Sample records for air-water gas exchange

  1. Continuous measurement of air-water gas exchange by underwater eddy covariance

    NASA Astrophysics Data System (ADS)

    Berg, Peter; Pace, Michael L.

    2017-12-01

    Exchange of gases, such as O2, CO2, and CH4, over the air-water interface is an important component in aquatic ecosystem studies, but exchange rates are typically measured or estimated with substantial uncertainties. This diminishes the precision of common ecosystem assessments associated with gas exchanges such as primary production, respiration, and greenhouse gas emission. Here, we used the aquatic eddy covariance technique - originally developed for benthic O2 flux measurements - right below the air-water interface (˜ 4 cm) to determine gas exchange rates and coefficients. Using an acoustic Doppler velocimeter and a fast-responding dual O2-temperature sensor mounted on a floating platform the 3-D water velocity, O2 concentration, and temperature were measured at high-speed (64 Hz). By combining these data, concurrent vertical fluxes of O2 and heat across the air-water interface were derived, and gas exchange coefficients were calculated from the former. Proof-of-concept deployments at different river sites gave standard gas exchange coefficients (k600) in the range of published values. A 40 h long deployment revealed a distinct diurnal pattern in air-water exchange of O2 that was controlled largely by physical processes (e.g., diurnal variations in air temperature and associated air-water heat fluxes) and not by biological activity (primary production and respiration). This physical control of gas exchange can be prevalent in lotic systems and adds uncertainty to assessments of biological activity that are based on measured water column O2 concentration changes. For example, in the 40 h deployment, there was near-constant river flow and insignificant winds - two main drivers of lotic gas exchange - but we found gas exchange coefficients that varied by several fold. This was presumably caused by the formation and erosion of vertical temperature-density gradients in the surface water driven by the heat flux into or out of the river that affected the turbulent

  2. On factors influencing air-water gas exchange in emergent wetlands

    USGS Publications Warehouse

    Ho, David T.; Engel, Victor C.; Ferron, Sara; Hickman, Benjamin; Choi, Jay; Harvey, Judson W.

    2018-01-01

    Knowledge of gas exchange in wetlands is important in order to determine fluxes of climatically and biogeochemically important trace gases and to conduct mass balances for metabolism studies. Very few studies have been conducted to quantify gas transfer velocities in wetlands, and many wind speed/gas exchange parameterizations used in oceanographic or limnological settings are inappropriate under conditions found in wetlands. Here six measurements of gas transfer velocities are made with SF6 tracer release experiments in three different years in the Everglades, a subtropical peatland with surface water flowing through emergent vegetation. The experiments were conducted under different flow conditions and with different amounts of emergent vegetation to determine the influence of wind, rain, water flow, waterside thermal convection, and vegetation on air-water gas exchange in wetlands. Measured gas transfer velocities under the different conditions ranged from 1.1 cm h−1 during baseline conditions to 3.2 cm h−1 when rain and water flow rates were high. Commonly used wind speed/gas exchange relationships would overestimate the gas transfer velocity by a factor of 1.2 to 6.8. Gas exchange due to thermal convection was relatively constant and accounted for 14 to 51% of the total measured gas exchange. Differences in rain and water flow among the different years were responsible for the variability in gas exchange, with flow accounting for 37 to 77% of the gas exchange, and rain responsible for up to 40%.

  3. Turbulence and wave breaking effects on air-water gas exchange

    PubMed

    Boettcher; Fineberg; Lathrop

    2000-08-28

    We present an experimental characterization of the effects of turbulence and breaking gravity waves on air-water gas exchange in standing waves. We identify two regimes that govern aeration rates: turbulent transport when no wave breaking occurs and bubble dominated transport when wave breaking occurs. In both regimes, we correlate the qualitative changes in the aeration rate with corresponding changes in the wave dynamics. In the latter regime, the strongly enhanced aeration rate is correlated with measured acoustic emissions, indicating that bubble creation and dynamics dominate air-water exchange.

  4. Gas exchange rates across the sediment-water and air-water interfaces in south San Francisco Bay

    USGS Publications Warehouse

    Hartman, Blayne; Hammond, Douglas E.

    1984-01-01

    Radon 222 concentrations in the water and sedimentary columns and radon exchange rates across the sediment-water and air-water interfaces have been measured in a section of south San Francisco Bay. Two independent methods have been used to determine sediment-water exchange rates, and the annual averages of these methods agree within the uncertainty of the determinations, about 20%. The annual average of benthic fluxes from shoal areas is nearly a factor of 2 greater than fluxes from the channel areas. Fluxes from the shoal and channel areas exceed those expected from simple molecular diffusion by factors of 4 and 2, respectively, apparently due to macrofaunal irrigation. Values of the gas transfer coefficient for radon exchange across the air-water interface were determined by constructing a radon mass balance for the water column and by direct measurement using floating chambers. The chamber method appears to yield results which are too high. Transfer coefficients computed using the mass balance method range from 0.4 m/day to 1.8 m/day, with a 6-year average of 1.0 m/day. Gas exchange is linearly dependent upon wind speed over a wind speed range of 3.2–6.4 m/s, but shows no dependence upon current velocity. Gas transfer coefficients predicted from an empirical relationship between gas exchange rates and wind speed observed in lakes and the oceans are within 30% of the coefficients determined from the radon mass balance and are considerably more accurate than coefficients predicted from theoretical gas exchange models.

  5. Air-water gas exchange and CO2 flux in a mangrove-dominated estuary

    USGS Publications Warehouse

    Ho, David T.; Ferrón, Sara; Engel, Victor C.; Larsen, Laurel G.; Barr, Jordan G.

    2014-01-01

    Mangrove forests are highly productive ecosystems, but the fate of mangrove-derived carbon remains uncertain. Part of that uncertainty stems from the fact that gas transfer velocities in mangrove-surrounded waters are not well determined, leading to uncertainty in air-water CO2 fluxes. Two SF6 tracer release experiments were conducted to determine gas transfer velocities (k(600) = 8.3 ± 0.4 and 8.1 ± 0.6 cm h−1), along with simultaneous measurements of pCO2 to determine the air-water CO2 fluxes from Shark River, Florida (232.11 ± 23.69 and 171.13 ± 20.28 mmol C m−2 d−1), an estuary within the largest contiguous mangrove forest in North America. The gas transfer velocity results are consistent with turbulent kinetic energy dissipation measurements, indicating a higher rate of turbulence and gas exchange than predicted by commonly used wind speed/gas exchange parameterizations. The results have important implications for carbon fluxes in mangrove ecosystems.

  6. Air-water Gas Exchange Rates on a Large Impounded River Measured Using Floating Domes (Poster)

    EPA Science Inventory

    Mass balance models of dissolved gases in rivers typically serve as the basis for whole-system estimates of greenhouse gas emission rates. An important component of these models is the exchange of dissolved gases between air and water. Controls on gas exchange rates (K) have be...

  7. Wind driven vertical transport in a vegetated, wetland water column with air-water gas exchange

    NASA Astrophysics Data System (ADS)

    Poindexter, C.; Variano, E. A.

    2010-12-01

    Flow around arrays of cylinders at low and intermediate Reynolds numbers has been studied numerically, analytically and experimentally. Early results demonstrated that at flow around randomly oriented cylinders exhibits reduced turbulent length scales and reduced diffusivity when compared to similarly forced, unimpeded flows (Nepf 1999). While horizontal dispersion in flows through cylinder arrays has received considerable research attention, the case of vertical dispersion of reactive constituents has not. This case is relevant to the vertical transfer of dissolved gases in wetlands with emergent vegetation. We present results showing that the presence of vegetation can significantly enhance vertical transport, including gas transfer across the air-water interface. Specifically, we study a wind-sheared air-water interface in which randomly arrayed cylinders represent emergent vegetation. Wind is one of several processes that may govern physical dispersion of dissolved gases in wetlands. Wind represents the dominant force for gas transfer across the air-water interface in the ocean. Empirical relationships between wind and the gas transfer coefficient, k, have been used to estimate spatial variability of CO2 exchange across the worlds’ oceans. Because wetlands with emergent vegetation are different from oceans, different model of wind effects is needed. We investigated the vertical transport of dissolved oxygen in a scaled wetland model built inside a laboratory tank equipped with an open-ended wind tunnel. Plastic tubing immersed in water to a depth of approximately 40 cm represented emergent vegetation of cylindrical form such as hard-stem bulrush (Schoenoplectus acutus). After partially removing the oxygen from the tank water via reaction with sodium sulfite, we used an optical probe to measure dissolved oxygen at mid-depth as the tank water re-equilibrated with the air above. We used dissolved oxygen time-series for a range of mean wind speeds to estimate the

  8. Aqueous turbulence structure immediately adjacent to the air - water interface and interfacial gas exchange

    NASA Astrophysics Data System (ADS)

    Wang, Binbin

    Air-sea interaction and the interfacial exchange of gas across the air-water interface are of great importance in coupled atmospheric-oceanic environmental systems. Aqueous turbulence structure immediately adjacent to the air-water interface is the combined result of wind, surface waves, currents and other environmental forces and plays a key role in energy budgets, gas fluxes and hence the global climate system. However, the quantification of turbulence structure sufficiently close to the air-water interface is extremely difficult. The physical relationship between interfacial gas exchange and near surface turbulence remains insufficiently investigated. This dissertation aims to measure turbulence in situ in a complex environmental forcing system on Lake Michigan and to reveal the relationship between turbulent statistics and the CO2 flux across the air-water interface. The major objective of this dissertation is to investigate the physical control of the interfacial gas exchange and to provide a universal parameterization of gas transfer velocity from environmental factors, as well as to propose a mechanistic model for the global CO2 flux that can be applied in three dimensional climate-ocean models. Firstly, this dissertation presents an advanced measurement instrument, an in situ free floating Particle Image Velocimetry (FPIV) system, designed and developed to investigate the small scale turbulence structure immediately below the air-water interface. Description of hardware components, design of the system, measurement theory, data analysis procedure and estimation of measurement error were provided. Secondly, with the FPIV system, statistics of small scale turbulence immediately below the air-water interface were investigated under a variety of environmental conditions. One dimensional wave-number spectrum and structure function sufficiently close to the water surface were examined. The vertical profiles of turbulent dissipation rate were intensively studied

  9. The Effect of Rain on Air-Water Gas Exchange

    NASA Technical Reports Server (NTRS)

    Ho, David T.; Bliven, Larry F.; Wanninkhof, Rik; Schlosser, Peter

    1997-01-01

    The relationship between gas transfer velocity and rain rate was investigated at NASA's Rain-Sea Interaction Facility (RSIF) using several SF, evasion experiments. During each experiment, a water tank below the rain simulator was supersaturated with SF6, a synthetic gas, and the gas transfer velocities were calculated from the measured decrease in SF6 concentration with time. The results from experiments with IS different rain rates (7 to 10 mm/h) and 1 of 2 drop sizes (2.8 or 4.2 mm diameter) confirm a significant and systematic enhancement of air-water gas exchange by rainfall. The gas transfer velocities derived from our experiment were related to the kinetic energy flux calculated from the rain rate and drop size. The relationship obtained for mono-dropsize rain at the RSIF was extrapolated to natural rain using the kinetic energy flux of natural rain calculated from the Marshall-Palmer raindrop size distribution. Results of laboratory experiments at RSIF were compared to field observations made during a tropical rainstorm in Miami, Florida and show good agreement between laboratory and field data.

  10. Reprint of: A numerical modelling of gas exchange mechanisms between air and turbulent water with an aquarium chemical reaction

    NASA Astrophysics Data System (ADS)

    Nagaosa, Ryuichi S.

    2014-08-01

    This paper proposes a new numerical modelling to examine environmental chemodynamics of a gaseous material exchanged between the air and turbulent water phases across a gas-liquid interface, followed by an aquarium chemical reaction. This study uses an extended concept of a two-compartment model, and assumes two physicochemical substeps to approximate the gas exchange processes. The first substep is the gas-liquid equilibrium between the air and water phases, A(g)⇌A(aq), with Henry's law constant H. The second is a first-order irreversible chemical reaction in turbulent water, A(aq)+H2O→B(aq)+H+ with a chemical reaction rate κA. A direct numerical simulation (DNS) technique has been employed to obtain details of the gas exchange mechanisms and the chemical reaction in the water compartment, while zero velocity and uniform concentration of A is considered in the air compartment. The study uses the different Schmidt numbers between 1 and 8, and six nondimensional chemical reaction rates between 10(≈0) to 101 at a fixed Reynolds number. It focuses on the effects of the Schmidt number and the chemical reaction rate on fundamental mechanisms of the gas exchange processes across the interface.

  11. Surfactant control of air-sea gas exchange across contrasting biogeochemical regimes

    NASA Astrophysics Data System (ADS)

    Pereira, Ryan; Schneider-Zapp, Klaus; Upstill-Goddard, Robert

    2014-05-01

    Air-sea gas exchange is important to the global partitioning of CO2.Exchange fluxes are products of an air-sea gas concentration difference, ΔC, and a gas transfer velocity, kw. The latter is controlled by the rate of turbulent diffusion at the air-sea interface but it cannot be directly measured and has a high uncertainty that is now considered one of the greatest challenges to quantifying net global air-sea CO2 exchange ...(Takahashi et al., 2009). One important control on kw is exerted by sea surface surfactants that arise both naturally from biological processes and through anthropogenic activity. They influence gas exchange in two fundamental ways: as a monolayer physical barrier and through modifying sea surface hydrodynamics and hence turbulent energy transfer. These effects have been demonstrated in the laboratory with artificial surfactants ...(Bock et al., 1999; Goldman et al., 1988) and through purposeful surfactant releases in coastal waters .(.).........().(Brockmann et al., 1982) and in the open ocean (Salter et al., 2011). Suppression of kwin these field experiments was ~5-55%. While changes in both total surfactant concentration and the composition of the natural surfactant pool might be expected to impact kw, the required in-situ studies are lacking. New data collected from the coastal North Sea in 2012-2013 shows significant spatio-temporal variability in the surfactant activity of organic matter within the sea surface microlayer that ranges from 0.07-0.94 mg/L T-X-100 (AC voltammetry). The surfactant activities show a strong winter/summer seasonal bias and general decrease in concentration with increasing distance from the coastline possibly associated with changing terrestrial vs. phytoplankton sources. Gas exchange experiments of this seawater using a novel laboratory tank and gas tracers (CH4 and SF6) demonstrate a 12-45% reduction in kw compared to surfactant-free water. Seasonally there is higher gas exchange suppression in the summer

  12. Influence of current velocity and wind speed on air-water gas exchange in a mangrove estuary

    NASA Astrophysics Data System (ADS)

    Ho, David T.; Coffineau, Nathalie; Hickman, Benjamin; Chow, Nicholas; Koffman, Tobias; Schlosser, Peter

    2016-04-01

    Knowledge of air-water gas transfer velocities and water residence times is necessary to study the fate of mangrove derived carbon exported into surrounding estuaries and ultimately to determine carbon balances in mangrove ecosystems. For the first time, the 3He/SF6 dual tracer technique, which has been proven to be a powerful tool to determine gas transfer velocities in the ocean, is applied to Shark River, an estuary situated in the largest contiguous mangrove forest in North America. The mean gas transfer velocity was 3.3 ± 0.2 cm h-1 during the experiment, with a water residence time of 16.5 ± 2.0 days. We propose a gas exchange parameterization that takes into account the major sources of turbulence in the estuary (i.e., bottom generated shear and wind stress).

  13. On the parameters influencing air-water gas exchange

    NASA Astrophysics Data System (ADS)

    JäHne, Bernd; Münnich, Karl Otto; BöSinger, Rainer; Dutzi, Alfred; Huber, Werner; Libner, Peter

    1987-02-01

    Detailed gas exchange measurements from two circular and one linear wind/wave tunnels are presented. Heat, He, CH4, CO2, Kr, and Xe have been used as tracers. The experiments show the central importance of waves for the water-side transfer process. With the onset of waves the Schmidt number dependence of the transfer velocity k changes from k ∝ Sc-⅔ to k ∝ Sc-½indicating a change in the boundary conditions at the surface. Moreover, energy put into the wave field by wind is transferred to near-surface turbulence enhancing gas transfer. The data show that the mean square slope of the waves is the best parameter to characterize the free wavy surface with respect to water-side transfer processes.

  14. Air-water gas exchange of chlorinated pesticides in four lakes spanning a 1,205 meter elevation range in the Canadian Rocky Mountains.

    PubMed

    Wilkinson, Andrew C; Kimpe, Lynda E; Blais, Jules M

    2005-01-01

    Concentrations of selected persistent organic pollutants (POPs) in air and water were measured from four lakes that transect the Canadian Rocky Mountains. These data were used in combination with wind velocity and temperature-adjusted Henry's law constants to estimate the direction and magnitude of chemical exchange across the air-water interface of these lakes. Bow Lake (1,975 m above sea level [masl]) was studied during the summers of 1998 through 2000; Donald (770 masl) was studied during the summer of 1999; Dixon Dam Lake (946 masl) and Kananaskis Lake (1,667 masl) were studied during the summer of 2000. Hexachlorobenzene (HCB) and dieldrin volatilized from Bow Lake in spring and summer of 1998 to 2000 at a rate of 0.92 +/-1.1 and 0.55+/-0.37 ng m(-2) d(-1), respectively. The alpha-endosulfan deposited to Bow Lake at a rate of 3.4+/-2.2 ng m(-2) d(-1). Direction of gas exchange for gamma-hexachlorocyclohexane (gamma-HCH) changed from net deposition in 1998 to net volatilization in 1999, partly because of a surge in y-HCH concentrations in the water at Bow Lake in 1999. Average gamma-HCH concentrations in air declined steadily over the three-year period, from 0.021 ng m(-3) in 1998, to 0.0023 ng m(-3) in 2000, and to volatilization in 1999 and 2000. Neither the concentrations of organochlorine compounds (OCs) in air and water, nor the direction and rate of air-water gas exchange correlate with temperature or elevation. In general, losses of pesticides by outflow were greater than the amount exchanged across the air-water interface in these lakes.

  15. Spume Drops: Their Potential Role in Air-Sea Gas Exchange

    NASA Astrophysics Data System (ADS)

