Vahid; Salehi, Mohammed</p> <p>2016-01-01</p> <p>In the current study, a series of V2<span class="hlt">O</span>5-Mo<span class="hlt">O</span><span class="hlt">3</span> catalyst supported on γ-Al2<span class="hlt">O</span><span class="hlt">3</span> with various V2<span class="hlt">O</span>5 and Mo<span class="hlt">O</span><span class="hlt">3</span> loadings was prepared by wet impregnation technique. The characterization of prepared catalysts includes BET surface area, powder X-ray diffraction (XRD), and oxygen chemisorptions. The partial oxidation of n-hexane by air over V2<span class="hlt">O</span>5-Mo<span class="hlt">O</span><span class="hlt">3</span>/γ-Al2<span class="hlt">O</span><span class="hlt">3</span> catalysts was carried out under flow condition in a fixed bed glass reactor. The effect of V2<span class="hlt">O</span>5 loading, temperature, Mo<span class="hlt">O</span><span class="hlt">3</span> loading, and n-hexane LHSV on the n-hexane conversion and the product selectivity were investigated. The partial oxygenated products of n-hexane oxidation were ethanol, acetic anhydride, acetic acid, and acetaldehyde. The 10% V2<span class="hlt">O</span>5-1%Mo<span class="hlt">O</span><span class="hlt">3</span>/γ-Al2<span class="hlt">O</span><span class="hlt">3</span> was found in most active and selective catalyst during partial oxidation of n-hexane. The results indicated that by increasing the temperature, the n-hexane conversion increases as well, although the selectivity of the products passes through a maximum by increasing the temperature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhCS.992a2040S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhCS.992a2040S"><span>Bi12Ti<span class="hlt">O</span>20 crystallization in a Bi2<span class="hlt">O</span><span class="hlt">3</span>-Ti<span class="hlt">O</span>2-Si<span class="hlt">O</span>2-Nd2<span class="hlt">O</span><span class="hlt">3</span> system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Slavov, S.; Jiao, Z.</p> <p>2018-03-01</p> <p>Polycrystalline mono-phase bismuth titanate was produced by free cooling from melts heated to 1170 °C. The control over the initial amounts in the starting compositions in the system Bi2<span class="hlt">O</span><span class="hlt">3</span>/Ti<span class="hlt">O</span>2/Si<span class="hlt">O</span>2/Nd2<span class="hlt">O</span><span class="hlt">3</span> and over the thermal gradient of the heat process resulted in the formation of specific structures and microstructures of monophase sillenite ceramics. The main phase Bi12Ti<span class="hlt">O</span>20 belongs to the amorphous network groups based on oxides of silicon, bismuth and titanium. In this work, we demonstrated a way to control the crystalline and amorphous phase formation in bulk poly-crystalline materials in the selected system.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JSSCh.258..776F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JSSCh.258..776F"><span>Ba<span class="hlt">3</span>CuOs2<span class="hlt">O</span>9 and Ba<span class="hlt">3</span>ZnOs2<span class="hlt">O</span>9, a comparative study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Feng, Hai L.; Jansen, Martin</p> <p>2018-02-01</p> <p>Polycrystalline samples of Ba<span class="hlt">3</span>CuOs2<span class="hlt">O</span>9 and Ba<span class="hlt">3</span>ZnOs2<span class="hlt">O</span>9 were synthesized by solid-state reactions. Ba<span class="hlt">3</span>CuOs2<span class="hlt">O</span>9 crystallizes in Cmcm, while Ba<span class="hlt">3</span>ZnOs2<span class="hlt">O</span>9 adopts the hexagonal space group P63/mmc. Both the crystal structures consist of face-sharing Os-centered octahedra forming dimer-like Os2<span class="hlt">O</span>9 units, which are interconnected by corner-sharing Cu<span class="hlt">O</span>6, or Zn<span class="hlt">O</span>6 octahedra, respectively. In Ba<span class="hlt">3</span>CuOs2<span class="hlt">O</span>9, the Cu<span class="hlt">O</span>6 octahedra show a characteristic Jahn-Teller distortion. Both, Ba<span class="hlt">3</span>CuOs2<span class="hlt">O</span>9 and Ba<span class="hlt">3</span>ZnOs2<span class="hlt">O</span>9, are electrically insulating. Magnetic and specific heat measurements confirm that Ba<span class="hlt">3</span>CuOs2<span class="hlt">O</span>9 is antiferromagnetically ordered below 47 K. Analysis of the magnetic data indicated that its magnetic properties are dominated by Cu2+ ions. The magnetic susceptibility of Ba<span class="hlt">3</span>ZnOs2<span class="hlt">O</span>9 is weakly temperature-dependent with a broad maximum ≈ 280 K, indicating the presence of strong exchange interactions within the Os2<span class="hlt">O</span>9 dimer. The residual magnetic susceptibility at low temperatures also suggests the presence of appreciable exchange coupling between the dimers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20090024811&hterms=isoprene&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Disoprene','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20090024811&hterms=isoprene&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Disoprene"><span>Formaldehyde Distribution over North America: Implications for Satellite Retrievals of Formaldehyde Columns and Isoprene Emission</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Millet, Dylan B.; Jacob, Daniel J.; Turquety, Solene; Hudman, Rynda C.; Wu, Shiliang; Anderson, Bruce E.; Fried, Alan; Walega, James; Heikes, Brian G.; Blake, Donald R.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20090024811'); toggleEditAbsImage('author_20090024811_show'); toggleEditAbsImage('author_20090024811_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20090024811_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20090024811_hide"></p> <p>2006-01-01</p> <p>Formaldehyde (<span class="hlt">HCHO</span>) columns measured from space provide constraints on emissions of volatile organic compounds (VOCs). Quantitative interpretation requires characterization of errors in <span class="hlt">HCHO</span> column retrievals and relating these columns to VOC emissions. Retrieval error is mainly in the air mass factor (AMF) which relates fitted backscattered radiances to vertical columns and requires external information on <span class="hlt">HCHO</span>, aerosols, and clouds. Here we use aircraft data collected over North America and the Atlantic to determine the local relationships between <span class="hlt">HCHO</span> columns and VOC emissions, calculate AMFs for <span class="hlt">HCHO</span> retrievals, assess the errors in deriving AMFs with a chemical transport model (GEOS-Chem), and draw conclusions regarding space-based mapping of VOC emissions. We show that isoprene drives observed <span class="hlt">HCHO</span> column variability over North America; <span class="hlt">HCHO</span> column data from space can thus be used effectively as a proxy for isoprene emission. From observed <span class="hlt">HCHO</span> and isoprene profiles we find an <span class="hlt">HCHO</span> molar yield from isoprene oxidation of 1.6 +/- 0.5, consistent with current chemical mechanisms. Clouds are the primary error source in the AMF calculation; errors in the <span class="hlt">HCHO</span> vertical profile and aerosols have comparatively little effect. The mean bias and 1Q uncertainty in the GEOS-Chem AMF calculation increase from <1% and 15% for clear skies to 17% and 24% for half-cloudy scenes. With fitting errors, this gives an overall 1 Q error in <span class="hlt">HCHO</span> satellite measurements of 25-31%. Retrieval errors, combined with uncertainties in the <span class="hlt">HCHO</span> yield from isoprene oxidation, result in a 40% (1sigma) error in inferring isoprene emissions from <span class="hlt">HCHO</span> satellite measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1989PhRvB..39.6690Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1989PhRvB..39.6690Y"><span>Oxygen isotope effect in YBa2Cu<span class="hlt">3</span><span class="hlt">O</span>7 prepared by burning YBa2Cu<span class="hlt">3</span> in 16<span class="hlt">O</span> and 18<span class="hlt">O</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yvon, Pascal J.; Schwarz, R. B.; Pierce, C. B.; Bernardez, L.; Conners, A.; Meisenheimer, R.</p> <p>1989-04-01</p> <p>We prepared YBa2Cu<span class="hlt">3</span> powder by ball milling a 2:1 molar mixture of the intermetallics BaCu and CuY. We synthesized YBa2Cu<span class="hlt">3</span>(16<span class="hlt">O</span>)7-x and YBa2Cu<span class="hlt">3</span>(18<span class="hlt">O</span>)7-x by oxidizing the YBa2Cu<span class="hlt">3</span> powder in 16<span class="hlt">O</span> and 18<span class="hlt">O</span>. The 16<span class="hlt">O</span>/18<span class="hlt">O</span> ratios were determined by laser-ionization and sputtering-ionization mass spectroscopy. The YBa2Cu<span class="hlt">3</span>(160)7-x sample had 99.8 at. %16<span class="hlt">O</span>, and the YBa2Cu<span class="hlt">3</span>(18<span class="hlt">O</span>)7-x sample had 96.5 at. %18<span class="hlt">O</span>. Susceptibility measurements of the superconducting transition temperature (Tc=91.7 K for 16<span class="hlt">O</span>; half-point transition at 84 K show an isotope effect of 0.4+/-0.1 K.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21353500','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21353500"><span>Preparation of surface plasmon resonance biosensor based on magnetic core/shell Fe<span class="hlt">3</span><span class="hlt">O</span>4/Si<span class="hlt">O</span>2 and Fe<span class="hlt">3</span><span class="hlt">O</span>4/Ag/Si<span class="hlt">O</span>2 nanoparticles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Liying; Sun, Ying; Wang, Jing; Wang, Jian; Yu, Aimin; Zhang, Hanqi; Song, Daqian</p> <p>2011-06-01</p> <p>In this paper, surface plasmon resonance biosensors based on magnetic core/shell Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Si<span class="hlt">O</span>(2) and Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Ag/Si<span class="hlt">O</span>(2) nanoparticles were developed for immunoassay. With Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4) and Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Ag nanoparticles being used as seeding materials, Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Si<span class="hlt">O</span>(2) and Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Ag/Si<span class="hlt">O</span>(2) nanoparticles were formed by hydrolysis of tetraethyl orthosilicate. The aldehyde group functionalized magnetic nanoparticles provide organic functionality for bioconjugation. The products were characterized by scanning electronic microscopy (SEM), transmission electronic microscopy (TEM), FTIR and UV-vis absorption spectrometry. The magnetic nanoparticles possess the unique superparamagnetism property, exceptional optical properties and good compatibilities, and could be used as immobilization matrix for goat anti-rabbit IgG. The magnetic nanoparticles can be easily immobilized on the surface of SPR biosensor chip by a magnetic pillar. The effects of Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Si<span class="hlt">O</span>(2) and Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Ag/Si<span class="hlt">O</span>(2) nanoparticles on the sensitivity of SPR biosensors were also investigated. As a result, the SPR biosensors based on Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Si<span class="hlt">O</span>(2) nanoparticles and Fe(<span class="hlt">3</span>)<span class="hlt">O</span>(4)/Ag/Si<span class="hlt">O</span>(2) nanoparticles exhibit a response for rabbit IgG in the concentration range of 1.25-20.00 μg ml(-1) and 0.30-20.00 μg ml(-1), respectively. Copyright © 2011 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003JaJAP..42.6102K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003JaJAP..42.6102K"><span>Ferroelectric