    Monahan, Edward C.; Staniec, Allison; Vlahos, Penny

    2017-12-01

    After summarizing the time scales defining the change of the physical properties of spume and other droplets cast up from the sea surface, the time scales governing drop-atmosphere gas exchange are compared. Following a broad review of the spume drop production functions described in the literature, a subset of these functions is selected via objective criteria, to represent typical, upper bound, and lower bound production functions. Three complementary mechanisms driving spume-atmosphere gas exchange are described, and one is then used to estimate the relative importance, over a broad range of wind speeds, of this spume drop mechanism compared to the conventional, diffusional, sea surface mechanism in air-sea gas exchange. While remaining uncertainties in the wind dependence of the spume drop production flux, and in the immediate sea surface gas flux, preclude a definitive conclusion, the findings of this study strongly suggest that, at high wind speeds (>20 m s-1 for dimethyl sulfide and >30 m s-1 for gases such a carbon dioxide), spume drops do make a significant contribution to air-sea gas exchange.Plain Language SummaryThis paper evaluates the existing spume drop generation functions available to date and selects a reasonable upper, lower and mid range function that are reasonable for use in <span class="hlt">air</span> sea <span class="hlt">exchange</span> models. Based on these the contribution of spume drops to overall <span class="hlt">air</span> sea <span class="hlt">gas</span> <span class="hlt">exchange</span> at different wind speeds is then evaluated to determine the % contribution of spume. Generally below 20ms-1 spume drops contribute <1% of <span class="hlt">gas</span> <span class="hlt">exchange</span> but may account for a significant amount of <span class="hlt">gas</span> <span class="hlt">exchange</span> at higher wind speeds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002EGSGA..27..874S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002EGSGA..27..874S"><span>Observational Studies of Parameters Influencing <span class="hlt">Air</span>-sea <span class="hlt">Gas</span> <span class="hlt">Exchange</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schimpf, U.; Frew, N. M.; Bock, E. J.; Hara, T.; Garbe, C. S.; Jaehne, B.</p> <p></p> <p>A physically-based modeling of the <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer that can be used to predict the <span class="hlt">gas</span> transfer rates with sufficient accuracy as a function of micrometeorological parameters is still lacking. State of the art are still simple <span class="hlt">gas</span> transfer rate/wind speed relationships. Previous measurements from Coastal Ocean Experiment in the Atlantic revealed positive correlations between mean square slope, near surface turbulent dis- sipation, and wind stress. It also demonstrated a strong negative correlation between mean square slope and the fluorescence of surface-enriched colored dissolved organic matter. Using heat as a proxy tracer for gases the <span class="hlt">exchange</span> process at the <span class="hlt">air/water</span> interface and the micro turbulence at the <span class="hlt">water</span> surface can be investigated. The anal- ysis of infrared image sequences allow the determination of the net heat flux at the ocean surface, the temperature gradient across the <span class="hlt">air</span>/sea interface and thus the heat transfer velocity and <span class="hlt">gas</span> transfer velocity respectively. Laboratory studies were carried out in the new Heidelberg wind-wave facility AELOTRON. Direct measurements of the Schmidt number exponent were done in conjunction with classical mass balance methods to estimate the transfer velocity. The laboratory results allowed to validate the basic assumptions of the so called controlled flux technique by applying differ- ent tracers for the <span class="hlt">gas</span> <span class="hlt">exchange</span> in a large Schmidt number regime. Thus a modeling of the Schmidt number exponent is able to fill the gap between laboratory and field measurements field. Both, the results from the laboratory and the field measurements should be able to give a further understanding of the mechanisms controlling the trans- port processes across the aqueous boundary layer and to relate the forcing functions to parameters measured by remote sensing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21141036','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21141036"><span>Advances in quantifying <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> and environmental forcing.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wanninkhof, Rik; Asher, William E; Ho, David T; Sweeney, Colm; McGillis, Wade R</p> <p>2009-01-01</p> <p>The past decade has seen a substantial amount of research on <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> and its environmental controls. These studies have significantly advanced the understanding of processes that control <span class="hlt">gas</span> transfer, led to higher quality field measurements, and improved estimates of the flux of climate-relevant gases between the ocean and atmosphere. This review discusses the fundamental principles of <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer and recent developments in <span class="hlt">gas</span> transfer theory, parameterizations, and measurement techniques in the context of the <span class="hlt">exchange</span> of carbon dioxide. However, much of this discussion is applicable to any sparingly soluble, non-reactive <span class="hlt">gas</span>. We show how the use of global variables of environmental forcing that have recently become available and <span class="hlt">gas</span> <span class="hlt">exchange</span> relationships that incorporate the main forcing factors will lead to improved estimates of global and regional <span class="hlt">air</span>-sea <span class="hlt">gas</span> fluxes based on better fundamental physical, chemical, and biological foundations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22418709','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22418709"><span>Discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span>, <span class="hlt">water</span> loss, and metabolism in Protaetia cretica (Cetoniinae, Scarabaeidae).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matthews, Philip G D; White, Craig R</p> <p>2012-01-01</p> <p>Insects are at high risk of desiccation because of their small size, high surface-area-to-volume ratio, and <span class="hlt">air</span>-filled tracheal system that ramifies throughout their bodies to transport O(2) and CO(2) to and from respiring cells. Although the tracheal system offers a high-conductance pathway for the movement of respiratory gases, it has the unintended consequence of allowing respiratory transpiration to the atmosphere. When resting, many species <span class="hlt">exchange</span> respiratory gases discontinuously, and an early hypothesis for the origin of these discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> cycles (DGCs) is that they serve to reduce respiratory <span class="hlt">water</span> loss. In this study, we test this "hygric" hypothesis by comparing rates of CO(2) <span class="hlt">exchange</span> and <span class="hlt">water</span> loss among flower beetles Protaetia cretica (Cetoniinae, Scarabaeidae) breathing either continuously or discontinuously. We show that, consistent with the expectations of the hygric hypothesis, rates of total <span class="hlt">water</span> loss are higher during continuous <span class="hlt">gas</span> <span class="hlt">exchange</span> than during discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> and that the ratio of respiratory <span class="hlt">water</span> loss to CO(2) <span class="hlt">exchange</span> is lower during discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span>. This conclusion is in agreement with other studies of beetles and cockroaches that also support the hygric hypothesis. However, this result does not exclude other adaptive hypotheses supported by work on ants and moth pupae. This ambiguity may arise because there are multiple independent evolutionary origins of DGCs and no single adaptive function underlying their genesis. Alternatively, the observed reduction in <span class="hlt">water</span> loss during DGCs may be a side effect of a nonadaptive <span class="hlt">gas</span> <span class="hlt">exchange</span> pattern that is elicited during periods of inactivity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20000740-coupling-phytoplankton-uptake-air-water-exchange-persistent-organic-pollutants','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20000740-coupling-phytoplankton-uptake-air-water-exchange-persistent-organic-pollutants"><span>Coupling of phytoplankton uptake and <span class="hlt">air-water</span> <span class="hlt">exchange</span> of persistent organic pollutants</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>Dachs, J.; Eisenreich, S.J.; Baker, J.E.</p> <p>1999-10-15</p> <p>A dynamic model that couples <span class="hlt">air-water</span> <span class="hlt">exchange</span> and phytoplankton uptake of persistent organic pollutants has been developed and then applied to PCB data from a small experimental lake. A sensitivity analysis of the model, taking into account the influence of physical environmental conditions such as temperature, wind speed, and mixing depth as well as plankton-related parameters such as biomass and growth rate was carried out for a number of PCBs with different physical-chemical properties. The results indicate that <span class="hlt">air-water</span> <span class="hlt">exchange</span> dynamics are influenced not only by physical parameters but also by phytoplankton biomass and growth rate. New phytoplankton production resultsmore » in substantially longer times to reach equilibrium. Phytoplankton uptake-induced depletion of the dissolved phase concentration maintains <span class="hlt">air</span> and <span class="hlt">water</span> phases out of equilibrium. Furthermore, PCBs in phytoplankton also take longer times to reach equilibrium with the dissolved <span class="hlt">water</span> phase when the latter is supported by diffusive <span class="hlt">air-water</span> <span class="hlt">exchange</span>. However, both model analysis and model application to the Experimental Lakes Area of northwestern Ontario (Canada) suggest that the <span class="hlt">gas</span> phase supports the concentrations of persistent organic pollutants, such as PCBs, in atmospherically driven aquatic environments.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFM.B53A0937M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFM.B53A0937M"><span>Using Riverboat-Mounted Eddy Covariance for Direct Measurements of <span class="hlt">Air-water</span> <span class="hlt">Gas</span> <span class="hlt">Exchange</span> in Amazonia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Miller, S. D.; Freitas, H.; Read, E.; Goulden, M. L.; Rocha, H.</p> <p>2007-12-01</p> <p><span class="hlt">Gas</span> evasion from Amazonian rivers and lakes to the atmosphere has been estimated to play an important role in the regional budget of carbon dioxide (Richey et al., 2002) and the global budget of methane (Melack et al., 2004). These flux estimates were calculated by combining remote sensing estimates of inundation area with <span class="hlt">water</span>-side concentration gradients and <span class="hlt">gas</span> transfer rates (piston velocities) estimated primarily from floating chamber measurements (footprint ~1 m2). The uncertainty in these fluxes was large, attributed primarily to uncertainty in the <span class="hlt">gas</span> <span class="hlt">exchange</span> parameterization. Direct measurements of the <span class="hlt">gas</span> <span class="hlt">exchange</span> coefficient are needed to improve the parameterizations in these environments, and therefore reduce the uncertainty in fluxes. The micrometeorological technique of eddy covariance is attractive since it is a direct measurement of <span class="hlt">gas</span> <span class="hlt">exchange</span> that samples over a much larger area than floating chambers, and is amenable to use from a moving platform. We present eddy covariance carbon dioxide <span class="hlt">exchange</span> measurements made using a small riverboat in rivers and lakes in the central Amazon near Santarem, Para, Brazil. <span class="hlt">Water</span>-side carbon dioxide concentration was measured in situ, and the <span class="hlt">gas</span> <span class="hlt">exchange</span> coefficient was calculated. We found the piston velocity at a site on the Amazon River to be similar to existing ocean-based parameterizations, whereas the piston velocity at a site on the Tapajos River was roughly a factor 5 higher. We hypothesize that the enhanced <span class="hlt">gas</span> <span class="hlt">exchange</span> at the Tapajos site was due to a shallow upwind fetch. Our results demonstrate the feasibility of boat-based eddy covariance on these rivers, and also the utility of a mobile platform to investigate spatial variability of <span class="hlt">gas</span> <span class="hlt">exchange</span>.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li class="active"><span>1</span></li> <li><a href="#" onclick='return showDiv("page_2");'>2</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><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_1 --> <div id="page_2" 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_1");'>1</a></li> <li class="active"><span>2</span></li> <li><a href="#" onclick='return showDiv("page_3");'>3</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><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="21"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26196214','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26196214"><span>20 Years of <span class="hlt">Air-Water</span> <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Observations for Pesticides in the Western Arctic Ocean.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jantunen, Liisa M; Wong, Fiona; Gawor, Anya; Kylin, Henrik; Helm, Paul A; Stern, Gary A; Strachan, William M J; Burniston, Deborah A; Bidleman, Terry F</p> <p>2015-12-01</p> <p>The Arctic has been contaminated by legacy organochlorine pesticides (OCPs) and currently used pesticides (CUPs) through atmospheric transport and oceanic currents. Here we report the time trends and <span class="hlt">air-water</span> <span class="hlt">exchange</span> of OCPs and CUPs from research expeditions conducted between 1993 and 2013. Compounds determined in both <span class="hlt">air</span> and <span class="hlt">water</span> were trans- and cis-chlordanes (TC, CC), trans- and cis-nonachlors (TN, CN), heptachlor exo-epoxide (HEPX), dieldrin (DIEL), chlorobornanes (ΣCHBs and toxaphene), dacthal (DAC), endosulfans and metabolite endosulfan sulfate (ENDO-I, ENDO-II, and ENDO SUL), chlorothalonil (CHT), chlorpyrifos (CPF), and trifluralin (TFN). Pentachloronitrobenzene (PCNB and quintozene) and its soil metabolite pentachlorothianisole (PCTA) were also found in <span class="hlt">air</span>. Concentrations of most OCPs declined in surface <span class="hlt">water</span>, whereas some CUPs increased (ENDO-I, CHT, and TFN) or showed no significant change (CPF and DAC), and most compounds declined in <span class="hlt">air</span>. Chlordane compound fractions TC/(TC + CC) and TC/(TC + CC + TN) decreased in <span class="hlt">water</span> and <span class="hlt">air</span>, while CC/(TC + CC + TN) increased. TN/(TC + CC + TN) also increased in <span class="hlt">air</span> and slightly, but not significantly, in <span class="hlt">water</span>. These changes suggest selective removal of more labile TC and/or a shift in chlordane sources. <span class="hlt">Water-air</span> fugacity ratios indicated net volatilization (FR > 1.0) or near equilibrium (FR not significantly different from 1.0) for most OCPs but net deposition (FR < 1.0) for ΣCHBs. Net deposition was shown for ENDO-I on all expeditions, while the net <span class="hlt">exchange</span> direction of other CUPs varied. Understanding the processes and current state of <span class="hlt">air</span>-surface <span class="hlt">exchange</span> helps to interpret environmental exposure and evaluate the effectiveness of international protocols and provides insights for the environmental fate of new and emerging chemicals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://lofe.dukejournals.org/content/2/1.abstract','USGSPUBS'); return false;" href="http://lofe.dukejournals.org/content/2/1.abstract"><span><span class="hlt">Air-water</span> oxygen <span class="hlt">exchange</span> in a large whitewater river</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hall, Robert O.; Kennedy, Theodore A.; Rosi-Marshall, Emma J.</p> <p>2012-01-01</p> <p><span class="hlt">Air-water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> governs fluxes of <span class="hlt">gas</span> into and out of aquatic ecosystems. Knowing this flux is necessary to calculate <span class="hlt">gas</span> budgets (i.e., O2) to estimate whole-ecosystem metabolism and basin-scale carbon budgets. Empirical data on rates of <span class="hlt">gas</span> <span class="hlt">exchange</span> for streams, estuaries, and oceans are readily available. However, there are few data from large rivers and no data from whitewater rapids. We measured <span class="hlt">gas</span> transfer velocity in the Colorado River, Grand Canyon, as decline in O2 saturation deficit, 7 times in a 28-km segment spanning 7 rapids. The O2 saturation deficit exists because of hypolimnetic discharge from Glen Canyon Dam, located 25 km upriver from Lees Ferry. <span class="hlt">Gas</span> transfer velocity (k600) increased with slope of the immediate reach. k600 was -1 in flat reaches, while k600 for the steepest rapid ranged 3600-7700 cm h-1, an extremely high value of k600. Using the rate of <span class="hlt">gas</span> <span class="hlt">exchange</span> per unit length of <span class="hlt">water</span> surface elevation (Kdrop, m-1), segment-integrated k600 varied between 74 and 101 cm h-1. Using Kdrop we scaled k600 to the remainder of the Colorado River in Grand Canyon. At the scale corresponding to the segment length where 80% of the O2 <span class="hlt">exchanged</span> with the atmosphere (mean length = 26.1 km), k600 varied 4.5-fold between 56 and 272 cm h-1 with a mean of 113 cm h-1. <span class="hlt">Gas</span> transfer velocity for the Colorado River was higher than those from other aquatic ecosystems because of large rapids. Our approach of scaling k600 based on Kdrop allows comparing <span class="hlt">gas</span> transfer velocity across rivers with spatially heterogeneous morphology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/203613-gaseous-exchange-polycyclic-aromatic-hydrocarbons-across-air-water-interface-lower-chesapeake-bay','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/203613-gaseous-exchange-polycyclic-aromatic-hydrocarbons-across-air-water-interface-lower-chesapeake-bay"><span>Gaseous <span class="hlt">exchange</span> of polycyclic aromatic hydrocarbons across the <span class="hlt">air-water</span> interface of lower Chesapeake Bay</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>Gustafson, K.E.; Dickhut, R.M.</p> <p>1995-12-31</p> <p>The gaseous <span class="hlt">exchange</span> fluxes of polycyclic aromatic hydrocarbons (PAHs) across the <span class="hlt">air-water</span> interface of lower Chesapeake Bay were determined using a modified two-film <span class="hlt">exchange</span> model. Sampling covered the period January 1994 to June 1995 for five sites on lower Chesapeake Bay ranging from rural to urban and highly industrialized. Simultaneous <span class="hlt">air</span> and <span class="hlt">water</span> samples were collected and the atmospheric <span class="hlt">gas</span> phase and <span class="hlt">water</span> column dissolved phase analyzed via GC/MS for 17 PAHs. The direction and magnitude of flux for each PAH was calculated using Henry`s law constants, hydrological and meteorological parameters, Temperature was observed to be an important environmental factormore » in determining both the direction and magnitude of PAH <span class="hlt">gas</span> <span class="hlt">exchange</span>. Nonetheless, wind speed significantly impacts mass transfer coefficients, and therefore was found to control the magnitude of flux. Spatial and temporal variation of PAH gaseous <span class="hlt">exchange</span> fluxes were examined. Fluxes were determined to be both into and out of Chesapeake Bay. The range of <span class="hlt">gas</span> <span class="hlt">exchange</span> fluxes ({minus}560 to 600{micro}g/M{sup 2}*Mo) is of the same order to 10X greater than atmospheric wet and dry depositional fluxes to lower Chesapeake Bay. The results of this study support the hypothesis that <span class="hlt">gas</span> <span class="hlt">exchange</span> is a major transport process affecting the net loadings of PAHs in lower Chesapeake Bay.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRD..122.7664L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRD..122.7664L"><span>Atmospheric deposition and <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> fluxes of DDT and HCH in the Yangtze River Estuary, East China Sea</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Zhongxia; Lin, Tian; Li, Yuanyuan; Jiang, Yuqing; Guo, Zhigang</p> <p>2017-07-01</p> <p>The Yangtze River Estuary (YRE) is strongly influenced by the Yangtze River and lies on the pathway of the East Asian Monsoon. This study examined atmospheric deposition and <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> fluxes of dichlorodiphenyltrichloroethane (DDT) and hexachlorocyclohexane (HCH) to determine whether the YRE is a sink or source of selected pesticides at the <span class="hlt">air-water</span> interface under the influences of river input and atmospheric transport. The <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> of DDT was characterized by net volatilization with a marked difference in its fluxes between summer (140 ng/m2/d) and the other three seasons (12 ng/m2/d), possibly due to the high surface seawater temperatures and larger riverine input in summer. However, there was no obvious seasonal variation in the atmospheric HCH deposition, and the <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> reached equilibrium because of low HCH levels in the <span class="hlt">air</span> and seawater after the long-term banning of HCH and the degradation. The <span class="hlt">gas</span> <span class="hlt">exchange</span> flux of HCH was comparable to the dry and wet deposition fluxes at the <span class="hlt">air-water</span> interface. This suggests that the influences from the Yangtze River input and East Asian continental outflow on the fate of HCH in the YRE were limited. The <span class="hlt">gas</span> <span class="hlt">exchange</span> flux of DDT was about fivefold higher than the total dry and wet deposition fluxes. DDT residues in agricultural soil transported by enhanced riverine runoff were responsible for sustaining such a high net volatilization in summer. Moreover, our results indicated that there were fresh sources of DDT from the local environment to sustain net volatilization throughout the year.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1980Tell...32..470H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1980Tell...32..470H"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> across the <span class="hlt">air</span>-sea interface</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hasse, L.; Liss, P. S.</p> <p>1980-10-01</p> <p>The physics of <span class="hlt">gas</span> <span class="hlt">exchange</span> at the <span class="hlt">air</span>-sea interface are reviewed. In order to describe the transfer of gases in the liquid near the boundary, a molecular plus eddy diffusivity concept is used, which has been found useful for smooth flow over solid surfaces. From consideration of the boundary conditions, a similar dependence of eddy diffusivity on distance from the interface can be derived for the flow beneath a <span class="hlt">gas</span>/liquid interface, at least in the absence of waves. The influence of waves is then discussed. It is evident from scale considerations that the effect of gravity waves is small. It is known from wind tunnel work that capillary waves enhance <span class="hlt">gas</span> transfer considerably. The existing hypotheses are apparently not sufficient to explain the observations. Examination of field data is even more frustrating since the data do not show the expected increase of <span class="hlt">gas</span> <span class="hlt">exchange</span> with wind speed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOS.A24C2606P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOS.A24C2606P"><span>Surfactant control of <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> from North Sea coastal <span class="hlt">waters</span> and the Atlantic Meridional Transect</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pereira, R.</p> <p>2016-02-01</p> <p> suppression and SA is much weaker (r2 = <0.01, n = 22). While organic matter composition and sources may have variable control on <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> between the provinces, the poor relationship observed between SA and k660 suggests that other environmental factors maybe more influential on <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> in the open ocean compared to North Sea coastal <span class="hlt">waters</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21917934','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21917934"><span><span class="hlt">Air-gas</span> <span class="hlt">exchange</span> reevaluated: clinically important results of a computer simulation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shunmugam, Manoharan; Shunmugam, Sudhakaran; Williamson, Tom H; Laidlaw, D Alistair</p> <p>2011-10-21</p> <p>The primary aim of this study was to evaluate the efficiency of <span class="hlt">air-gas</span> <span class="hlt">exchange</span> techniques and the factors that influence the final concentration of an intraocular <span class="hlt">gas</span> tamponade. Parameters were varied to find the optimum method of performing an <span class="hlt">air-gas</span> <span class="hlt">exchange</span> in ideal circumstances. A computer model of the eye was designed using 3D software with fluid flow analysis capabilities. Factors such as angular distance between ports, <span class="hlt">gas</span> infusion gauge, exhaust vent gauge and depth were varied in the model. Flow rate and axial length were also modulated to simulate faster injections and more myopic eyes, respectively. The flush volume of <span class="hlt">gas</span> required to achieve a 97% intraocular <span class="hlt">gas</span> fraction concentration were compared. Modulating individual factors did not reveal any clinically significant difference in the angular distance between ports, exhaust vent size, and depth or rate of <span class="hlt">gas</span> injection. In combination, however, there was a 28% increase in <span class="hlt">air-gas</span> <span class="hlt">exchange</span> efficiency comparing the most efficient with the least efficient studied parameters in this model. The <span class="hlt">gas</span> flush volume required to achieve a 97% <span class="hlt">gas</span> fill also increased proportionately at a ratio of 5.5 to 6.2 times the volume of the eye. A 35-mL flush is adequate for eyes up to 25 mm in axial length; however, eyes longer than this would require a much greater flush volume, and surgeons should consider using two separate 50-mL <span class="hlt">gas</span> syringes to ensure optimal <span class="hlt">gas</span> concentration for eyes greater than 25 mm in axial length.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1570919','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1570919"><span>Structure, function and evolution of the <span class="hlt">gas</span> <span class="hlt">exchangers</span>: comparative perspectives</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Maina, JN</p> <p>2002-01-01</p> <p>Over the evolutionary continuum, animals have faced similar fundamental challenges of acquiring molecular oxygen for aerobic metabolism. Under limitations and constraints imposed by factors such as phylogeny, behaviour, body size and environment, they have responded differently in founding optimal respiratory structures. A quintessence of the aphorism that ‘necessity is the mother of invention’, <span class="hlt">gas</span> <span class="hlt">exchangers</span> have been inaugurated through stiff cost–benefit analyses that have evoked transaction of trade-offs and compromises. Cogent structural–functional correlations occur in constructions of <span class="hlt">gas</span> <span class="hlt">exchangers</span>: within and between taxa, morphological complexity and respiratory efficiency increase with metabolic capacities and oxygen needs. Highly active, small endotherms have relatively better-refined <span class="hlt">gas</span> <span class="hlt">exchangers</span> compared with large, inactive ectotherms. Respiratory structures have developed from the plain cell membrane of the primeval prokaryotic unicells to complex multifunctional ones ofthe modern Metazoa. Regarding the respiratory medium used to extract oxygen from, animal life has had only two choices – <span class="hlt">water</span> or <span class="hlt">air</span> – within the biological range of temperature and pressure the only naturally occurring respirable fluids. In rarer cases, certain animalshave adapted to using both media. Gills (evaginated <span class="hlt">gas</span> <span class="hlt">exchangers</span>) are the primordial respiratory organs: they are the archetypal <span class="hlt">water</span> breathing organs. Lungs (invaginated <span class="hlt">gas</span> <span class="hlt">exchangers</span>) are the model <span class="hlt">air</span> breathing organs. Bimodal (transitional) breathers occupy the water–<span class="hlt">air</span> interface. Presentation and exposure of external (<span class="hlt">water/air</span>) and internal (haemolymph/blood) respiratory media, features determined by geometric arrangement of the conduits, are important features for <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency: counter-current, cross-current, uniform pool and infinite pool designs have variably developed. PMID:12430953</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2889562','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2889562"><span>Universal model for <span class="hlt">water</span> costs of <span class="hlt">gas</span> <span class="hlt">exchange</span> by animals and plants</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Woods, H. Arthur; Smith, Jennifer N.</p> <p>2010-01-01</p> <p>For terrestrial animals and plants, a fundamental cost of living is <span class="hlt">water</span> vapor lost to the atmosphere during <span class="hlt">exchange</span> of metabolic gases. Here, by bringing together previously developed models for specific taxa, we integrate properties common to all terrestrial <span class="hlt">gas</span> <span class="hlt">exchangers</span> into a universal model of <span class="hlt">water</span> loss. The model predicts that <span class="hlt">water</span> loss scales to <span class="hlt">gas</span> <span class="hlt">exchange</span> with an exponent of 1 and that the amount of <span class="hlt">water</span> lost per unit of <span class="hlt">gas</span> <span class="hlt">exchanged</span> depends on several factors: the surface temperature of the respiratory system near the outside of the organism, the <span class="hlt">gas</span> consumed (oxygen or carbon dioxide), the steepness of the gradients for <span class="hlt">gas</span> and vapor, and the transport mode (convective or diffusive). Model predictions were largely confirmed by data on 202 species in five taxa—insects, birds, bird eggs, mammals, and plants—spanning nine orders of magnitude in rate of <span class="hlt">gas</span> <span class="hlt">exchange</span>. Discrepancies between model predictions and data seemed to arise from biologically interesting violations of model assumptions, which emphasizes how poorly we understand <span class="hlt">gas</span> <span class="hlt">exchange</span> in some taxa. The universal model provides a unified conceptual framework for analyzing <span class="hlt">exchange</span>-associated <span class="hlt">water</span> losses across taxa with radically different metabolic and <span class="hlt">exchange</span> systems. PMID:20404161</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20404161','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20404161"><span>Universal model for <span class="hlt">water</span> costs of <span class="hlt">gas</span> <span class="hlt">exchange</span> by animals and plants.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Woods, H Arthur; Smith, Jennifer N</p> <p>2010-05-04</p> <p>For terrestrial animals and plants, a fundamental cost of living is <span class="hlt">water</span> vapor lost to the atmosphere during <span class="hlt">exchange</span> of metabolic gases. Here, by bringing together previously developed models for specific taxa, we integrate properties common to all terrestrial <span class="hlt">gas</span> <span class="hlt">exchangers</span> into a universal model of <span class="hlt">water</span> loss. The model predicts that <span class="hlt">water</span> loss scales to <span class="hlt">gas</span> <span class="hlt">exchange</span> with an exponent of 1 and that the amount of <span class="hlt">water</span> lost per unit of <span class="hlt">gas</span> <span class="hlt">exchanged</span> depends on several factors: the surface temperature of the respiratory system near the outside of the organism, the <span class="hlt">gas</span> consumed (oxygen or carbon dioxide), the steepness of the gradients for <span class="hlt">gas</span> and vapor, and the transport mode (convective or diffusive). Model predictions were largely confirmed by data on 202 species in five taxa--insects, birds, bird eggs, mammals, and plants--spanning nine orders of magnitude in rate of <span class="hlt">gas</span> <span class="hlt">exchange</span>. Discrepancies between model predictions and data seemed to arise from biologically interesting violations of model assumptions, which emphasizes how poorly we understand <span class="hlt">gas</span> <span class="hlt">exchange</span> in some taxa. The universal model provides a unified conceptual framework for analyzing <span class="hlt">exchange</span>-associated <span class="hlt">water</span> losses across taxa with radically different metabolic and <span class="hlt">exchange</span> systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013GeoRL..40.5683H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013GeoRL..40.5683H"><span>Efficient <span class="hlt">gas</span> <span class="hlt">exchange</span> between a boreal river and the atmosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huotari, Jussi; Haapanala, Sami; Pumpanen, Jukka; Vesala, Timo; Ojala, Anne</p> <p>2013-11-01</p> <p>largest uncertainties in accurately resolving the role of rivers and streams in carbon cycling stem from difficulties in determining <span class="hlt">gas</span> <span class="hlt">exchange</span> between <span class="hlt">water</span> and the atmosphere. So far, estimates for river-atmosphere <span class="hlt">gas</span> <span class="hlt">exchange</span> have lacked direct ecosystem-scale flux measurements not disturbing <span class="hlt">gas</span> <span class="hlt">exchange</span> across the <span class="hlt">air-water</span> interface. We conducted the first direct riverine <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements with eddy covariance in tandem with continuous surface <span class="hlt">water</span> CO2 measurements in a large boreal river for 30 days. Our measured <span class="hlt">gas</span> transfer velocity was, on average, 20.8 cm h-1, which is clearly higher than the model estimates based on river channel morphology and <span class="hlt">water</span> velocity, whereas our floating chambers gave comparable values at 17.3 cm h-1. These results demonstrate that present estimates for riverine CO2 emissions are very likely too low. This result is also relevant to any other gases emitted, as their diffusive <span class="hlt">exchange</span> rates are similarly proportional to <span class="hlt">gas</span> transfer velocity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=245933','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=245933"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> relations responses of spring wheat to full-season infrared warming</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p><span class="hlt">Gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> relations responses to full-season in situ infrared (IR) warming were evaluated for hard red spring wheat (Triticum aestivum L. cv. Yecora Rojo) grown in an open field in a semi-arid desert region of the Southwest USA. A Temperature Free-<span class="hlt">Air</span> Controlled Enhancement (T-FACE) ap...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=276837','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=276837"><span><span class="hlt">Gas</span> <span class="hlt">Exchange</span> and <span class="hlt">Water</span> Relations Responses of Spring Wheat to Full-Season Infrared Warming</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p><span class="hlt">Gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> relations were evaluated under full-season in situ infrared (IR) warming for hard red spring wheat (Triticum aestivum L. cv. Yecora Rojo) grown in an open field in a semiarid desert region of the southwest USA. A temperature free-<span class="hlt">air</span> controlled enhancement (T-FACE) apparatus u...