and Piezoelectric Properties of KNb<span class="hlt">O</span><span class="hlt">3</span> Ceramics Containing Small Amounts of LaFe<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kakimoto, Ken-ichi; Masuda, Izumi; Ohsato, Hitoshi</p> <p>2003-09-01</p> <p>Dense KNb<span class="hlt">O</span><span class="hlt">3</span> ceramics have been successfully synthesized by pressure-less sintering under optimized heat-treatment conditions using a small amount of La2<span class="hlt">O</span><span class="hlt">3</span> and Fe2<span class="hlt">O</span><span class="hlt">3</span> additives. KNb<span class="hlt">O</span><span class="hlt">3</span> forms (K1-xLax)(Nb1-xFex)<span class="hlt">O</span><span class="hlt">3</span> solid solutions and changes in the crystal system, depending on the additive content, from orthorhombic to tetragonal at x of 0.020, and from tetragonal to cubic at x of 0.200 or higher. When only 0.002 mol of La2<span class="hlt">O</span><span class="hlt">3</span> and Fe2<span class="hlt">O</span><span class="hlt">3</span> (x=0.002) was added into KNb<span class="hlt">O</span><span class="hlt">3</span>, the highest value (98.8%) of the theoretical density was obtained. This specimen showed orthorhombic symmetry with a high Curie temperature of 420°C, and demonstrated a well-saturated ferroelectric hysteresis loop with large remanent polarization (Pr) of 18 μC/cm2, which is comparable to the value reported for pure KNb<span class="hlt">O</span><span class="hlt">3</span> ceramics fabricated by hot pressing. Furthermore, the x=0.002 specimen showed a planar electromechanical coupling ratio (kp) of 0.17 and piezoelectric d33 constant of 98 pC/N, regardless of the unsaturated poling state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SuMi..112..262K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SuMi..112..262K"><span>Transparent Cu4<span class="hlt">O</span><span class="hlt">3</span>/Zn<span class="hlt">O</span> heterojunction photoelectric devices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Hong-Sik; Yadav, Pankaj; Patel, Malkeshkumar; Kim, Joondong; Pandey, Kavita; Lim, Donggun; Jeong, Chaehwan</p> <p>2017-12-01</p> <p>The present article reports the development of flexible, self-biased, broadband, high speed and transparent heterojunction photodiode, which is essentially important for the next generation electronic devices. We grow semitransparent p-type Cu4<span class="hlt">O</span><span class="hlt">3</span> using the reactive sputtering method at room temperature. The structural and optical properties of the Cu4<span class="hlt">O</span><span class="hlt">3</span> film were investigated by using the X-ray diffraction and UV-visible spectroscopy, respectively. The p-Cu4<span class="hlt">O</span><span class="hlt">3</span>/n-Zn<span class="hlt">O</span> heterojunction diode under dark condition yields rectification behavior with an extremely low saturation current value of 1.8 × 10-10 A and a zero bias photocurrent under illumination condition. The transparent p-Cu4<span class="hlt">O</span><span class="hlt">3</span>/n-Zn<span class="hlt">O</span> heterojunction photodetector can be operated without an external bias, due to the light-induced voltage production. The metal oxide heterojunction based on Cu4<span class="hlt">O</span><span class="hlt">3</span>/Zn<span class="hlt">O</span> would provide a route for the transparent and flexible photoelectric devices, including photodetectors and photovoltaics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29282869','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29282869"><span>Emission analysis of Tb<span class="hlt">3</span>+ -and Sm<span class="hlt">3</span>+ -ion-doped (Li2 <span class="hlt">O</span>/Na2 <span class="hlt">O</span>/K2 <span class="hlt">O</span>) and (Li2 <span class="hlt">O</span> + Na2 <span class="hlt">O</span>/Li2 <span class="hlt">O</span> + K2 <span class="hlt">O</span>/K2 <span class="hlt">O</span> + Na2 <span class="hlt">O</span>)-modified borosilicate glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Naveen Kumar Reddy, B; Sailaja, S; Thyagarajan, K; Jho, Young Dahl; Sudhakar Reddy, B</p> <p>2018-05-01</p> <p>Four series of borosilicate glasses modified by alkali oxides and doped with Tb <span class="hlt">3</span>+ and Sm <span class="hlt">3</span>+ ions were prepared using the conventional melt quenching technique, with the chemical composition 74.5B 2 <span class="hlt">O</span> <span class="hlt">3</span> + 10Si<span class="hlt">O</span> 2 + 5Mg<span class="hlt">O</span> + R + 0.5(Tb 2 <span class="hlt">O</span> <span class="hlt">3</span> /Sm 2 <span class="hlt">O</span> <span class="hlt">3</span> ) [where R = 10(Li 2 <span class="hlt">O</span> /Na 2 <span class="hlt">O</span>/K 2 <span class="hlt">O</span>) for series A and C, and R = 5(Li 2 <span class="hlt">O</span> + Na 2 <span class="hlt">O</span>/Li 2 <span class="hlt">O</span> + K 2 <span class="hlt">O</span>/K 2 <span class="hlt">O</span> + Na 2 <span class="hlt">O</span>) for series B and D]. The X-ray diffraction (XRD) patterns of all the prepared glasses indicate their amorphous nature. The spectroscopic properties of the prepared glasses were studied by optical absorption analysis, photoluminescence excitation (PLE) and photoluminescence (PL) analysis. A green emission corresponding to the 5 D 4 → 7 F 5 (543 nm) transition of the Tb <span class="hlt">3</span>+ ions was registered under excitation at 379 nm for series A and B glasses. The emission spectra of the Sm <span class="hlt">3</span>+ ions with the series C and D glasses showed strong reddish-orange emission at 600 nm ( 4 G 5/2 → 6 H 7/2 ) with an excitation wavelength λ exci = 404 nm ( 6 H 5/2 → 4 F 7/2 ). Furthermore, the change in the luminescence intensity with the addition of an alkali oxide and combinations of these alkali oxides to borosilicate glasses doped with Tb <span class="hlt">3</span>+ and Sm <span class="hlt">3</span>+ ions was studied to optimize the potential alkali-oxide-modified borosilicate glass. Copyright © 2017 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1389309-structural-studies-bi-nb-teo-glasses','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1389309-structural-studies-bi-nb-teo-glasses"><span>Structural studies of Bi 2 <span class="hlt">O</span> <span class="hlt">3</span> -Nb 2 <span class="hlt">O</span> 5 -Te<span class="hlt">O</span> 2 glasses</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Wilding, Martin C.; Delaizir, Gaelle; Benmore, Chris J.; ...</p> <p>2016-07-25</p> <p>Bi 2<span class="hlt">O</span> <span class="hlt">3</span>-Nb 2<span class="hlt">O</span> 5-Te<span class="hlt">O</span> 2 glasses show unusual annealing behavior with appearance of spherulites within the matrix glass structure for the Bi 0.5Nb 0.5Te <span class="hlt">3</span><span class="hlt">O</span> 8 composition. The textures resemble those found previously among polyamorphic Al 2<span class="hlt">O</span> <span class="hlt">3</span>-Y 2<span class="hlt">O</span> <span class="hlt">3</span> glasses containing metastably co-existing high- and low-density phases produced during quenching. However the spherulites produced within the Bi 2<span class="hlt">O</span> <span class="hlt">3</span>-Nb 2<span class="hlt">O</span> 5-Te<span class="hlt">O</span> 2 glass are crystalline and can be identified as an “anti-glass” phase related to β-Bi 2Te 4<span class="hlt">O</span> 11. Here, we used high energy synchrotron X-ray diffraction data to study structures of binary and ternary glasses quenched frommore » liquids within the Bi 2<span class="hlt">O</span> <span class="hlt">3</span>-Nb 2<span class="hlt">O</span> 5-Te<span class="hlt">O</span> 2 system. These reveal a glassy network based on interconnected Te<span class="hlt">O</span> 4 and Te<span class="hlt">O</span> <span class="hlt">3</span> units that is related to Te<span class="hlt">O</span> 2 crystalline materials but with larger Te…Te separations due to the presence of Te<span class="hlt">O</span> <span class="hlt">3</span> groups and non-bridging oxygens linked to modifier (Bi <span class="hlt">3</span> +, Nb 5 +) cations. Analysis of the viscosity-temperature relations indicates that the glass-forming liquids are “fragile” and there is no evidence for a LLPT occurring in the supercooled liquid. The glasses obtained by quenching likely correspond to a high-density amorphous (HDA) state. Subsequent annealing above T g shows mainly evidence for direct crystallization of the “anti-glass” tellurite phase. But, some evidence may exist for simultaneous formation of nanoscale amorphous spherulites that could correspond to the LDA polyamorph. The quenching and annealing behavior of Bi 2<span class="hlt">O</span> <span class="hlt">3</span>-Nb 2<span class="hlt">O</span> 5-Te<span class="hlt">O</span> 2 supercooled liquids and glasses is compared with similar materials in the Al 2<span class="hlt">O</span> <span class="hlt">3</span>-Y 2<span class="hlt">O</span> <span class="hlt">3</span> system.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/AD0750930','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/AD0750930"><span>Results of Research in the Field of Alkylphenol Additives</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p></p> <p>The preparation of Ba and Ca salts of alkylphenols and bis(2- hydroxyphenylalkyl) sulfides are discussed along with the condensation products of... alkylphenols with <span class="hlt">HCHO</span> or <span class="hlt">HCHO</span> and urea, an amine or NH<span class="hlt">3</span> as well as the dithiophosphates of the latter. The properties and performance of lubricating oils</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MMTB..tmp..936H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MMTB..tmp..936H"><span>Viscosity of Ti<span class="hlt">O</span>2-Fe<span class="hlt">O</span>-Ti2<span class="hlt">O</span><span class="hlt">3</span>-Si<span class="hlt">O</span>2-Mg<span class="hlt">O-CaO</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> for High-Titania Slag Smelting Process</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hu, Kai; Lv, Xuewei; Li, Shengping; Lv, Wei; Song, Bing; Han, Kexi</p> <p>2018-05-01</p> <p>The present study demonstrates the dependence of viscosity on chemical composition and temperature of high-titania slag, a very important raw material for producing titanium dioxide. The results indicated that completely molten high-titania slag exhibits a viscosity of less than 1 dPa s with negligible dependence on temperature. However, it increases dramatically with decreasing temperature slightly below the critical temperature, i.e., the solidus temperature of the slag. Above the critical temperature, the slag samples displayed the same order of viscosity at 0.6 dPa s, regardless of their compositional variation. However, the Fe<span class="hlt">O</span>, Ca<span class="hlt">O</span>, and Mg<span class="hlt">O</span> were confirmed to decrease viscosity, while Si<span class="hlt">O</span>2 and Ti2<span class="hlt">O</span><span class="hlt">3</span> increase it. The apparent activation energy for viscosity-temperature relation and liquidus temperature based on experiments and thermodynamic calculations are also presented. Conclusively, the critical temperatures of the slags are on average 15 K below their corresponding calculated liquidus temperatures. The increase in Fe<span class="hlt">O</span> content was found to considerably lower the critical temperature, while the increase in both Ti2<span class="hlt">O</span><span class="hlt">3</span> and Ti<span class="hlt">O</span>2 contents increases it. The main phases of the slag in solid state, as indicated by X-ray diffraction, are (Fe, Mg) x Ti y <span class="hlt">O</span>5 (x + y = <span class="hlt">3</span>, pseudobrookite) and rutile.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010064404','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010064404"><span>The Effect of Microstructure on Mechanical Properties of Directionally Solidified Al2<span class="hlt">O</span><span class="hlt">3</span>/Zr<span class="hlt">O</span>2(Y2<span class="hlt">O</span><span class="hlt">3</span>) Eutectic</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sayir, Ali; Farmer, Serene C.</p> <p>1999-01-01</p> <p>The eutectic architecture of a continuous reinforcing phase within a higher volume fraction phase or matrix can be described as a naturally occurring in-situ composite. Here we report the results of experiments aimed at identifying the sources of high temperature creep resistance and high levels of strength in a two phase Al2<span class="hlt">O</span><span class="hlt">3</span>/Zr<span class="hlt">O</span>2(Y2<span class="hlt">O</span><span class="hlt">3</span>) system. The mechanical properties of two phase Al2<span class="hlt">O</span><span class="hlt">3</span>/Zr<span class="hlt">O</span>2(Y2<span class="hlt">O</span><span class="hlt">3</span>) eutectic are superior to those of either constituent alone due to strong constraining effects provided by the coherent interfaces and microstructure. The Al<span class="hlt">O</span><span class="hlt">3</span>/Zr<span class="hlt">O</span>2(Y2<span class="hlt">O</span><span class="hlt">3</span>) eutectic maintains a low energy interface resulting from directional solidification and can produce strong and stable reinforcing phase/matrix bonding. The phases comprising a eutectic are thermodynamically compatible at higher homologous temperatures than man-made composites and as such offer the potential for superior high temperature properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..MARS32009D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..MARS32009D"><span>Strain dependence of interfacial antiferromagnetic coupling in La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span>/SrRu<span class="hlt">O</span><span class="hlt">3</span> superlattices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Das, Sujit; Herklotz, Andreas; Pippel, Eckhard; Guo, Er-Jia; Rata, Diana; Dörr, Kathrin</p> <p>2015-03-01</p> <p>We have investigated the magnetic response of La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span>/SrRu<span class="hlt">O</span><span class="hlt">3</span> superlattices to biaxial in-plane strain applied in-situ. Superlattices grown on piezoelectric substrates of 0.72PbMg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span><span class="hlt">O</span><span class="hlt">3</span>-0.28PbTi<span class="hlt">O</span><span class="hlt">3</span>(001) (PMN-PT) show strong antiferromagnetic coupling of the two ferromagnetic components. The coupling field of μ0HAF = 1.8 T is found to change by μ0 ΔHAF / Δɛ ~ -520 mT %-1 under reversible biaxial strain (Δɛ) at 80 K in a [La0.7Sr0.<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span>(22 Å)/SrRu<span class="hlt">O</span><span class="hlt">3</span>(55 Å)]15 superlattice. This reveals a significant strain effect on interfacial coupling. The applied in-plane compression enhances the ferromagnetic order in the manganite layers which are under as-grown tensile strain. It is thus difficult to disentangle the contributions from strain-dependent antiferromagnetic Mn-<span class="hlt">O</span>-Ru interface coupling and Mn-<span class="hlt">O</span>-Mn ferromagnetic double exchange near the interface, since the enhanced magnetic order of Mn spins leads to a larger net coupling of SrRu<span class="hlt">O</span><span class="hlt">3</span> layers at the interface. We discuss our experimental findings taken into account both the strain-dependent orbital occupation in a single-ion picture and the enhanced Mn order at the interface. This work was supported by the DFG within the Collaborative Research Center SFB 762 ``Functionality of Oxide Interfaces.''</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008SPIE.7056E..0YZ','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008SPIE.7056E..0YZ"><span>Ultraviolet-infrared laser-induced domain inversion in Mg<span class="hlt">O</span>-doped congruent LiNb<span class="hlt">O</span><span class="hlt">3</span> and near stoichiometric LiTa<span class="hlt">O</span><span class="hlt">3</span> crystals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhi, Ya'nan; Qu, Weijuan; Liu, De'an; Sun, Jianfeng; Yan, Aimin; Liu, Liren</p> <p>2008-08-01</p> <p>Laser-induced domain inversion is a promising technique for domain engineering in LiNb<span class="hlt">O</span><span class="hlt">3</span> and LiTa<span class="hlt">O</span><span class="hlt">3</span>. The ultraviolet-infrared laser induced domain inversions in Mg<span class="hlt">O</span>-doped congruent LiNb<span class="hlt">O</span><span class="hlt">3</span> and near stoichiometric LiTa<span class="hlt">O</span><span class="hlt">3</span> crystals are investigated for the first time here. Within the wavelength range from 351 to 799 nm, the different reductions of nucleation field induced by the focused continuous laser irradiation are systematically investigated in the Mg<span class="hlt">O</span>-doped congruent LiNb<span class="hlt">O</span><span class="hlt">3</span> crystals. The investigation of ultrashort-pulse laser-induced domain inversion in Mg<span class="hlt">O</span>-doped congruent LiNb<span class="hlt">O</span><span class="hlt">3</span> is performed with 800 nm wavelength irradiation. The focused continuous ultraviolet laser-induced ferroelectric domain inversion in the near stoichiometric LiTa<span class="hlt">O</span><span class="hlt">3</span> is also investigated. The different physical explanations, based on space charge field and defect formation, are presented for the laser-induced domain inversion, and the solid experimental proofs are also presented. The results provide the solid experimental proofs and feasible schemes for the further investigation of laser-induced domain engineering in Mg<span class="hlt">O</span>-doped LiNb<span class="hlt">O</span><span class="hlt">3</span> and near stoichiometric LiTa<span class="hlt">O</span><span class="hlt">3</span> crystals. The important characteristics of domain inversion, including domain wall and internal field, in LiNb<span class="hlt">O</span><span class="hlt">3</span> crystals are also investigated by the digital holographic interferometry with an improved reconstruction method, and some creative experimental results and conclusions are achieved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1832g0021S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1832g0021S"><span>Structure of Te<span class="hlt">O</span>2 - LiNb<span class="hlt">O</span><span class="hlt">3</span> glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shinde, A. B.; Krishna, P. S. R.; Rao, Rekha</p> <p>2017-05-01</p> <p>Tellurite based lithium niobate glasses with composition (100-x)Te<span class="hlt">O</span>2-xLiNb<span class="hlt">O</span><span class="hlt">3</span> (x=0.1,0.2 & 0.<span class="hlt">3</span>) were prepared by conventional melt quenching method. The microscopic structural investigation of these glasses is carried out by means of neutron diffraction and Raman scattering measurements. It is found that the basic structural units in these glasses are Te<span class="hlt">O</span>4 trigonal bipyramids(TBP), Te<span class="hlt">O</span><span class="hlt">3</span> trigonal pyramids(TP) and Nb<span class="hlt">O</span>6 Octahedra depending on the composition. It is evident from Raman studies that TBPs decreases, TPs increases and Nb<span class="hlt">O</span>6 Octahedra increases with increasing x. From Neutron diffraction studies it is found that network is comprised of TBPs and TPs along with Te<span class="hlt">O</span><span class="hlt">3</span>+1 structural units. Distorted Nb<span class="hlt">O</span>6 octahedral units are present and also increase with the increase in x.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22475973-hydrothermal-synthesis-sub-sub-coated-sub-sub-eu-sup-nanotubes-enhanced-photoluminescence-properties','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22475973-hydrothermal-synthesis-sub-sub-coated-sub-sub-eu-sup-nanotubes-enhanced-photoluminescence-properties"><span>Hydrothermal synthesis of Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} coated Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} nanotubes for enhanced photoluminescence properties</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gao, Linhui, E-mail: gaolhui@zstu.edu.cn; Wang, Guangfa; Zhu, Hongliang</p> <p></p> <p>Highlights: • Eu{sup <span class="hlt">3</span>+} doped Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} nanotubes. • Hydrothermal synthesis of Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} coated Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} nanostructures assissted with a further heat treatment. • Tunable coating ratios of Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} coated Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} nanophosphor. • Enhanced photoluminescence intensity of Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} more than 60% by Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} surface coating. - Abstract: Novel Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} coated Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} nanotubes with different coating ratios were synthesized successfully by a facile two-step process, including hydrothermal synthesis of Y(OH){sub <span class="hlt">3</span>} coated Y(OH){sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} as precursors and then calcination ofmore » them at 1000 °C for 2 h. X-ray diffraction patterns and field emission scanning electron microscope images indicated these Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} coated Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} phosphors possess tubular nanostructures. The photoluminescence properties of Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>} coated Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} were systematically investigated by photoluminescence spectra, and photoluminescence enhancement was observed after proper coating. In other words, the coating ratio played a crucial role in photoluminescence efficiency. When it was 1/9, the photoluminescence intensity of {sup 5}D{sub 0} → {sup 7}F{sub 2} emission (about 613 nm) was 60% higher than that of Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}: Eu{sup <span class="hlt">3</span>+} phosphors under 255 nm excitation. Therefore, surface coating may be an alternative route for enhanced photoluminescence properties of the Y{sub 2}<span class="hlt">O</span>{sub <span class="hlt">3</span>}:Eu{sup <span class="hlt">3</span>+} red-emitting phosphor.