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1084027','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1084027"><span>Recovery of <span class="hlt">Water</span> from Boiler Flue <span class="hlt">Gas</span> Using Condensing Heat <span class="hlt">Exchangers</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>Levy, Edward; Bilirgen, Harun; DuPont, John</p> <p>2011-03-31</p> <p>Most of the <span class="hlt">water</span> used in a thermoelectric power plant is used for cooling, and DOE has been focusing on possible techniques to reduce the amount of fresh <span class="hlt">water</span> needed for cooling. DOE has also been placing emphasis on recovery of usable <span class="hlt">water</span> from sources not generally considered, such as mine <span class="hlt">water</span>, <span class="hlt">water</span> produced from oil and <span class="hlt">gas</span> extraction, and <span class="hlt">water</span> contained in boiler flue <span class="hlt">gas</span>. This report deals with development of condensing heat <span class="hlt">exchanger</span> technology for recovering moisture from flue <span class="hlt">gas</span> from coal-fired power plants. The report describes: • An expanded data base on <span class="hlt">water</span> and acid condensation characteristicsmore » of condensing heat <span class="hlt">exchangers</span> in coal-fired units. This data base was generated by performing slip stream tests at a power plant with high sulfur bituminous coal and a wet FGD scrubber and at a power plant firing highmoisture, low rank coals. • Data on typical concentrations of HCl, HNO{sub 3} and H{sub 2}SO{sub 4} in low temperature condensed flue <span class="hlt">gas</span> moisture, and mercury capture efficiencies as functions of process conditions in power plant field tests. • Theoretical predictions for sulfuric acid concentrations on tube surfaces at temperatures above the <span class="hlt">water</span> vapor dewpoint temperature and below the sulfuric acid dew point temperature. • Data on corrosion rates of candidate heat <span class="hlt">exchanger</span> tube materials for the different regions of the heat <span class="hlt">exchanger</span> system as functions of acid concentration and temperature. • Data on effectiveness of acid traps in reducing sulfuric acid concentrations in a heat <span class="hlt">exchanger</span> tube bundle. • Condensed flue <span class="hlt">gas</span> <span class="hlt">water</span> treatment needs and costs. • Condensing heat <span class="hlt">exchanger</span> designs and installed capital costs for full-scale applications, both for installation immediately downstream of an ESP or baghouse and for installation downstream of a wet SO{sub 2} scrubber. • Results of cost-benefit studies of condensing heat <span class="hlt">exchangers</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1037725','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1037725"><span>Recovery of <span class="hlt">Water</span> from Boiler Flue <span class="hlt">Gas</span> Using Condensing Heat <span class="hlt">Exchangers</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>Edward Levy; Harun Bilirgen; John DuPoint</p> <p>2011-03-31</p> <p>Most of the <span class="hlt">water</span> used in a thermoelectric power plant is used for cooling, and DOE has been focusing on possible techniques to reduce the amount of fresh <span class="hlt">water</span> needed for cooling. DOE has also been placing emphasis on recovery of usable <span class="hlt">water</span> from sources not generally considered, such as mine <span class="hlt">water</span>, <span class="hlt">water</span> produced from oil and <span class="hlt">gas</span> extraction, and <span class="hlt">water</span> contained in boiler flue <span class="hlt">gas</span>. This report deals with development of condensing heat <span class="hlt">exchanger</span> technology for recovering moisture from flue <span class="hlt">gas</span> from coal-fired power plants. The report describes: (1) An expanded data base on <span class="hlt">water</span> and acid condensation characteristicsmore » of condensing heat <span class="hlt">exchangers</span> in coal-fired units. This data base was generated by performing slip stream tests at a power plant with high sulfur bituminous coal and a wet FGD scrubber and at a power plant firing high-moisture, low rank coals. (2) Data on typical concentrations of HCl, HNO{sub 3} and H{sub 2}SO{sub 4} in low temperature condensed flue <span class="hlt">gas</span> moisture, and mercury capture efficiencies as functions of process conditions in power plant field tests. (3) Theoretical predictions for sulfuric acid concentrations on tube surfaces at temperatures above the <span class="hlt">water</span> vapor dewpoint temperature and below the sulfuric acid dew point temperature. (4) Data on corrosion rates of candidate heat <span class="hlt">exchanger</span> tube materials for the different regions of the heat <span class="hlt">exchanger</span> system as functions of acid concentration and temperature. (5) Data on effectiveness of acid traps in reducing sulfuric acid concentrations in a heat <span class="hlt">exchanger</span> tube bundle. (6) Condensed flue <span class="hlt">gas</span> <span class="hlt">water</span> treatment needs and costs. (7) Condensing heat <span class="hlt">exchanger</span> designs and installed capital costs for full-scale applications, both for installation immediately downstream of an ESP or baghouse and for installation downstream of a wet SO{sub 2} scrubber. (8) Results of cost-benefit studies of condensing heat <span class="hlt">exchangers</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/864049','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/864049"><span>Fluidized bed heat <span class="hlt">exchanger</span> with <span class="hlt">water</span> cooled <span class="hlt">air</span> distributor and dust hopper</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Jukkola, Walfred W.; Leon, Albert M.; Van Dyk, Jr., Garritt C.; McCoy, Daniel E.; Fisher, Barry L.; Saiers, Timothy L.; Karstetter, Marlin E.</p> <p>1981-11-24</p> <p>A fluidized bed heat <span class="hlt">exchanger</span> is provided in which <span class="hlt">air</span> is passed through a bed of particulate material containing fuel. A steam-<span class="hlt">water</span> natural circulation system is provided for heat <span class="hlt">exchange</span> and the housing of the heat <span class="hlt">exchanger</span> has a <span class="hlt">water</span>-wall type construction. Vertical in-bed heat <span class="hlt">exchange</span> tubes are provided and the <span class="hlt">air</span> distributor is <span class="hlt">water</span>-cooled. A <span class="hlt">water</span>-cooled dust hopper is provided in the housing to collect particulates from the combustion gases and separate the combustion zone from a volume within said housing in which convection heat <span class="hlt">exchange</span> tubes are provided to extract heat from the exiting combustion gases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006GeoRL..3314803Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006GeoRL..3314803Z"><span>Impacts of winter storms on <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Weiqing; Perrie, Will; Vagle, Svein</p> <p>2006-07-01</p> <p>The objective of this study is to investigate <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> during winter storms, using field measurements from Ocean Station Papa in the Northeast Pacific (50°N, 145°W). We show that increasing <span class="hlt">gas</span> transfer rates are coincident with increasing winds and deepening depth of bubble penetration, and that this process depends on sea state. Wave-breaking is shown to be an important factor in the <span class="hlt">gas</span> transfer velocity during the peaks of the storms, increasing the flux rates by up to 20%. <span class="hlt">Gas</span> transfer rates and concentrations can exhibit asymmetry, reflecting a sudden increase with the onset of a storm, and gradual recovery stages.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFMOS31B1280P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFMOS31B1280P"><span>Micrometeorological Measurement of Fetch- and Atmospheric Stability-Dependent <span class="hlt">Air</span>- <span class="hlt">Water</span> <span class="hlt">Exchange</span> of Legacy Semivolatile Organic Contaminants in Lake Superior</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Perlinger, J. A.; Tobias, D. E.; Rowe, M. D.</p> <p>2008-12-01</p> <p>Coastal <span class="hlt">waters</span> including the Laurentian Great Lakes are particularly susceptible to local, regional, and long- range transport and deposition of semivolatile organic contaminants (SOCs) as gases and/or associated with particles. Recently-marketed SOCs can be expected to undergo net deposition in surface <span class="hlt">waters</span>, whereas legacy SOCs such as polychlorinated biphenyls (PCBs) are likely to be at equilibrium with respect to <span class="hlt">air-water</span> <span class="hlt">exchange</span>, or, if atmospheric concentrations decrease through, e.g., policy implementation, to undergo net <span class="hlt">gas</span> emission. SOC <span class="hlt">air-water</span> <span class="hlt">exchange</span> flux is usually estimated using the two-film model. This model describes molecular diffusion through the <span class="hlt">air</span> and <span class="hlt">water</span> films adjacent to the <span class="hlt">air-water</span> interface. <span class="hlt">Air-water</span> <span class="hlt">exchange</span> flux is estimated as the product of SOC fugacity, typically based on on-shore gaseous concentration measurements, and a transfer coefficient, the latter which is estimated from SOC properties and environmental conditions. The transfer coefficient formulation commonly applied neglects resistance to <span class="hlt">exchange</span> in the internal boundary layer under atmospherically stable conditions, and the use of on-shore gaseous concentration neglects fetch-dependent equilibration, both of which will tend to cause overestimation of flux magnitude. Thus, for legacy chemicals or in any highly contaminated surface <span class="hlt">water</span>, the rate at which the <span class="hlt">water</span> is cleansed through <span class="hlt">gas</span> emission tends to be over-predicted using this approach. Micrometeorological measurement of <span class="hlt">air-water</span> <span class="hlt">exchange</span> rates of legacy SOCs was carried out on ships during four transect experiments during off-shore flow in Lake Superior using novel multicapillary collection devices and thermal extraction technology to measure parts-per-quadrillion SOC levels. Employing sensible heat in the modified Bowen ratio, fluxes at three over-<span class="hlt">water</span> stations along the transects were measured, along with up-wind, onshore gaseous concentration and aqueous concentration. The atmosphere was unstable for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JGRC..117.5035A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JGRC..117.5035A"><span>Statistics of surface divergence and their relation to <span class="hlt">air-water</span> <span class="hlt">gas</span> transfer velocity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Asher, William E.; Liang, Hanzhuang; Zappa, Christopher J.; Loewen, Mark R.; Mukto, Moniz A.; Litchendorf, Trina M.; Jessup, Andrew T.</p> <p>2012-05-01</p> <p><span class="hlt">Air</span>-sea <span class="hlt">gas</span> fluxes are generally defined in terms of the <span class="hlt">air/water</span> concentration difference of the <span class="hlt">gas</span> and the <span class="hlt">gas</span> transfer velocity,kL. Because it is difficult to measure kLin the ocean, it is often parameterized using more easily measured physical properties. Surface divergence theory suggests that infrared (IR) images of the <span class="hlt">water</span> surface, which contain information concerning the movement of <span class="hlt">water</span> very near the <span class="hlt">air-water</span> interface, might be used to estimatekL. Therefore, a series of experiments testing whether IR imagery could provide a convenient means for estimating the surface divergence applicable to <span class="hlt">air</span>-sea <span class="hlt">exchange</span> were conducted in a synthetic jet array tank embedded in a wind tunnel. <span class="hlt">Gas</span> transfer velocities were measured as a function of wind stress and mechanically generated turbulence; laser-induced fluorescence was used to measure the concentration of carbon dioxide in the top 300 μm of the <span class="hlt">water</span> surface; IR imagery was used to measure the spatial and temporal distribution of the aqueous skin temperature; and particle image velocimetry was used to measure turbulence at a depth of 1 cm below the <span class="hlt">air-water</span> interface. It is shown that an estimate of the surface divergence for both wind-shear driven turbulence and mechanically generated turbulence can be derived from the surface skin temperature. The estimates derived from the IR images are compared to velocity field divergences measured by the PIV and to independent estimates of the divergence made using the laser-induced fluorescence data. Divergence is shown to scale withkLvalues measured using gaseous tracers as predicted by conceptual models for both wind-driven and mechanically generated turbulence.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17674350','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17674350"><span><span class="hlt">Air</span> bells of <span class="hlt">water</span> spiders are an extended phenotype modified in response to <span class="hlt">gas</span> composition.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schütz, Dolores; Taborsky, Michael; Drapela, Thomas</p> <p>2007-10-01</p> <p>The <span class="hlt">water</span> spider Argyroneta aquatica (Clerck) is the only spider that spends its whole life under <span class="hlt">water</span>. <span class="hlt">Water</span> spiders keep an <span class="hlt">air</span> bubble around their body for breathing and build under-<span class="hlt">water</span> <span class="hlt">air</span> bells, which they use for shelter and raising offspring, digesting and consuming prey, moulting, depositing eggs and sperm, and copulating. It is unclear whether these bells are an important oxygen reservoir for breathing under <span class="hlt">water</span>, or whether they serve mainly to create <span class="hlt">water</span>-free space for feeding and reproduction. In this study, we manipulated the composition of the <span class="hlt">gas</span> inside the bell of female <span class="hlt">water</span> spiders to test whether they monitor the quality of this <span class="hlt">gas</span>, and replenish oxygen if required. We <span class="hlt">exchanged</span> the entire <span class="hlt">gas</span> in the bell either with pure O2, pure CO2, or with ambient <span class="hlt">air</span> as control, and monitored behavioural responses. The test spiders surfaced and replenished <span class="hlt">air</span> more often in the CO2 treatment than in the O2 treatment, and they increased bell building behaviour. In addition to active oxygen regulation, they monitored and adjusted the bells by adding silk. These results show that <span class="hlt">water</span> spiders use the <span class="hlt">air</span> bell as an oxygen reservoir, and that it functions as an external lung, which renders it essential for living under <span class="hlt">water</span> permanently. A. aquatica is the only animal that collects, transports, and stores <span class="hlt">air</span>, and monitors its property for breathing, which is an adaptive response of a terrestrial animal to the colonization of an aquatic habitat.</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_1");'>1</a></li> <li class="active"><span>2</span></li> <li><a href="#" onclick='return showDiv("page_3");'>3</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><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_2 --> <div id="page_3" 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_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</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="41"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26910987','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26910987"><span>[Summer Greenhouse Gases <span class="hlt">Exchange</span> Flux Across <span class="hlt">Water-air</span> Interface in Three <span class="hlt">Water</span> Reservoirs Located in Different Geologic Setting in Guangxi, China].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Jian-hong; Pu, Jun-bing; Sun, Ping-an; Yuan, Dao-xian; Liu, Wen; Zhang, Tao; Mo, Xue</p> <p>2015-11-01</p> <p>Due to special hydrogeochemical characteristics of calcium-rich, alkaline and DIC-rich ( dissolved inorganic carbon) environment controlled by the weathering products from carbonate rock, the <span class="hlt">exchange</span> characteristics, processes and controlling factors of greenhouse <span class="hlt">gas</span> (CO2 and CH4) across <span class="hlt">water-air</span> interface in karst <span class="hlt">water</span> reservoir show obvious differences from those of non-karst <span class="hlt">water</span> reservoir. Three <span class="hlt">water</span> reservoirs (Dalongdong reservoir-karst reservoir, Wulixia reservoir--semi karst reservoir, Si'anjiang reservoir-non-karst reservoir) located in different geologic setting in Guangxi Zhuang Autonomous Region, China were chosen to reveal characteristics and controlling factors of greenhouse <span class="hlt">gas</span> <span class="hlt">exchange</span> flux across <span class="hlt">water-air</span> interface. Two common approaches, floating chamber (FC) and thin boundary layer models (TBL), were employed to research and contrast greenhouse <span class="hlt">gas</span> <span class="hlt">exchange</span> flux across <span class="hlt">water-air</span> interface from three reservoirs. The results showed that: (1) surface-layer <span class="hlt">water</span> in reservoir area and discharging <span class="hlt">water</span> under dam in Dalongdong <span class="hlt">water</span> reservoir were the source of atmospheric CO2 and CH4. Surface-layer <span class="hlt">water</span> in reservoir area in Wulixia <span class="hlt">water</span> reservoir was the sink of atmospheric CO2 and the source of atmospheric CH4, while discharging <span class="hlt">water</span> under dam was the source of atmospheric CO2 and CH4. Surface-layer <span class="hlt">water</span> in Si'anjiang <span class="hlt">water</span> reservoir was the sink of atmospheric CO2 and source of atmospheric CH4. (2) CO2 and CH4 effluxes in discharging <span class="hlt">water</span> under dam were much more than those in surface-layer <span class="hlt">water</span> in reservoir area regardless of karst reservoir or non karst reservoir. Accordingly, more attention should be paid to the CO2 and CH4 emission from discharging <span class="hlt">water</span> under dam. (3) In the absence of submerged soil organic matters and plants, the difference of CH4 effluxes between karst groundwater-fed reservoir ( Dalongdong <span class="hlt">water</span> reservoir) and non-karst area ( Wulixia <span class="hlt">water</span> reservoir and Si'anjiang <span class="hlt">water</span> reservoir) was less. However, CO2</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014OcSci..10..587S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014OcSci..10..587S"><span>An automated <span class="hlt">gas</span> <span class="hlt">exchange</span> tank for determining <span class="hlt">gas</span> transfer velocities in natural seawater samples</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schneider-Zapp, K.; Salter, M. E.; Upstill-Goddard, R. C.</p> <p>2014-07-01</p> <p>In order to advance understanding of the role of seawater surfactants in the <span class="hlt">air</span>-sea <span class="hlt">exchange</span> of climatically active trace gases via suppression of the <span class="hlt">gas</span> transfer velocity (kw), we constructed a fully automated, closed <span class="hlt">air-water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> tank and coupled analytical system. The system allows <span class="hlt">water</span>-side turbulence in the tank to be precisely controlled with an electronically operated baffle. Two coupled <span class="hlt">gas</span> chromatographs and an integral equilibrator, connected to the tank in a continuous <span class="hlt">gas</span>-tight system, allow temporal changes in the partial pressures of SF6, CH4 and N2O to be measured simultaneously in the tank <span class="hlt">water</span> and headspace at multiple turbulence settings, during a typical experimental run of 3.25 h. PC software developed by the authors controls all operations and data acquisition, enabling the optimisation of experimental conditions with high reproducibility. The use of three gases allows three independent estimates of kw for each turbulence setting; these values are subsequently normalised to a constant Schmidt number for direct comparison. The normalised kw estimates show close agreement. Repeated experiments with Milli-Q <span class="hlt">water</span> demonstrate a typical measurement accuracy of 4% for kw. Experiments with natural seawater show that the system clearly resolves the effects on kw of spatial and temporal trends in natural surfactant activity. The system is an effective tool with which to probe the relationships between kw, surfactant activity and biogeochemical indices of primary productivity, and should assist in providing valuable new insights into the <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014OcScD..11..693S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014OcScD..11..693S"><span>An automated <span class="hlt">gas</span> <span class="hlt">exchange</span> tank for determining <span class="hlt">gas</span> transfer velocities in natural seawater samples</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schneider-Zapp, K.; Salter, M. E.; Upstill-Goddard, R. C.</p> <p>2014-02-01</p> <p>In order to advance understanding of the role of seawater surfactants in the <span class="hlt">air</span>-sea <span class="hlt">exchange</span> of climatically active trace gases via suppression of the <span class="hlt">gas</span> transfer velocity (kw), we constructed a fully automated, closed <span class="hlt">air-water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> tank and coupled analytical system. The system allows <span class="hlt">water</span>-side turbulence in the tank to be precisely controlled with an electronically operated baffle. Two coupled <span class="hlt">gas</span> chromatographs and an integral equilibrator, connected to the tank in a continuous <span class="hlt">gas</span>-tight system, allow temporal changes in the partial pressures of SF6, CH4 and N2O to be measured simultaneously in the tank <span class="hlt">water</span> and headspace at multiple turbulence settings, during a typical experimental run of 3.25 h. PC software developed by the authors controls all operations and data acquisition, enabling the optimisation of experimental conditions with high reproducibility. The use of three gases allows three independent estimates of kw for each turbulence setting; these values are subsequently normalised to a constant Schmidt number for direct comparison. The normalised kw estimates show close agreement. Repeated experiments with MilliQ <span class="hlt">water</span> demonstrate a typical measurement accuracy of 4% for kw. Experiments with natural seawater show that the system clearly resolves the effects on kw of spatial and temporal trends in natural surfactant activity. The system is an effective tool with which to probe the relationships between kw, surfactant activity and biogeochemical indices of primary productivity, and should assist in providing valuable new insights into the <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ACPD...1313285B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ACPD...1313285B"><span><span class="hlt">Air</span>/sea DMS <span class="hlt">gas</span> transfer in the North Atlantic: evidence for limited interfacial <span class="hlt">gas</span> <span class="hlt">exchange</span> at high wind speed</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bell, T. G.; De Bruyn, W.; Miller, S. D.; Ward, B.; Christensen, K.; Saltzman, E. S.</p> <p>2013-05-01</p> <p>Shipboard measurements of eddy covariance DMS <span class="hlt">air</span>/sea fluxes and seawater concentration were carried out in the North Atlantic bloom region in June/July 2011. <span class="hlt">Gas</span> transfer coefficients (k660) show a linear dependence on mean horizontal wind speed at wind speeds up to 11 m s-1. At higher wind speeds the relationship between k660 and wind speed weakens. At high winds, measured DMS fluxes were lower than predicted based on the linear relationship between wind speed and interfacial stress extrapolated from low to intermediate wind speeds. In contrast, the transfer coefficient for sensible heat did not exhibit this effect. The apparent suppression of <span class="hlt">air</span>/sea <span class="hlt">gas</span> flux at higher wind speeds appears to be related to sea state, as determined from shipboard wave measurements. These observations are consistent with the idea that long waves suppress near surface <span class="hlt">water</span> side turbulence, and decrease interfacial <span class="hlt">gas</span> transfer. This effect may be more easily observed for DMS than for less soluble gases, such as CO2, because the <span class="hlt">air</span>/sea <span class="hlt">exchange</span> of DMS is controlled by interfacial rather than bubble-mediated <span class="hlt">gas</span> transfer under high wind speed conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=260737','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=260737"><span>Ecosystem Warming Affects Vertical Distribution of Leaf <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Properties and <span class="hlt">Water</span> Relations of Spring Wheat</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>The vertical distribution of <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> relations responses to full-season in situ infrared (IR) warming were evaluated for hard red spring wheat (Triticum aestivum L. cv. Yecora Rojo) grown in an open field in a semiarid desert region of the Southwest USA. A Temperature Free-<span class="hlt">Air</span> Contro...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=68872&keyword=water+AND+gas+AND+exchange&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=68872&keyword=water+AND+gas+AND+exchange&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>DIFFUSIVE <span class="hlt">EXCHANGE</span> OF GASEOUS POLYCYCLIC AROMATIC HYDROCARBONS AND POLYCHLORINATED BIPHENYLS ACROSS THE <span class="hlt">AIR-WATER</span> INTERFACE OF THE CHESAPEAKE BAY. (R825245)</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>Dissolved and <span class="hlt">gas</span>-phase concentrations of nine polycyclic aromatic hydrocarbons and 46 polychlorinated biphenyl congeners were measured at eight sites on the Chesapeake Bay at four different times of the year to estimate net diffusive <span class="hlt">air-water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> rates. Gaseous PAHs ar...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27461227','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27461227"><span><span class="hlt">Air</span> breathing and aquatic <span class="hlt">gas</span> <span class="hlt">exchange</span> during hypoxia in armoured catfish.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Scott, Graham R; Matey, Victoria; Mendoza, Julie-Anne; Gilmour, Kathleen M; Perry, Steve F; Almeida-Val, Vera M F; Val, Adalberto L</p> <p>2017-01-01</p> <p><span class="hlt">Air</span> breathing in fish is commonly believed to have arisen as an adaptation to aquatic hypoxia. The effectiveness of <span class="hlt">air</span> breathing for tissue O 2 supply depends on the ability to avoid O 2 loss as oxygenated blood from the <span class="hlt">air</span>-breathing organ passes through the gills. Here, we evaluated whether the armoured catfish (Hypostomus aff. pyreneusi)-a facultative <span class="hlt">air</span> breather-can avoid branchial O 2 loss while <span class="hlt">air</span> breathing in aquatic hypoxia, and we measured various other respiratory and metabolic traits important for O 2 supply and utilization. Fish were instrumented with opercular catheters to measure the O 2 tension (PO 2 ) of expired <span class="hlt">water</span>, and <span class="hlt">air</span> breathing and aquatic respiration were measured during progressive stepwise hypoxia in the <span class="hlt">water</span>. Armoured catfish exhibited relatively low rates of O 2 consumption and gill ventilation, and gill ventilation increased in hypoxia due primarily to increases in ventilatory stroke volume. Armoured catfish began <span class="hlt">air</span> breathing at a <span class="hlt">water</span> PO 2 of 2.5 kPa, and both <span class="hlt">air</span>-breathing frequency and hypoxia tolerance (as reflected by PO 2 at loss of equilibrium, LOE) was greater in individuals with a larger body mass. Branchial O 2 loss, as reflected by higher PO 2 in expired than in inspired <span class="hlt">water</span>, was observed in a minority (4/11) of individuals as <span class="hlt">water</span> PO 2 approached that at LOE. Armoured catfish also exhibited a gill morphology characterized by short filaments bearing short fused lamellae, large interlamellar cell masses, low surface area, and a thick epithelium that increased <span class="hlt">water</span>-to-blood diffusion distance. Armoured catfish had a relatively low blood-O 2 binding affinity when sampled in normoxia (P 50 of 3.1 kPa at pH 7.4), but were able to rapidly increase binding affinity during progressive hypoxia exposure (to a P 50 of 1.8 kPa). Armoured catfish also had low activities of several metabolic enzymes in white muscle, liver, and brain. Therefore, low rates of metabolism and gill ventilation, and a reduction in branchial <span class="hlt">gas-exchange</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A22A..08Q','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A22A..08Q"><span>The <span class="hlt">air</span>, carbon, <span class="hlt">water</span> synergies and trade-offs in China's natural <span class="hlt">gas</span> industry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Qin, Y.; Mauzerall, D. L.; Höglund-Isaksson, L.; Wagner, F.; Byers, E.</p> <p>2017-12-01</p> <p>Both energy production and consumption can simultaneously affect regional <span class="hlt">air</span> quality, local <span class="hlt">water</span> stress, and the global climate. Identifying <span class="hlt">air</span>, carbon and <span class="hlt">water</span> impacts of various energy sources and end-uses is important in determining the relative merits of various energy policies. Here, we examine the <span class="hlt">air-carbon-water</span> interdependencies of China's six major natural <span class="hlt">gas</span> source choices (domestic conventional natural <span class="hlt">gas</span>, domestic coal-based synthetic natural <span class="hlt">gas</span> (SNG), domestic shale <span class="hlt">gas</span>, imported liquefied natural <span class="hlt">gas</span>, imported Russian pipeline <span class="hlt">gas</span>, and imported Central Asian pipeline <span class="hlt">gas</span>) and three end-use coal-to-<span class="hlt">gas</span> deployment strategies (with substitution strategies that focus in turn on <span class="hlt">air</span> quality, carbon, and <span class="hlt">water</span>) in 2020. On the supply side, we find that <span class="hlt">gas</span> sources other than SNG offer national <span class="hlt">air-carbon-water</span> co-benefits. However, we find striking <span class="hlt">air-carbon/water</span> trade-offs for SNG at the national scale. Moreover, the use of SNG significantly increases <span class="hlt">water</span> demand and carbon emissions in regions already suffering from the most severe <span class="hlt">water</span> stress and the highest per capita carbon footprint. On the end-use side, <span class="hlt">gas</span> substitution for coal can result in enormous variations in <span class="hlt">air</span> quality, carbon, and <span class="hlt">water</span> impacts, with notable <span class="hlt">air</span>-carbon synergies but <span class="hlt">air-water</span> trade-offs. Our study finds that, except for SNG, end-use choices generally have a much larger influence on <span class="hlt">air</span> quality, carbon emissions and <span class="hlt">water</span> use than do <span class="hlt">gas</span> source choices. Simultaneous consideration of <span class="hlt">air</span>, carbon, and <span class="hlt">water</span> impacts is necessary in designing both beneficial energy development and deployment policies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ACP....1311073B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ACP....1311073B"><span><span class="hlt">Air</span>-sea dimethylsulfide (DMS) <span class="hlt">gas</span> transfer in the North Atlantic: evidence for limited interfacial <span class="hlt">gas</span> <span class="hlt">exchange</span> at high wind speed</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bell, T. G.; De Bruyn, W.; Miller, S. D.; Ward, B.; Christensen, K.; Saltzman, E. S.</p> <p>2013-11-01</p> <p>Shipboard measurements of eddy covariance dimethylsulfide (DMS) <span class="hlt">air</span>-sea fluxes and seawater concentration were carried out in the North Atlantic bloom region in June/July 2011. <span class="hlt">Gas</span> transfer coefficients (k660) show a linear dependence on mean horizontal wind speed at wind speeds up to 11 m s-1. At higher wind speeds the relationship between k660 and wind speed weakens. At high winds, measured DMS fluxes were lower than predicted based on the linear relationship between wind speed and interfacial stress extrapolated from low to intermediate wind speeds. In contrast, the transfer coefficient for sensible heat did not exhibit this effect. The apparent suppression of <span class="hlt">air</span>-sea <span class="hlt">gas</span> flux at higher wind speeds appears to be related to sea state, as determined from shipboard wave measurements. These observations are consistent with the idea that long waves suppress near-surface <span class="hlt">water</span>-side turbulence, and decrease interfacial <span class="hlt">gas</span> transfer. This effect may be more easily observed for DMS than for less soluble gases, such as CO2, because the <span class="hlt">air</span>-sea <span class="hlt">exchange</span> of DMS is controlled by interfacial rather than bubble-mediated <span class="hlt">gas</span> transfer under high wind speed conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17328184','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17328184"><span><span class="hlt">Air-water</span> <span class="hlt">exchange</span> and dry deposition of polybrominated diphenyl ethers at a coastal site in Izmir Bay, Turkey.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cetin, Banu; Odabasi, Mustafa</p> <p>2007-02-01</p> <p>The <span class="hlt">air-water</span> <span class="hlt">exchange</span> of polybrominated diphenyl ethers (PBDEs), an emerging class of persistent organic pollutants (POPs), was investigated using paired <span class="hlt">air-water</span> samples (n = 15) collected in July and December, 2005 from Guzelyali Port in Izmir Bay, Turkey. Total dissolved-phase <span class="hlt">water</span> concentrations of PBDEs (sigma7PBDEs) were 212 +/- 65 and 87 +/- 57 pg L(-1) (average +/- SD) in summer and winter, respectively. BDE-209 was the most abundant congener in all samples, followed by BDE-99 and -47. Average ambient <span class="hlt">gas</span>-phase sigma7PBDE concentrations were between 189 +/- 61 (summer) and 76 +/- 65 pg m(-3) (winter). Net <span class="hlt">air-water</span> <span class="hlt">exchange</span> fluxes ranged from -0.9 +/- 1.0 (BDE-28) (volatilization) to 11.1 +/- 5.4 (BDE-209) ng m(-2) day(-1) (deposition). The BDE-28 fluxes were mainly volatilization while the other congeners were deposited. <span class="hlt">Gas</span>- and dissolved-phase concentrations were significantly correlated (P = 0.33-0.55, p < 0.05, except for BDE-209, r = 0.05, p > 0.05) indicating thatthe atmosphere controls the surface <span class="hlt">water</span> PBDE levels in this coastal environment. Estimated particulate dry deposition fluxes ranged between 2.7 +/- 1.9 (BDE-154) and 116 +/- 84 ng m(-2) day(-1) (BDE-209) indicating that dry deposition is also a significant input to surface <span class="hlt">waters</span> in the study area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018HMT....54.1951C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018HMT....54.1951C"><span>Performance of casting aluminum-silicon alloy condensing heating <span class="hlt">exchanger</span> for <span class="hlt">gas</span>-fired boiler</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cao, Weixue; Liu, Fengguo; You, Xue-yi</p> <p>2018-07-01</p> <p>Condensing <span class="hlt">gas</span> boilers are widely used due to their high heat efficiency, which comes from their ability to use the recoverable sensible heat and latent heat in flue <span class="hlt">gas</span>. The condensed <span class="hlt">water</span> of the boiler exhaust has strong corrosion effect on the heat <span class="hlt">exchanger</span>, which restricts the further application of the condensing <span class="hlt">gas</span> boiler. In recent years, a casting aluminum-silicon alloy (CASA), which boasts good anti-corrosion properties, has been introduced to condensing hot <span class="hlt">water</span> boilers. In this paper, the heat transfer performance, CO and NOx emission concentrations and CASA corrosion resistance of a heat <span class="hlt">exchanger</span> are studied by an efficiency bench test of the <span class="hlt">gas</span>-fired boiler. The experimental results are compared with heat <span class="hlt">exchangers</span> produced by Honeywell and Beka. The results show that the excess <span class="hlt">air</span> coefficient has a significant effect on the heat efficiency and CO and NOx emission of the CASA <span class="hlt">water</span> heater. When the excess <span class="hlt">air</span> coefficient of the CASA <span class="hlt">gas</span> boiler is 1.3, the CO and NOx emission concentration of the flue <span class="hlt">gas</span> satisfies the design requirements, and the heat efficiency of <span class="hlt">water</span> heater is 90.8%. In addition, with the increase of heat load rate, the heat transfer coefficient of the heat <span class="hlt">exchanger</span> and the heat efficiency of the <span class="hlt">water</span> heater are increased. However, when the heat load rate is at 90%, the NOx emission in the exhaust <span class="hlt">gas</span> is the highest. Furthermore, when the temperature of flue <span class="hlt">gas</span> is below 57 °C, the condensation of <span class="hlt">water</span> vapor occurs, and the pH of condensed <span class="hlt">water</span> is in the 2.5 5.5 range. The study shows that CASA <span class="hlt">water</span> heater has good corrosion resistance and a high heat efficiency of 88%. Compared with the heat <span class="hlt">exchangers</span> produced by Honeywell and Beka, there is still much work to do in optimizing and improving the <span class="hlt">water</span> heater.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018HMT...tmp...25C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018HMT...tmp...25C"><span>Performance of casting aluminum-silicon alloy condensing heating <span class="hlt">exchanger</span> for <span class="hlt">gas</span>-fired boiler</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cao, Weixue; Liu, Fengguo; You, Xue-yi</p> <p>2018-01-01</p> <p>Condensing <span class="hlt">gas</span> boilers are widely used due to their high heat efficiency, which comes from their ability to use the recoverable sensible heat and latent heat in flue <span class="hlt">gas</span>. The condensed <span class="hlt">water</span> of the boiler exhaust has strong corrosion effect on the heat <span class="hlt">exchanger</span>, which restricts the further application of the condensing <span class="hlt">gas</span> boiler. In recent years, a casting aluminum-silicon alloy (CASA), which boasts good anti-corrosion properties, has been introduced to condensing hot <span class="hlt">water</span> boilers. In this paper, the heat transfer performance, CO and NOx emission concentrations and CASA corrosion resistance of a heat <span class="hlt">exchanger</span> are studied by an efficiency bench test of the <span class="hlt">gas</span>-fired boiler. The experimental results are compared with heat <span class="hlt">exchangers</span> produced by Honeywell and Beka. The results show that the excess <span class="hlt">air</span> coefficient has a significant effect on the heat efficiency and CO and NOx emission of the CASA <span class="hlt">water</span> heater. When the excess <span class="hlt">air</span> coefficient of the CASA <span class="hlt">gas</span> boiler is 1.3, the CO and NOx emission concentration of the flue <span class="hlt">gas</span> satisfies the design requirements, and the heat efficiency of <span class="hlt">water</span> heater is 90.8%. In addition, with the increase of heat load rate, the heat transfer coefficient of the heat <span class="hlt">exchanger</span> and the heat efficiency of the <span class="hlt">water</span> heater are increased. However, when the heat load rate is at 90%, the NOx emission in the exhaust <span class="hlt">gas</span> is the highest. Furthermore, when the temperature of flue <span class="hlt">gas</span> is below 57 °C, the condensation of <span class="hlt">water</span> vapor occurs, and the pH of condensed <span class="hlt">water</span> is in the 2.5 5.5 range. The study shows that CASA <span class="hlt">water</span> heater has good corrosion resistance and a high heat efficiency of 88%. Compared with the heat <span class="hlt">exchangers</span> produced by Honeywell and Beka, there is still much work to do in optimizing and improving the <span class="hlt">water</span> heater.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26868055','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26868055"><span>A meta-analysis of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status responses to drought.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yan, Weiming; Zhong, Yangquanwei; Shangguan, Zhouping</p> <p>2016-02-12</p> <p>Drought is considered to be one of the most devastating natural hazards, and it is predicted to become increasingly frequent and severe in the future. Understanding the plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status response to drought is very important with regard to future climate change. We conducted a meta-analysis based on studies of plants worldwide and aimed to determine the changes in <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status under different drought intensities (mild, moderate and severe), different photosynthetic pathways (C3 and C4) and growth forms (herbs, shrubs, trees and lianas). Our results were as follows: 1) drought negatively impacted <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status, and stomatal conductance (gs) decreased more than other physiological traits and declined to the greatest extent in shrubs and C3 plants. Furthermore, C4 plants had an advantage compared to C3 plants under the same drought conditions. 