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvB..96t1406Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvB..96t1406Z"><span>La interstitial defect-induced insulator-metal transition in the oxide heterostructures LaAl <span class="hlt">O</span><span class="hlt">3</span> /SrTi <span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, Jun; Yang, Ming; Feng, Yuan Ping; Rusydi, Andrivo</p> <p>2017-11-01</p> <p>Perovskite oxide interfaces have attracted tremendous research interest for their fundamental physics and promising all-oxide electronic applications. Here, based on first-principles calculations, we propose a surface La interstitial promoted interface insulator-metal transition in LaAl <span class="hlt">O</span><span class="hlt">3</span> /SrTi <span class="hlt">O</span><span class="hlt">3</span> (110). Compared with surface oxygen vacancies, which play a determining role on the insulator-metal transition of LaAl <span class="hlt">O</span><span class="hlt">3</span> /SrTi <span class="hlt">O</span><span class="hlt">3</span> (001) interfaces, we find that surface La interstitials can be more experimentally realistic and accessible for manipulation and more stable in an ambient atmospheric environment. Interestingly, these surface La interstitials also induce significant spin-splitting states with a Ti dy z/dx z character at a conducting LaAl <span class="hlt">O</span><span class="hlt">3</span> /SrTi <span class="hlt">O</span><span class="hlt">3</span> (110) interface. On the other hand, for insulating LaAl <span class="hlt">O</span><span class="hlt">3</span> /SrTi <span class="hlt">O</span><span class="hlt">3</span> (110) (<4 unit cells LaAl <span class="hlt">O</span><span class="hlt">3</span> thickness), a distortion between La (Al) and <span class="hlt">O</span> atoms is found at the LaAl <span class="hlt">O</span><span class="hlt">3</span> side, partially compensating the polarization divergence. Our results reveal the origin of the metal-insulator transition in LaAl <span class="hlt">O</span><span class="hlt">3</span> /SrTi <span class="hlt">O</span><span class="hlt">3</span> (110) heterostructures, and also shed light on the manipulation of the superior properties of LaAl <span class="hlt">O</span><span class="hlt">3</span> /SrTi <span class="hlt">O</span><span class="hlt">3</span> (110) for different possibilities in electronic and magnetic applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MetRT.115..304L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MetRT.115..304L"><span>Investigation the influences of B2<span class="hlt">O</span><span class="hlt">3</span> and R2<span class="hlt">O</span> on the structure and crystallization behaviors of Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> based F-free mold flux</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Jiangling; Kong, Bowen; Gao, Xiangyu; Liu, Qingcai; Shu, Qifeng; Chou, Kuochih</p> <p>2018-04-01</p> <p>The influences of B2<span class="hlt">O</span><span class="hlt">3</span> and R2<span class="hlt">O</span> on the structure and crystallization of Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> based F-free mold flux were investigated by Raman Spectroscopy and Differential Scanning Calorimetry Technique, respectively, for developing a new type of F-free mold flux. The results of structural investigations showed that B<span class="hlt">3</span>+ is mainly in the form of [BO<span class="hlt">3</span>]. And [BO<span class="hlt">3</span>] appears to form BIII-<span class="hlt">O</span>-Al linkage which will produce a positive effect on forming [Al<span class="hlt">O</span>4] network. The number of bridging oxygen and the degree of polymerization of [Al<span class="hlt">O</span>4] network structure for Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> system were also increased with the increasing of B2<span class="hlt">O</span><span class="hlt">3</span>. On the contrary, with the addition of R2<span class="hlt">O</span> into Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-B2<span class="hlt">O</span><span class="hlt">3</span> system, the number of bridging oxygen and the degree of polymerization of [Al<span class="hlt">O</span>4] network were decreased. DSC results showed that the crystallization process became more sluggish with the increase of B2<span class="hlt">O</span><span class="hlt">3</span>, which indicated that the crystallization ability was weakened. While the quenched mold fluxes crystallized more rapidly when introducing R2<span class="hlt">O</span>. In other word, the crystallization rates of Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> based slags were accelerated by the introduction of R2<span class="hlt">O</span>. The liquidus temperature and crystallization temperature were decreased with the increasing amount of B2<span class="hlt">O</span><span class="hlt">3</span> or by addition of R2<span class="hlt">O</span> into Ca<span class="hlt">O</span>-Al2<span class="hlt">O</span><span class="hlt">3</span> system. Only one kind of crystal was precipitated in 8% B2<span class="hlt">O</span><span class="hlt">3</span> and %R2<span class="hlt">O</span>-containing samples, which was CaAl2<span class="hlt">O</span>4 identified by SEM-EDS. When the content of B2<span class="hlt">O</span><span class="hlt">3</span> increased from 8% to 16%, Ca<span class="hlt">3</span>B2<span class="hlt">O</span>6 is clearly observed, demonstrating that the crystallization ability of Ca<span class="hlt">3</span>B2<span class="hlt">O</span>6 is enhanced by the increasing concentration of B2<span class="hlt">O</span><span class="hlt">3</span> in mold flux. The Ca/Al ratio of the generated calcium aluminate has been altered from 1:2 to 1:4 with the increasing of B2<span class="hlt">O</span><span class="hlt">3</span>. The size of CaAl2<span class="hlt">O</span>4 crystal is gradually increased with the addition of R2<span class="hlt">O</span>. The crystallization ability of CaAl2<span class="hlt">O</span>4 is promoted by R2<span class="hlt">O</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPCM...29W5601P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPCM...29W5601P"><span>Structural, magnetic and electronic properties of pulsed-laser-deposition grown SrFe<span class="hlt">O</span><span class="hlt">3</span>-δ thin films and SrFe<span class="hlt">O</span><span class="hlt">3</span>-δ /La2/<span class="hlt">3</span>Ca1/<span class="hlt">3</span>Mn<span class="hlt">O</span><span class="hlt">3</span> multilayers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Perret, E.; Sen, K.; Khmaladze, J.; Mallett, B. P. P.; Yazdi-Rizi, M.; Marsik, P.; Das, S.; Marozau, I.; Uribe-Laverde, M. A.; de Andrés Prada, R.; Strempfer, J.; Döbeli, M.; Biškup, N.; Varela, M.; Mathis, Y.-L.; Bernhard, C.</p> <p>2017-12-01</p> <p>We studied the structural, magnetic and electronic properties of SrFe<span class="hlt">O</span><span class="hlt">3</span>-δ (SFO) thin films and SrFe<span class="hlt">O</span><span class="hlt">3</span>-δ /La2/<span class="hlt">3</span> Ca1/<span class="hlt">3</span> Mn<span class="hlt">O</span><span class="hlt">3</span> (LCMO) superlattices that have been grown with pulsed laser deposition (PLD) on La0.<span class="hlt">3</span> Sr0.7 Al0.65 Ta0.35 <span class="hlt">O</span><span class="hlt">3</span> (LSAT) substrates. X-ray reflectometry and scanning transmission electron microscopy (STEM) confirm the high structural quality of the films and flat and atomically sharp interfaces of the superlattices. The STEM data also reveal a difference in the interfacial layer stacking with a Sr<span class="hlt">O</span> layer at the LCMO/SFO and a La<span class="hlt">O</span> layer at the SFO/LCMO interfaces along the PLD growth direction. The x-ray diffraction (XRD) data suggest that the as grown SFO films and SFO/LCMO superlattices have an oxygen-deficient SrFe<span class="hlt">O</span><span class="hlt">3</span>-δ structure with I4/ mmm space group symmetry (δ≤slant 0.2 ). Subsequent ozone annealed SFO films are consistent with an almost oxygen stoichiometric structure (δ ≈ 0 ). The electronic and magnetic properties of these SFO films are similar to the ones of corresponding single crystals. In particular, the as grown SrFe<span class="hlt">O</span><span class="hlt">3</span>-δ films are insulating whereas the ozone annealed films are metallic. The magneto-resistance effects of the as grown SFO films have a similar magnitude as in the single crystals, but extend over a much wider temperature range. Last but not least, for the SFO/LCMO superlattices we observe a rather large exchange bias effect that varies as a function of the cooling field.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1338347-pressure-induced-ferroelectric-paraelectric-transition-litao3-li-mg-tao3','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1338347-pressure-induced-ferroelectric-paraelectric-transition-litao3-li-mg-tao3"><span>Pressure-induced ferroelectric to paraelectric transition in LiTa<span class="hlt">O</span> <span class="hlt">3</span> and (Li,Mg)Ta<span class="hlt">O</span> <span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Yamanaka, Takamitsu; Nakamoto, Yuki; Takei, Fumihiko; ...</p> <p>2016-02-16</p> <p>X-ray powder diffraction and Raman scattering of LiTa<span class="hlt">O</span> <span class="hlt">3</span> (LT) and (Li,Mg)Ta<span class="hlt">O</span> <span class="hlt">3</span> (LMT) have been measured under pressure up to 46 GPa. Above 30 GPa, the ferroelectric rhombohedral phase (R<span class="hlt">3</span>c, Z – 6) of LiTa<span class="hlt">O</span> <span class="hlt">3</span> transforms to a paraelectric orthorhombic phase (Pnma with Z – 4) with a large hysteresis. Rietveld profile fitting analysis shows that the Li-<span class="hlt">O</span> bond is compressed and approaches that of Ta-<span class="hlt">O</span> with pressure. The cation distribution analysis of the orthorhombic perovskite structure shows that Li and Ta are located in the octahedral 8-fold coordination sites. Difference Fourier |F obs(hkl)| - |F cal(hkl)| mapsmore » of LiTa<span class="hlt">O</span> <span class="hlt">3</span> and (Li,Mg)Ta<span class="hlt">O</span> <span class="hlt">3</span> indicate polarization in the c axis direction and a more distinct electron density distribution around the Ta position for (Li,Mg)Ta<span class="hlt">O</span> <span class="hlt">3</span> compared to LiTa<span class="hlt">O</span> <span class="hlt">3</span>. The observed effective charges indicate that for (Li,Mg)Ta<span class="hlt">O</span> <span class="hlt">3</span> without vacancies Ta 5+ becomes less ionized as a function of Mg substitution. Considering both site occupancy and effective charge analysis, Ta 5+ is reduced to Ta 4.13+. Mg 2+ and <span class="hlt">O</span> 2- change to Mg 1.643+ and <span class="hlt">O</span> 1.732 -, respectively. The space- and time-averaged structures of the dynamical vibration of atoms can be elucidated from the electron density analysis by difference Fourier and temperature factors T(hkl) in the structure refinement. The refinement of the temperature factor is consistent with the cation distribution assuming full stoichiometry. The residual electron density induced from the excess electron in (Li,Mg)Ta<span class="hlt">O</span> <span class="hlt">3</span> indicates more electrons around the Ta site, as confirmed by the effective charge analysis. Raman spectra of LiTa<span class="hlt">O</span> <span class="hlt">3</span> and (Li,Mg)Ta<span class="hlt">O</span> <span class="hlt">3</span> show notable changes over the measured pressure range. Raman peaks centered at 250 cm –1 and 350 cm –1 at ambient pressure merge above 8 GPa, which we associate with the diminishing of difference in distances between Li-<span class="hlt">O</span> and Ta-<span class="hlt">O</span> bonds with pressure in both materials. Finally, Raman spectra show significant changes at 28 GPa and 33 GPa for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApPhA.124..279M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApPhA.124..279M"><span>Thermal conversion of Cu4<span class="hlt">O</span><span class="hlt">3</span> into Cu<span class="hlt">O</span> and Cu2<span class="hlt">O</span> and the electrical properties of magnetron sputtered Cu4<span class="hlt">O</span><span class="hlt">3</span> thin films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Murali, Dhanya S.; Aryasomayajula, Subrahmanyam</p> <p>2018-03-01</p> <p>Among the three oxides of copper (Cu<span class="hlt">O</span>, Cu2<span class="hlt">O</span>, and Cu4<span class="hlt">O</span><span class="hlt">3</span>), Cu4<span class="hlt">O</span><span class="hlt">3</span> phase (paramelaconite is a natural, and very scarce mineral) is very difficult to synthesize. It contains copper in both + 1 and + 2 valence states, with an average composition Cu2 1+Cu2 2+<span class="hlt">O</span><span class="hlt">3</span>. We have successfully synthesized Cu4<span class="hlt">O</span><span class="hlt">3</span> phase at room temperature (300 K) by reactive DC magnetron sputtering by controlling the oxygen flow rate (Murali and Subrahmanyam in J Phys D Appl Phys 49:375102, 2016). In the present communication, Cu4<span class="hlt">O</span><span class="hlt">3</span> thin films are converted to Cu<span class="hlt">O</span> phases by annealing in the air at 680 K and to Cu2<span class="hlt">O</span> phase when annealed in argon at 720 K; these phase changes are confirmed by temperature-dependent Raman spectroscopy studies. Probably, this is the first report of the conversion of Cu4<span class="hlt">O</span><span class="hlt">3</span>-Cu<span class="hlt">O</span> and Cu2<span class="hlt">O</span> by thermal annealing. The temperature-dependent (300-200 K) electrical transport properties of Cu4<span class="hlt">O</span><span class="hlt">3</span> thin films show that the charge transport above 190 K follows Arrhenius-type behavior with activation energy of 0.14 eV. From photo-electron spectroscopy and electrical transport measurements of Cu4<span class="hlt">O</span><span class="hlt">3</span> thin films, a downward band bending is observed at the surface of the thin film, which shows its p-type semiconducting nature. The successful preparation of phase pure p-type semiconducting Cu4<span class="hlt">O</span><span class="hlt">3</span> could provide opportunities to further explore its potential applications.