2) The decrease in gs mainly reduced the transpiration rate (Tr), and gs could explain 55% of the decrease in the photosynthesis (A) and 74% of the decline in Tr. 3). Finally, <span class="hlt">gas</span> <span class="hlt">exchange</span> showed a close relationship with the leaf <span class="hlt">water</span> status. Our study provides comprehensive information about the changes in plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status under drought.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4751433','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4751433"><span>A meta-analysis of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status responses to drought</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yan, Weiming; Zhong, Yangquanwei; Shangguan, Zhouping</p> <p>2016-01-01</p> <p>Drought is considered to be one of the most devastating natural hazards, and it is predicted to become increasingly frequent and severe in the future. Understanding the plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status response to drought is very important with regard to future climate change. We conducted a meta-analysis based on studies of plants worldwide and aimed to determine the changes in <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status under different drought intensities (mild, moderate and severe), different photosynthetic pathways (C3 and C4) and growth forms (herbs, shrubs, trees and lianas). Our results were as follows: 1) drought negatively impacted <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status, and stomatal conductance (gs) decreased more than other physiological traits and declined to the greatest extent in shrubs and C3 plants. Furthermore, C4 plants had an advantage compared to C3 plants under the same drought conditions. 2) The decrease in gs mainly reduced the transpiration rate (Tr), and gs could explain 55% of the decrease in the photosynthesis (A) and 74% of the decline in Tr. 3). Finally, <span class="hlt">gas</span> <span class="hlt">exchange</span> showed a close relationship with the leaf <span class="hlt">water</span> status. Our study provides comprehensive information about the changes in plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status under drought. PMID:26868055</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19778365','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19778365"><span>Why and how terrestrial plants <span class="hlt">exchange</span> gases with <span class="hlt">air</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cieslik, S; Omasa, K; Paoletti, E</p> <p>2009-11-01</p> <p>This work is intended as a review of <span class="hlt">gas</span> <span class="hlt">exchange</span> processes between the atmosphere and the terrestrial vegetation, which have been known for more than two centuries since the discovery of photosynthesis. The physical and biological mechanisms of <span class="hlt">exchange</span> of carbon dioxide, <span class="hlt">water</span> vapour, volatile organic compounds emitted by plants and <span class="hlt">air</span> pollutants taken up by them, is critically reviewed. The role of stomatal physiology is emphasised, as it controls most of these processes. The techniques used for measurement of <span class="hlt">gas</span> <span class="hlt">exchange</span> fluxes between the atmosphere and vegetation are outlined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998PhDT........22V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998PhDT........22V"><span>Photochemical influences on the <span class="hlt">air-water</span> <span class="hlt">exchange</span> of mercury</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vette, Alan Frederic</p> <p></p> <p>The formation of dissolved gaseous mercury (DGM) in natural <span class="hlt">waters</span> is an important component in the biogeochemical cycle of mercury (Hg). The predominate form of DGM in natural <span class="hlt">waters</span>, gaseous elemental Hg (Hg0), may be transferred from the <span class="hlt">water</span> to the atmosphere. <span class="hlt">Gas</span> <span class="hlt">exchange</span> may reduce the amount of Hg available for methyl-Hg formation, the most toxic form of Hg that bioaccumulates in the food chain. Determining the mechanisms and rates of DGM formation is essential in understanding the fate and cycling of Hg in aquatic ecosystems. Field and laboratory experiments were conducted to evaluate the effect of light on DGM formation in surface <span class="hlt">waters</span> containing different levels of dissolved organic carbon (DOC). <span class="hlt">Water</span> samples collected from the Tahqwamenon River and Whitefish Bay on Lake Superior were amended with divalent Hg (Hg2+) and irradiated under a variety of reaction conditions to determine rates of DGM formation. The <span class="hlt">water</span> samples were also analyzed for various Hg species (total, filtered, easily reducible and dissolved gaseous Hg), DOC and light attenuation. Additional field studies were conducted on Lake Michigan to measure gaseous Hg in <span class="hlt">air</span> and <span class="hlt">water</span>. These data were used to develop a mechanistic model to estimate <span class="hlt">air-water</span> <span class="hlt">exchange</span> of gaseous Hg. This research found that photochemical formation of DGM was affected by penetration of UV A radiation (320-400 nm). Formation of DGM was enhanced at higher DOC concentrations, indicating DOC photosensitized the reduction of Hg2+ to Hg0. Wavelength studies determined that formation of DGM was significantly reduced in the absence of UV A. Field studies showed DGM concentrations were highest near the <span class="hlt">water</span> surface and peaked at mid-day, indicating a photo-induced source of DGM. The conversion of reducible Hg2+ to Hg0 was suppressed in high DOC <span class="hlt">waters</span> where UV A penetration was limited. The mechanistic model predicted similar DGM concentrations to the observed values and demonstrated that deposition and emission</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15557031','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15557031"><span>The hyperoxic switch: assessing respiratory <span class="hlt">water</span> loss rates in tracheate arthropods with continuous <span class="hlt">gas</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lighton, John R B; Schilman, Pablo E; Holway, David A</p> <p>2004-12-01</p> <p>Partitioning the relative contributions of cuticular and respiratory <span class="hlt">water</span> loss in a tracheate arthropod is relatively easy if it undergoes discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> cycles or DGCs, leaving its rate of cuticular <span class="hlt">water</span> loss in primary evidence while its spiracles are closed. Many arthropods are not so obliging and emit CO(2) continuously, making cuticular and respiratory <span class="hlt">water</span> losses difficult or impossible to partition. We report here that by switching ambient <span class="hlt">air</span> from 21 to 100% O(2), marked spiracular constriction takes place, causing a transient but substantial - up to 90% - reduction in CO(2) output. A reduction in <span class="hlt">water</span> loss rate occurs at the same time. Using this approach, we investigated respiratory <span class="hlt">water</span> loss in Drosophila melanogaster and in two ant species, Forelius mccooki and Pogonomyrmex californicus. Our results - respiratory <span class="hlt">water</span> loss estimates of 23%, 7.6% and 5.6% of total <span class="hlt">water</span> loss rates, respectively - are reasonable in light of literature estimates, and suggest that the 'hyperoxic switch' may allow straightforward estimation of respiratory <span class="hlt">water</span> loss rates in arthropods lacking discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span>. In P. californicus, which we were able to measure with and without a DGC, presence or absence of a DGC did not affect respiratory vs total <span class="hlt">water</span> loss rates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26538177','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26538177"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> and dive characteristics of the free-swimming backswimmer Anisops deanei.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jones, Karl K; Snelling, Edward P; Watson, Amy P; Seymour, Roger S</p> <p>2015-11-01</p> <p>Many aquatic insects utilise <span class="hlt">air</span> bubbles on the surface of their bodies to supply O2 while they dive. The bubbles can simply store O2, as in the case of an '<span class="hlt">air</span> store', or they can act as a physical '<span class="hlt">gas</span> gill', extracting O2 from the <span class="hlt">water</span>. Backswimmers of the genus Anisops augment their <span class="hlt">air</span> store with O2 from haemoglobin cells located in the abdomen. The O2 release from the haemoglobin helps stabilise bubble volume, enabling backswimmers to remain near neutrally buoyant for a period of the dive. It is generally assumed that the backswimmer <span class="hlt">air</span> store does not act as a <span class="hlt">gas</span> gill and that <span class="hlt">gas</span> <span class="hlt">exchange</span> with the <span class="hlt">water</span> is negligible. This study combines measurements of dive characteristics under different exotic gases (N2, He, SF6, CO) with mathematical modelling, to show that the <span class="hlt">air</span> store of the backswimmer Anisops deanei does <span class="hlt">exchange</span> gases with the <span class="hlt">water</span>. Our results indicate that approximately 20% of O2 consumed during a dive is obtained directly from the <span class="hlt">water</span>. Oxygen from the <span class="hlt">water</span> complements that released from the haemoglobin, extending the period of near-neutral buoyancy and increasing dive duration. © 2015. Published by The Company of Biologists Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12578005','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12578005"><span>Effects of <span class="hlt">air</span> current speed on <span class="hlt">gas</span> <span class="hlt">exchange</span> in plant leaves and plant canopies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kitaya, Y; Tsuruyama, J; Shibuya, T; Yoshida, M; Kiyota, M</p> <p>2003-01-01</p> <p>To obtain basic data on adequate <span class="hlt">air</span> circulation to enhance plant growth in a closed plant culture system in a controlled ecological life support system (CELSS), an investigation was made of the effects of the <span class="hlt">air</span> current speed ranging from 0.01 to 1.0 m s-1 on photosynthesis and transpiration in sweetpotato leaves and photosynthesis in tomato seedlings canopies. The <span class="hlt">gas</span> <span class="hlt">exchange</span> rates in leaves and canopies were determined by using a chamber method with an infrared <span class="hlt">gas</span> analyzer. The net photosynthetic rate and the transpiration rate increased significantly as the <span class="hlt">air</span> current speeds increased from 0.01 to 0.2 m s-1. The transpiration rate increased gradually at <span class="hlt">air</span> current speeds ranging from 0.2 to 1.0 m s-1 while the net photosynthetic rate was almost constant at <span class="hlt">air</span> current speeds ranging from 0.5 to 1.0 m s-1. The increase in the net photosynthetic and transpiration rates were strongly dependent on decreased boundary-layer resistances against <span class="hlt">gas</span> diffusion. The net photosynthetic rate of the plant canopy was doubled by an increased <span class="hlt">air</span> current speed from 0.1 to 1.0 m s-1 above the plant canopy. The results demonstrate the importance of <span class="hlt">air</span> movement around plants for enhancing the <span class="hlt">gas</span> <span class="hlt">exchange</span> in the leaf, especially in plant canopies in the CELSS. c2002 COSPAR. Published by Elsevier Science Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23821716','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23821716"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> patterns and <span class="hlt">water</span> loss rates in the Table Mountain cockroach, Aptera fusca (Blattodea: Blaberidae).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Groenewald, Berlizé; Bazelet, Corinna S; Potter, C Paige; Terblanche, John S</p> <p>2013-10-15</p> <p>The importance of metabolic rate and/or spiracle modulation for saving respiratory <span class="hlt">water</span> is contentious. One major explanation for <span class="hlt">gas</span> <span class="hlt">exchange</span> pattern variation in terrestrial insects is to effect a respiratory <span class="hlt">water</span> loss (RWL) saving. To test this, we measured the rates of CO2 and H2O release ( and , respectively) in a previously unstudied, mesic cockroach, Aptera fusca, and compared <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> loss parameters among the major <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns (continuous, cyclic, discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span>) at a range of temperatures. Mean , and per unit did not differ among the <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns at all temperatures (P>0.09). There was no significant association between temperature and <span class="hlt">gas</span> <span class="hlt">exchange</span> pattern type (P=0.63). Percentage of RWL (relative to total <span class="hlt">water</span> loss) was typically low (9.79±1.84%) and did not differ significantly among <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns at 15°C (P=0.26). The method of estimation had a large impact on the percentage of RWL, and of the three techniques investigated (traditional, regression and hyperoxic switch), the traditional method generally performed best. In many respects, A. fusca has typical <span class="hlt">gas</span> <span class="hlt">exchange</span> for what might be expected from other insects studied to date (e.g. , , RWL and cuticular <span class="hlt">water</span> loss). However, we found for A. fusca that expressed as a function of metabolic rate was significantly higher than the expected consensus relationship for insects, suggesting it is under considerable pressure to save <span class="hlt">water</span>. Despite this, we found no consistent evidence supporting the conclusion that transitions in pattern type yield reductions in RWL in this mesic cockroach.</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_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_3 --> <div id="page_4" 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_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</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="61"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A43G2558W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A43G2558W"><span><span class="hlt">Air</span>-sea <span class="hlt">exchange</span> and <span class="hlt">gas</span>-particle partitioning of polycyclic aromatic hydrocarbons over the northwestern Pacific Ocean: Role of East Asian continental outflow</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Z.; Guo, Z.</p> <p>2017-12-01</p> <p>We measured 15 parent polycyclic aromatic hydrocarbons (PAHs) in atmosphere and <span class="hlt">water</span> during a research cruise from the East China Sea (ECS) to the northwestern Pacific Ocean (NWP) in the spring of 2015 to investigate the occurrence, <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span>, and <span class="hlt">gas</span>-particle partitioning of PAHs with a particular focus on the influence of East Asian continental outflow. The gaseous PAH composition and identification of sources were consistent with PAHs from the upwind area, indicating that the gaseous PAHs (three- to five-ring PAHs) were influenced by upwind land pollution. In addition, <span class="hlt">air</span>-sea <span class="hlt">exchange</span> fluxes of gaseous PAHs were estimated to be -54.2 to 107.4 ng m-2 d-1, and was indicative of variations of land-based PAH inputs. The logarithmic <span class="hlt">gas</span>-particle partition coefficient (logKp) of PAHs regressed linearly against the logarithmic subcooled liquid vapor pressure, with a slope of -0.25. This was significantly larger than the theoretical value (-1), implying disequilibrium between the gaseous and particulate PAHs over the NWP. The non-equilibrium of PAH <span class="hlt">gas</span>-particle partitioning was shielded from the volatilization of three-ring gaseous PAHs from seawater and lower soot concentrations in particular when the oceanic <span class="hlt">air</span> masses prevailed. Modeling PAH absorption into organic matter and adsorption onto soot carbon revealed that the status of PAH <span class="hlt">gas</span>-particle partitioning deviated more from the modeling Kp for oceanic <span class="hlt">air</span> masses than those for continental <span class="hlt">air</span> masses, which coincided with higher volatilization of three-ring PAHs and confirmed the influence of <span class="hlt">air</span>-sea <span class="hlt">exchange</span>. Meanwhile, significant linear regressions between logKp and logKoa (logKsa) for PAHs were observed for continental <span class="hlt">air</span> masses, suggesting the dominant effect of East Asian continental outflow on atmospheric PAHs over the NWP during the sampling campaign.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23238597','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23238597"><span><span class="hlt">Water-gas</span> <span class="hlt">exchange</span> of organochlorine pesticides at Lake Chaohu, a large Chinese lake.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ouyang, Hui-Ling; He, Wei; Qin, Ning; Kong, Xiang-Zhen; Liu, Wen-Xiu; He, Qi-Shuang; Yang, Chen; Jiang, Yu-Jiao; Wang, Qing-Mei; Yang, Bin; Xu, Fu-Liu</p> <p>2013-04-01</p> <p>Organochlorine pesticides (OCPs), a potential threat to ecosystems and human health, are still widely residual in the environment. The residual levels of OCPs in the <span class="hlt">water</span> and <span class="hlt">gas</span> phase were monitored in Lake Chaohu, a large Chinese lake, from March 2010 to February 2011. Nineteen types of OCPs were detected in the <span class="hlt">water</span> with a total concentration of 7.27 ± 3.32 ng/l. Aldrin, DDTs and HCHs were the major OCPs in the <span class="hlt">water</span>, accounting for 38.3%, 28.9% and 23.6% of the total, respectively. The highest mean concentration (12.32 ng/l) in the <span class="hlt">water</span> was found in September, while the lowest (1.74 ng/l) was found in November. Twenty types of gaseous OCPs were detected in the atmosphere with a total concentration of 542.0 ± 636.5 pg/m(3). Endosulfan, DDTs and chlordane were the major gaseous OCPs in the atmosphere, accounting for 48.9%, 22.5% and 14.4% of the total, respectively. The mean concentration of gaseous OCPs was significantly higher in summer than in winter. o,p'-DDE was the main metabolite of DDT in both the <span class="hlt">water</span> and <span class="hlt">gas</span> phase. Of the HCHs, 52.3% existed as β-HCH in the <span class="hlt">water</span>, while α-HCH (37.9%) and γ-HCH (30.9%) were dominant isomers in the <span class="hlt">gas</span> phase. The average fluxes were -21.11, -3.30, -152.41, -35.50 and -1314.15 ng/(m(2) day) for α-HCH, γ-HCH, HCB, DDT and DDE, respectively. The <span class="hlt">water-gas</span> <span class="hlt">exchanges</span> of the five types of OCPs indicate that <span class="hlt">water</span> was the main potential source of gaseous OCPs in the atmosphere. A sensitivity analysis indicated that the <span class="hlt">water-gas</span> flux of α-HCH, γ-HCH and DDT is more vulnerable than that of HCB and DDE to the variation of the parameters. The possible source of the HCHs in the <span class="hlt">water</span> was from the historical usage of lindane; however, that in the <span class="hlt">air</span> was mainly from the recent usage of lindane. The technical DDT and dicofol might be the source of DDTs in the <span class="hlt">water</span> and <span class="hlt">air</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013BGeo...10.1379C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013BGeo...10.1379C"><span>Technical Note: A simple method for <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements in mesocosms and its application in carbon budgeting</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Czerny, J.; Schulz, K. G.; Ludwig, A.; Riebesell, U.</p> <p>2013-03-01</p> <p>Mesocosms as large experimental units provide the opportunity to perform elemental mass balance calculations, e.g. to derive net biological turnover rates. However, the system is in most cases not closed at the <span class="hlt">water</span> surface and gases <span class="hlt">exchange</span> with the atmosphere. Previous attempts to budget carbon pools in mesocosms relied on educated guesses concerning the <span class="hlt">exchange</span> of CO2 with the atmosphere. Here, we present a simple method for precise determination of <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> in mesocosms using N2O as a deliberate tracer. Beside the application for carbon budgeting, transfer velocities can be used to calculate <span class="hlt">exchange</span> rates of any <span class="hlt">gas</span> of known concentration, e.g. to calculate aquatic production rates of climate relevant trace gases. Using an arctic KOSMOS (Kiel Off Shore Mesocosms for future Ocean Simulation) experiment as an exemplary dataset, it is shown that the presented method improves accuracy of carbon budget estimates substantially. Methodology of manipulation, measurement, data processing and conversion to CO2 fluxes are explained. A theoretical discussion of prerequisites for precise <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements provides a guideline for the applicability of the method under various experimental conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013BGeo...10.5793S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013BGeo...10.5793S"><span>Biology and <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> controls on the distribution of carbon isotope ratios (δ13C) in the ocean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmittner, A.; Gruber, N.; Mix, A. C.; Key, R. M.; Tagliabue, A.; Westberry, T. K.</p> <p>2013-09-01</p> <p>Analysis of observations and sensitivity experiments with a new three-dimensional global model of stable carbon isotope cycling elucidate processes that control the distribution of δ13C of dissolved inorganic carbon (DIC) in the contemporary and preindustrial ocean. Biological fractionation and the sinking of isotopically light δ13C organic matter from the surface into the interior ocean leads to low δ13CDIC values at depths and in high latitude surface <span class="hlt">waters</span> and high values in the upper ocean at low latitudes with maxima in the subtropics. <span class="hlt">Air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> has two effects. First, it acts to reduce the spatial gradients created by biology. Second, the associated temperature-dependent fractionation tends to increase (decrease) δ13CDIC values of colder (warmer) <span class="hlt">water</span>, which generates gradients that oppose those arising from biology. Our model results suggest that both effects are similarly important in influencing surface and interior δ13CDIC distributions. However, since <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> is slow in the modern ocean, the biological effect dominates spatial δ13CDIC gradients both in the interior and at the surface, in contrast to conclusions from some previous studies. Calcium carbonate cycling, pH dependency of fractionation during <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span>, and kinetic fractionation have minor effects on δ13CDIC. Accumulation of isotopically light carbon from anthropogenic fossil fuel burning has decreased the spatial variability of surface and deep δ13CDIC since the industrial revolution in our model simulations. Analysis of a new synthesis of δ13CDIC measurements from years 1990 to 2005 is used to quantify preformed and remineralized contributions as well as the effects of biology and <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span>. The model reproduces major features of the observed large-scale distribution of δ13CDIC as well as the individual contributions and effects. Residual misfits are documented and analyzed. Simulated surface and subsurface δ13CDIC are influenced by</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/13382','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/13382"><span><span class="hlt">Water</span> use in forest canopy black cherry trees and its relationship to leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and environment</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>B. J. Joyce; K. C. Steiner; J. M. Skelly</p> <p>1996-01-01</p> <p>Models of canopy <span class="hlt">gas</span> <span class="hlt">exchange</span> are needed to connect leaf-level measurement to higher scales. Because of the correspondence between leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> use, it may be possible to predict variation in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> at the canopy level by monitoring rates of branch <span class="hlt">water</span> use.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=197971&Lab=NERL&keyword=gas+AND+behaviour&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=197971&Lab=NERL&keyword=gas+AND+behaviour&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>Simulating the Vapour Phase <span class="hlt">Air/Water</span> <span class="hlt">Exchange</span> of p,p′-DDE, p,p′-DDT, Lindane, and 2,3,7,8-Tetrachlorodibenzodioxin</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>Uncertainties in our understanding of gaseous <span class="hlt">air/water</span> <span class="hlt">exchange</span> have emerged as major sources of concern in efforts to construct global and regional mass balances of both the green house <span class="hlt">gas</span> carbon dioxide and semi-volatile persistent, bioaccumulative and toxic chemicals. Hoff e...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26359720','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26359720"><span>Different Apparent <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Coefficients for CO2 and CH4: Comparing a Brown-<span class="hlt">Water</span> and a Clear-<span class="hlt">Water</span> Lake in the Boreal Zone during the Whole Growing Season.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rantakari, Miitta; Heiskanen, Jouni; Mammarella, Ivan; Tulonen, Tiina; Linnaluoma, Jessica; Kankaala, Paula; Ojala, Anne</p> <p>2015-10-06</p> <p>The <span class="hlt">air-water</span> <span class="hlt">exchange</span> of carbon dioxide (CO2) and methane (CH4) is a central process during attempts to establish carbon budgets for lakes and landscapes containing lakes. Lake-atmosphere diffusive <span class="hlt">gas</span> <span class="hlt">exchange</span> is dependent on the concentration gradient between <span class="hlt">air</span> and surface <span class="hlt">water</span> and also on the <span class="hlt">gas</span> transfer velocity, often described with the <span class="hlt">gas</span> transfer coefficient k. We used the floating-chamber method in connection with surface <span class="hlt">water</span> <span class="hlt">gas</span> concentration measurements to estimate the <span class="hlt">gas</span> transfer velocity of CO2 (kCO2) and CH4 (kCH4) weekly throughout the entire growing season in two contrasting boreal lakes, a humic oligotrophic lake and a clear-<span class="hlt">water</span> productive lake, in order to investigate the earlier observed differences between kCO2 and kCH4. We found that the seasonally averaged <span class="hlt">gas</span> transfer velocity of CH4 was the same for both lakes. When the lakes were sources of CO2, the <span class="hlt">gas</span> transfer velocity of CO2 was also similar between the two study lakes. The <span class="hlt">gas</span> transfer velocity of CH4 was constantly higher than that of CO2 in both lakes, a result also found in other studies but for reasons not yet fully understood. We found no differences between the lakes, demonstrating that the difference between kCO2 and kCH4 is not dependent on season or the characteristics of the lake.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24484174','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24484174"><span>Oxo-<span class="hlt">exchange</span> of <span class="hlt">gas</span>-phase uranyl, neptunyl, and plutonyl with <span class="hlt">water</span> and methanol.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lucena, Ana F; Odoh, Samuel O; Zhao, Jing; Marçalo, Joaquim; Schreckenbach, Georg; Gibson, John K</p> <p>2014-02-17</p> <p>A challenge in actinide chemistry is activation of the strong bonds in the actinyl ions, AnO2(+) and AnO2(2+), where An = U, Np, or Pu. Actinyl activation in oxo-<span class="hlt">exchange</span> with <span class="hlt">water</span> in solution is well established, but the <span class="hlt">exchange</span> mechanisms are unknown. <span class="hlt">Gas</span>-phase actinyl oxo-<span class="hlt">exchange</span> is a means to probe these processes in detail for simple systems, which are amenable to computational modeling. <span class="hlt">Gas</span>-phase <span class="hlt">exchange</span> reactions of UO2(+), NpO2(+), PuO2(+), and UO2(2+) with <span class="hlt">water</span> and methanol were studied by experiment and density functional theory (DFT); reported for the first time are experimental results for UO2(2+) and for methanol <span class="hlt">exchange</span>, as well as <span class="hlt">exchange</span> rate constants. Key findings are faster <span class="hlt">exchange</span> of UO2(2+) versus UO2(+) and faster <span class="hlt">exchange</span> with methanol versus <span class="hlt">water</span>; faster <span class="hlt">exchange</span> of UO2(+) versus PuO2(+) was quantified. Computed potential energy profiles (PEPs) are in accord with the observed kinetics, validating the utility of DFT to model these <span class="hlt">exchange</span> processes. The seemingly enigmatic result of faster <span class="hlt">exchange</span> for uranyl, which has the strongest oxo-bonds, may reflect reduced covalency in uranyl as compared with plutonyl.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17936295','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17936295"><span>Breathe softly, beetle: continuous <span class="hlt">gas</span> <span class="hlt">exchange</span>, <span class="hlt">water</span> loss and the role of the subelytral space in the tenebrionid beetle, Eleodes obscura.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schilman, Pablo E; Kaiser, Alexander; Lighton, John R B</p> <p>2008-01-01</p> <p>Flightless, diurnal tenebrionid beetles are commonly found in deserts. They possess a curious morphological adaptation, the subelytral cavity (an <span class="hlt">air</span> space beneath the fused elytra) the function of which is not completely understood. In the tenebrionid beetle Eleodes obscura, we measured abdominal movements within the subelytral cavity, and the activity of the pygidial cleft (which seals or unseals the subelytral cavity), simultaneously with total CO2 release rate and <span class="hlt">water</span> loss rate. First, we found that E. obscura has the lowest cuticular permeability measured in flow-through respirometry in an insect (0.90 microg H2O cm(-2) Torr(-1) h(-1)). Second, it does not exhibit a discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> cycle. Third, we describe the temporal coupling between <span class="hlt">gas</span> <span class="hlt">exchange</span>, <span class="hlt">water</span> loss, subelytral space volume, and the capacity of the subelytral space to <span class="hlt">exchange</span> gases with its surroundings as indicated by pygidial cleft state. Fourth, we suggest possible mechanisms that may reduce respiratory <span class="hlt">water</span> loss rates in E. obscura. Finally, we suggest that E. obscura cannot <span class="hlt">exchange</span> respiratory gases discontinuously because of a morphological constraint (small tracheal or spiracular conductance). This "conductance constraint hypothesis" may help to explain the otherwise puzzling phylogenetic patterns of continuous vs. discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> observed in tracheate arthropods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12806145','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12806145"><span>Mercury <span class="hlt">exchange</span> at the <span class="hlt">air-water</span>-soil interface: an overview of methods.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fang, Fengman; Wang, Qichao; Liu, Ruhai</p> <p>2002-06-12</p> <p>An attempt is made to assess the present knowledge about the methods of determining mercury (Hg) <span class="hlt">exchange</span> at the <span class="hlt">air-water</span>-soil interface during the past 20 years. Methods determining processes of wet and dry removal/deposition of atmospheric Hg to aquatic and terrestrial ecosystems, as well as methods determining Hg emission fluxes to the atmosphere from natural surfaces (soil and <span class="hlt">water</span>) are discussed. On the basis of the impressive advances that have been made in the areas relating to Hg <span class="hlt">exchange</span> among <span class="hlt">air-soil-water</span> interfaces, we analyzed existing problems and shortcomings in our current knowledge. In addition, some important fields worth further research are discussed and proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhDT........15Q','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhDT........15Q"><span>The <span class="hlt">Air-Carbon-Water</span> Synergies and Trade-Offs in China's Natural <span class="hlt">Gas</span> Industry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Qin, Yue</p> <p></p> <p>China's coal-dominated energy structure is partly responsible for its domestic <span class="hlt">air</span> pollution, local <span class="hlt">water</span> stress, and the global climate change. Primarily to tackle the haze issue, China has been actively promoting a nationwide coal to natural <span class="hlt">gas</span> end-use switch. My dissertation focuses on evaluating the <span class="hlt">air</span> quality, carbon, and <span class="hlt">water</span> impacts and their interactions in China's natural <span class="hlt">gas</span> industry. Chapter 2 assesses the lifecycle climate performance of China's shale <span class="hlt">gas</span> in comparison to coal based on stage-level energy consumption and methane leakage rates. I find the mean lifecycle carbon footprint of shale <span class="hlt">gas</span> is about 30-50% lower than that of coal under both 20 year and 100 year global warming potentials (GWP20 and GWP100). However, primarily due to large uncertainties in methane leakage, the lifecycle carbon footprint of shale <span class="hlt">gas</span> in China could be 15-60% higher than that of coal across sectors under GWP20. Chapter 3 evaluates the <span class="hlt">air</span> quality, human health, and the climate impacts of China's coal-based synthetic natural <span class="hlt">gas</span> (SNG) development. Based on earlier 2020 SNG production targets, I conduct an integrated assessment to identify production technologies and end-use applications that will bring as large <span class="hlt">air</span> quality and health benefits as possible while keeping carbon penalties as small as possible. I find that, due to inefficient and uncontrolled coal combustion in households, allocating currently available SNG to the residential sector proves to be the best SNG allocation option. Chapter 4 compares the <span class="hlt">air</span> quality, carbon, and <span class="hlt">water</span> impacts of China's six major <span class="hlt">gas</span> sources under three end-use substitution scenarios, which are focused on maximizing <span class="hlt">air</span> pollutant emission reductions, CO 2 emission reductions, and <span class="hlt">water</span> stress index (WSI)-weighted <span class="hlt">water</span> consumption reductions, respectively. I find striking national <span class="hlt">air-carbon/water</span> trade-offs due to SNG, which also significantly increases <span class="hlt">water</span> demands and carbon emissions in regions already suffering from</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.B34A..02L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.B34A..02L"><span><span class="hlt">Air-Water</span> <span class="hlt">Exchange</span> of Legacy and Emerging Organic Pollutants across the Great Lakes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lohmann, R.; Ruge, Z.; Khairy, M.; Muir, D.; Helm, P.</p> <p>2014-12-01</p> <p>Organochlorine pesticides (OCPs) and polychlorinated biphenyls (PCBs) are transported to great <span class="hlt">water</span> bodies via long-range atmospheric transport and released from the surface <span class="hlt">water</span> as <span class="hlt">air</span> concentrations continue to diminish. As the largest fresh <span class="hlt">water</span> bodies in North America, the Great Lakes have both the potential to accumulate and serve as a secondary source of persistent bioaccumulative toxins. OCP and PCB concentrations were sampled at 30+ sites across Lake Superior, Ontario and Erie in the summer of 2011. Polyethylene passive samplers (PEs) were simultaneously deployed in surface <span class="hlt">water</span> and near surface atmosphere to determine <span class="hlt">air-water</span> gaseous <span class="hlt">exchange</span> of OCPs and PCBs. In Lake Superior, surface <span class="hlt">water</span> and atmospheric concentrations were dominated by α-HCH (average 250 pg/L and 4.2 pg/m3, respectively), followed by HCB (average 17 pg/L and 89 pg/m3, respectively). <span class="hlt">Air-water</span> <span class="hlt">exchange</span> varied greatly between sites and individual OCPs, however α-endosulfan was consistently deposited into the surface <span class="hlt">water</span> (average 19 pg/m2/day). PCBs in the <span class="hlt">air</span> and <span class="hlt">water</span> were characterized by penta- and hexachlorobiphenyls with distribution along the coast correlated with proximity to developed areas. <span class="hlt">Air-water</span> <span class="hlt">exchange</span> gradients generally yielded net volatilization of PCBs out of Lake Superior. Gaseous concentrations of hexachlorobenzene, dieldrin and chlordanes were significantly higher (p < 0.05) at Lake Erie than Lake Ontario. A multiple linear regression that incorporated meteorological, landuse and population data was used to explain variability in the atmospheric concentrations. Results indicated that landuse (urban and/or cropland) greatly explained the variability in the data. Freely dissolved concentrations of OCPs (<LOD-114 pg/L) were lower than previously detected concentrations. Nonetheless, concentrations of p,p'-DDE and chlordanes were higher than <span class="hlt">water</span> quality guidelines for the protection of human health from the consumption of fish. Spatial distributions of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130011141','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130011141"><span>Prototype Vent <span class="hlt">Gas</span> Heat <span class="hlt">Exchanger</span> for Exploration EVA - Performance and Manufacturing Characteristics</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Quinn, Gregory J.; Strange, Jeremy; Jennings, Mallory</p> <p>2013-01-01</p> <p>NASA is developing new portable life support system (PLSS) technologies, which it is demonstrating in an unmanned ground based prototype unit called PLSS 2.0. One set of technologies within the PLSS provides suitable ventilation to an astronaut while on an EVA. A new component within the ventilation <span class="hlt">gas</span> loop is a liquid-to-<span class="hlt">gas</span> heat <span class="hlt">exchanger</span> to transfer excess heat from the <span class="hlt">gas</span> to the thermal control system s liquid coolant loop. A unique bench top prototype heat <span class="hlt">exchanger</span> was built and tested for use in PLSS 2.0. The heat <span class="hlt">exchanger</span> was designed as a counter-flow, compact plate fin type using stainless steel. Its design was based on previous compact heat <span class="hlt">exchangers</span> manufactured by United Technologies Aerospace Systems (UTAS), but was half the size of any previous heat <span class="hlt">exchanger</span> model and one third the size of previous liquid-to-<span class="hlt">gas</span> heat <span class="hlt">exchangers</span>. The prototype heat <span class="hlt">exchanger</span> was less than 40 cubic inches and weighed 2.57 lb. Performance of the heat <span class="hlt">exchanger</span> met the requirements and the model predictions. The <span class="hlt">water</span> side and <span class="hlt">gas</span> side pressure drops were less 0.8 psid and 0.5 inches of <span class="hlt">water</span>, respectively, and an effectiveness of 94% was measured at the nominal <span class="hlt">air</span> side pressure of 4.1 psia.