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5466374','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5466374"><span>MnTi<span class="hlt">O</span><span class="hlt">3</span>-driven low-temperature oxidative coupling of methane over Ti<span class="hlt">O</span>2-doped Mn2<span class="hlt">O</span><span class="hlt">3</span>-Na2WO4/Si<span class="hlt">O</span>2 catalyst</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wang, Pengwei; Zhao, Guofeng; Wang, Yu; Lu, Yong</p> <p>2017-01-01</p> <p>Oxidative coupling of methane (OCM) is a promising method for the direct conversion of methane to ethene and ethane (C2 products). Among the catalysts reported previously, Mn2<span class="hlt">O</span><span class="hlt">3</span>-Na2WO4/Si<span class="hlt">O</span>2 showed the highest conversion and selectivity, but only at 800° to 900°C, which represents a substantial challenge for commercialization. We report a Ti<span class="hlt">O</span>2-doped Mn2<span class="hlt">O</span><span class="hlt">3</span>-Na2WO4/Si<span class="hlt">O</span>2 catalyst by using Ti-MWW zeolite as Ti<span class="hlt">O</span>2 dopant as well as Si<span class="hlt">O</span>2 support, enabling OCM with 26% conversion and 76% C2-C<span class="hlt">3</span> selectivity at 720°C because of MnTi<span class="hlt">O</span><span class="hlt">3</span> formation. MnTi<span class="hlt">O</span><span class="hlt">3</span> triggers the low-temperature Mn2+↔Mn<span class="hlt">3</span>+ cycle for <span class="hlt">O</span>2 activation while working synergistically with Na2WO4 to selectively convert methane to C2-C<span class="hlt">3</span>. We also prepared a practical Mn2<span class="hlt">O</span><span class="hlt">3</span>-Ti<span class="hlt">O</span>2-Na2WO4/Si<span class="hlt">O</span>2 catalyst in a ball mill. This catalyst can be transformed in situ into MnTi<span class="hlt">O</span><span class="hlt">3</span>-Na2WO4/Si<span class="hlt">O</span>2, yielding 22% conversion and 62% selectivity at 650°C. Our results will stimulate attempts to understand more fully the chemistry of MnTi<span class="hlt">O</span><span class="hlt">3</span>-governed low-temperature activity, which might lead to commercial exploitation of a low-temperature OCM process. PMID:28630917</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JCrGr.479...67N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JCrGr.479...67N"><span>Crystal orientation of monoclinic β-Ga2<span class="hlt">O</span><span class="hlt">3</span> thin films formed on cubic Mg<span class="hlt">O</span> substrates with a γ-Ga2<span class="hlt">O</span><span class="hlt">3</span> interfacial layer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakagomi, Shinji; Kokubun, Yoshihiro</p> <p>2017-12-01</p> <p>The crystal orientation relationship between β-Ga2<span class="hlt">O</span><span class="hlt">3</span> and Mg<span class="hlt">O</span> in β-Ga2<span class="hlt">O</span><span class="hlt">3</span> thin films prepared on (1 0 0), (1 1 1), and (1 1 0) Mg<span class="hlt">O</span> substrates was investigated by X-ray diffraction measurements and cross-sectional transmission electron microscopy images. The γ-Ga2<span class="hlt">O</span><span class="hlt">3</span> interfacial layer was present between β-Ga2<span class="hlt">O</span><span class="hlt">3</span> and Mg<span class="hlt">O</span> acted as a buffer to connect β-Ga2<span class="hlt">O</span><span class="hlt">3</span> on Mg<span class="hlt">O</span>. The following conditions were satisfied under each case: β-Ga2<span class="hlt">O</span><span class="hlt">3</span> (1 0 0)||Mg<span class="hlt">O</span> (1 0 0) and β-Ga2<span class="hlt">O</span><span class="hlt">3</span> [0 0 1]||Mg<span class="hlt">O</span> 〈0 1 1〉 for the formation of β-Ga2<span class="hlt">O</span><span class="hlt">3</span> on (1 0 0) Mg<span class="hlt">O</span>, and β-Ga2<span class="hlt">O</span><span class="hlt">3</span> (2 bar 0 1)||Mg<span class="hlt">O</span> (1 1 1) for the formation of β-Ga2<span class="hlt">O</span><span class="hlt">3</span> on (1 1 1) Mg<span class="hlt">O</span>, as well as each condition of β-Ga2<span class="hlt">O</span><span class="hlt">3</span> [0 1 0] (1 0 0)||Mg<span class="hlt">O</span> [ 1 bar 1 0 ] (0 0 1), β-Ga2<span class="hlt">O</span><span class="hlt">3</span> [0 1 0] (1 0 0)||Mg<span class="hlt">O</span> [ 0 1 bar 1 ] (1 0 0), and β-Ga2<span class="hlt">O</span><span class="hlt">3</span> [0 1 0] (1 0 0)||Mg<span class="hlt">O</span> [ 1 0 1 bar ] (0 1 0). β-Ga2<span class="hlt">O</span><span class="hlt">3</span> (1 bar 0 2)||Mg<span class="hlt">O</span>(1 1 0) and β-Ga2<span class="hlt">O</span><span class="hlt">3</span> [0 1 0] ⊥ Mg<span class="hlt">O</span> [0 0 1] for β-Ga2<span class="hlt">O</span><span class="hlt">3</span> formed on (1 1 0) Mg<span class="hlt">O</span>. The β-Ga2<span class="hlt">O</span><span class="hlt">3</span> formed on (1 1 1) Mg<span class="hlt">O</span> at 800 °C exhibited a threefold structure. The β-Ga2<span class="hlt">O</span><span class="hlt">3</span> formed on (1 1 0) Mg<span class="hlt">O</span> had a twofold structure but different by 90° from the result reported previously.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1344666-effects-al2o3-b2o3-li2o-na2o-sio2-nepheline-crystallization-hanford-high-level-waste-glasses','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1344666-effects-al2o3-b2o3-li2o-na2o-sio2-nepheline-crystallization-hanford-high-level-waste-glasses"><span>Effects of Al2<span class="hlt">O</span><span class="hlt">3</span>, B2<span class="hlt">O</span><span class="hlt">3</span>, Li2<span class="hlt">O</span>, Na2<span class="hlt">O</span>, and Si<span class="hlt">O</span>2 on Nepheline Crystallization in Hanford High Level Waste Glasses</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kroll, Jared O.; Vienna, John D.; Schweiger, Michael J.</p> <p>2016-09-15</p> <p>Nepheline (nominally NaAlSi<span class="hlt">O</span>4) formation during slow cooling of high-alumina (25.4 - 34.5 mass% Al2<span class="hlt">O</span><span class="hlt">3</span>) Hanford high level waste glasses may significantly reduce product durability. To investigate the effects of composition on nepheline crystallization, 29 compositions were formulated by adjusting Al2<span class="hlt">O</span><span class="hlt">3</span>, B2<span class="hlt">O</span><span class="hlt">3</span>, Li2<span class="hlt">O</span>, Na2<span class="hlt">O</span>, and Si<span class="hlt">O</span>2 around a baseline glass that precipitated 12 mass% nepheline. Thirteen of these compositions were generated by adjusting one-component-at-a-time, while two or three components were adjusted to produce the other 16 (with all remaining components staying in the same relative proportions). Quantitative X-ray diffraction was used to determine nepheline concentration in each sample. Twenty two glassesmore » precipitated nepheline, two of which also precipitated eucryptite (nominally LiAlSi<span class="hlt">O</span>4), and one glass formed only eucryptite upon slow cooling. Increasing Na2<span class="hlt">O</span> and Li2<span class="hlt">O</span> had the strongest effect in promoting nepheline formation. Increasing B2<span class="hlt">O</span><span class="hlt">3</span> inhibited nepheline formation. Si<span class="hlt">O</span>2 and Al2<span class="hlt">O</span><span class="hlt">3</span> showed non-linear behavior related to nepheline formation. The composition effects on nepheline formation in these glasses are reported.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11800614','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11800614"><span>The Os<span class="hlt">O</span>(<span class="hlt">3</span>)F(+) and mu-F(Os<span class="hlt">O</span>(<span class="hlt">3</span>)F)(2)(+) cations: their syntheses and study by Raman and (19)F NMR spectroscopy and electron structure calculations and X-ray crystal structures of [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][PnF(6)] (Pn = As, Sb), [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][HF](2)[AsF(6)], [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][HF][SbF(6)], and [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][Sb(<span class="hlt">3</span>)F(16)].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gerken, Michael; Dixon, David A; Schrobilgen, Gary J</p> <p>2002-01-28</p> <p>The fluoride ion donor properties of Os<span class="hlt">O</span>(<span class="hlt">3</span>)F(2) have been investigated. The salts [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][AsF(6)], [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][HF](2)[AsF(6)], mu-F(Os<span class="hlt">O</span>(<span class="hlt">3</span>)F)(2)[AsF(6)], [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][HF](2)[SbF(6)], and [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][HF][SbF(6)] have been prepared by reaction of Os<span class="hlt">O</span>(<span class="hlt">3</span>)F(2) with AsF(5) and SbF(5) in HF solvent and have been characterized in the solid state by Raman spectroscopy. The single-crystal X-ray diffraction studies of [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][AsF(6)] (P2(1)/n, a = 7.0001(11) A, c = 8.8629(13) A, beta = 92.270(7) degrees, Z = 4, and R(1) = 0.0401 at -126 degrees C), [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][SbF(6)] (P2(1)/c, a = 5.4772(14) A, b = 10.115(<span class="hlt">3</span>) A, c = 12.234(<span class="hlt">3</span>) A, beta = 99.321(5) degrees, Z = 4, and R(1) = 0.0325 at -173 degrees C), [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][HF](2)[AsF(6)] (P2(1)/n, a = 5.1491(9) A, b = 8.129(2) A, c = 19.636(7) A, beta = 95.099(7) degrees, Z = 4, and R(1) = 0.0348 at -117 degrees C), and [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][HF][SbF(6)] (Pc, a = 5.244(4) A, b = 9.646(6) A, c = 15.269(10) A, beta = 97.154(13) degrees, Z = 4, and R(1) = 0.0558 at -133 degrees C) have shown that the Os<span class="hlt">O</span>(<span class="hlt">3</span>)F(+) cations exhibit strong contacts to the anions and HF solvent molecules giving rise to cyclic, dimeric structures in which the osmium atoms have coordination numbers of 6. The reaction of Os<span class="hlt">O</span>(<span class="hlt">3</span>)F(2) with neat SbF(5) yielded [Os<span class="hlt">O</span>(<span class="hlt">3</span>)F][Sb(<span class="hlt">3</span>)F(16)], which has been characterized by (19)F NMR spectroscopy in SbF(5) and SO(2)ClF solvents and by Raman spectroscopy and single-crystal X-ray diffraction in the solid state (P4(1)m, a = 10.076(6) A, c = 7.585(8) A, Z = 2, and R(1) = 0.0858 at -113 degrees C). The weak fluoride ion basicity of the Sb(<span class="hlt">3</span>)F(16)(-) anion resulted in an Os<span class="hlt">O</span>(<span class="hlt">3</span>)F(+) cation (C(<span class="hlt">3</span>)(v) point symmetry) that is well isolated from the anion and in which the osmium is four-coordinate. The geometrical parameters and vibrational frequencies of Os<span class="hlt">O</span>(<span class="hlt">3</span>)F(+), Re<span class="hlt">O</span>(<span class="hlt">3</span>)F, mu-F(Os<span class="hlt">O</span>(<span class="hlt">3</span>)F)(2)(+), (FO(<span class="hlt">3</span>)Os--FPnF(5))(2), and (FO(<span class="hlt">3</span>)Os--(HF)(2)--FPnF(5))(2) (Pn = As, Sb) have been calculated using density functional theory methods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhD...51o5205M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhD...51o5205M"><span>On the quantification of the dissolved hydroxyl radicals in the plasma-liquid system using the molecular probe method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ma, Yupengxue; Gong, Xinning; He, Bangbang; Li, Xiaofei; Cao, Dianyu; Li, Junshuai; Xiong, Qing; Chen, Qiang; Chen, Bing Hui; Huo Liu, Qing</p> <p>2018-04-01</p> <p>Hydroxyl (OH) radical is one of the most important reactive species produced by plasma-liquid interactions, and the OH in liquid phase (dissolved OH radical, OHdis) takes effect in many plasma-based applications due to its high reactivity. Therefore, the quantification of the OHdis in a plasma-liquid system is of great importance, and a molecular probe method usually used for the OHdis detection might be applied. Herein, we investigate the validity of using the molecular probe method to estimate the [OHdis] in the plasma-liquid system. Dimethyl sulfoxide is used as the molecular probe to estimate the [OHdis] in an air plasma-liquid system, and usually the estimation of [OHdis] is deduced by quantifying the OHdis-induced derivative, the formaldehyde (<span class="hlt">HCHO</span>). The analysis indicates that the true concentration of the OHdis should be estimated from the sum of three terms: the formed <span class="hlt">HCHO</span>, the existing OH scavengers, and the H2<span class="hlt">O</span>2 formed from the OHdis. The results show that the measured [<span class="hlt">HCHO</span>] needs to be corrected since the <span class="hlt">HCHO</span> consumption is not negligible in the plasma-liquid system. We conclude from the results and the analysis that the molecular probe method generally underestimates the [OHdis] in the plasma-liquid system. If one wants to obtain the true concentration of the OHdis in the plasma-liquid system, one needs to know the consumption behavior of the OHdis-induced derivatives, the information of the OH scavengers (such as hydrated electron, atomic hydrogen besides the molecular probe), and also the knowledge of the H2<span class="hlt">O</span>2 formed from the OHdis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JAP...120n3102S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JAP...120n3102S"><span>Europium gallium garnet (Eu<span class="hlt">3</span>Ga5<span class="hlt">O</span>12) and Eu<span class="hlt">3</span>Ga<span class="hlt">O</span>6: Synthesis and material properties</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sawada, Kenji; Nakamura, Toshihiro; Adachi, Sadao</p> <p>2016-10-01</p> <p>Eu-Ga-<span class="hlt">O</span> ternary compounds were synthesized from a mixture of cubic (c-) Eu2<span class="hlt">O</span><span class="hlt">3</span> and monoclinic Ga2<span class="hlt">O</span><span class="hlt">3</span> (β-Ga2<span class="hlt">O</span><span class="hlt">3</span>) raw powders using the solid-state reaction method by calcination at Tc = 1200 °C. The structural and optical properties of the Eu-Ga-<span class="hlt">O</span> ternary compounds were investigated using X-ray diffraction analysis, photoluminescence (PL) analysis, PL excitation (PLE) spectroscopy, and Raman scattering measurements. Stoichiometric compounds such as cubic Eu<span class="hlt">3</span>Ga5<span class="hlt">O</span>12 (EGG) and orthorhombic Eu<span class="hlt">3</span>Ga<span class="hlt">O</span>6 were synthesized using molar ratios of x = 0.375 and 0.75 [x≡Eu2<span class="hlt">O</span><span class="hlt">3</span>/(Eu2<span class="hlt">O</span><span class="hlt">3</span> + Ga2<span class="hlt">O</span><span class="hlt">3</span>)], respectively, together with the end-point binary compounds β-Ga2<span class="hlt">O</span><span class="hlt">3</span> (x = 0) and monoclinic (m-) Eu2<span class="hlt">O</span><span class="hlt">3</span> (x = 1.0). The structural change from "cubic" to "monoclinic" in Eu2<span class="hlt">O</span><span class="hlt">3</span> is due to the structural phase transition occurring at Tc ≥ 1050 °C. In principle, the perovskite-type EuGa<span class="hlt">O</span><span class="hlt">3</span> and monoclinic Eu4Ga2<span class="hlt">O</span>9 can also be synthesized at x = 0.5 and 0.667, respectively; however, such stoichiometric compounds could not be synthesized in this study. The PL and PLE properties of EGG and Eu<span class="hlt">3</span>Ga<span class="hlt">O</span>6 were studied in detail. The temperature dependence of the PL spectra was observed through measurements carried out between T = 20 and 300 K and explained using a newly developed model. Raman scattering measurements were also performed on the Eu-Ga-<span class="hlt">O</span> ternary systems over the entire composition range from x = 0 (β-Ga2<span class="hlt">O</span><span class="hlt">3</span>) to 1.0 (m-Eu2<span class="hlt">O</span><span class="hlt">3</span>).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..MARB12005A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..MARB12005A"><span>Incipient 2D Mott insulators in extreme high electron density, ultra-thin GdTi<span class="hlt">O</span><span class="hlt">3</span>/SrTi<span class="hlt">O</span><span class="hlt">3</span>/GdTi<span class="hlt">O</span><span class="hlt">3</span> quantum wells</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Allen, S. James; Ouellette, Daniel G.; Moetakef, Pouya; Cain, Tyler; Chen, Ru; Balents, Leon; Stemmer, Susanne</p> <p>2013-03-01</p> <p>By reducing the number of Sr<span class="hlt">O</span> planes in a GdTi<span class="hlt">O</span><span class="hlt">3</span> /SrTi<span class="hlt">O</span><span class="hlt">3</span>/ GdTi<span class="hlt">O</span><span class="hlt">3</span> quantum well heterostructure, an electron gas with ~ fixed 2D electron density can be driven close to the Mott metal insulator transition - a quantum critical point at ~1 electron per unit cell. A single interface between the Mott insulator GdTi<span class="hlt">O</span><span class="hlt">3</span> and band insulator SrTi<span class="hlt">O</span><span class="hlt">3</span> has been shown to introduce ~ 1/2 electron per interface unit cell. Two interfaces produce a quantum well with ~ 7 1014 cm-2 electrons: at the limit of a single Sr<span class="hlt">O</span> layer it may produce a 2D magnetic Mott insulator. We use temperature and frequency dependent (DC - <span class="hlt">3</span>eV) conductivity and temperature dependent magneto-transport to understand the relative importance of electron-electron interactions, electron-phonon interactions, and surface roughness scattering as the electron gas is compressed toward the quantum critical point. Terahertz time-domain and FTIR spectroscopies, measure the frequency dependent carrier mass and scattering rate, and the mid-IR polaron absorption as a function of quantum well thickness. At the extreme limit of a single Sr<span class="hlt">O</span> plane, we observe insulating behavior with an optical gap substantially less than that of the surrounding GdTi<span class="hlt">O</span><span class="hlt">3</span>, suggesting a novel 2D Mott insulator. MURI program of the Army Research Office - Grant No. W911-NF-09-1-0398</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ApSS..266...17L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ApSS..266...17L"><span>Reaction between Ni<span class="hlt">O</span> and Al2<span class="hlt">O</span><span class="hlt">3</span> in Ni<span class="hlt">O</span>/γ-Al2<span class="hlt">O</span><span class="hlt">3</span> catalysts probed by positronium atom</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, C. Y.; Zhang, H. J.; Chen, Z. Q.</p> <p>2013-02-01</p> <p>Ni<span class="hlt">O</span>/γ-Al2<span class="hlt">O</span><span class="hlt">3</span> catalysts with Ni<span class="hlt">O</span> content of 9 wt% and 24 wt% were prepared by solid state reaction method. They are annealed in air at temperatures from 100 °C to 1000 °C. Positron lifetime spectra were measured to study the microstructure variation during annealing process. Four positron lifetime components were resolved with two long lifetime τ<span class="hlt">3</span> and τ4, which can be attributed to the ortho-positronium lifetime in microvoids and large pores, respectively. It was found that the longest lifetime τ4 is rather sensitive to the chemical environment of the large pores. The Ni<span class="hlt">O</span> active centers in the catalysts cause decrease of both τ4 and its intensity I4, which is due to the spin-conversion of positronium induced by Ni<span class="hlt">O</span>. However, after heating the catalysts above 600 °C, abnormal increase of the lifetime τ4 is observed. This is due to the formation of NiAl2<span class="hlt">O</span>4 spinel from the reaction of Ni<span class="hlt">O</span> and γ-Al2<span class="hlt">O</span><span class="hlt">3</span>. The generated NiAl2<span class="hlt">O</span>4 weakens the spin-conversion effect of positronium, thus leads to the increase of <span class="hlt">o</span>-Ps lifetime τ4. Formation of NiAl2<span class="hlt">O</span>4 is further confirmed by both X-ray diffraction and X-ray photoelectron spectroscopy measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006JPS...158.1538J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006JPS...158.1538J"><span>Nano Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span> mixed oxide and Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span>-carbon composite oxides as new and novel electrodes for supercapacitor applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jayalakshmi, M.; Venugopal, N.; Raja, K. Phani; Rao, M. Mohan</p> <p></p> <p>New nano-materials like Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span> and Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span>-carbon were synthesized by a single step hydrothermal method in searching for novel mixed oxides with high electrochemical double layer capacitance. A Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span>-carbon sample was calcined at 600 °C and tested for its performance. The source of carbon was tetrapropyl ammonium hydroxide. The capacitive behavior of Sn<span class="hlt">O</span> 2 was compared to the performance of Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span>, Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span>-carbon and calcined Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span>-carbon using the techniques of cyclic voltammetry, double potential step, chronopotentiometry and E-log I polarization. In 0.1 M NaCl solutions, Sn<span class="hlt">O</span> 2-Al 2<span class="hlt">O</span> <span class="hlt">3</span> gave the best performance with a value of 119 Fg -1 and cycled 1000 times. The nano-material mixed oxides were characterized by TEM, XRD, ICP-AES and SEM-EDAX.