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24643387','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24643387"><span>Concentrations, atmospheric partitioning, and <span class="hlt">air-water</span>/soil surface <span class="hlt">exchange</span> of polychlorinated dibenzo-p-dioxin and dibenzofuran along the upper reaches of the Haihe River basin, North China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nie, Zhiqiang; Die, Qingqi; Yang, Yufei; Tang, Zhenwu; Wang, Qi; Huang, Qifei</p> <p>2014-01-01</p> <p>Polychlorinated dibenzo-p-dioxin and dibenzofuran (PCDD/PCDF) were overall measured and compared in ambient <span class="hlt">air</span>, <span class="hlt">water</span>, soils, and sediments along the upper reaches of the Haihe River of North China, so as to evaluate their concentrations, profiles, and to understand the processes of <span class="hlt">gas</span>-particle partitioning and <span class="hlt">air-water</span>/soil <span class="hlt">exchange</span>. The following results were obtained: (1) The average concentrations (toxic equivalents, TEQs) of 2,3,7,8-PCDD/PCDF in <span class="hlt">air</span>, <span class="hlt">water</span>, sediment, and soil samples were 4,855 fg/m(3), 9.5 pg/L, 99.2 pg/g dry weight (dw), and 56.4 pg/g (203 fg TEQ/m(3), 0.46 pg TEQ/L, 2.2 pg TEQ/g dw, and 1.3 pg TEQ/g, respectively), respectively. (2) Although OCDF, 1,2,3,4,6,7,8-HpCDF, OCDD, and 1,2,3,4,6,7,8-HpCDD were the dominant congeners among four environmental sinks, obvious discrepancies of these congener and homologue patterns of PCDD/PCDF were observed still. (3) Significant linear correlations for PCDD/PCDF were observed between the <span class="hlt">gas</span>-particle partition coefficient (K p) and the subcooled liquid vapor pressure (P L (0)) and octanol-<span class="hlt">air</span> partition coefficient (K oa). (4) Fugacity fraction values of <span class="hlt">air-water</span> <span class="hlt">exchange</span> indicated that most of PCDD/PCDF homologues were dominated by net volatilization from <span class="hlt">water</span> into <span class="hlt">air</span>. The low-chlorinated PCDD/PCDF (tetra- to hexa-) presented a strong net volatilization from the soil into <span class="hlt">air</span>, while high-chlorinated PCDD/PCDF (hepta- to octa-) were mainly close to equilibrium for <span class="hlt">air</span>-soil <span class="hlt">exchange</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28675854','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28675854"><span><span class="hlt">Air</span>-sea <span class="hlt">exchange</span> and <span class="hlt">gas</span>-particle partitioning of polycyclic aromatic hydrocarbons over the northwestern Pacific Ocean: Role of East Asian continental outflow.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Zilan; Lin, Tian; Li, Zhongxia; Jiang, Yuqing; Li, Yuanyuan; Yao, Xiaohong; Gao, Huiwang; Guo, Zhigang</p> <p>2017-11-01</p> <p>We measured 15 parent polycyclic aromatic hydrocarbons (PAHs) in atmosphere and <span class="hlt">water</span> during a research cruise from the East China Sea (ECS) to the northwestern Pacific Ocean (NWP) in the spring of 2015 to investigate the occurrence, <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span>, and <span class="hlt">gas</span>-particle partitioning of PAHs with a particular focus on the influence of East Asian continental outflow. The gaseous PAH composition and identification of sources were consistent with PAHs from the upwind area, indicating that the gaseous PAHs (three-to five-ring PAHs) were influenced by upwind land pollution. In addition, <span class="hlt">air</span>-sea <span class="hlt">exchange</span> fluxes of gaseous PAHs were estimated to be -54.2-107.4 ng m -2 d -1 , and was indicative of variations of land-based PAH inputs. The logarithmic <span class="hlt">gas</span>-particle partition coefficient (logK p ) of PAHs regressed linearly against the logarithmic subcooled liquid vapor pressure (logP L 0 ), with a slope of -0.25. This was significantly larger than the theoretical value (-1), implying disequilibrium between the gaseous and particulate PAHs over the NWP. The non-equilibrium of PAH <span class="hlt">gas</span>-particle partitioning was shielded from the volatilization of three-ring gaseous PAHs from seawater and lower soot concentrations in particular when the oceanic <span class="hlt">air</span> masses prevailed. Modeling PAH absorption into organic matter and adsorption onto soot carbon revealed that the status of PAH <span class="hlt">gas</span>-particle partitioning deviated more from the modeling K p for oceanic <span class="hlt">air</span> masses than those for continental <span class="hlt">air</span> masses, which coincided with higher volatilization of three-ring PAHs and confirmed the influence of <span class="hlt">air</span>-sea <span class="hlt">exchange</span>. Meanwhile, significant linear regressions between logK p and logK oa (logK sa ) for PAHs were observed for continental <span class="hlt">air</span> masses, suggesting the dominant effect of East Asian continental outflow on atmospheric PAHs over the NWP during the sampling campaign. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23957244','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23957244"><span>Differences in <span class="hlt">gas</span> <span class="hlt">exchange</span> contribute to habitat differentiation in Iberian columbines from contrasting light and <span class="hlt">water</span> environments.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jaime, R; Serichol, C; Alcántara, J M; Rey, P J</p> <p>2014-03-01</p> <p>During photosynthesis, respiration and transpiration, <span class="hlt">gas</span> <span class="hlt">exchange</span> occurs via the stomata and so plants face a trade-off between maximising photosynthesis while minimising transpiration (expressed as <span class="hlt">water</span> use efficiency, WUE). The ability to cope with this trade-off and regulate photosynthetic rate and stomatal conductance may be related to niche differentiation between closely related species. The present study explored this as a possible mechanism for habitat differentiation in Iberian columbines. The roles of irradiance and <span class="hlt">water</span> stress were assessed to determine niche differentiation among Iberian columbines via distinct <span class="hlt">gas</span> <span class="hlt">exchange</span> processes. Photosynthesis-irradiance curves (P-I curves) were obtained for four taxa, and common garden experiments were conducted to examine plant responses to <span class="hlt">water</span> and irradiance stress, by measuring instantaneous <span class="hlt">gas</span> <span class="hlt">exchange</span> and plant performance. <span class="hlt">Gas</span> <span class="hlt">exchange</span> was also measured in ten individuals using two to four field populations per taxon. The taxa had different P-I curves and <span class="hlt">gas</span> <span class="hlt">exchange</span> in the field. At the species level, <span class="hlt">water</span> stress and irradiance explained habitat differentiation. Within each species, a combination of irradiance and <span class="hlt">water</span> stress explained the between-subspecies habitat differentiation. Despite differences in stomatal conductance and CO2 assimilation, taxa did not have different WUE under field conditions, which suggests that the environment equally modifies photosynthesis and transpiration. The P-I curves, <span class="hlt">gas</span> <span class="hlt">exchange</span> in the field and plant responses to experimental <span class="hlt">water</span> and irradiance stresses support the hypothesis that habitat differentiation is associated with differences among taxa in tolerance to abiotic stress mediated by distinct <span class="hlt">gas</span> <span class="hlt">exchange</span> responses. © 2013 German Botanical Society and The Royal Botanical Society of the Netherlands.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=104827&keyword=applications+AND+thermodynamic&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=104827&keyword=applications+AND+thermodynamic&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>A CRITICAL ASSESSMENT OF ELEMENTAL MERCURY <span class="hlt">AIR/WATER</span> <span class="hlt">EXCHANGE</span> PARTNERS</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>Although evasion of elemental mercury from aquatic systems can significantly deplete net mercury accumulation resulting from atmospheric deposition, the current ability to model elemental mercury <span class="hlt">air/water</span> <span class="hlt">exchange</span> is limited by uncertainties in our understanding of all gaseous a...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOS.A34C2670V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOS.A34C2670V"><span>Setting an Upper Limit on <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Through Sea-Spray</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vlahos, P.; Monahan, E. C.; Andreas, E. L.</p> <p>2016-02-01</p> <p><span class="hlt">Air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> parameterization is critical to understanding both climate forcing and feedbacks and is key in biogeochemistry cycles. Models based on wind speed have provided empirical estimates of <span class="hlt">gas</span> <span class="hlt">exchange</span> that are useful though it is likely that at high wind speeds of over 10 m/s there are important <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters including bubbles and sea spray that have not been well constrained. Here we address the sea-spray component of <span class="hlt">gas</span> <span class="hlt">exchange</span> at these high wind speeds to set sn upper boundary condition for the <span class="hlt">gas</span> <span class="hlt">exchange</span> of the six model gases including; nobel gases helium, neon and argon, diatomic gases nitrogen and oxygen and finally, the more complex <span class="hlt">gas</span> carbon dioxide. Estimates are based on the spray generation function of Andreas and Monahan and the gases are tested under three scenarios including 100 percent saturation and complete droplet evaporation, 100 percent saturation and a more realistic scenario in which a fraction of droplets evaporate completely, a fraction evaporate to some degree and a fraction returns to the <span class="hlt">water</span> side without significant evaporation. Finally the latter scenario is applied to representative under saturated concentrations of the gases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=59799&keyword=Exchange+AND+gaseous&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=59799&keyword=Exchange+AND+gaseous&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>THE EFFECT OF SALINITY ON RATES OF ELEMENTAL MERCURY <span class="hlt">AIR/WATER</span> <span class="hlt">EXCHANGE</span></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>The U.S. EPA laboratory in Athens, Georgia i spursuing the goal of developing a model for describing toxicant vapor phase <span class="hlt">air/water</span> <span class="hlt">exchange</span> under all relevant environmental conditions. To date, the two-layer <span class="hlt">exchange</span> model (suitable for low wind speed conditions) has been modif...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25743409','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25743409"><span>Estimation of <span class="hlt">air-water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> coefficient in a shallow lagoon based on 222Rn mass balance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cockenpot, S; Claude, C; Radakovitch, O</p> <p>2015-05-01</p> <p>The radon-222 mass balance is now commonly used to quantify <span class="hlt">water</span> fluxes due to Submarine Groundwater Discharge (SGD) in coastal areas. One of the main loss terms of this mass balance, the radon evasion to the atmosphere, is based on empirical equations. This term is generally estimated using one among the many empirical equations describing the <span class="hlt">gas</span> transfer velocity as a function of wind speed that have been proposed in the literature. These equations were, however, mainly obtained from areas of deep <span class="hlt">water</span> and may be less appropriate for shallow areas. Here, we calculate the radon mass balance for a windy shallow coastal lagoon (mean depth of 6m and surface area of 1.55*10(8) m(2)) and use these data to estimate the radon loss to the atmosphere and the corresponding <span class="hlt">gas</span> transfer velocity. We present new equations, adapted to our shallow <span class="hlt">water</span> body, to express the <span class="hlt">gas</span> transfer velocity as a function of wind speed at 10 m height (wind range from 2 to 12.5 m/s). When compared with those from the literature, these equations fit particularly well with the one of Kremer et al. (2003). Finally, we emphasize that some <span class="hlt">gas</span> transfer <span class="hlt">exchange</span> may always occur, even for conditions without wind. Copyright © 2015 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_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_4 --> <div id="page_5" 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_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> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="81"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=546779','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=546779"><span><span class="hlt">Gas</span> <span class="hlt">Exchange</span> of Algae</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ammann, Elizabeth C. B.; Lynch, Victoria H.</p> <p>1966-01-01</p> <p>Changes in the oxygen partial pressure of <span class="hlt">air</span> over the range of 8 to 258 mm of Hg did not adversely affect the photosynthetic capacity of Chlorella pyrenoidosa. <span class="hlt">Gas</span> <span class="hlt">exchange</span> and growth measurements remained constant for 3-week periods and were similar to <span class="hlt">air</span> controls (oxygen pressure of 160 mm of Hg). Oxygen partial pressures of 532 and 745 mm of Hg had an adverse effect on algal metabolism. Carbon dioxide consumption was 24% lower in the <span class="hlt">gas</span> mixture containing oxygen at a pressure 532 mm of Hg than in the <span class="hlt">air</span> control, and the growth rate was slightly reduced. Oxygen at a partial pressure of 745 mm of Hg decreased the photosynthetic rate 39% and the growth rate 37% over the corresponding rates in <span class="hlt">air</span>. The lowered metabolic rates remained constant during 14 days of measurements, and the effect was reversible after this time. Substitution of helium or argon for the nitrogen in <span class="hlt">air</span> had no effect on oxygen production, carbon dioxide consumption, or growth rate for 3-week periods. All measurements were made at a total pressure of 760 mm of Hg, and all <span class="hlt">gas</span> mixtures were enriched with 2% carbon dioxide. Thus, the physiological functioning and reliability of a photosynthetic <span class="hlt">gas</span> <span class="hlt">exchanger</span> should not be adversely affected by: (i) oxygen partial pressures ranging from 8 to 258 mm of Hg; (ii) the use of pure oxygen at reduced total pressure (155 to 258 mm of Hg) unless pressure per se affects photosynthesis, or (iii) the inclusion of helium or argon in the <span class="hlt">gas</span> environment (up to a partial pressure of 595 mm of Hg). PMID:5927028</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11967744','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11967744"><span>[Phylogeny of <span class="hlt">gas</span> <span class="hlt">exchange</span> systems].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jürgens, K D; Gros, G</p> <p>2002-04-01</p> <p> found in the respiration via the skin, which is of significance in some amphibians, but is limited by the thickness of the skin that constitutes a substantial diffusion path for O2 and CO2. The thick skin, on the other hand, provides mechanical protection as well as flexibility for the animals' body and helps avoid massive <span class="hlt">water</span> loss via the body surface. The gills of fishes, in contrast, exhibit rather short diffusion distances, are located in a mechanically protected space, and the problem of <span class="hlt">water</span> loss does not exist. The flows of blood and <span class="hlt">water</span> occur in opposite direction (countercurrent flow) and this situation makes an arterial PO2 approaching the environmental PO2 possible. A major disadvantage is constituted by the environmental medium since <span class="hlt">water</span> contains little O2 compared to <span class="hlt">air</span> and, to compensate this, much energy is expended to maintain a high flow rate of <span class="hlt">water</span> through the gills. In the mammalian lung ("pool system"), the presence of a dead space and the rhythmic ventilation that replaces only a small fraction of the <span class="hlt">gas</span> volume of the lung per breath, are responsible for an arterial PO2 (2/3 of the atmospheric PO2) that cannot reach the expiratory PO2. However, an advantage of this feature is the constantly high alveolar and arterial PCO2, which provides a highly effective H(+) buffer system in the entire body. The apparent disadvantage of the mammalian lung is avoided by the avian lung, which uses an extended system of airways to establish continuous equilibration of a part of the capillary blood with fresh <span class="hlt">air</span> (cross current system), during inspiration as well as during expiration. In this system, arterial PO2 can significantly exceed expiratory PO2. A disadvantage here is the enormous amount of space taken up by the avian lung, in animals of 1 kg body weight three times as much as taken up by the mammalian lung. All respiratory <span class="hlt">exchange</span> systems considered here exhibit high degrees of optimization - yet follow highly diverse construction principles</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28416704','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28416704"><span>Diurnal Variation in <span class="hlt">Gas</span> <span class="hlt">Exchange</span>: The Balance between Carbon Fixation and <span class="hlt">Water</span> Loss.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matthews, Jack S A; Vialet-Chabrand, Silvere R M; Lawson, Tracy</p> <p>2017-06-01</p> <p>Stomatal control of transpiration is critical for maintaining important processes, such as plant <span class="hlt">water</span> status, leaf temperature, as well as permitting sufficient CO 2 diffusion into the leaf to maintain photosynthetic rates ( A ). Stomatal conductance often closely correlates with A and is thought to control the balance between <span class="hlt">water</span> loss and carbon gain. It has been suggested that a mesophyll-driven signal coordinates A and stomatal conductance responses to maintain this relationship; however, the signal has yet to be fully elucidated. Despite this correlation under stable environmental conditions, the responses of both parameters vary spatially and temporally and are dependent on species, environment, and plant <span class="hlt">water</span> status. Most current models neglect these aspects of <span class="hlt">gas</span> <span class="hlt">exchange</span>, although it is clear that they play a vital role in the balance of carbon fixation and <span class="hlt">water</span> loss. Future efforts should consider the dynamic nature of whole-plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and how it represents much more than the sum of its individual leaf-level components, and they should take into consideration the long-term effect on <span class="hlt">gas</span> <span class="hlt">exchange</span> over time. © 2017 American Society of Plant Biologists. All Rights Reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850005886&hterms=heat+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dheat%2Bexchange','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850005886&hterms=heat+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dheat%2Bexchange"><span><span class="hlt">Air</span>-sea heat <span class="hlt">exchange</span>, an element of the <span class="hlt">water</span> cycle</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chahine, M. T.</p> <p>1984-01-01</p> <p>The distribution and variation of <span class="hlt">water</span> vapor, clouds and precipitation are examined. Principal driving forces for these distributions are energy <span class="hlt">exchange</span> and evaporation at the <span class="hlt">air</span>-sea interface, which are also important elements of <span class="hlt">air</span>-sea interaction studies. The overall aim of <span class="hlt">air</span>-sea interaction studies is to quantitatively determine mass, momentum and energy fluxes, with the goal of understanding the mechanisms controlling them. The results of general circulation simulations indicate that the atmosphere in mid-latitudes responds to changes in the oceanic surface conditions in the tropics. This correlation reflects the strong interaction between tropical and mid-latitude conditions caused by the transport of heat and momentum from the tropics. Studies of <span class="hlt">air</span>-sea <span class="hlt">exchanges</span> involve a large number of physica, chemical and dynamical processes including heat flux, radiation, sea-surface temperature, precipitation, winds and ocean currents. The fluxes of latent heat are studied and the potential use of satellite data in determining them evaluated. Alternative ways of inferring heat fluxes will be considered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=180684&keyword=global+AND+water+AND+issues&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=180684&keyword=global+AND+water+AND+issues&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>ISSUES IN SIMULATING ELEMENTAL MERCURY <span class="hlt">AIR/WATER</span> <span class="hlt">EXCHANGE</span> AND AQUEOUS MONOMETHYLMERCURY SPECIATION</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>This presentation focuses on two areas relevant to assessing the global fate and bioavailability of mercury: elemental mercury <span class="hlt">air/water</span> <span class="hlt">exchange</span> and aqueous environmental monomethylmercury speciation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18640753','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18640753"><span>Dry deposition and soil-<span class="hlt">air</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> of polychlorinated biphenyls (PCBs) in an industrial area.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bozlaker, Ayse; Odabasi, Mustafa; Muezzinoglu, Aysen</p> <p>2008-12-01</p> <p>Ambient <span class="hlt">air</span> and dry deposition, and soil samples were collected at the Aliaga industrial site in Izmir, Turkey. Atmospheric total (particle+<span class="hlt">gas</span>) Sigma(41)-PCB concentrations were higher in summer (3370+/-1617 pg m(-3), average+SD) than in winter (1164+/-618 pg m(-3)), probably due to increased volatilization with temperature. Average particulate Sigma(41)-PCBs dry deposition fluxes were 349+/-183 and 469+/-328 ng m(-2) day(-1) in summer and winter, respectively. Overall average particulate deposition velocity was 5.5+/-3.5 cm s(-1). The spatial distribution of Sigma(41)-PCB soil concentrations (n=48) showed that the iron-steel plants, ship dismantling facilities, refinery and petrochemicals complex are the major sources in the area. Calculated <span class="hlt">air</span>-soil <span class="hlt">exchange</span> fluxes indicated that the contaminated soil is a secondary source to the atmosphere for lighter PCBs and as a sink for heavier ones. Comparable magnitude of <span class="hlt">gas</span> <span class="hlt">exchange</span> and dry particle deposition fluxes indicated that both mechanisms are equally important for PCB movement between <span class="hlt">air</span> and soil in Aliaga.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4510840','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4510840"><span>Oxygen-limited thermal tolerance is seen in a plastron-breathing insect and can be induced in a bimodal <span class="hlt">gas</span> <span class="hlt">exchanger</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>Verberk, Wilco C. E. P.; Bilton, David T.</p> <p>2015-01-01</p> <p>ABSTRACT Thermal tolerance has been hypothesized to result from a mismatch between oxygen supply and demand. However, the generality of this hypothesis has been challenged by studies on various animal groups, including <span class="hlt">air</span>-breathing adult insects. Recently, comparisons across taxa have suggested that differences in <span class="hlt">gas</span> <span class="hlt">exchange</span> mechanisms could reconcile the discrepancies found in previous studies. Here, we test this suggestion by comparing the behaviour of related insect taxa with different <span class="hlt">gas</span> <span class="hlt">exchange</span> mechanisms, with and without access to <span class="hlt">air</span>. We demonstrate oxygen-limited thermal tolerance in <span class="hlt">air</span>-breathing adults of the plastron-<span class="hlt">exchanging</span> <span class="hlt">water</span> bug Aphelocheirus aestivalis. Ilyocoris cimicoides, a related, bimodal <span class="hlt">gas</span> <span class="hlt">exchanger</span>, did not exhibit such oxygen-limited thermal tolerance and relied increasingly on aerial <span class="hlt">gas</span> <span class="hlt">exchange</span> with warming. Intriguingly, however, when denied access to <span class="hlt">air</span>, oxygen-limited thermal tolerance could also be induced in this species. Patterns in oxygen-limited thermal tolerance were found to be consistent across life-history stages in these insects, with nymphs employing the same <span class="hlt">gas</span> <span class="hlt">exchange</span> mechanisms as adults. These results advance our understanding of oxygen limitation at high temperatures; differences in the degree of respiratory control appear to modulate the importance of oxygen in setting tolerance limits. PMID:25964420</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/11366','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/11366"><span>A Controlled Environment System For Measuring Plant-Atmosphere <span class="hlt">Gas</span> <span class="hlt">Exchange</span></span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>James M. Brown</p> <p>1975-01-01</p> <p>Describes an inexpensive, efficient system for measuring plant-atmosphere <span class="hlt">gas</span> <span class="hlt">exchange</span>. Designed to measure transpiration from potted tree seedlings, it is readily adaptable for measuring other <span class="hlt">gas</span> <span class="hlt">exchanges</span> or <span class="hlt">gas</span> <span class="hlt">exchange</span> by plant parts. Light level, <span class="hlt">air</span> and root temperature can be precisely controlled at minimum cost.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830046452&hterms=water+gas+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dwater%2Bgas%2Bexchange','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830046452&hterms=water+gas+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dwater%2Bgas%2Bexchange"><span>Methane flux across the <span class="hlt">air-water</span> interface - <span class="hlt">Air</span> velocity effects</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sebacher, D. I.; Harriss, R. C.; Bartlett, K. B.</p> <p>1983-01-01</p> <p>Methane loss to the atmosphere from flooded wetlands is influenced by the degree of supersaturation and wind stress at the <span class="hlt">water</span> surface. Measurements in freshwater ponds in the St. Marks Wildlife Refuge, Florida, demonstrated that for the combined variability of CH4 concentrations in surface <span class="hlt">water</span> and <span class="hlt">air</span> velocity over the <span class="hlt">water</span> surface, CH4 flux varied from 0.01 to 1.22 g/sq m/day. The liquid <span class="hlt">exchange</span> coefficient for a two-layer model of the <span class="hlt">gas</span>-liquid interface was calculated as 1.7 cm/h for CH4 at <span class="hlt">air</span> velocity of zero and as 1.1 + 1.2 v to the 1.96th power cm/h for <span class="hlt">air</span> velocities from 1.4 to 3.5 m/s and <span class="hlt">water</span> temperatures of 20 C.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21822726','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21822726"><span>Seasonal photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span>-use efficiency in a constitutive CAM plant, the giant saguaro cactus (Carnegiea gigantea).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bronson, Dustin R; English, Nathan B; Dettman, David L; Williams, David G</p> <p>2011-11-01</p> <p>Crassulacean acid metabolism (CAM) and the capacity to store large quantities of <span class="hlt">water</span> are thought to confer high <span class="hlt">water</span> use efficiency (WUE) and survival of succulent plants in warm desert environments. Yet the highly variable precipitation, temperature and humidity conditions in these environments likely have unique impacts on underlying processes regulating photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> and WUE, limiting our ability to predict growth and survival responses of desert CAM plants to climate change. We monitored net CO(2) assimilation (A(net)), stomatal conductance (g(s)), and transpiration (E) rates periodically over 2 years in a natural population of the giant columnar cactus Carnegiea gigantea (saguaro) near Tucson, Arizona USA to investigate environmental and physiological controls over carbon gain and <span class="hlt">water</span> loss in this ecologically important plant. We hypothesized that seasonal changes in daily integrated <span class="hlt">water</span> use efficiency (WUE(day)) in this constitutive CAM species would be driven largely by stomatal regulation of nighttime transpiration and CO(2) uptake responding to shifts in nighttime <span class="hlt">air</span> temperature and humidity. The lowest WUE(day) occurred during time periods with extreme high and low <span class="hlt">air</span> vapor pressure deficit (D(a)). The diurnal with the highest D(a) had low WUE(day) due to minimal net carbon gain across the 24 h period. Low WUE(day) was also observed under conditions of low D(a); however, it was due to significant transpiration losses. <span class="hlt">Gas</span> <span class="hlt">exchange</span> measurements on potted saguaro plants exposed to experimental changes in D(a) confirmed the relationship between D(a) and g(s). Our results suggest that climatic changes involving shifts in <span class="hlt">air</span> temperature and humidity will have large impacts on the <span class="hlt">water</span> and carbon economy of the giant saguaro and potentially other succulent CAM plants of warm desert environments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170009534','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170009534"><span><span class="hlt">Gas</span> Turbine Engine with <span class="hlt">Air</span>/Fuel Heat <span class="hlt">Exchanger</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Krautheim, Michael Stephen (Inventor); Chouinard, Donald G. (Inventor); Donovan, Eric Sean (Inventor); Karam, Michael Abraham (Inventor); Vetters, Daniel Kent (Inventor)</p> <p>2017-01-01</p> <p>One embodiment of the present invention is a unique aircraft propulsion <span class="hlt">gas</span> turbine engine. Another embodiment is a unique <span class="hlt">gas</span> turbine engine. Another embodiment is a unique <span class="hlt">gas</span> turbine engine. Other embodiments include apparatuses, systems, devices, hardware, methods, and combinations for <span class="hlt">gas</span> turbine engines with heat <span class="hlt">exchange</span> systems. Further embodiments, forms, features, aspects, benefits, and advantages of the present application will become apparent from the description and figures provided herewith.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25964420','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25964420"><span>Oxygen-limited thermal tolerance is seen in a plastron-breathing insect and can be induced in a bimodal <span class="hlt">gas</span> <span class="hlt">exchanger</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Verberk, Wilco C E P; Bilton, David T</p> <p>2015-07-01</p> <p>Thermal tolerance has been hypothesized to result from a mismatch between oxygen supply and demand. However, the generality of this hypothesis has been challenged by studies on various animal groups, including <span class="hlt">air</span>-breathing adult insects. Recently, comparisons across taxa have suggested that differences in <span class="hlt">gas</span> <span class="hlt">exchange</span> mechanisms could reconcile the discrepancies found in previous studies. Here, we test this suggestion by comparing the behaviour of related insect taxa with different <span class="hlt">gas</span> <span class="hlt">exchange</span> mechanisms, with and without access to <span class="hlt">air</span>. We demonstrate oxygen-limited thermal tolerance in <span class="hlt">air</span>-breathing adults of the plastron-<span class="hlt">exchanging</span> <span class="hlt">water</span> bug Aphelocheirus aestivalis. Ilyocoris cimicoides, a related, bimodal <span class="hlt">gas</span> <span class="hlt">exchanger</span>, did not exhibit such oxygen-limited thermal tolerance and relied increasingly on aerial <span class="hlt">gas</span> <span class="hlt">exchange</span> with warming. Intriguingly, however, when denied access to <span class="hlt">air</span>, oxygen-limited thermal tolerance could also be induced in this species. Patterns in oxygen-limited thermal tolerance were found to be consistent across life-history stages in these insects, with nymphs employing the same <span class="hlt">gas</span> <span class="hlt">exchange</span> mechanisms as adults. These results advance our understanding of oxygen limitation at high temperatures; differences in the degree of respiratory control appear to modulate the importance of oxygen in setting tolerance limits. © 2015. Published by The Company of Biologists Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18805813','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18805813"><span>Investigating onychophoran <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> balance as a means to inform current controversies in arthropod physiology.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Clusella-Trullas, Susana; Chown, Steven L</p> <p>2008-10-01</p> <p>Several controversies currently dominate the fields of arthropod metabolic rate, <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> balance, including the extent to which modulation of <span class="hlt">gas</span> <span class="hlt">exchange</span> reduces <span class="hlt">water</span> loss, the origins of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span>, the relationship between metabolic rate and life-history strategies, and the causes of Palaeozoic gigantism. In all of these areas, repeated calls have been made for the investigation of groups that might most inform the debates, especially of taxa in key phylogenetic positions. Here we respond to this call by investigating metabolic rate, respiratory <span class="hlt">water</span> loss and critical oxygen partial pressure (Pc) in the onychophoran Peripatopsis capensis, a member of a group basal to the arthropods, and by synthesizing the available data on the Onychophora. The rate of carbon dioxide release (VCO2) at 20 degrees C in P. capensis is 0.043 ml CO2 h(-1), in keeping with other onychophoran species; suggesting that low metabolic rates in some arthropod groups are derived. Continuous <span class="hlt">gas</span> <span class="hlt">exchange</span> suggests that more complex <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns are also derived. Total <span class="hlt">water</span> loss in P. capensis is 57 mg H2O h(-1) at 20 degrees C, similar to modern estimates for another onychophoran species. High relative respiratory <span class="hlt">water</span> loss rates ( approximately 34%; estimated using a regression technique) suggest that the basal condition in arthropods may be a high respiratory <span class="hlt">water</span> loss rate. Relatively high Pc values (5-10% O2) suggest that substantial safety margins in insects are also a derived condition. Curling behaviour in P. capensis appears to be a strategy to lower energetic costs when resting, and the concomitant depression of <span class="hlt">water</span> loss is a proximate consequence of this behaviour.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20516484','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20516484"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> and hydraulics in seedlings of Hevea brasiliensis during <span class="hlt">water</span> stress and recovery.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chen, Jun-Wen; Zhang, Qiang; Li, Xiao-Shuang; Cao, Kun-Fang</p> <p>2010-07-01</p> <p>The response of plants to drought has received significant attention, but far less attention has been given to the dynamic response of plants during recovery from drought. Photosynthetic performance and hydraulic capacity were monitored in seedlings of Hevea brasiliensis under <span class="hlt">water</span> stress and during recovery following rewatering. Leaf <span class="hlt">water</span> relation, <span class="hlt">gas</span> <span class="hlt">exchange</span> rate and hydraulic conductivity decreased gradually after <span class="hlt">water</span> stress fell below a threshold, whereas instantaneous <span class="hlt">water</span> use efficiency and osmolytes increased significantly. After 5 days of rewatering, leaf <span class="hlt">water</span> relation, maximum stomatal conductance (g(s-max)) and plant hydraulic conductivity had recovered to the control levels except for sapwood area-specific hydraulic conductivity, photosynthetic assimilation rate and osmolytes. During the phase of <span class="hlt">water</span> stress, stomata were almost completely closed before <span class="hlt">water</span> transport efficiency decreased substantially, and moreover, the leaf hydraulic pathway was more vulnerable to <span class="hlt">water</span> stress-induced embolism than the stem hydraulic pathway. Meanwhile, g(s-max) was linearly correlated with hydraulic capacity when <span class="hlt">water</span> stress exceeded a threshold. In addition, a positive relationship was shown to occur between the recovery of g(s-max) and of hydraulic capacity during the phase of rewatering. Our results suggest (i) that stomatal closure effectively reduces the risk of xylem dysfunction in <span class="hlt">water</span>-stressed plants at the cost of <span class="hlt">gas</span> <span class="hlt">exchange</span>, (ii) that the leaf functions as a safety valve to protect the hydraulic pathway from <span class="hlt">water</span> stress-induced dysfunction to a larger extent than does the stem and (iii) that the full drought recovery of <span class="hlt">gas</span> <span class="hlt">exchange</span> is restricted by not only hydraulic factors but also non-hydraulic factors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28426140','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28426140"><span>Stomatal kinetics and photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> along a continuum of isohydric to anisohydric regulation of plant <span class="hlt">water</span> status.