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003JAP....93..933K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003JAP....93..933K"><span>Infrared and Raman spectroscopy of [Pb(Zn1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>]0.92-[PbTi<span class="hlt">O</span><span class="hlt">3</span>]0.08 and [Pb(Mg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>]0.71-[PbTi<span class="hlt">O</span><span class="hlt">3</span>]0.29 single crystals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kamba, S.; Buixaderas, E.; Petzelt, J.; Fousek, J.; Nosek, J.; Bridenbaugh, P.</p> <p>2003-01-01</p> <p>Far-infrared reflectivity spectra of [Pb(Zn1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>]0.92-[PbTi<span class="hlt">O</span><span class="hlt">3</span>]0.08 and [Pb(Mg1/<span class="hlt">3</span>Nb2/<span class="hlt">3</span>)<span class="hlt">O</span><span class="hlt">3</span>]0.71-[PbTi<span class="hlt">O</span><span class="hlt">3</span>]0.29 single crystals were investigated between 10 and 530 K, micro-Raman spectra were recorded between 300 and 800 K. No phonon softening was observed near either of the ferroelectric phase transitions. The low-frequency dielectric anomaly in the paraelectric phase is caused by contribution of dynamic polar nanoclusters with the main dispersion in the microwave range. Infrared and Raman spectra confirm the locally doubled unit cell (Zprim=2) in the paraelectric and ferroelectric phases due to the ordering in the perovskite B sites and occurrence of polar nanoclusters in the paraelectric phase. The lowest-frequency transverse optical (TO1) phonon mode active in the infrared spectra is underdamped in contrast to the recent result of inelastic neutron scattering, where no TO1 mode could be observed for the wave vectors q⩽0.2 Å-1. This discrepancy was explained by different q vectors probed in infrared and neutron experiments. The infrared probe couples with very long-wavelength phonons (q≈10-5 Å-1) which see the homogeneous medium averaged over the nanoclusters, whereas the neutron probe couples with phonons whose wavelength is comparable to the nanocluster size (q⩾10-2 Å-1).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003PSSAR.200..346K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003PSSAR.200..346K"><span>Epitaxial structure and transport in LaTi<span class="hlt">O</span><span class="hlt">3</span>+x films on (001) SrTi<span class="hlt">O</span><span class="hlt">3</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, K. H.; Norton, D. P.; Budai, J. D.; Chisholm, M. F.; Sales, B. C.; Christen, D. K.; Cantoni, C.</p> <p>2003-12-01</p> <p>The structure and transport properties of LaTi<span class="hlt">O</span><span class="hlt">3</span>+x epitaxial thin films grown on (001) SrTi<span class="hlt">O</span><span class="hlt">3</span> by pulsed-laser deposition is examined. Four-circle X-ray diffraction indicates that the films possess the defect perovskite LaTi<span class="hlt">O</span><span class="hlt">3</span> structure when deposited in vacuum, with the higher X compounds forming at moderate oxygen pressures. The crystal structure of the LaTi<span class="hlt">O</span><span class="hlt">3</span> films is tetragonal in the epitaxial films, in contrast to the orthorhombic structure observed in bulk materials. A domain structure is observed in the films, consisting of LaTi<span class="hlt">O</span><span class="hlt">3</span> oriented either with the [110] or [001] directions perpendicular to the substrate surface. Z-contrast scanning transmission electron microscopy reveals that this domain structure is not present in the first few unit cells of the film, but emerges approximately 2-<span class="hlt">3</span> nm from the SrTi<span class="hlt">O</span><span class="hlt">3</span>/LaTi<span class="hlt">O</span><span class="hlt">3</span> interface. Upon increasing the oxygen pressure during growth, a shift in the lattice d-spacing parallel to the substrate surface is observed, and is consistent with the growth of the La2Ti2<span class="hlt">O</span>7 phase. However, van der Pauw measurements show that the films with the larger d-spacing remain conductive, albeit with a resistivity that is significantly higher than that for the perovskite LaTi<span class="hlt">O</span><span class="hlt">3</span> films. The transport behavior suggests that the films grown at higher oxygen pressures are LaTi<span class="hlt">O</span><span class="hlt">3</span>+x with 0.4 < x < 0.5. (</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JApSp..84..866V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JApSp..84..866V"><span>Luminescence of Eu:Y<span class="hlt">3</span>Al5<span class="hlt">O</span>12, Eu:Lu<span class="hlt">3</span>Al5<span class="hlt">O</span>12, and Eu:GdAl<span class="hlt">O</span><span class="hlt">3</span> Nanocrystals Synthesized by Solution Combustion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vilejshikova, E. V.; Khort, A. A.; Podbolotov, K. B.; Loiko, P. A.; Shimanski, V. I.; Shashkov, S. N.; Yumashev, K. V.</p> <p>2017-11-01</p> <p>Nanocrystals of rare-earth garnets Y<span class="hlt">3</span>Al5<span class="hlt">O</span>12 and Lu<span class="hlt">3</span>Al5<span class="hlt">O</span>12 and perovskite GdAl<span class="hlt">O</span><span class="hlt">3</span> highly doped (10-20 at%) with Eu<span class="hlt">3</span>+ are synthesized by the solution combustion technique and subsequent annealing in air at 800 and 1300<span class="hlt">o</span>C. Their structure, morphology, and phase composition are studied. These materials exhibit intense red luminescence under UV excitation. Eu:GdAl<span class="hlt">O</span><span class="hlt">3</span> luminescence has CIE 1931 color coordinates (0.632, 0.368); dominant wavelength, 599.6 nm; and color purity, >99%. Judd-Ofelt parameters, luminescence branching ratios, and lifetimes of the Eu<span class="hlt">3</span>+ 5D0 state are determined. The luminescence quantum yield for Eu:GdAl<span class="hlt">O</span><span class="hlt">3</span> (10 at%) reaches 74% with a lifetime of 1.4 ms for the 5D0 state. The synthesized materials are promising for red ceramic phosphors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26439469','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26439469"><span>Atmospheric Chemistry of (CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span>, (CF<span class="hlt">3</span>)2CHOCHO, and CF<span class="hlt">3</span>C(<span class="hlt">O</span>)OCH<span class="hlt">3</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Østerstrøm, Freja From; Wallington, Timothy J; Sulbaek Andersen, Mads P; Nielsen, Ole John</p> <p>2015-10-22</p> <p>Smog chambers with in situ FTIR detection were used to measure rate coefficients in 700 Torr of air and 296 ± 2 K of: k(Cl+(CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span>) = (5.41 ± 1.63) × 10(-12), k(Cl+(CF<span class="hlt">3</span>)2CHOCHO) = (9.44 ± 1.81) × 10(-15), k(Cl+CF<span class="hlt">3</span>C(<span class="hlt">O</span>)OCH<span class="hlt">3</span>) = (6.28 ± 0.98) × 10(-14), k(OH+(CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span>) = (1.86 ± 0.41) × 10(-13), and k(OH+(CF<span class="hlt">3</span>)2CHOCHO) = (2.08 ± 0.63) × 10(-14) cm(<span class="hlt">3</span>) molecule(-1) s(-1). The Cl atom initiated oxidation of (CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span> gives (CF<span class="hlt">3</span>)2CHOCHO in a yield indistinguishable from 100%. The OH radical initiated oxidation of (CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span> gives the following products (molar yields): (CF<span class="hlt">3</span>)2CHOCHO (76 ± 8)%, CF<span class="hlt">3</span>C(<span class="hlt">O</span>)OCH<span class="hlt">3</span> (16 ± 2)%, CF<span class="hlt">3</span>C(<span class="hlt">O</span>)CF<span class="hlt">3</span> (4 ± 1)%, and C(<span class="hlt">O</span>)F2 (45 ± 5)%. The primary oxidation product (CF<span class="hlt">3</span>)2CHOCHO reacts with Cl atoms to give secondary products (molar yields): CF<span class="hlt">3</span>C(<span class="hlt">O</span>)CF<span class="hlt">3</span> (67 ± 7)%, CF<span class="hlt">3</span>C(<span class="hlt">O</span>)OCHO (28 ± <span class="hlt">3</span>)%, and C(<span class="hlt">O</span>)F2 (118 ± 12)%. CF<span class="hlt">3</span>C(<span class="hlt">O</span>)OCH<span class="hlt">3</span> reacts with Cl atoms to give: CF<span class="hlt">3</span>C(<span class="hlt">O</span>)OCHO (80 ± 8)% and C(<span class="hlt">O</span>)F2 (6 ± 1)%. Atmospheric lifetimes of (CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span>, (CF<span class="hlt">3</span>)2CHOCHO, and CF<span class="hlt">3</span>C(<span class="hlt">O</span>)OCH<span class="hlt">3</span> were estimated to be 62 days, 1.5 years, and 220 days, respectively. The 100-year global warming potentials (GWPs) for (CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span>, (CF<span class="hlt">3</span>)2CHOCHO, and CF<span class="hlt">3</span>C(<span class="hlt">O</span>)OCH<span class="hlt">3</span> are estimated to be 6, 121, and 46, respectively. A comprehensive description of the atmospheric fate of (CF<span class="hlt">3</span>)2CHOCH<span class="hlt">3</span> is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AIPC.1278..175A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AIPC.1278..175A"><span><span class="hlt">O</span>2(a1Δ) Quenching In The <span class="hlt">O/O</span>2/<span class="hlt">O</span><span class="hlt">3</span> System</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Azyazov, V. N.; Mikheyev, P. A.; Postell, D.; Heaven, M. C.</p> <p>2010-10-01</p> <p>The development of discharge singlet oxygen generators (DSOG's) that can operate at high pressures is required for the power scaling of the discharge oxygen iodine laser. In order to achieve efficient high-pressure DSOG operation it is important to understand the mechanisms by which singlet oxygen (<span class="hlt">O</span>2(a1Δ)) is quenched in these devices. It has been proposed that three-body deactivation processes of the type <span class="hlt">O</span>2(a1Δ)+<span class="hlt">O</span>+M→2<span class="hlt">O</span>2+M provide significant energy loss channels. To further explore these reactions the physical and reactive quenching of <span class="hlt">O</span>2(a1Δ) in <span class="hlt">O</span>(<span class="hlt">3</span>P)/<span class="hlt">O</span>2/<span class="hlt">O</span><span class="hlt">3</span>/CO2/He/Ar mixtures has been investigated. Oxygen atoms and singlet oxygen molecules were produced by the 248 nm laser photolysis of ozone. The kinetics of <span class="hlt">O</span>2(a1Δ) quenching were followed by observing the 1268 nm fluorescence of the <span class="hlt">O</span>2a1Δ-X<span class="hlt">3</span>∑ transition. Fast quenching of <span class="hlt">O</span>2(a1Δ) in the presence of oxygen atoms and molecules was observed. The mechanism of the process has been examined using kinetic models, which indicate that quenching by vibrationally excited ozone is the dominant reaction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16397824','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16397824"><span>Upconversion properties of Er<span class="hlt">3</span>+/Yb<span class="hlt">3</span>+ co-doped Te<span class="hlt">O</span>2-Ti<span class="hlt">O</span>2-K2<span class="hlt">O</span> glasses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Su, Fangning; Deng, Zaide</p> <p>2006-01-01</p> <p>The Er<span class="hlt">3</span>+/Yb<span class="hlt">3</span>+ co-doped Te<span class="hlt">O</span>2-Ti<span class="hlt">O</span>2-K2<span class="hlt">O</span> glasses were prepared by conventional melting procedures, and their upconversion spectra were performed. The dependence of luminescence intensity on the ratio of Yb<span class="hlt">3</span>+/Er<span class="hlt">3</span>+ was studied, and the relationship between green upconversion luminescence intensity and Er<span class="hlt">3</span>+ concentration is discussed in detail. The 546 nm green upconversion luminescence intensity is optimised in the studied glasses either when the Yb<span class="hlt">3</span>+/Er<span class="hlt">3</span>+ ratio is 25/1 and Er<span class="hlt">3</span>+ concentration is 0.1 mol%, or when the Yb<span class="hlt">3</span>+/Er<span class="hlt">3</span>+ ratio is 10/1 and Er<span class="hlt">3</span>+ concentration is 0.15 mol%. These glasses could be one of the potential candidates for LD pumping microchip solid-state lasers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19070431','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19070431"><span>Preparation and catalytic activities of LaFe<span class="hlt">O</span><span class="hlt">3</span> and Fe2<span class="hlt">O</span><span class="hlt">3</span> for HMX thermal decomposition.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wei, Zhi-Xian; Xu, Yan-Qing; Liu, Hai-Yan; Hu, Chang-Wen</p> <p>2009-06-15</p> <p>Perovskite-type LaFe<span class="hlt">O</span>(<span class="hlt">3</span>) and alpha-Fe(2)<span class="hlt">O</span>(<span class="hlt">3</span>) with high specific surface areas were directly prepared with appropriate stearic acid-nitrates ratios by a novel stearic acid solution combustion method. The obtained powders were characterized by XRD, FT-IR and XPS techniques. The catalytic activities of perovskite-type LaFe<span class="hlt">O</span>(<span class="hlt">3</span>) and alpha-Fe(2)<span class="hlt">O</span>(<span class="hlt">3</span>) for the thermal decomposition of octahydro-1,<span class="hlt">3</span>,5,7-tetranitro-1,<span class="hlt">3</span>,5,7-tetrazocine (HMX) were investigated by TG and TG-EGA techniques. The experimental results show that the catalytic activity of perovskite-type LaFe<span class="hlt">O</span>(<span class="hlt">3</span>) was much higher than that of alpha-Fe(2)<span class="hlt">O</span>(<span class="hlt">3</span>) because of higher concentration of surface-adsorbed oxygen (<span class="hlt">O</span>(ad)) and hydroxyl of LaFe<span class="hlt">O</span>(<span class="hlt">3</span>). The study points out a potential way to develop new and more active perovskite-type catalysts for the HMX thermal decomposition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25754819','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25754819"><span>Faraday rotation and photoluminescence in heavily Tb(<span class="hlt">3</span>+)-doped Ge<span class="hlt">O</span>2-B2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses for fiber-integrated magneto-optics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gao, Guojun; Winterstein-Beckmann, Anja; Surzhenko, Oleksii; Dubs, Carsten; Dellith, Jan; Schmidt, Markus A; Wondraczek, Lothar</p> <p>2015-03-10</p> <p>We report on the magneto-optical (MO) properties of heavily Tb(<span class="hlt">3</span>+)-doped Ge<span class="hlt">O</span>2-B2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses towards fiber-integrated paramagnetic MO devices. For a Tb(<span class="hlt">3</span>+) ion concentration of up to 9.7 × 10(21) cm(-<span class="hlt">3</span>), the reported glass exhibits an absolute negative Faraday rotation of ~120 rad/T/m at 632.8 nm. The optimum spectral ratio between Verdet constant and light transmittance over the spectral window of 400-1500 nm is found for a Tb(<span class="hlt">3</span>+) concentration of ~6.5 × 10(21) cm(-<span class="hlt">3</span>). For this glass, the crystallization stability, expressed as the difference between glass transition temperature and onset temperature of melt crystallization exceeds 100 K, which is a prerequisite for fiber drawing. In addition, a high activation energy of crystallization is achieved at this composition. Optical absorption occurs in the NUV and blue spectral region, accompanied by Tb(<span class="hlt">3</span>+) photoluminescence. In the heavily doped materials, a UV/blue-to-green photo-conversion gain of ~43% is achieved. The lifetime of photoluminescence is ~2.2 ms at a stimulated emission cross-section σem of ~1.1 × 10(-21) cm(2) for ~ 5.0 × 10(21) cm(-<span class="hlt">3</span>) Tb(<span class="hlt">3</span>+). This results in an optical gain parameter σem*τ of ~2.5 × 10(-24) cm(2)s, what could be of interest for implementation of a Tb(<span class="hlt">3</span>+) fiber laser.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4354094','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4354094"><span>Faraday rotation and photoluminescence in heavily Tb<span class="hlt">3</span>+-doped Ge<span class="hlt">O</span>2-B2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses for fiber-integrated magneto-optics</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gao, Guojun; Winterstein-Beckmann, Anja; Surzhenko, Oleksii; Dubs, Carsten; Dellith, Jan; Schmidt, Markus A.; Wondraczek, Lothar</p> <p>2015-01-01</p> <p>We report on the magneto-optical (MO) properties of heavily Tb<span class="hlt">3</span>+-doped Ge<span class="hlt">O</span>2-B2<span class="hlt">O</span><span class="hlt">3</span>-Al2<span class="hlt">O</span><span class="hlt">3</span>-Ga2<span class="hlt">O</span><span class="hlt">3</span> glasses towards fiber-integrated paramagnetic MO devices. For a Tb<span class="hlt">3</span>+ ion concentration of up to 9.7 × 1021 cm−<span class="hlt">3</span>, the reported glass exhibits an absolute negative Faraday rotation of ~120 rad/T/m at 632.8 nm. The optimum spectral ratio between Verdet constant and light transmittance over the spectral window of 400–1500 nm is found for a Tb<span class="hlt">3</span>+ concentration of ~6.5 × 1021 cm−<span class="hlt">3</span>. For this glass, the crystallization stability, expressed as the difference between glass transition temperature and onset temperature of melt crystallization exceeds 100 K, which is a prerequisite for fiber drawing. In addition, a high activation energy of crystallization is achieved at this composition. Optical absorption occurs in the NUV and blue spectral region, accompanied by Tb<span class="hlt">3</span>+ photoluminescence. In the heavily doped materials, a UV/blue-to-green photo-conversion gain of ~43% is achieved. The lifetime of photoluminescence is ~2.2 ms at a stimulated emission cross-section σem of ~1.1 × 10−21 cm2 for ~ 5.0 × 1021 cm−<span class="hlt">3</span> Tb<span class="hlt">3</span>+. This results in an optical gain parameter σem*τ of ~2.5 × 10−24 cm2s, what could be of interest for implementation of a Tb<span class="hlt">3</span>+ fiber laser. PMID:25754819</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AIPC.1657d0016S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AIPC.1657d0016S"><span>Effects of frit addition on the surface morphology and structural properties of Zn<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-Mn2<span class="hlt">O</span><span class="hlt">3</span> discs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shahardin, Ahmad Hajidi; Mahmud, Shahrom; Sendi, Rabab Khalid</p> <p>2015-04-01</p> <p>Zn<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-Mn2<span class="hlt">O</span><span class="hlt">3</span> discs were prepared using conventional ceramic processing method and sintered at 1000°C. The different percentages of frit on the Zn<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-Mn2<span class="hlt">O</span><span class="hlt">3</span> discs were 0.0%, 0.5%, 1.0% and <span class="hlt">3</span>.0%. From FESEM observation, the grain structure and grain growth were more uniformly constructed and well distributed. Frit addition was found to cause a big drop in the average grain size from 4.59 µm to 2.76 µm even with an addition of 0.5 mol%. The Si and Al content in the frit recipe might have played a role as inhibiting agents in grain growth during sintering. RAMAN intensity and phase shifting were not affected by frit addition except at <span class="hlt">3</span> mol%. Frit addition did not affect the formation of secondary phases. Frit addition below <span class="hlt">3</span> mol% in Zn<span class="hlt">O</span>-Bi2<span class="hlt">O</span><span class="hlt">3</span>-Mn2<span class="hlt">O</span><span class="hlt">3</span> varistor discs can be used as a method in controlling grain size without affecting other properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008APS..MARV31007K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008APS..MARV31007K"><span>Fabrication and characterization of complex oxide RENi<span class="hlt">O</span><span class="hlt">3</span>/LaAl<span class="hlt">O</span><span class="hlt">3</span> superlattices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kareev, M.; Freeland, J. W.; Liu, J.; Kirby, B.; Keimer, B.; Chakhalian, J.</p> <p>2008-03-01</p> <p>Nowadays there has been growing interest to synthesis of atomically thin complex oxide superlattices which can result in novel electronic and magnetic properties at the interface. Here we report on digital synthesis of single unit cell nickel based heterostructures of RENi<span class="hlt">O</span><span class="hlt">3</span>/LaAl<span class="hlt">O</span><span class="hlt">3</span> (RE = La, Nd and Pr) superlattices on SrTi<span class="hlt">O</span><span class="hlt">3</span> and LaAl<span class="hlt">O</span><span class="hlt">3</span> by laser MBE. RHEED analysis, grazing angle XRD and AFM imaging have confirmed the high quality of the epitaxially grown superlattices. The magnetic and electronic properties of the superlattices have been elucidated by polarized X-ray spectroscopies, which show a non-trivial evolution of magnetism and charge of the LNO layer with increasing LNO layer thickness. The work has been supported by U.S. DOD-ARO under Contract No. 0402-17291.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_25 --> <div class="footer-extlink text-muted" style="margin-bottom:1rem; text-align:center;">Some links on this page may take you to non-federal websites. Their policies may differ from this site.</div> </div><!-- container --> <footer><a id="backToTop" href="#top"> </a><nav><a id="backToTop" href="#top"> </a><ul class="links"><a id="backToTop" href="#top"> </a><li><a id="backToTop" href="#top"></a><a href="/sitemap.html">Site Map</a></li> <li><a href="/members/index.html">Members Only</a></li> <li><a href="/website-policies.html">Website Policies</a></li> <li><a href="https://doe.responsibledisclosure.com/hc/en-us" target="_blank">Vulnerability Disclosure Program</a></li> <li><a href="/contact.html">Contact Us</a></li> </ul> <div class="small">Science.gov is maintained by the U.S. Department of Energy's <a href="https://www.osti.gov/" target="_blank">Office of Scientific and Technical Information</a>, in partnership with <a href="https://www.cendi.gov/" target="_blank">CENDI</a>.</div> </nav> </footer> <script type="text/javascript"><!-- // var lastDiv = ""; function showDiv(divName) { // hide last div if (lastDiv) { document.getElementById(lastDiv).className = "hiddenDiv"; } //if value of the box is not nothing and an object with that name exists, then change the class if (divName && document.getElementById(divName)) { document.getElementById(divName).className = "visibleDiv"; lastDiv = divName; } } //--> </script> <script> /** * Function that tracks a click on an outbound link in Google Analytics. * This function takes a valid URL string as an argument, and uses that URL string * as the event label. */ var trackOutboundLink = function(url,collectionCode) { try { h = window.open(url); setTimeout(function() { ga('send', 'event', 'topic-page-click-through', collectionCode, url); }, 1000); } catch(err){} }; </script> <!-- Google Analytics --> <script> (function(i,s,o,g,r,a,m){i['GoogleAnalyticsObject']=r;i[r]=i[r]||function(){ (i[r].q=i[r].q||[]).push(arguments)},i[r].l=1*new Date();a=s.createElement(o), m=s.getElementsByTagName(o)[0];a.async=1;a.src=g;m.parentNode.insertBefore(a,m) })(window,document,'script','//www.google-analytics.com/analytics.js','ga'); ga('create', 'UA-1122789-34', 'auto'); ga('send', 'pageview'); </script> <!-- End Google Analytics --> <script> showDiv('page_1') </script> </body> </html>