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Meinzer, Frederick C; Smith, Duncan D; Woodruff, David R; Marias, Danielle E; McCulloh, Katherine A; Howard, Ava R; Magedman, Alicia L</p> <p>2017-08-01</p> <p>Species' differences in the stringency of stomatal control of plant <span class="hlt">water</span> potential represent a continuum of isohydric to anisohydric behaviours. However, little is known about how quasi-steady-state stomatal regulation of <span class="hlt">water</span> potential may relate to dynamic behaviour of stomata and photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> in species operating at different positions along this continuum. Here, we evaluated kinetics of light-induced stomatal opening, activation of photosynthesis and features of quasi-steady-state photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> in 10 woody species selected to represent different degrees of anisohydry. Based on a previously developed proxy for the degree of anisohydry, species' leaf <span class="hlt">water</span> potentials at turgor loss, we found consistent trends in photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> traits across a spectrum of isohydry to anisohydry. More anisohydric species had faster kinetics of stomatal opening and activation of photosynthesis, and these kinetics were closely coordinated within species. Quasi-steady-state stomatal conductance and measures of photosynthetic capacity and performance were also greater in more anisohydric species. Intrinsic <span class="hlt">water</span>-use efficiency estimated from leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and stable carbon isotope ratios was lowest in the most anisohydric species. In comparisons between <span class="hlt">gas</span> <span class="hlt">exchange</span> traits, species rankings were highly consistent, leading to species-independent scaling relationships over the range of isohydry to anisohydry observed. © 2017 John Wiley & Sons Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5388528-conversion-deuterium-gas-heavy-water-catalytic-isotopic-exchange-using-wetproof-catalyst','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5388528-conversion-deuterium-gas-heavy-water-catalytic-isotopic-exchange-using-wetproof-catalyst"><span>Conversion of deuterium <span class="hlt">gas</span> to heavy <span class="hlt">water</span> by catalytic isotopic <span class="hlt">exchange</span> using wetproof catalyst</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>Quaiattini, R.J.; McGauley, M.P.; Burns, D.L.</p> <p></p> <p>The invention at Chalk River Nuclear Laboratories of a simple method of wetproofing platinum catalysts allows them to retain their activity in liquid <span class="hlt">water</span>. High performance catalysts for the hydrogen-<span class="hlt">water</span> isotope <span class="hlt">exchange</span> reaction that remain active for years can now be routinely produced. The first commercial application using the ordered-bed-type wetproofed isotope <span class="hlt">exchange</span> catalyst developed and patented by Atomic Energy of Canada Ltd. has been successfully completed. Approximately 9100 m/sup 3/ of deuterium <span class="hlt">gas</span> stored at Brookhaven National Laboratory was converted to high grade heavy <span class="hlt">water</span>. Conversion efficiency exceeded 99.8%. The product D/sub 2/O concentration was 6.7 percentage points highermore » than the feed D/sub 2/ <span class="hlt">gas</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013BGD....10.8415S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013BGD....10.8415S"><span>Biology and <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> controls on the distribution of carbon isotope ratios (δ13C) in the ocean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmittner, A.; Gruber, N.; Mix, A. C.; Key, R. M.; Tagliabue, A.; Westberry, T. K.</p> <p>2013-05-01</p> <p>Analysis of observations and sensitivity experiments with a new three-dimensional global model of stable carbon isotope cycling elucidate the processes that control the distribution of δ13C in the contemporary and preindustrial ocean. Biological fractionation dominates the distribution of δ13CDIC of dissolved inorganic carbon (DIC) due to the sinking of isotopically light δ13C organic matter from the surface into the interior ocean. This process leads to low δ13CDIC values at dephs and in high latitude surface <span class="hlt">waters</span> and high values in the upper ocean at low latitudes with maxima in the subtropics. <span class="hlt">Air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> provides an important secondary influence due to two effects. First, it acts to reduce the spatial gradients created by biology. Second, the associated temperature dependent fractionation tends to increase (decrease) δ13CDIC values of colder (warmer) <span class="hlt">water</span>, which generates gradients that oppose those arising from biology. Our model results suggest that both effects are similarly important in influencing surface and interior δ13CDIC distributions. However, <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> is slow, so biological effect dominate spatial δ13CDIC gradients both in the interior and at the surface, in constrast to conclusions from some previous studies. Analysis of a new synthesis of δ13CDIC measurements from years 1990 to 2005 is used to quantify preformed (δ13Cpre) and remineralized (δ13Crem) contributions as well as the effects of biology (Δδ13Cbio) and <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> (δ13C*). The model reproduces major features of the observed large-scale distribution of δ13CDIC, δ13Cpre, δ13Crem, δ13C*, and Δδ13Cbio. Residual misfits are documented and analyzed. Simulated surface and subsurface δ13CDIC are influenced by details of the ecosystem model formulation. For example, inclusion of a simple parameterization of iron limitation of phytoplankton growth rates and temperature-dependent zooplankton grazing rates improves the agreement with δ13CDIC</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C31D..01L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C31D..01L"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> in the ice zone: the role of small waves and big animals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Loose, B.; Takahashi, A.; Bigdeli, A.</p> <p>2016-12-01</p> <p>The balance of <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> and net biological carbon fixation determine the transport and transformation of carbon dioxide and methane in the ocean. <span class="hlt">Air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> is mostly driven by upper ocean physics, but biology can also play a role. In the open ocean, <span class="hlt">gas</span> <span class="hlt">exchange</span> increases proportionate to the square of wind speed. When sea ice is present, this dependence breaks down in part because breaking waves and <span class="hlt">air</span> bubble entrainment are damped out by interactions between sea ice and the wave field. At the same time, sea ice motions, formation, melt, and even sea ice-associated organisms can act to introduce turbulence and <span class="hlt">air</span> bubbles into the upper ocean, thereby enhancing <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span>. We take advantage of the knowledge advances of upper ocean physics including bubble dynamics to formulate a model for <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> in the sea ice zone. Here, we use the model to examine the role of small-scale waves and diving animals that trap <span class="hlt">air</span> for insulation, including penguins, seals and polar bears. We compare these processes to existing parameterizations of wave and bubble dynamics in the open ocean, to observe how sea ice both mitigates and locally enhances <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3976162','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3976162"><span>Rapid and long-term effects of <span class="hlt">water</span> deficit on <span class="hlt">gas</span> <span class="hlt">exchange</span> and hydraulic conductance of silver birch trees grown under varying atmospheric humidity</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2014-01-01</p> <p>Background Effects of <span class="hlt">water</span> deficit on plant <span class="hlt">water</span> status, <span class="hlt">gas</span> <span class="hlt">exchange</span> and hydraulic conductance were investigated in Betula pendula under artificially manipulated <span class="hlt">air</span> humidity in Eastern Estonia. The study was aimed to broaden an understanding of the ability of trees to acclimate with the increasing atmospheric humidity predicted for northern Europe. Rapidly-induced <span class="hlt">water</span> deficit was imposed by dehydrating cut branches in open-<span class="hlt">air</span> conditions; long-term <span class="hlt">water</span> deficit was generated by seasonal drought. Results The rapid <span class="hlt">water</span> deficit quantified by leaf (ΨL) and branch <span class="hlt">water</span> potentials (ΨB) had a significant (P < 0.001) effect on <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters, while inclusion of ΨB in models resulted in a considerably better fit than those including ΨL, which supports the idea that stomatal openness is regulated to prevent stem rather than leaf xylem dysfunction. Under moderate <span class="hlt">water</span> deficit (ΨL≥-1.55 MPa), leaf conductance to <span class="hlt">water</span> vapour (gL), transpiration rate and leaf hydraulic conductance (KL) were higher (P < 0.05) and leaf temperature lower in trees grown in elevated <span class="hlt">air</span> humidity (H treatment) than in control trees (C treatment). Under severe <span class="hlt">water</span> deficit (ΨL<-1.55 MPa), the treatments showed no difference. The humidification manipulation influenced most of the studied characteristics, while the effect was to a great extent realized through changes in soil <span class="hlt">water</span> availability, i.e. due to higher soil <span class="hlt">water</span> potential in H treatment. Two functional characteristics (gL, KL) exhibited higher (P < 0.05) sensitivity to <span class="hlt">water</span> deficit in trees grown under increased <span class="hlt">air</span> humidity. Conclusions The experiment supported the hypothesis that physiological traits in trees acclimated to higher <span class="hlt">air</span> humidity exhibit higher sensitivity to rapid <span class="hlt">water</span> deficit with respect to two characteristics - leaf conductance to <span class="hlt">water</span> vapour and leaf hydraulic conductance. Disproportionate changes in sensitivity of stomatal versus leaf hydraulic conductance to <span class="hlt">water</span> deficit</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24655599','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24655599"><span>Rapid and long-term effects of <span class="hlt">water</span> deficit on <span class="hlt">gas</span> <span class="hlt">exchange</span> and hydraulic conductance of silver birch trees grown under varying atmospheric humidity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sellin, Arne; Niglas, Aigar; Õunapuu-Pikas, Eele; Kupper, Priit</p> <p>2014-03-24</p> <p>Effects of <span class="hlt">water</span> deficit on plant <span class="hlt">water</span> status, <span class="hlt">gas</span> <span class="hlt">exchange</span> and hydraulic conductance were investigated in Betula pendula under artificially manipulated <span class="hlt">air</span> humidity in Eastern Estonia. The study was aimed to broaden an understanding of the ability of trees to acclimate with the increasing atmospheric humidity predicted for northern Europe. Rapidly-induced <span class="hlt">water</span> deficit was imposed by dehydrating cut branches in open-<span class="hlt">air</span> conditions; long-term <span class="hlt">water</span> deficit was generated by seasonal drought. The rapid <span class="hlt">water</span> deficit quantified by leaf (ΨL) and branch <span class="hlt">water</span> potentials (ΨB) had a significant (P < 0.001) effect on <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters, while inclusion of ΨB in models resulted in a considerably better fit than those including ΨL, which supports the idea that stomatal openness is regulated to prevent stem rather than leaf xylem dysfunction. Under moderate <span class="hlt">water</span> deficit (ΨL≥-1.55 MPa), leaf conductance to <span class="hlt">water</span> vapour (gL), transpiration rate and leaf hydraulic conductance (KL) were higher (P < 0.05) and leaf temperature lower in trees grown in elevated <span class="hlt">air</span> humidity (H treatment) than in control trees (C treatment). Under severe <span class="hlt">water</span> deficit (ΨL<-1.55 MPa), the treatments showed no difference. The humidification manipulation influenced most of the studied characteristics, while the effect was to a great extent realized through changes in soil <span class="hlt">water</span> availability, i.e. due to higher soil <span class="hlt">water</span> potential in H treatment. Two functional characteristics (gL, KL) exhibited higher (P < 0.05) sensitivity to <span class="hlt">water</span> deficit in trees grown under increased <span class="hlt">air</span> humidity. The experiment supported the hypothesis that physiological traits in trees acclimated to higher <span class="hlt">air</span> humidity exhibit higher sensitivity to rapid <span class="hlt">water</span> deficit with respect to two characteristics - leaf conductance to <span class="hlt">water</span> vapour and leaf hydraulic conductance. Disproportionate changes in sensitivity of stomatal versus leaf hydraulic conductance to <span class="hlt">water</span> deficit will impose greater risk of desiccation</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/2017AIPC.1879b0006C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1879b0006C"><span>Devise of an exhaust <span class="hlt">gas</span> heat <span class="hlt">exchanger</span> for a thermal oil heater in a palm oil refinery plant</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chucherd, Panom; Kittisupakorn, Paisan</p> <p>2017-08-01</p> <p>This paper presents the devise of an exhaust <span class="hlt">gas</span> heat <span class="hlt">exchanger</span> for waste heat recovery of the exhausted flue <span class="hlt">gas</span> of palm oil refinery plant. This waste heat can be recovered by installing an economizer to heat the feed <span class="hlt">water</span> which can save the fuel consumption of the coal fired steam boiler and the outlet temperature of flue <span class="hlt">gas</span> will be controlled in order to avoid the acid dew point temperature and protect the filter bag. The decrease of energy used leads to the reduction of CO2 emission. Two designed economizer studied in this paper are <span class="hlt">gas</span> in tube and <span class="hlt">water</span> in tube. The <span class="hlt">gas</span> in tube <span class="hlt">exchanger</span> refers to the shell and tube heat <span class="hlt">exchanger</span> which the flue <span class="hlt">gas</span> flows in tube; this designed <span class="hlt">exchanger</span> is used in the existing unit. The new designed <span class="hlt">water</span> in tube refers to the shell and tube heat <span class="hlt">exchanger</span> which the <span class="hlt">water</span> flows in the tube; this designed <span class="hlt">exchanger</span> is proposed for new implementation. New economizer has the overall coefficient of heat transfer of 19.03 W/m2.K and the surface heat transfer area of 122 m2 in the optimized case. Experimental results show that it is feasible to install economizer in the exhaust flue <span class="hlt">gas</span> system between the <span class="hlt">air</span> preheater and the bag filter, which has slightly disadvantage effect in the system. The system can raise the feed <span class="hlt">water</span> temperature from 40 to 104°C and flow rate 3.31 m3/h, the outlet temperature of flue <span class="hlt">gas</span> is maintained about 130 °C.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016E%26ES...35a2003A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016E%26ES...35a2003A"><span>The potential role of sea spray droplets in facilitating <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Andreas, E. L.; Vlahos, P.; Monahan, E. C.</p> <p>2016-05-01</p> <p>For over 30 years, <span class="hlt">air</span>-sea interaction specialists have been evaluating and parameterizing the role of whitecap bubbles in <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span>. To our knowledge, no one, however, has studied the mirror image process of whether sea spray droplets can facilitate <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span>. We are therefore using theory, data analysis, and numerical modeling to quantify the role of spray on <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer. In this, our first formal work on this subject, we seek the rate-limiting step in spray-mediated <span class="hlt">gas</span> transfer by evaluating the three time scales that govern the <span class="hlt">exchange</span>: τ <span class="hlt">air</span> , which quantifies the rate of transfer between the atmospheric <span class="hlt">gas</span> reservoir and the surface of the droplet; τ int , which quantifies the <span class="hlt">exchange</span> rate across the <span class="hlt">air</span>-droplet interface; and τ aq , which quantifies <span class="hlt">gas</span> mixing within the aqueous solution droplet.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=59576&keyword=film+AND+analysis&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=59576&keyword=film+AND+analysis&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>THE ROLE OF AQUEOUS THIN FILM EVAPORATIVE COOLING ON RATES OF ELEMENTAL MERCURY <span class="hlt">AIR-WATER</span> <span class="hlt">EXCHANGE</span> UNDER TEMPERATURE DISEQUILIBRIUM CONDITIONS</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>The technical conununity has only recently addressed the role of atmospheric temperature variations on rates of <span class="hlt">air-water</span> vapor phase toxicant <span class="hlt">exchange</span>. The technical literature has documented that: 1) day time rates of elemental mercury vapor phase <span class="hlt">air-water</span> <span class="hlt">exchange</span> can exceed ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21669328','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21669328"><span>Atmospheric concentrations and <span class="hlt">air</span>-soil <span class="hlt">gas</span> <span class="hlt">exchange</span> of polycyclic aromatic hydrocarbons (PAHs) in remote, rural village and urban areas of Beijing-Tianjin region, North China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Wentao; Simonich, Staci; Giri, Basant; Chang, Ying; Zhang, Yuguang; Jia, Yuling; Tao, Shu; Wang, Rong; Wang, Bin; Li, Wei; Cao, Jun; Lu, Xiaoxia</p> <p>2011-07-01</p> <p>Forty passive <span class="hlt">air</span> samplers were deployed to study the occurrence of <span class="hlt">gas</span> and particulate phase PAHs in remote, rural village and urban areas of Beijing-Tianjin region, North China for four seasons (spring, summer, fall and winter) from 2007 to 2008. The influence of emissions on the spatial distribution pattern of <span class="hlt">air</span> PAH concentrations was addressed. In addition, the <span class="hlt">air</span>-soil <span class="hlt">gas</span> <span class="hlt">exchange</span> of PAHs was studied using fugacity calculations. The median gaseous and particulate phase PAH concentrations were 222 ng/m³ and 114 ng/m³, respectively, with a median total PAH concentration of 349 ng/m³. Higher PAH concentrations were measured in winter than in other seasons. <span class="hlt">Air</span> PAH concentrations measured at the rural villages and urban sites in the northern mountain region were significantly lower than those measured at sites in the southern plain during all seasons. However, there was no significant difference in PAH concentrations between the rural villages and urban sites in the northern and southern areas. This urban-rural PAH distribution pattern was related to the location of PAH emission sources and the population distribution. The location of PAH emission sources explained 56%-77% of the spatial variation in ambient <span class="hlt">air</span> PAH concentrations. The annual median <span class="hlt">air</span>-soil <span class="hlt">gas</span> <span class="hlt">exchange</span> flux of PAHs was 42.2 ng/m²/day from soil to <span class="hlt">air</span>. Among the 15 PAHs measured, acenaphthylene (ACY) and acenaphthene (ACE) contributed to more than half of the total <span class="hlt">exchange</span> flux. Furthermore, the <span class="hlt">air</span>-soil <span class="hlt">gas</span> <span class="hlt">exchange</span> fluxes of PAHs at the urban sites were higher than those at the remote and rural sites. In summer, more gaseous PAHs volatilized from soil to <span class="hlt">air</span> because of higher temperatures and increased rainfall. However, in winter, more gaseous PAHs deposited from <span class="hlt">air</span> to soil due to higher PAH emissions and lower temperatures. The soil TOC concentration had no significant influence on the <span class="hlt">air</span>-soil <span class="hlt">gas</span> <span class="hlt">exchange</span> of PAHs. Copyright © 2011 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25399878','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25399878"><span>Estimation of bias with the single-zone assumption in measurement of residential <span class="hlt">air</span> <span class="hlt">exchange</span> using the perfluorocarbon tracer <span class="hlt">gas</span> method.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Van Ryswyk, K; Wallace, L; Fugler, D; MacNeill, M; Héroux, M È; Gibson, M D; Guernsey, J R; Kindzierski, W; Wheeler, A J</p> <p>2015-12-01</p> <p>Residential <span class="hlt">air</span> <span class="hlt">exchange</span> rates (AERs) are vital in understanding the temporal and spatial drivers of indoor <span class="hlt">air</span> quality (IAQ). Several methods to quantify AERs have been used in IAQ research, often with the assumption that the home is a single, well-mixed <span class="hlt">air</span> zone. Since 2005, Health Canada has conducted IAQ studies across Canada in which AERs were measured using the perfluorocarbon tracer (PFT) <span class="hlt">gas</span> method. Emitters and detectors of a single PFT <span class="hlt">gas</span> were placed on the main floor to estimate a single-zone AER (AER(1z)). In three of these studies, a second set of emitters and detectors were deployed in the basement or second floor in approximately 10% of homes for a two-zone AER estimate (AER(2z)). In total, 287 daily pairs of AER(2z) and AER(1z) estimates were made from 35 homes across three cities. In 87% of the cases, AER(2z) was higher than AER(1z). Overall, the AER(1z) estimates underestimated AER(2z) by approximately 16% (IQR: 5-32%). This underestimate occurred in all cities and seasons and varied in magnitude seasonally, between homes, and daily, indicating that when measuring residential <span class="hlt">air</span> <span class="hlt">exchange</span> using a single PFT <span class="hlt">gas</span>, the assumption of a single well-mixed <span class="hlt">air</span> zone very likely results in an under prediction of the AER. The results of this study suggest that the long-standing assumption that a home represents a single well-mixed <span class="hlt">air</span> zone may result in a substantial negative bias in <span class="hlt">air</span> <span class="hlt">exchange</span> estimates. Indoor <span class="hlt">air</span> quality professionals should take this finding into consideration when developing study designs or making decisions related to the recommendation and installation of residential ventilation systems. © 2014 Her Majesty the Queen in Right of Canada. Indoor <span class="hlt">Air</span> published by John Wiley & Sons Ltd Reproduced with the permission of the Minister of Health Canada.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28526196','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28526196"><span>Use of a numerical simulation approach to improve the estimation of <span class="hlt">air-water</span> <span class="hlt">exchange</span> fluxes of polycyclic aromatic hydrocarbons in a coastal zone.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lai, I-Chien; Lee, Chon-Lin; Ko, Fung-Chi; Lin, Ju-Chieh; Huang, Hu-Ching; Shiu, Ruei-Feng</p> <p>2017-07-15</p> <p>The <span class="hlt">air-water</span> <span class="hlt">exchange</span> is important for determining the transport, fate, and chemical loading of polycyclic aromatic hydrocarbons (PAHs) in the atmosphere and in aquatic systems. Investigations of PAH <span class="hlt">air-water</span> <span class="hlt">exchange</span> are mostly based on observational data obtained using complicated field sampling processes. This study proposes a new approach to improve the estimation of long-term PAH <span class="hlt">air-water</span> <span class="hlt">exchange</span> fluxes by using a multivariate regression model to simulate hourly gaseous PAH concentrations. Model performance analysis and the benefits from this approach indicate its effectiveness at improving the flux estimations and at decreasing the field sampling difficulty. The proposed GIS mapping approach is useful for box model establishment and is tested for visualization of the spatiotemporal variations of <span class="hlt">air-water</span> <span class="hlt">exchange</span> fluxes in a coastal zone. The <span class="hlt">air-water</span> <span class="hlt">exchange</span> fluxes illustrated by contour maps suggest that the atmospheric PAHs might have greater impacts on offshore sites than on the coastal area in this study. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/952467','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/952467"><span>Recovery of <span class="hlt">Water</span> from Boiler Flue <span class="hlt">Gas</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>Edward Levy; Harun Bilirgen; Kwangkook Jeong</p> <p>2008-09-30</p> <p>This project dealt with use of condensing heat <span class="hlt">exchangers</span> to recover <span class="hlt">water</span> vapor from flue <span class="hlt">gas</span> at coal-fired power plants. Pilot-scale heat transfer tests were performed to determine the relationship between flue <span class="hlt">gas</span> moisture concentration, heat <span class="hlt">exchanger</span> design and operating conditions, and <span class="hlt">water</span> vapor condensation rate. The tests also determined the extent to which the condensation processes for <span class="hlt">water</span> and acid vapors in flue <span class="hlt">gas</span> can be made to occur separately in different heat transfer sections. The results showed flue <span class="hlt">gas</span> <span class="hlt">water</span> vapor condensed in the low temperature region of the heat <span class="hlt">exchanger</span> system, with <span class="hlt">water</span> capture efficiencies depending stronglymore » on flue <span class="hlt">gas</span> moisture content, cooling <span class="hlt">water</span> inlet temperature, heat <span class="hlt">exchanger</span> design and flue <span class="hlt">gas</span> and cooling <span class="hlt">water</span> flow rates. Sulfuric acid vapor condensed in both the high temperature and low temperature regions of the heat transfer apparatus, while hydrochloric and nitric acid vapors condensed with the <span class="hlt">water</span> vapor in the low temperature region. Measurements made of flue <span class="hlt">gas</span> mercury concentrations upstream and downstream of the heat <span class="hlt">exchangers</span> showed a significant reduction in flue <span class="hlt">gas</span> mercury concentration within the heat <span class="hlt">exchangers</span>. A theoretical heat and mass transfer model was developed for predicting rates of heat transfer and <span class="hlt">water</span> vapor condensation and comparisons were made with pilot scale measurements. Analyses were also carried out to estimate how much flue <span class="hlt">gas</span> moisture it would be practical to recover from boiler flue <span class="hlt">gas</span> and the magnitude of the heat rate improvements which could be made by recovering sensible and latent heat from flue <span class="hlt">gas</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25993893','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25993893"><span>Modelling non-steady-state isotope enrichment of leaf <span class="hlt">water</span> in a <span class="hlt">gas-exchange</span> cuvette environment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Song, Xin; Simonin, Kevin A; Loucos, Karen E; Barbour, Margaret M</p> <p>2015-12-01</p> <p>The combined use of a <span class="hlt">gas-exchange</span> system and laser-based isotope measurement is a tool of growing interest in plant ecophysiological studies, owing to its relevance for assessing isotopic variability in leaf <span class="hlt">water</span> and/or transpiration under non-steady-state (NSS) conditions. However, the current Farquhar & Cernusak (F&C) NSS leaf <span class="hlt">water</span> model, originally developed for open-field scenarios, is unsuited for use in a <span class="hlt">gas-exchange</span> cuvette environment where isotope composition of <span class="hlt">water</span> vapour (δv ) is intrinsically linked to that of transpiration (δE ). Here, we modified the F&C model to make it directly compatible with the δv -δE dynamic characteristic of a typical cuvette setting. The resultant new model suggests a role of 'net-flux' (rather than 'gross-flux' as suggested by the original F&C model)-based leaf <span class="hlt">water</span> turnover rate in controlling the time constant (τ) for the approach to steady sate. The validity of the new model was subsequently confirmed in a cuvette experiment involving cotton leaves, for which we demonstrated close agreement between τ values predicted from the model and those measured from NSS variations in isotope enrichment of transpiration. Hence, we recommend that our new model be incorporated into future isotope studies involving a cuvette condition where the transpiration flux directly influences δv . There is an increasing popularity among plant ecophysiologists to use a <span class="hlt">gas-exchange</span> system coupled to laser-based isotope measurement for investigating non-steady state (NSS) isotopic variability in leaf <span class="hlt">water</span> (and/or transpiration); however, the current Farquhar & Cernusak (F&C) NSS leaf <span class="hlt">water</span> model is unsuited for use in a <span class="hlt">gas-exchange</span> cuvette environment due to its implicit assumption of isotope composition of <span class="hlt">water</span> vapor (δv ) being constant and independent of that of transpiration (δE ). In the present study, we modified the F&C model to make it compatible with the dynamic relationship between δv and δE as is typically associated</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009HMT....46..175M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009HMT....46..175M"><span>High temperature heat <span class="hlt">exchanger</span> studies for applications to <span class="hlt">gas</span> turbines</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Min, June Kee; Jeong, Ji Hwan; Ha, Man Yeong; Kim, Kui Soon</p> <p>2009-12-01</p> <p>Growing demand for environmentally friendly aero <span class="hlt">gas</span>-turbine engines with lower emissions and improved specific fuel consumption can be met by incorporating heat <span class="hlt">exchangers</span> into <span class="hlt">gas</span> turbines. Relevant researches in such areas as the design of a heat <span class="hlt">exchanger</span> matrix, materials selection, manufacturing technology, and optimization by a variety of researchers have been reviewed in this paper. Based on results reported in previous studies, potential heat <span class="hlt">exchanger</span> designs for an aero <span class="hlt">gas</span> turbine recuperator, intercooler, and cooling-<span class="hlt">air</span> cooler are suggested.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26642083','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26642083"><span>Gaseous and Freely-Dissolved PCBs in the Lower Great Lakes Based on Passive Sampling: Spatial Trends and <span class="hlt">Air-Water</span> <span class="hlt">Exchange</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, Ying; Wang, Siyao; McDonough, Carrie A; Khairy, Mohammed; Muir, Derek C G; Helm, Paul A; Lohmann, Rainer</p> <p>2016-05-17</p> <p>Polyethylene passive sampling was performed to quantify gaseous and freely dissolved polychlorinated biphenyls (PCBs) in the <span class="hlt">air</span> and <span class="hlt">water</span> of Lakes Erie and Ontario during 2011-2012. In view of differing physical characteristics and the impacts of historical contamination by PCBs within these lakes, spatial variation of PCB concentrations and <span class="hlt">air-water</span> <span class="hlt">exchange</span> across these lakes may be expected. Both lakes displayed statistically similar aqueous and atmospheric PCB concentrations. Total aqueous concentrations of 29 PCBs ranged from 1.5 pg L(-1) in the open lake of Lake Erie (site E02) in 2011 spring to 105 pg L(-1) in Niagara (site On05) in 2012 summer, while total atmospheric concentrations were 7.7-634 pg m(-3) across both lakes. A west-to-east gradient was observed for aqueous PCBs in Lake Erie. River discharge and localized influences (e.g., sediment resuspension and regional alongshore transport) likely dominated spatial trends of aqueous PCBs in both lakes. <span class="hlt">Air-water</span> <span class="hlt">exchange</span> fluxes of Σ7PCBs ranged from -2.4 (±1.9) ng m(-2) day(-1) (deposition) in Sheffield (site E03) to 9.0 (±3.1) ng m(-2) day(-1) (volatilization) in Niagara (site On05). Net volatilization of PCBs was the primary trend across most sites and periods. Almost half of variation in <span class="hlt">air-water</span> <span class="hlt">exchange</span> fluxes was attributed to the difference in aqueous concentrations of PCBs. Uncertainty analysis in fugacity ratios and mass fluxes in <span class="hlt">air-water</span> <span class="hlt">exchange</span> of PCBs indicated that PCBs have reached or approached equilibrium only at the eastern Lake Erie and along the Canadian shore of Lake Ontario sites, where <span class="hlt">air-water</span> <span class="hlt">exchange</span> fluxes dominated atmospheric concentrations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22319207','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22319207"><span>Quantitative variation in <span class="hlt">water</span>-use efficiency across <span class="hlt">water</span> regimes and its relationship with circadian, vegetative, reproductive, and leaf <span class="hlt">gas-exchange</span> traits.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Edwards, Christine E; Ewers, Brent E; McClung, C Robertson; Lou, Ping; Weinig, Cynthia</p> <p>2012-05-01</p> <p>Drought limits light harvesting, resulting in lower plant growth and reproduction. One trait important for plant drought response is <span class="hlt">water</span>-use efficiency (WUE). We investigated (1) how the joint genetic architecture of WUE, reproductive characters, and vegetative traits changed across drought and well-<span class="hlt">watered</span> conditions, (2) whether traits with distinct developmental bases (e.g. leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> versus reproduction) differed in the environmental sensitivity of their genetic architecture, and (3) whether quantitative variation in circadian period was related to drought response in Brassica rapa. Overall, WUE increased in drought, primarily because stomatal conductance, and thus <span class="hlt">water</span> loss, declined more than carbon fixation. Genotypes with the highest WUE in drought expressed the lowest WUE in well-<span class="hlt">watered</span> conditions, and had the largest vegetative and floral organs in both treatments. Thus, large changes in WUE enabled some genotypes to approach vegetative and reproductive trait optima across environments. The genetic architecture differed for <span class="hlt">gas-exchange</span> and vegetative traits across drought and well-<span class="hlt">watered</span> conditions, but not for floral traits. Correlations between circadian and leaf <span class="hlt">gas-exchange</span> traits were significant but did not vary across treatments, indicating that circadian period affects physiological function regardless of <span class="hlt">water</span> availability. These results suggest that WUE is important for drought tolerance in Brassica rapa and that artificial selection for increased WUE in drought will not result in maladaptive expression of other traits that are correlated with WUE.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19203976','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19203976"><span>The hydraulic conductance of Fraxinus ornus leaves is constrained by soil <span class="hlt">water</span> availability and coordinated with <span class="hlt">gas</span> <span class="hlt">exchange</span> rates.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gortan, Emmanuelle; Nardini, Andrea; Gascó, Antonio; Salleo, Sebastiano</p> <p>2009-04-01</p> <p>Leaf hydraulic conductance (Kleaf) is known to be an important determinant of plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and photosynthesis. Little is known about the long-term impact of different environmental factors on the hydraulic construction of leaves and its eventual consequences on leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>. In this study, we investigate the impact of soil <span class="hlt">water</span> availability on Kleaf of Fraxinus ornus L. as well as the influence of Kleaf on <span class="hlt">gas</span> <span class="hlt">exchange</span> rates and plant <span class="hlt">water</span> status. With this aim, Kleaf, leaf conductance to <span class="hlt">water</span> vapour (gL), leaf <span class="hlt">water</span> potential (Psileaf) and leaf mass per area (LMA) were measured in F. ornus trees, growing in 21 different sites with contrasting <span class="hlt">water</span> availability. Plants growing in arid sites had lower Kleaf, gL and Psileaf than those growing in sites with higher <span class="hlt">water</span> availability. On the contrary, LMA was similar in the two groups. The Kleaf values recorded in sites with two different levels of soil <span class="hlt">water</span> availability were constantly different from each other regardless of the amount of precipitation recorded over 20 days before measurements. Moreover, Kleaf was correlated with gL values. Our data suggest that down-regulation of Kleaf is a component of adaptation of plants to drought-prone habitats. Low Kleaf implies reduced <span class="hlt">gas</span> <span class="hlt">exchange</span> which may, in turn, influence the climatic conditions on a local/regional scale. It is concluded that leaf hydraulics and its changes in response to resource availability should receive greater attention in studies aimed at modelling biosphere-atmosphere interactions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=315495','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=315495"><span>Deficit irrigation: Arriving at the crop <span class="hlt">water</span> stress index via <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>Plant <span class="hlt">gas</span> <span class="hlt">exchange</span> provides a highly sensitive measure of the degree of drought stress. Canopy temperature (Tc) provides a much easier to acquire indication of crop <span class="hlt">water</span> deficit that has been used in irrigation scheduling systems, but interpretation of this measurement has proven difficult. Our goa...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8143714','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8143714"><span>Pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> in acute respiratory failure.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rodriguez-Roisin, R</p> <p>1994-01-01</p> <p>The principal function of the lung is to facilitate the <span class="hlt">exchange</span> of the respiratory gases, oxygen (O2) and carbon dioxide (CO2). When the lung fails as a <span class="hlt">gas</span> <span class="hlt">exchanger</span> respiratory failure ensues. Clinically, it is generally accepted that an arterial oxygen tension (PaO2) of less than 60 mmHg or a PaCO2 of greater than 50 mmHg, or both, whilst breathing room <span class="hlt">air</span> are values consistent with the concept of respiratory failure. This article will deal, firstly, with some basic aspects of the physiology of pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> and more specifically on the measurement of ventilation-perfusion (VA/Q) relationships, the most influential factor determining hypoxaemia. The second part highlights the most important findings on pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> in the adult respiratory distress syndrome (ARDS) and other common acute respiratory failure conditions, such as pneumonia, acute exacerbation of chronic obstructive pulmonary disease (COPD) and status asthmaticus, based on the data obtained by means of the multiple inert <span class="hlt">gas</span> elimination approach, a technique which gives a detailed picture of VA/Q ratio distributions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4674977','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4674977"><span>Estimation of bias with the single-zone assumption in measurement of residential <span class="hlt">air</span> <span class="hlt">exchange</span> using the perfluorocarbon tracer <span class="hlt">gas</span> method</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Van Ryswyk, K; Wallace, L; Fugler, D; MacNeill, M; Héroux, M È; Gibson, M D; Guernsey, J R; Kindzierski, W; Wheeler, A J</p> <p>2015-01-01</p> <p>Residential <span class="hlt">air</span> <span class="hlt">exchange</span> rates (AERs) are vital in understanding the temporal and spatial drivers of indoor <span class="hlt">air</span> quality (IAQ). Several methods to quantify AERs have been used in IAQ research, often with the assumption that the home is a single, well-mixed <span class="hlt">air</span> zone. Since 2005, Health Canada has conducted IAQ studies across Canada in which AERs were measured using the perfluorocarbon tracer (PFT) <span class="hlt">gas</span> method. Emitters and detectors of a single PFT <span class="hlt">gas</span> were placed on the main floor to estimate a single-zone AER (AER1z). In three of these studies, a second set of emitters and detectors were deployed in the basement or second floor in approximately 10% of homes for a two-zone AER estimate (AER2z). In total, 287 daily pairs of AER2z and AER1z estimates were made from 35 homes across three cities. In 87% of the cases, AER2z was higher than AER1z. Overall, the AER1z estimates underestimated AER2z by approximately 16% (IQR: 5–32%). This underestimate occurred in all cities and seasons and varied in magnitude seasonally, between homes, and daily, indicating that when measuring residential <span class="hlt">air</span> <span class="hlt">exchange</span> using a single PFT <span class="hlt">gas</span>, the assumption of a single well-mixed <span class="hlt">air</span> zone very likely results in an under prediction of the AER. PMID:25399878</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27209375','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27209375"><span><span class="hlt">Gas</span> <span class="hlt">Exchange</span> Models for a Flexible Insect Tracheal System.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Simelane, S M; Abelman, S; Duncan, F D</p> <p>2016-06-01</p> <p>In this paper two models for movement of respiratory gases in the insect trachea are presented. One model considers the tracheal system as a single flexible compartment while the other model considers the trachea as a single flexible compartment with <span class="hlt">gas</span> <span class="hlt">exchange</span>. This work represents an extension of Ben-Tal's work on compartmental <span class="hlt">gas</span> <span class="hlt">exchange</span> in human lungs and is applied to the insect tracheal system. The purpose of the work is to study nonlinear phenomena seen in the insect respiratory system. It is assumed that the flow inside the trachea is laminar, and that the <span class="hlt">air</span> inside the chamber behaves as an ideal <span class="hlt">gas</span>. Further, with the isothermal assumption, the expressions for the tracheal partial pressures of oxygen and carbon dioxide, rate of volume change, and the rates of change of oxygen concentration and carbon dioxide concentration are derived. The effects of some flow parameters such as diffusion capacities, reaction rates and <span class="hlt">air</span> concentrations on net flow are studied. Numerical simulations of the tracheal flow characteristics are performed. The models developed provide a mathematical framework to further investigate <span class="hlt">gas</span> <span class="hlt">exchange</span> in insects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25761782','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25761782"><span>Protein structural dynamics at the <span class="hlt">gas/water</span> interface examined by hydrogen <span class="hlt">exchange</span> mass spectrometry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xiao, Yiming; Konermann, Lars</p> <p>2015-08-01</p> <p><span class="hlt">Gas/water</span> interfaces (such as <span class="hlt">air</span> bubbles or foam) are detrimental to the stability of proteins, often causing aggregation. This represents a potential problem for industrial processes, for example, the production and handling of protein drugs. Proteins possess surfactant-like properties, resulting in a high affinity for <span class="hlt">gas/water</span> interfaces. The tendency of previously buried nonpolar residues to maximize contact with the <span class="hlt">gas</span> phase can cause significant structural distortion. Most earlier studies in this area employed spectroscopic tools that could only provide limited information. Here we use hydrogen/deuterium <span class="hlt">exchange</span> (HDX) mass spectrometry (MS) for probing the conformational dynamics of the model protein myoglobin (Mb) in the presence of N(2) bubbles. HDX/MS relies on the principle that unfolded and/or highly dynamic regions undergo faster deuteration than tightly folded segments. In bubble-free solution Mb displays EX2 behavior, reflecting the occurrence of short-lived excursions to partially unfolded conformers. A dramatically different behavior is seen in the presence of N(2) bubbles; EX2 dynamics still take place, but in addition the protein shows EX1 behavior. The latter results from interconversion of the native state with conformers that are globally unfolded and long-lived. These unfolded species likely correspond to Mb that is adsorbed to the surface of <span class="hlt">gas</span> bubbles. N(2) sparging also induces aggregation. To explain the observed behavior we propose a simple model, that is, "semi-unfolded" ↔ "native" ↔ "globally unfolded" → "aggregated". This model quantitatively reproduces the experimentally observed kinetics. To the best of our knowledge, the current study marks the first exploration of surface denaturation phenomena by HDX/MS. © 2015 The Protein Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27458653','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27458653"><span>Spatial Distribution and <span class="hlt">Air-Water</span> <span class="hlt">Exchange</span> of Organic Flame Retardants in the Lower Great Lakes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>McDonough, Carrie A; Puggioni, Gavino; Helm, Paul A; Muir, Derek; Lohmann, Rainer</p> <p>2016-09-06</p> <p>Organic flame retardants (OFRs) such as polybrominated diphenyl ethers (PBDEs) and novel halogenated flame retardants (NHFRs) are ubiquitous, persistent, and bioaccumulative contaminants that have been used in consumer goods to slow combustion. In this study, polyethylene passive samplers (PEs) were deployed throughout the lower Great Lakes (Lake Erie and Lake Ontario) to measure OFRs in <span class="hlt">air</span> and <span class="hlt">water</span>, calculate <span class="hlt">air-water</span> <span class="hlt">exchange</span> fluxes, and investigate spatial trends. Dissolved Σ12BDE was greatest in Lake Ontario near Toronto (18 pg/L), whereas gaseous Σ12BDE was greatest on the southern shoreline of Lake Erie (11 pg/m(3)). NHFRs were generally below detection limits. <span class="hlt">Air-water</span> <span class="hlt">exchange</span> was dominated by absorption of BDEs 47 and 99, ranging from -964 pg/m(2)/day to -30 pg/m(2)/day. Σ12BDE in <span class="hlt">air</span> and <span class="hlt">water</span> was significantly correlated with surrounding population density, suggesting that phased-out PBDEs continued to be emitted from population centers along the Great Lakes shoreline in 2012. Correlation with dissolved Σ12BDE was strongest when considering population within 25 km while correlation with gaseous Σ12BDE was strongest when using population within 3 km to the south of each site. Bayesian kriging was used to predict dissolved Σ12BDE over the lakes, illustrating the utility of relatively highly spatially resolved measurements in identifying potential hot spots for future study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21692813','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21692813"><span>Responses to <span class="hlt">water</span> stress of <span class="hlt">gas</span> <span class="hlt">exchange</span> and metabolites in Eucalyptus and Acacia spp.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Warren, Charles R; Aranda, Ismael; Cano, F Javier</p> <p>2011-10-01</p> <p>Studies of <span class="hlt">water</span> stress commonly examine either <span class="hlt">gas</span> <span class="hlt">exchange</span> or leaf metabolites, and many fail to quantify the concentration of CO₂ in the chloroplasts (C(c)). We redress these limitations by quantifying C(c) from discrimination against ¹³CO₂ and using <span class="hlt">gas</span> chromatography-mass spectrometry (GC-MS) for leaf metabolite profiling. Five Eucalyptus and two Acacia species from semi-arid to mesic habitats were subjected to a 2 month <span class="hlt">water</span> stress treatment (Ψ(pre-dawn) = -1.7 to -2.3 MPa). Carbohydrates dominated the leaf metabolite profiles of species from dry areas, whereas organic acids dominated the metabolite profiles of species from wet areas. <span class="hlt">Water</span> stress caused large decreases in photosynthesis and C(c), increases in 17-33 metabolites and decreases in 0-9 metabolites. In most species, fructose, glucose and sucrose made major contributions to osmotic adjustment. In Acacia, significant osmotic adjustment was also caused by increases in pinitol, pipecolic acid and trans-4-hydroxypipecolic acid. There were also increases in low-abundance metabolites (e.g. proline and erythritol), and metabolites that are indicative of stress-induced changes in metabolism [e.g. γ-aminobutyric acid (GABA) shunt, photorespiration, phenylpropanoid pathway]. The response of <span class="hlt">gas</span> <span class="hlt">exchange</span> to <span class="hlt">water</span> stress and rewatering is rather consistent among species originating from mesic to semi-arid habitats, and the general response of metabolites to <span class="hlt">water</span> stress is rather similar, although the specific metabolites involved may vary. © 2011 Blackwell Publishing Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUSM.H33C..06B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUSM.H33C..06B"><span><span class="hlt">Air</span> Sparging Versus <span class="hlt">Gas</span> Saturated <span class="hlt">Water</span> Injection for Remediation of Volatile LNAPL in the Borden Aquifer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barker, J.; Nelson, L.; Doughty, C.; Thomson, N.; Lambert, J.</p> <p>2009-05-01</p> <p>In the shallow, rather homogeneous, unconfined Borden sand aquifer, field trials of <span class="hlt">air</span> sparging (Tomlinson et al., 2003) and pulsed <span class="hlt">air</span> sparging (Lambert et al., 2009) have been conducted, the latter to remediate a residual gasoline source emplaced below the <span class="hlt">water</span> table. As well, a supersaturated (with CO2) <span class="hlt">water</span> injection (SWI) technology, using the inVentures inFusion system, has been trialed in two phases: 1. in the uncontaminated sand aquifer to evaluate the radius of influence, extent of lateral <span class="hlt">gas</span> movement and <span class="hlt">gas</span> saturation below the <span class="hlt">water</span> table, and 2. in a sheet pile cell in the Borden aquifer to evaluate the recovery of volatile hydrocarbon components (pentane and hexane) of an LNAPL emplaced below the <span class="hlt">water</span> table (Nelson et al., 2008). The SWI injects <span class="hlt">water</span> supersaturated with CO2. The supersaturated injected <span class="hlt">water</span> moves laterally away from the sparge point, releasing CO2 over a wider area than does <span class="hlt">gas</span> sparging from a single well screen. This presentation compares these two techniques in terms of their potential for remediating volatile NAPL components occurring below the <span class="hlt">water</span> table in a rather homogeneous sand aquifer. <span class="hlt">Air</span> sparging created a significantly greater <span class="hlt">air</span> saturation in the vicinity of the sparge well than did the CO2 system (60 percent versus 16 percent) in the uncontaminated Borden aquifer. However, SWI pushed <span class="hlt">water</span>, still supersaturated with CO2, up to about 2.5 m from the injection well. This would seem to provide a considerable advantage over <span class="hlt">air</span> sparging from a point, in that <span class="hlt">gas</span> bubbles are generated at a much larger radius from the point of injection with SWI and so should involve additional <span class="hlt">gas</span> pathways through a residual NAPL. Overall, <span class="hlt">air</span> sparging created a greater area of influence, defined by measurable <span class="hlt">air</span> saturation in the aquifer, but <span class="hlt">air</span> sparging also injected about 12 times more <span class="hlt">gas</span> than was injected in the SWI trials. The pulsed <span class="hlt">air</span> sparging at Borden (Lambert et al.) removed about 20 percent (4.6 kg) of gasoline</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('http://adsabs.harvard.edu/abs/2018ThEng..65..300O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ThEng..65..300O"><span>Study of Cycling <span class="hlt">Air</span>-Cooling System with a Cold Accumulator for Micro <span class="hlt">Gas</span>-Turbine Installations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ochkov, V. F.; Stepanova, T. A.; Katenev, G. M.; Tumanovskii, V. A.; Borisova, P. N.</p> <p>2018-05-01</p> <p>Using the cycling <span class="hlt">air</span>-cooling systems of the CTIC type (Combustion Turbine Inlet Cooling) with a cold accumulator in a micro <span class="hlt">gas</span>-turbine installation (micro-GTI) to preserve its capacity under the seasonal temperature rise of outside <span class="hlt">air</span> is described. <span class="hlt">Water</span> ice is used as the body-storage in the accumulators, and ice <span class="hlt">water</span> (<span class="hlt">water</span> at 0.5-1.0°C) is used as the body that cools <span class="hlt">air</span>. The ice <span class="hlt">water</span> circulates between the accumulator and the <span class="hlt">air-water</span> heat <span class="hlt">exchanger</span>. The cold accumulator model with renewable ice resources is considered. The model contains the heat-<span class="hlt">exchanging</span> tube lattice-evaporator covered with ice. The lattice is cross-flowed with <span class="hlt">water</span>. The criterion heat <span class="hlt">exchange</span> equation that describes the process in the cold accumulator under consideration is presented. The calculations of duration of its active operation were performed. The dependence of cold accumulator service life on <span class="hlt">water</span> circulation rate was evaluated. The adequacy of the design model was confirmed experimentally in the mock-up of the cold accumulator with a refrigerating machine periodically creating a 200 kg ice reserve in the reservoir-storage. The design model makes it possible to determine the weight of ice reserve of the discharged cold accumulator for cooling the cycle <span class="hlt">air</span> in the operation of a C-30 type micro- GTI produced by the Capstone Company or micro-GTIs of other capacities. Recommendations for increasing the working capacity of cold accumulators of CTIC-systems of a micro-GTI were made.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12010472','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12010472"><span><span class="hlt">Water</span> relations and <span class="hlt">gas</span> <span class="hlt">exchange</span> in poplar and willow under <span class="hlt">water</span> stress and elevated atmospheric CO2.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Johnson, Jon D; Tognetti, Roberto; Paris, Piero</p> <p>2002-05-01</p> <p>Predictions of shifts in rainfall patterns as atmospheric [CO2] increases could impact the growth of fast growing trees such as Populus spp. and Salix spp. and the interaction between elevated CO2 and <span class="hlt">water</span> stress in these species is unknown. The objectives of this study were to characterize the responses to elevated CO2 and <span class="hlt">water</span> stress in these two species, and to determine if elevated CO2 mitigated drought stress effects. <span class="hlt">Gas</span> <span class="hlt">exchange</span>, <span class="hlt">water</span> potential components, whole plant transpiration and growth response to soil drying and recovery were assessed in hybrid poplar (clone 53-246) and willow (Salix sagitta) rooted cuttings growing in either ambient (350 &mgr;mol mol-1) or elevated (700 &mgr;mol mol-1) atmospheric CO2 concentration ([CO2]). Predawn <span class="hlt">water</span> potential decreased with increasing <span class="hlt">water</span> stress while midday <span class="hlt">water</span> potentials remained unchanged (isohydric response). Turgor potentials at both predawn and midday increased in elevated [CO2], indicative of osmotic adjustment. <span class="hlt">Gas</span> <span class="hlt">exchange</span> was reduced by <span class="hlt">water</span> stress while elevated [CO2] increased photosynthetic rates, reduced leaf conductance and nearly doubled instantaneous transpiration efficiency in both species. Dark respiration decreased in elevated [CO2] and <span class="hlt">water</span> stress reduced Rd in the trees growing in ambient [CO2]. Willow had 56% lower whole plant hydraulic conductivity than poplar, and showed a 14% increase in elevated [CO2] while poplar was unresponsive. The physiological responses exhibited by poplar and willow to elevated [CO2] and <span class="hlt">water</span> stress, singly, suggest that these species respond like other tree species. The interaction of [CO2] and <span class="hlt">water</span> stress suggests that elevated [CO2] did mitigate the effects of <span class="hlt">water</span> stress in willow, but not in poplar.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSEC54B1327H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSEC54B1327H"><span>First System-Wide Estimates of <span class="hlt">Air</span>-Sea <span class="hlt">Exchange</span> of Carbon Dioxide in the Chesapeake Bay</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Herrmann, M.; Najjar, R.; Menendez, A.</p> <p>2016-02-01</p> <p>Estuaries are estimated to play a major role in the global carbon cycle by degassing between 0.25 and 0.4 Pg C y-1, comparable to the uptake of atmospheric CO2 by continental shelf <span class="hlt">waters</span> and as much as one quarter of the uptake of atmospheric CO2 by the open ocean. However, the global estimates of estuarine CO2 <span class="hlt">gas</span> <span class="hlt">exchange</span> are highly uncertain mostly due to limited data availability and extreme heterogeneity of coastal systems. Notably, the <span class="hlt">air-water</span> CO2 flux for the largest U.S. estuary, the Chesapeake Bay, is yet unknown. Here we provide first system-level CO2 <span class="hlt">gas</span> <span class="hlt">exchange</span> estimates for the Chesapeake Bay, using data from the Chesapeake Bay <span class="hlt">Water</span> Quality Monitoring Program (CBWQMP) and other data sources. We focus on the main stem of the Chesapeake Bay; hence, tributaries, such as the tidal portions of the Potomac and James Rivers, are not included in this first estimation of the flux. The preliminary results show the Bay to be a net source of CO2 to the atmosphere, outgassing on average 0.2 Tg C yr-1 over the study period, between 1985 and 2013. The spatial and temporal variability of the <span class="hlt">gas</span> <span class="hlt">exchange</span> will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22292498','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22292498"><span>Comparison of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and stable isotope signature of <span class="hlt">water</span>-soluble compounds along canopy gradients of co-occurring Douglas-fir and European beech.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bögelein, Rebekka; Hassdenteufel, Martin; Thomas, Frank M; Werner, Willy</p> <p>2012-07-01</p> <p>Combined δ(13) C and δ(18) O analyses of <span class="hlt">water</span>-soluble leaf and twig phloem material were used to determine intrinsic <span class="hlt">water</span>-use efficiency (iWUE) and variability of stomatal conductance at different crown positions in adult European beech (Fagus sylvatica) and Douglas-fir (Pseudotsuga menziesii) trees. Simultaneous <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements allowed evaluation of the differences in calculating iWUE from leaf or phloem <span class="hlt">water</span>-soluble compounds, and comparison with a semi-quantitative dual isotope model to infer variability of net photosynthesis (A(n) ) between the investigated crown positions. Estimates of iWUE from δ(13) C of leaf <span class="hlt">water</span>-soluble organic matter (WSOM) outperformed the estimates from phloem compounds. In the beech crown, δ(13) C of leaf WSOM coincided clearly with <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements. The relationship was not as reliable in the Douglas-fir. The differences in δ(18) O between leaf and phloem material were found to correlate with stomatal conductance. The semi-quantitative model approach was applicable for comparisons of daily average A(n) between different crown positions and trees. Intracanopy gradients were more pronounced in the beech than in the Douglas-fir, which reached higher values of iWUE at the respective positions, particularly under dry <span class="hlt">air</span> conditions. © 2012 Blackwell Publishing Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20336837','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20336837"><span>Impact of airway <span class="hlt">gas</span> <span class="hlt">exchange</span> on the multiple inert <span class="hlt">gas</span> elimination technique: theory.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Anderson, Joseph C; Hlastala, Michael P</p> <p>2010-03-01</p> <p>The multiple inert <span class="hlt">gas</span> elimination technique (MIGET) provides a method for estimating alveolar <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency. Six soluble inert gases are infused into a peripheral vein. Measurements of these gases in breath, arterial blood, and venous blood are interpreted using a mathematical model of alveolar <span class="hlt">gas</span> <span class="hlt">exchange</span> (MIGET model) that neglects airway <span class="hlt">gas</span> <span class="hlt">exchange</span>. A mathematical model describing airway and alveolar <span class="hlt">gas</span> <span class="hlt">exchange</span> predicts that two of these gases, ether and acetone, <span class="hlt">exchange</span> primarily within the airways. To determine the effect of airway <span class="hlt">gas</span> <span class="hlt">exchange</span> on the MIGET, we selected two additional gases, toluene and m-dichlorobenzene, that have the same blood solubility as ether and acetone and minimize airway <span class="hlt">gas</span> <span class="hlt">exchange</span> via their low <span class="hlt">water</span> solubility. The airway-alveolar <span class="hlt">gas</span> <span class="hlt">exchange</span> model simulated the <span class="hlt">exchange</span> of toluene, m-dichlorobenzene, and the six MIGET gases under multiple conditions of alveolar ventilation-to-perfusion, VA/Q, heterogeneity. We increased the importance of airway <span class="hlt">gas</span> <span class="hlt">exchange</span> by changing bronchial blood flow, Qbr. From these simulations, we calculated the excretion and retention of the eight inert gases and divided the results into two groups: (1) the standard MIGET gases which included acetone and ether and (2) the modified MIGET gases which included toluene and m-dichlorobenzene. The MIGET mathematical model predicted distributions of ventilation and perfusion for each grouping of gases and multiple perturbations of VA/Q and Qbr. Using the modified MIGET gases, MIGET predicted a smaller dead space fraction, greater mean VA, greater log(SDVA), and more closely matched the imposed VA distribution than that using the standard MIGET gases. Perfusion distributions were relatively unaffected.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25830350','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25830350"><span>Peach <span class="hlt">water</span> relations, <span class="hlt">gas</span> <span class="hlt">exchange</span>, growth and shoot mortality under <span class="hlt">water</span> deficit in semi-arid weather conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rahmati, Mitra; Davarynejad, Gholam Hossein; Génard, Michel; Bannayan, Mohammad; Azizi, Majid; Vercambre, Gilles</p> <p>2015-01-01</p> <p>In this study the sensitivity of peach tree (Prunus persica L.) to three <span class="hlt">water</span> stress levels from mid-pit hardening until harvest was assessed. Seasonal patterns of shoot and fruit growth, <span class="hlt">gas</span> <span class="hlt">exchange</span> (leaf photosynthesis, stomatal conductance and transpiration) as well as carbon (C) storage/mobilization were evaluated in relation to plant <span class="hlt">water</span> status. A simple C balance model was also developed to investigate sink-source relationship in relation to plant <span class="hlt">water</span> status at the tree level. The C source was estimated through the leaf area dynamics and leaf photosynthesis rate along the season. The C sink was estimated for maintenance respiration and growth of shoots and fruits. <span class="hlt">Water</span> stress significantly reduced <span class="hlt">gas</span> <span class="hlt">exchange</span>, and fruit, and shoot growth, but increased fruit dry matter concentration. Growth was more affected by <span class="hlt">water</span> deficit than photosynthesis, and shoot growth was more sensitive to <span class="hlt">water</span> deficit than fruit growth. Reduction of shoot growth was associated with a decrease of shoot elongation, emergence, and high shoot mortality. <span class="hlt">Water</span> scarcity affected tree C assimilation due to two interacting factors: (i) reduction in leaf photosynthesis (-23% and -50% under moderate (MS) and severe (SS) <span class="hlt">water</span> stress compared to low (LS) stress during growth season) and (ii) reduction in total leaf area (-57% and -79% under MS and SS compared to LS at harvest). Our field data analysis suggested a Ψstem threshold of -1.5 MPa below which daily net C gain became negative, i.e. C assimilation became lower than C needed for respiration and growth. Negative C balance under MS and SS associated with decline of trunk carbohydrate reserves--may have led to drought-induced vegetative mortality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4382189','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4382189"><span>Peach <span class="hlt">Water</span> Relations, <span class="hlt">Gas</span> <span class="hlt">Exchange</span>, Growth and Shoot Mortality under <span class="hlt">Water</span> Deficit in Semi-Arid Weather Conditions</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Rahmati, Mitra; Davarynejad, Gholam Hossein; Génard, Michel; Bannayan, Mohammad; Azizi, Majid; Vercambre, Gilles</p> <p>2015-01-01</p> <p>In this study the sensitivity of peach tree (Prunus persica L.) to three <span class="hlt">water</span> stress levels from mid-pit hardening until harvest was assessed. Seasonal patterns of shoot and fruit growth, <span class="hlt">gas</span> <span class="hlt">exchange</span> (leaf photosynthesis, stomatal conductance and transpiration) as well as carbon (C) storage/mobilization were evaluated in relation to plant <span class="hlt">water</span> status. A simple C balance model was also developed to investigate sink-source relationship in relation to plant <span class="hlt">water</span> status at the tree level. The C source was estimated through the leaf area dynamics and leaf photosynthesis rate along the season. The C sink was estimated for maintenance respiration and growth of shoots and fruits. <span class="hlt">Water</span> stress significantly reduced <span class="hlt">gas</span> <span class="hlt">exchange</span>, and fruit, and shoot growth, but increased fruit dry matter concentration. Growth was more affected by <span class="hlt">water</span> deficit than photosynthesis, and shoot growth was more sensitive to <span class="hlt">water</span> deficit than fruit growth. Reduction of shoot growth was associated with a decrease of shoot elongation, emergence, and high shoot mortality. <span class="hlt">Water</span> scarcity affected tree C assimilation due to two interacting factors: (i) reduction in leaf photosynthesis (-23% and -50% under moderate (MS) and severe (SS) <span class="hlt">water</span> stress compared to low (LS) stress during growth season) and (ii) reduction in total leaf area (-57% and -79% under MS and SS compared to LS at harvest). Our field data analysis suggested a Ψstem threshold of -1.5 MPa below which daily net C gain became negative, i.e. C assimilation became lower than C needed for respiration and growth. Negative C balance under MS and SS associated with decline of trunk carbohydrate reserves – may have led to drought-induced vegetative mortality. PMID:25830350</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ThEng..64..680B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ThEng..64..680B"><span>Investigation and optimization of the depth of flue <span class="hlt">gas</span> heat recovery in surface heat <span class="hlt">exchangers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bespalov, V. V.; Bespalov, V. I.; Melnikov, D. V.</p> <p>2017-09-01</p> <p>Economic issues associated with designing deep flue <span class="hlt">gas</span> heat recovery units for natural <span class="hlt">gas</span>-fired boilers are examined. The governing parameter affecting the performance and cost of surface-type condensing heat recovery heat <span class="hlt">exchangers</span> is the heat transfer surface area. When firing natural <span class="hlt">gas</span>, the heat recovery depth depends on the flue <span class="hlt">gas</span> temperature at the condenser outlet and determines the amount of condensed <span class="hlt">water</span> vapor. The effect of the outlet flue <span class="hlt">gas</span> temperature in a heat recovery heat <span class="hlt">exchanger</span> on the additionally recovered heat power is studied. A correlation has been derived enabling one to determine the best heat recovery depth (or the final cooling temperature) maximizing the anticipated reduced annual profit of a power enterprise from implementation of energy-saving measures. Results of optimization are presented for a surface-type condensing <span class="hlt">gas-air</span> plate heat recovery heat <span class="hlt">exchanger</span> for the climatic conditions and the economic situation in Tomsk. The predictions demonstrate that it is economically feasible to design similar heat recovery heat <span class="hlt">exchangers</span> for a flue <span class="hlt">gas</span> outlet temperature of 10°C. In this case, the payback period for the investment in the heat recovery heat <span class="hlt">exchanger</span> will be 1.5 years. The effect of various factors on the optimal outlet flue <span class="hlt">gas</span> temperature was analyzed. Most climatic, economical, or technological factors have a minor effect on the best outlet temperature, which remains between 5 and 20°C when varying the affecting factors. The derived correlation enables us to preliminary estimate the outlet (final) flue <span class="hlt">gas</span> temperature that should be used in designing the heat transfer surface of a heat recovery heat <span class="hlt">exchanger</span> for a <span class="hlt">gas</span>-fired boiler as applied to the specific climatic conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1050299','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1050299"><span>PREDICTION OF TOTAL DISSOLVED <span class="hlt">GAS</span> <span class="hlt">EXCHANGE</span> AT HYDROPOWER DAMS</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>Hadjerioua, Boualem; Pasha, MD Fayzul K; Stewart, Kevin M</p> <p>2012-07-01</p> <p>Total dissolved <span class="hlt">gas</span> (TDG) supersaturation in <span class="hlt">waters</span> released at hydropower dams can cause <span class="hlt">gas</span> bubble trauma in fisheries resulting in physical injuries and eyeball protrusion that can lead to mortality. Elevated TDG pressures in hydropower releases are generally caused by the entrainment of <span class="hlt">air</span> in spillway releases and the subsequent <span class="hlt">exchange</span> of atmospheric gasses into solution during passage through the stilling basin. The network of dams throughout the Columbia River Basin (CRB) are managed for irrigation, hydropower production, flood control, navigation, and fish passage that frequently result in both voluntary and involuntary spillway releases. These dam operations are constrained bymore » state and federal <span class="hlt">water</span> quality standards for TDG saturation which balance the benefits of spillway operations designed for Endangered Species Act (ESA)-listed fisheries versus the degradation to <span class="hlt">water</span> quality as defined by TDG saturation. In the 1970s, the United States Environmental Protection Agency (USEPA), under the federal Clean <span class="hlt">Water</span> Act (Section 303(d)), established a criterion not to exceed the TDG saturation level of 110% in order to protect freshwater and marine aquatic life. The states of Washington and Oregon have adopted special <span class="hlt">water</span> quality standards for TDG saturation in the tailrace and forebays of hydropower facilities on the Columbia and Snake Rivers where spillway operations support fish passage objectives. The physical processes that affect TDG <span class="hlt">exchange</span> at hydropower facilities have been studied throughout the CRB in site-specific studies and routine <span class="hlt">water</span> quality monitoring programs. These data have been used to quantify the relationship between project operations, structural properties, and TDG <span class="hlt">exchange</span>. These data have also been used to develop predictive models of TDG <span class="hlt">exchange</span> to support real-time TDG management decisions. These empirically based predictive models have been developed for specific projects and account for both the fate of spillway</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/39792','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/39792"><span>Growth, <span class="hlt">gas</span> <span class="hlt">exchange</span>, foliar nitrogen content, and <span class="hlt">water</span> use of subirrigated and overhead irrigated Populus tremuloides Michx. seedlings</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Anthony S. Davis; Matthew M. Aghai; Jeremiah R. Pinto; Kent G. Apostal</p> <p>2011-01-01</p> <p>Because limitations on <span class="hlt">water</span> used by container nurseries has become commonplace, nursery growers will have to improve irrigation management. Subirrigation systems may provide an alternative to overhead irrigation systems by mitigating groundwater pollution and excessive <span class="hlt">water</span> consumption. Seedling growth, <span class="hlt">gas</span> <span class="hlt">exchange</span>, leaf nitrogen (N) content, and <span class="hlt">water</span> use were...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16246857','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16246857"><span><span class="hlt">Water</span> stress-induced modifications of leaf hydraulic architecture in sunflower: co-ordination with <span class="hlt">gas</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nardini, Andrea; Salleo, Sebastiano</p> <p>2005-12-01</p> <p>The hydraulic architecture, <span class="hlt">water</span> relationships, and <span class="hlt">gas</span> <span class="hlt">exchange</span> of leaves of sunflower plants, grown under different levels of <span class="hlt">water</span> stress, were measured. Plants were either irrigated with tap <span class="hlt">water</span> (controls) or with PEG600 solutions with osmotic potential of -0.4 and -0.8 MPa (PEG04 and PEG08 plants, respectively). Mature leaves were measured for hydraulic resistance (R(leaf)) before and after making several cuts across minor veins, thus getting the hydraulic resistance of the venation system (R(venation)). R(leaf) was nearly the same in controls and PEG04 plants but it was reduced by about 30% in PEG08 plants. On the contrary, R(venation) was lowest in controls and increased in PEG04 and PEG08 plants as a likely result of reduction in the diameter of the veins' conduits. As a consequence, the contribution of R(venation) to the overall R(leaf) markedly increased from controls to PEG08 plants. Leaf conductance to <span class="hlt">water</span> vapour (g(L)) was highest in controls and significantly lower in PEG04 and PEG08 plants. Moreover, g(L) was correlated to R(venation) and to leaf <span class="hlt">water</span> potential (psi(leaf)) with highly significant linear relationships. It is concluded that <span class="hlt">water</span> stress has an important effect on the hydraulic construction of leaves. This, in turn, might prove to be a crucial factor in plant-<span class="hlt">water</span> relationships and <span class="hlt">gas</span> <span class="hlt">exchange</span> under <span class="hlt">water</span> stress conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998EPJAP...3..295B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998EPJAP...3..295B"><span>Optimization of heat and mass transfers in counterflow corrugated-plate liquid-<span class="hlt">gas</span> <span class="hlt">exchangers</span> used in a greenhouse dehumidifier</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bentounes, N.; Jaffrin, A.</p> <p>1998-09-01</p> <p>Heat and mass transfers occuring in a counterflow direct contact liquid-<span class="hlt">gas</span> <span class="hlt">exchanger</span> determine the performance of a new greenhouse <span class="hlt">air</span> dehumidifier designed at INRA. This prototype uses triethylene glycol (TEG) as the desiccant fluid which extracts <span class="hlt">water</span> vapor from the <span class="hlt">air</span>. The regeneration of the TEG desiccant fluid is then performed by direct contact with combustion <span class="hlt">gas</span> from a high efficiency boiler equipped with a condensor. The heat and mass transfers between the thin film of diluted TEG and the hot <span class="hlt">gas</span> were simulated by a model which uses correlation formula from the literature specifically relevant to the present cross-corrugated plates geometry. A simple set of analytical solutions is first derived, which explains why some possible processes can clearly be far from optimal. Then, more exact numerical calculations confirm that some undesirable <span class="hlt">water</span> recondensations on the upper part of the <span class="hlt">exchanger</span> were limiting the performance of this prototype. More suitable conditions were defined for the process, which lead to a new design of the apparatus. In this second prototype, a <span class="hlt">gas-gas</span> <span class="hlt">exchanger</span> provides dryer and cooler <span class="hlt">gas</span> to the basis of the regenerators, while a warmer TEG is fed on the top. A whole range of operating conditions was experimented and measured parameters were compared with numerical simulations of this new configuration: recondensation did not occur any more. As a consequence, this second prototype was able to concentrate the desiccant fluid at the desired rate of 20 kg H_{2O}/hour, under temperature and humidity conditions which correspond to the dehumidification of a 1000 m2 greenhouse heated at night during the winter season.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24854169','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24854169"><span>Dynamics of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>, xylem and phloem transport, <span class="hlt">water</span> potential and carbohydrate concentration in a realistic 3-D model tree crown.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nikinmaa, Eero; Sievänen, Risto; Hölttä, Teemu</p> <p>2014-09-01</p> <p>Tree models simulate productivity using general <span class="hlt">gas</span> <span class="hlt">exchange</span> responses and structural relationships, but they rarely check whether leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and resulting <span class="hlt">water</span> and assimilate transport and driving pressure gradients remain within acceptable physical boundaries. This study presents an implementation of the cohesion-tension theory of xylem transport and the Münch hypothesis of phloem transport in a realistic 3-D tree structure and assesses the <span class="hlt">gas</span> <span class="hlt">exchange</span> and transport dynamics. A mechanistic model of xylem and phloem transport was used, together with a tested leaf assimilation and transpiration model in a realistic tree architecture to simulate leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> and carbohydrate transport within an 8-year-old Scots pine tree. The model solved the dynamics of the amounts of <span class="hlt">water</span> and sucrose solute in the xylem, cambium and phloem using a fine-grained mesh with a system of coupled ordinary differential equations. The simulations predicted the observed patterns of pressure gradients and sugar concentration. Diurnal variation of environmental conditions influenced tree-level gradients in turgor pressure and sugar concentration, which are important drivers of carbon allocation. The results and between-shoot variation were sensitive to structural and functional parameters such as tree-level scaling of conduit size and phloem unloading. Linking whole-tree-level <span class="hlt">water</span> and assimilate transport, <span class="hlt">gas</span> <span class="hlt">exchange</span> and sink activity opens a new avenue for plant studies, as features that are difficult to measure can be studied dynamically with the model. Tree-level responses to local and external conditions can be tested, thus making the approach described here a good test-bench for studies of whole-tree physiology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26388365','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26388365"><span>Direct analysis of ultra-trace semiconductor <span class="hlt">gas</span> by inductively coupled plasma mass spectrometry coupled with <span class="hlt">gas</span> to particle conversion-<span class="hlt">gas</span> <span class="hlt">exchange</span> technique.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ohata, Masaki; Sakurai, Hiromu; Nishiguchi, Kohei; Utani, Keisuke; Günther, Detlef</p> <p>2015-09-03</p> <p>An inductively coupled plasma mass spectrometry (ICPMS) coupled with <span class="hlt">gas</span> to particle conversion-<span class="hlt">gas</span> <span class="hlt">exchange</span> technique was applied to the direct analysis of ultra-trace semiconductor <span class="hlt">gas</span> in ambient <span class="hlt">air</span>. The ultra-trace semiconductor gases such as arsine (AsH3) and phosphine (PH3) were converted to particles by reaction with ozone (O3) and ammonia (NH3) gases within a <span class="hlt">gas</span> to particle conversion device (GPD). The converted particles were directly introduced and measured by ICPMS through a <span class="hlt">gas</span> <span class="hlt">exchange</span> device (GED), which could penetrate the particles as well as <span class="hlt">exchange</span> to Ar from either non-reacted gases such as an <span class="hlt">air</span> or remaining gases of O3 and NH3. The particle size distribution of converted particles was measured by scanning mobility particle sizer (SMPS) and the results supported the elucidation of particle agglomeration between the particle converted from semiconductor <span class="hlt">gas</span> and the particle of ammonium nitrate (NH4NO3) which was produced as major particle in GPD. Stable time-resolved signals from AsH3 and PH3 in <span class="hlt">air</span> were obtained by GPD-GED-ICPMS with continuous <span class="hlt">gas</span> introduction; however, the slightly larger fluctuation, which could be due to the ionization fluctuation of particles in ICP, was observed compared to that of metal carbonyl <span class="hlt">gas</span> in Ar introduced directly into ICPMS. The linear regression lines were obtained and the limits of detection (LODs) of 1.5 pL L(-1) and 2.4 nL L(-1) for AsH3 and PH3, respectively, were estimated. Since these LODs revealed sufficiently lower values than the measurement concentrations required from semiconductor industry such as 0.5 nL L(-1) and 30 nL L(-1) for AsH3 and PH3, respectively, the GPD-GED-ICPMS could be useful for direct and high sensitive analysis of ultra-trace semiconductor <span class="hlt">gas</span> in <span class="hlt">air</span>. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ERL....13d5004S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ERL....13d5004S"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> at whole plant level shows that a less conservative <span class="hlt">water</span> use is linked to a higher performance in three ecologically distinct pine species</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Salazar-Tortosa, D.; Castro, J.; Rubio de Casas, R.; Viñegla, B.; Sánchez-Cañete, E. P.; Villar-Salvador, P.</p> <p>2018-04-01</p> <p>Increasing temperatures and decreasing precipitation in large areas of the planet as a consequence of global warming will affect plant growth and survival. However, the impact of climatic conditions will differ across species depending on their stomatal response to increasing aridity, as this will ultimately affect the balance between carbon assimilation and <span class="hlt">water</span> loss. In this study, we monitored <span class="hlt">gas</span> <span class="hlt">exchange</span>, growth and survival in saplings of three widely distributed European pine species (Pinus halepensis, P. nigra and P. sylvestris) with contrasting distribution and ecological requirements in order to ascertain the relationship between stomatal control and plant performance. The experiment was conducted in a common garden environment resembling rainfall and temperature conditions that two of the three species are expected to encounter in the near future. In addition, <span class="hlt">gas</span> <span class="hlt">exchange</span> was monitored both at the leaf and at the whole-plant level using a transient-state closed chamber, which allowed us to model the response of the whole plant to increased <span class="hlt">air</span> evaporative demand (AED). P. sylvestris was the species with lowest survival and performance. By contrast, P. halepensis showed no mortality, much higher growth (two orders of magnitude), carbon assimilation (ca. 14 fold higher) and stomatal conductance and <span class="hlt">water</span> transpiration (ca. 4 fold higher) than the other two species. As a consequence, P. halepensis exhibited higher values of <span class="hlt">water</span>-use efficiency than the rest of the species even at the highest values of AED. Overall, the results strongly support that the weaker stomatal control of P. halepensis, which is linked to lower stem <span class="hlt">water</span> potential, enabled this species to maximize carbon uptake under drought stress and ultimately outperform the more <span class="hlt">water</span> conservative P. nigra and P. sylvestris. These results suggest that under a hotter drought scenario P. nigra and P. sylvestris would very likely suffer increased mortality, whereas P. halepensis could maintain</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900054400&hterms=Cotton&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DCotton','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900054400&hterms=Cotton&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DCotton"><span>The relationship between leaf <span class="hlt">water</span> status, <span class="hlt">gas</span> <span class="hlt">exchange</span>, and spectral reflectance in cotton leaves</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bowman, William D.</p> <p>1989-01-01</p> <p>Measurements of leaf spectral reflectance, the components of <span class="hlt">water</span> potential, and leaf <span class="hlt">gas</span> <span class="hlt">exchanges</span> as a function of leaf <span class="hlt">water</span> content were made to evaluate the use of NIR reflectance as an indicator of plant <span class="hlt">water</span> status. Significant correlations were determined between spectral reflectance at 810 nm, 1665 nm, and 2210 nm and leaf relative <span class="hlt">water</span> content, total <span class="hlt">water</span> potential, and turgor pressure. However, the slopes of these relationships were relatively shallow and, when evaluated over the range of leaf <span class="hlt">water</span> contents in which physiological activity occurs (e.g., photosynthesis), had lower r-squared values, and some relationships were not statistically significant. NIR reflectance varied primarily as a function of leaf <span class="hlt">water</span> content, and not independently as a function of turgor pressure, which is a sensitive indicator of leaf <span class="hlt">water</span> status. The limitations of this approach to measuring plant <span class="hlt">water</span> stress are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/862648','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/862648"><span>Process for hydrogen isotope concentration between liquid <span class="hlt">water</span> and hydrogen <span class="hlt">gas</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Stevens, William H.</p> <p>1976-09-21</p> <p>A process for hydrogen isotope <span class="hlt">exchange</span> and concentration between liquid <span class="hlt">water</span> and hydrogen <span class="hlt">gas</span>, wherein liquid <span class="hlt">water</span> and hydrogen <span class="hlt">gas</span> are contacted, in an <span class="hlt">exchange</span> section, with one another and with at least one catalyst body comprising at least one metal selected from Group VIII of the Periodic Table and preferably a support therefor, the catalyst body has a liquid-<span class="hlt">water</span>-repellent, <span class="hlt">gas</span> permeable polymer or organic resin coating, preferably a fluorinated olefin polymer or silicone coating, so that the isotope concentration takes place by two simultaneously occurring steps, namely, ##EQU1## WHILE THE HYDROGEN <span class="hlt">GAS</span> FED TO THE <span class="hlt">EXCHANGE</span> SECTION IS DERIVED IN A REACTOR VESSEL FROM LIQUID <span class="hlt">WATER</span> THAT HAS PASSED THROUGH THE <span class="hlt">EXCHANGE</span> SECTION.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMED13B0888G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMED13B0888G"><span>The <span class="hlt">AirWaterGas</span> Teacher Professional Development Program: Lessons Learned by Pairing Scientists and Teachers to Develop Curriculum on Global Climate Change and Regional Unconventional Oil and <span class="hlt">Gas</span> Development</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gardiner, L. S.; Hatheway, B.; Rogers, J. D.; Casey, J. G.; Lackey, G.; Birdsell, D.; Brown, K.; Polmear, M.; Capps, S.; Rosenblum, J.; Sitterley, K.; Hafich, K. A.; Hannigan, M.; Knight, D.</p> <p>2015-12-01</p> <p>The <span class="hlt">AirWaterGas</span> Teacher Professional Development Program, run by the UCAR Center for Science Education, brought together scientists and secondary science teachers in a yearlong program culminating in the development of curriculum related to the impacts of unconventional oil and <span class="hlt">gas</span> development. Graduate students and research scientists taught about their research area and its relationship to oil and <span class="hlt">gas</span> throughout three online courses during the 2015-16 school year, during which teachers and scientists engaged in active online discussions. Topics covered included climate change, oil and <span class="hlt">gas</span> infrastructure, <span class="hlt">air</span> quality, <span class="hlt">water</span> quality, public health, and practices and policies relating to oil and <span class="hlt">gas</span> development. Building upon their initial online interactions and a face-to-face meeting in March, teachers were paired with appropriate <span class="hlt">AirWaterGas</span> team members as science advisors during a month-long residency in Boulder, Colorado. During the residency, graduate student scientists provided resources and feedback as teachers developed curriculum projects in collaboration with each other and UCAR science educators. Additionally, teachers and <span class="hlt">AirWaterGas</span> researchers shared experiences on an oil and <span class="hlt">gas</span> well site tour, and a short course on drilling methods with a drilling rig simulator. Here, we share lessons learned from both sides of the aisle, including initial results from program assessment conducted with the participating teachers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16545998','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16545998"><span>Foliar trichomes, boundary layers, and <span class="hlt">gas</span> <span class="hlt">exchange</span> in 12 species of epiphytic Tillandsia (Bromeliaceae).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Benz, Brett W; Martin, Craig E</p> <p>2006-04-01</p> <p>We examined the relationships between H2O and CO2 <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters and leaf trichome cover in 12 species of Tillandsia that exhibit a wide range in trichome size and trichome cover. Previous investigations have hypothesized that trichomes function to enhance boundary layers around Tillandsioid leaves thereby buffering the evaporative demand of the atmosphere and retarding transpirational <span class="hlt">water</span> loss. Data presented herein suggest that trichome-enhanced boundary layers have negligible effects on Tillandsia <span class="hlt">gas</span> <span class="hlt">exchange</span>, as indicated by the lack of statistically significant relationships in regression analyses of <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters and trichome cover. We calculated trichome and leaf boundary layer components, and their associated effects on H2O and CO2 <span class="hlt">gas</span> <span class="hlt">exchange</span>. The results further indicate trichome-enhanced boundary layers do not significantly reduce transpirational <span class="hlt">water</span> loss. We conclude that although the trichomes undoubtedly increase the thickness of the boundary layer, the increase due to Tillandsioid trichomes is inconsequential in terms of whole leaf boundary layers, and any associated reduction in transpirational <span class="hlt">water</span> loss is also negligible within the whole plant <span class="hlt">gas</span> <span class="hlt">exchange</span> pathway.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AtmEn.178...31J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AtmEn.178...31J"><span>Seasonal atmospheric deposition and <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> of polycyclic aromatic hydrocarbons over the Yangtze River Estuary, East China Sea: Implications for source-sink processes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jiang, Yuqing; Lin, Tian; Wu, Zilan; Li, Yuanyuan; Li, Zhongxia; Guo, Zhigang; Yao, Xiaohong</p> <p>2018-04-01</p> <p>In this work, <span class="hlt">air</span> samples and surface seawater samples covering four seasons from March 2014 to January 2015 were collected from a background receptor site in the YRE to explore the seasonal fluxes of <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> and dry and wet deposition of 15 polycyclic aromatic hydrocarbons (PAHs) and their source-sink processes at the <span class="hlt">air</span>-sea interface. The average dry and wet deposition fluxes of 15 PAHs were estimated as 879 ± 1393 ng m-2 d-1 and 755 ± 545 ng m-2 d-1, respectively. Gaseous PAH release from seawater to the atmosphere averaged 3114 ± 1999 ng m-2 d-1 in a year round. The <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> of PAHs was the dominant process at the <span class="hlt">air</span>-sea interface in the YRE as the magnitude of volatilization flux of PAHs exceeded that of total dry and wet deposition. The <span class="hlt">gas</span> PAH <span class="hlt">exchange</span> flux was dominated by three-ring PAHs, with the highest value in summer and lowest in winter, indicating a marked seasonal variation owing to differences in Henry's law constants associated with temperature, as well as wind speed and gaseous-dissolved gradient among seasons. Based on the simplified mass balance estimation, a net 11 tons y-1 of PAHs (mainly three-ring PAHs) were volatilized from seawater to the atmosphere in a ∼20,000 km2 area in the YRE. Other than the year-round Yangtze River input and ocean ship emissions, the selective release of low-molecular-weight PAHs from bottom sediments in winter due to resuspension triggered by the East Asian winter monsoon is another potential source of PAHs. This work suggests that the source-sink processes of PAHs at the <span class="hlt">air</span>-sea interface in the YRE play a crucial role in regional cycling of PAHs.</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/2008AGUFMOS31B1259A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFMOS31B1259A"><span>The Relationship Between Temperature and <span class="hlt">Gas</span> Concentration Fluctuation Rates at an <span class="hlt">Air-Water</span> Interface</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Asher, W. E.; Jessup, A. T.; Liang, H.; Zappa, C. J.</p> <p>2008-12-01</p> <p>The <span class="hlt">air</span>-sea flux, F, of a sparingly soluble nonreactive <span class="hlt">gas</span> can be expressed as F = kG(CS-CW), where kG is the <span class="hlt">gas</span> transfer velocity, CS is the concentration of <span class="hlt">gas</span> that would be expected in the <span class="hlt">water</span> if the system were in Henry's <span class="hlt">Gas</span> Law equilibrium, and CW is the <span class="hlt">gas</span> concentration in the bulk <span class="hlt">water</span>. An analogous relationship for the net heat flux can also be written using the heat transfer velocity, kH, and the bulk-skin temperature difference in the aqueous phase. Surface divergence theory for the <span class="hlt">air-water</span> transfer of <span class="hlt">gas</span> and heat predicts that kG and kH will scale as the square root of the surface divergence rate, r. However, because of the interaction between diffusivity and the scale depth of the surface divergences, the scale factor for heat is likely to be different from the scale factor for gases. Infrared imagery was used to measure the timescales of variations in temperature at a <span class="hlt">water</span> surface and laser-induced fluorescence (LIF) was used to measure temporal fluctuations in aqueous-phase concentrations of carbon dioxide (CO2) at a <span class="hlt">water</span> surface. The rate at which these temperature and concentration fluctuations occur is then assumed to be related to r. The divergence rates derived for temperature from the IR images can be compared to the rates for <span class="hlt">gas</span> derived from the LIF measurements to understand how r estimated from the two measurements differ. The square root of r is compared to concurrently measured kG for helium and sulfur hexafluoride to test the assumption that r1/2 scales with kG. Additionally, we measured kH using the active controlled flux technique, and those heat transfer velocities can also be used to test for a r1/2 dependence. All measurements reported here were made in the APL-UW synthetic jet array facility.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/351688-sensing-flux-volatile-chemicals-through-air-water-interface','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/351688-sensing-flux-volatile-chemicals-through-air-water-interface"><span>Sensing the flux of volatile chemicals through the <span class="hlt">air-water</span> interface</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>Mackay, D.; Schroeder, W.H.; Ooijen, H. von</p> <p>1997-12-31</p> <p>There are several situations in which there is a need to assess the direction and magnitude of the flux across the <span class="hlt">air-water</span> interface. Contaminants may be evaporating or absorbing in wastewater treatment systems in natural lake, river, estuarine and marine systems, and any attempt to compile a mass balance must include this process. In this study the authors review the theory underlying <span class="hlt">air-water</span> <span class="hlt">exchange</span>, then describe and discuss a sparging approach by which the direction and magnitude of the flux can be ascertained. The principle of the method is that a known flow rate of <span class="hlt">air</span> is bubbled through themore » sparger and allowed to equilibrate with the <span class="hlt">water</span>. The <span class="hlt">gas</span> exiting the <span class="hlt">water</span> surface is passed through a sorbent trap and later analyzed. The concentration, and hence the fugacity, of the contaminant in the sparged <span class="hlt">air</span> can be deduced. In parallel, a similar flow of <span class="hlt">air</span> from the atmosphere above the <span class="hlt">water</span> is drawn through another sparger at a similar flow rate for a similar time and the trapped chemical analyzed giving the concentration and fugacity in the <span class="hlt">air</span>. These data show the direction of <span class="hlt">air-water</span> <span class="hlt">exchange</span> (i.e. from high to low fugacity) and with information on the mass transfer coefficients and area, the flux. Successful tests were conducted of the system in a laboratory tank, in Lake Ontario and in Hamilton Harbour. Analyses of the traps showed a large number of peaks on the chromatogram many of which are believed to be of petroleum origin from fuels and vessel exhaust. The system will perform best under conditions where concentrations of specific contaminants are large, as occurs in waste <span class="hlt">water</span> treatment systems. The approach has the potential to contribute to more accurate assessment of <span class="hlt">air-water</span> fluxes. It avoids the problems of different analytical methodologies and the effect of sorption in the <span class="hlt">water</span> column.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFD.E4003J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFD.E4003J"><span>Respiratory Mechanics and <span class="hlt">Gas</span> <span class="hlt">Exchange</span>: The Effect of Surfactants</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jbaily, Abdulrahman; Szeri, Andrew J.</p> <p>2017-11-01</p> <p>The purpose of the lung is to <span class="hlt">exchange</span> gases, primarily oxygen and carbon dioxide, between the atmosphere and the circulatory system. To enable this <span class="hlt">exchange</span>, the airways in the lungs terminate in some 300 million alveoli that provide adequate surface area for transport. During breathing, work must be done to stretch various tissues to accommodate a greater volume of <span class="hlt">gas</span>. Considerable work must also be done to expand the liquid lining (hypophase) that coats the interior surfaces of the alveoli. This is enabled by a surface active lipo-protein complex, known as pulmonary surfactant, that modifies the surface tension at the hypophase-<span class="hlt">air</span> interface. Surfactants also serve as physical barriers that modify the rate of <span class="hlt">gas</span> transfer across interfaces. We develop a mathematical model to study the action of pulmonary surfactant and its determinative contributions to breathing. The model is used to explore the influence of surfactants on alveolar mechanics and on <span class="hlt">gas</span> <span class="hlt">exchange</span>: it relates the work of respiration at the level of the alveolus to the <span class="hlt">gas</span> <span class="hlt">exchange</span> rate through the changing influence of pulmonary surfactant over the breathing cycle. This work is motivated by a need to develop improved surfactant replacement therapies to treat serious medical conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/55311','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/55311"><span>Stomatal kinetics and photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> along a continuum of isohydric to anisohydric regulation of plant <span class="hlt">water</span> status</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Frederick C. Meinzer; Duncan D. Smith; David R. Woodruff; Danielle E. Marias; Katherine A. McCulloh; Ava R. Howard; Alicia L. Magedman</p> <p>2017-01-01</p> <p>Species’ differences in the stringency of stomatal control of plant <span class="hlt">water</span> potential represent a continuum of isohydric to anisohydric behaviours. However, little is known about how quasi-steady-state stomatal regulation of <span class="hlt">water</span> potential may relate to dynamic behaviour of stomata and photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> in species operating at different positions along this...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007GBioC..21.2015S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007GBioC..21.2015S"><span>Constraining global <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> for CO2 with recent bomb 14C measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sweeney, Colm; Gloor, Emanuel; Jacobson, Andrew R.; Key, Robert M.; McKinley, Galen; Sarmiento, Jorge L.; Wanninkhof, Rik</p> <p>2007-06-01</p> <p>The 14CO2 released into the stratosphere during bomb testing in the early 1960s provides a global constraint on <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> of soluble atmospheric gases like CO2. Using the most complete database of dissolved inorganic radiocarbon, DI14C, available to date and a suite of ocean general circulation models in an inverse mode we recalculate the ocean inventory of bomb-produced DI14C in the global ocean and confirm that there is a 25% decrease from previous estimates using older DI14C data sets. Additionally, we find a 33% lower globally averaged <span class="hlt">gas</span> transfer velocity for CO2 compared to previous estimates (Wanninkhof, 1992) using the NCEP/NCAR Reanalysis 1 1954-2000 where the global mean winds are 6.9 m s-1. Unlike some earlier ocean radiocarbon studies, the implied <span class="hlt">gas</span> transfer velocity finally closes the gap between small-scale deliberate tracer studies and global-scale estimates. Additionally, the total inventory of bomb-produced radiocarbon in the ocean is now in agreement with global budgets based on radiocarbon measurements made in the stratosphere and troposphere. Using the implied relationship between wind speed and <span class="hlt">gas</span> transfer velocity ks = 0.27<u102>(Sc/660)-0.5 and standard partial pressure difference climatology of CO2 we obtain an net <span class="hlt">air</span>-sea flux estimate of 1.3 ± 0.5 PgCyr-1 for 1995. After accounting for the carbon transferred from rivers to the deep ocean, our estimate of oceanic uptake (1.8 ± 0.5 PgCyr-1) compares well with estimates based on ocean inventories, ocean transport inversions using ocean concentration data, and model simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17664033','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17664033"><span>Carbon source/sink function of a subtropical, eutrophic lake determined from an overall mass balance and a <span class="hlt">gas</span> <span class="hlt">exchange</span> and carbon burial balance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Hong; Xing, Yangping; Xie, Ping; Ni, Leyi; Rong, Kewen</p> <p>2008-02-01</p> <p>Although studies on carbon burial in lake sediments have shown that lakes are disproportionately important carbon sinks, many studies on gaseous carbon <span class="hlt">exchange</span> across the <span class="hlt">water-air</span> interface have demonstrated that lakes are supersaturated with CO(2) and CH(4) causing a net release of CO(2) and CH(4) to the atmosphere. In order to more accurately estimate the net carbon source/sink function of lake ecosystems, a more comprehensive carbon budget is needed, especially for gaseous carbon <span class="hlt">exchange</span> across the <span class="hlt">water-air</span> interface. Using two methods, overall mass balance and <span class="hlt">gas</span> <span class="hlt">exchange</span> and carbon burial balance, we assessed the carbon source/sink function of Lake Donghu, a subtropical, eutrophic lake, from April 2003 to March 2004. With the overall mass balance calculations, total carbon input was 14 905 t, total carbon output was 4950 t, and net carbon budget was +9955 t, suggesting that Lake Donghu was a great carbon sink. For the <span class="hlt">gas</span> <span class="hlt">exchange</span> and carbon burial balance, gaseous carbon (CO(2) and CH(4)) emission across the <span class="hlt">water-air</span> interface totaled 752 t while carbon burial in the lake sediment was 9477 t. The ratio of carbon emission into the atmosphere to carbon burial into the sediment was only 0.08. This low ratio indicates that Lake Donghu is a great carbon sink. Results showed good agreement between the two methods with both showing Lake Donghu to be a great carbon sink. This results from the high primary production of Lake Donghu, substantive allochthonous carbon inputs and intensive anthropogenic activity. Gaseous carbon emission accounted for about 15% of the total carbon output, indicating that the total output would be underestimated without including gaseous carbon <span class="hlt">exchange</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25813755','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25813755"><span><span class="hlt">Water</span> relations and <span class="hlt">gas</span> <span class="hlt">exchange</span> of fan bryophytes and their adaptations to microhabitats in an Asian subtropical montane cloud forest.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Song, Liang; Zhang, Yong-Jiang; Chen, Xi; Li, Su; Lu, Hua-Zheng; Wu, Chuan-Sheng; Tan, Zheng-Hong; Liu, Wen-Yao; Shi, Xian-Meng</p> <p>2015-07-01</p> <p>Fan life forms are bryophytes with shoots rising from vertical substratum that branch repeatedly in the horizontal plane to form flattened photosynthetic surfaces, which are well suited for intercepting <span class="hlt">water</span> from moving <span class="hlt">air</span>. However, detailed <span class="hlt">water</span> relations, <span class="hlt">gas</span> <span class="hlt">exchange</span> characteristics of fan bryophytes and their adaptations to particular microhabitats remain poorly understood. In this study, we measured and analyzed microclimatic data, as well as <span class="hlt">water</span> release curves, pressure-volume relationships and photosynthetic <span class="hlt">water</span> and light response curves for three common fan bryophytes in an Asian subtropical montane cloud forest (SMCF). Results demonstrate high relative humidity but low light levels and temperatures in the understory, and a strong effect of fog on <span class="hlt">water</span> availability for bryophytes in the SMCF. The facts that fan bryophytes in dry <span class="hlt">air</span> lose most of their free <span class="hlt">water</span> within 1 h, and a strong dependence of net photosynthesis rates on <span class="hlt">water</span> content, imply that the transition from a hydrated, photosynthetically active state to a dry, inactive state is rapid. In addition, fan bryophytes developed relatively high cell wall elasticity and the osmoregulatory capacity to tolerate desiccation. These fan bryophytes had low light saturation and compensation point of photosynthesis, indicating shade tolerance. It is likely that fan bryophytes can flourish on tree trunks in the SMCF because of substantial annual precipitation, average relative humidity, and frequent and persistent fog, which can provide continual <span class="hlt">water</span> sources for them to intercept. Nevertheless, the low <span class="hlt">water</span> retention capacity and strong dependence of net photosynthesis on <span class="hlt">water</span> content of fan bryophytes indicate a high risk of unbalanced carbon budget if the frequency and severity of drought increase in the future as predicted.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002GMS...127..141S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002GMS...127..141S"><span>A model of <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> incorporating the physics of the turbulent boundary layer and the properties of the sea surface</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Soloviev, Alexander; Schluessel, Peter</p> <p></p> <p>The model presented contains interfacial, bubble-mediated, ocean mixed layer, and remote sensing components. The interfacial (direct) <span class="hlt">gas</span> transfer dominates under conditions of low and—for quite soluble gases like CO2—moderate wind speeds. Due to the similarity between the <span class="hlt">gas</span> and heat transfer, the temperature difference, ΔT, across the thermal molecular boundary layer (cool skin of the ocean) and the interfacial <span class="hlt">gas</span> transfer coefficient, Kint are presumably interrelated. A coupled parameterization for ΔT and Kint has been derived in the context of a surface renewal model [Soloviev and Schluessel, 1994]. In addition to the Schmidt, Sc, and Prandtl, Pr, numbers, the important parameters are the surface Richardson number, Rƒ0, and the Keulegan number, Ke. The more readily available cool skin data are used to determine the coefficients that enter into both parameterizations. At high wind speeds, the Ke-number dependence is further verified with the formula for transformation of the surface wind stress to form drag and white capping, which follows from the renewal model. A further extension of the renewal model includes effects of solar radiation and rainfall. The bubble-mediated component incorporates the Merlivat et al. [1993] parameterization with the empirical coefficients estimated by Asher and Wanninkhof [1998]. The oceanic mixed layer component accounts for stratification effects on the <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span>. Based on the example of <span class="hlt">Gas</span>Ex-98, we demonstrate how the results of parameterization and modeling of the <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> can be extended to the global scale, using remote sensing techniques.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120017344','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120017344"><span>Prototype Vent <span class="hlt">Gas</span> Heat <span class="hlt">Exchanger</span> for Exploration EVA - Performance and Manufacturing Characteristics</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jennings, Mallory; Quinn, Gregory; Strange, Jeremy</p> <p>2012-01-01</p> <p>NASA is developing new portable life support system (PLSS) technologies, which it is demonstrating in an unmanned ground based prototype unit called PLSS 2.0. One set of technologies within the PLSS provides suitable ventilation to an astronaut while on an EVA. A new component within the ventilation <span class="hlt">gas</span> loop is a liquid-to-<span class="hlt">gas</span> heat <span class="hlt">exchanger</span> to transfer excess heat from the <span class="hlt">gas</span> to the thermal control system's liquid coolant loop. A unique bench top prototype heat <span class="hlt">exchanger</span> was built and tested for use in PLSS 2.0. The heat <span class="hlt">exchanger</span> was designed as a counter-flow, compact plate fin type using stainless steel. Its design was based on previous compact heat <span class="hlt">exchangers</span> manufactured by United Technologies Aerospace Systems, but was half the size of any previous heat <span class="hlt">exchanger</span> model and one third the size of previous liquid-to-<span class="hlt">gas</span> heat <span class="hlt">exchangers</span>. The prototype heat <span class="hlt">exchanger</span> was less than 40 cubic inches and weighed 2.6 lb. The <span class="hlt">water</span> side and <span class="hlt">gas</span> side pressure drops were 0.8 psid and 0.5 inches of <span class="hlt">water</span>, respectively. Performance of the heat <span class="hlt">exchanger</span> at the nominal pressure of 4.1 psia was measured at 94%, while a <span class="hlt">gas</span> inlet pressure of 25 psia resulted in an effectiveness of 84%. These results compared well with the model, which was scaled for the small size. Modeling of certain phenomena that affect performance, such as flow distribution in the headers was particularly difficult due to the small size of the heat <span class="hlt">exchanger</span>. Data from the tests has confirmed the correction factors that were used in these parts of the model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhCS1015c2072S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhCS1015c2072S"><span>Modeling of Hydrate Formation Mode in Raw Natural <span class="hlt">Gas</span> <span class="hlt">Air</span> Coolers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Scherbinin, S. V.; Prakhova, M. Yu; Krasnov, A. N.; Khoroshavina, E. A.</p> <p>2018-05-01</p> <p><span class="hlt">Air</span> cooling units (ACU) are used at all the <span class="hlt">gas</span> fields for cooling natural <span class="hlt">gas</span> after compressing. When using ACUs on raw (wet) <span class="hlt">gas</span> in a low temperature condition, there is a danger of hydrate plug formation in the heat <span class="hlt">exchanging</span> tubes of the ACU. To predict possible hydrate formation, a mathematical model of the <span class="hlt">air</span> cooler thermal behavior used in the control system shall adequately calculate not only <span class="hlt">gas</span> temperature at the cooler's outlet, but also a dew point value, a temperature at which condensation, as well as the <span class="hlt">gas</span> hydrate formation point, onsets. This paper proposes a mathematical model allowing one to determine the pressure in the <span class="hlt">air</span> cooler which makes hydrate formation for a given <span class="hlt">gas</span> composition possible.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008OSJ....43...17L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008OSJ....43...17L"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> rates measured using a dual-tracer (SF6 and3he) method in the coastal <span class="hlt">waters</span> of Korea</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Hyun-Woo; Lee, Kitack; Kaown, Duk-In</p> <p>2008-03-01</p> <p>Over a period of 5 days between August 12 and 17, 2005, we performed a <span class="hlt">gas</span> <span class="hlt">exchange</span> experiment using the dual tracer method in a tidal coastal ocean located off the southern coast of Korea. The <span class="hlt">gas</span> <span class="hlt">exchange</span> rate was determined from temporal changes in the ratio of3He to SF6 measured daily in the surface mixed layer. The measured <span class="hlt">gas</span> <span class="hlt">exchange</span> rate ( k CO 2), normalized to a Schmidt number of 600 for CO2 in fresh <span class="hlt">water</span> at 20°C, was approximately 5.0 cm h-1 at a mean wind speed of 3.9 m s-1 during the study period. This value is significantly less than those obtained from floating chamber-based experiments performed previously in estuarine environments, but is similar in magnitude to values obtained using the dual tracer method in river and tidal coastal <span class="hlt">waters</span> and values predicted on the basis of the relationship between the <span class="hlt">gas</span> <span class="hlt">exchange</span> rate and wind speed (Wanninkhof 1992), which is generally applicable to the open ocean. Our result is also consistent with the relationship of Raymond and Cole (2001), which was derived from experiments carried out in estuarine environments using222Rn and chlorofluorocarbons along with measurements undertaken in the Hudson River, Canada, using SF6 and3He. Our results indicate that tidal action in a microtidal region did not discernibly enhance the measured k CO 2 value.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010004231','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010004231"><span>BOREAS TE-10 Leaf <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hall, Forrest G. (Editor); Papagno, Andrea (Editor); Middleton, Elizabeth; Sullivan, Joseph</p> <p>2000-01-01</p> <p>The Boreal Ecosystem-Atmospheric Study (BOREAS) TE-10 (Terrestrial Ecology) team collected several data sets in support of its efforts to characterize and interpret information on the reflectance, transmittance, <span class="hlt">gas</span> <span class="hlt">exchange</span>, chlorophyll content, carbon content, hydrogen content, and nitrogen content of boreal vegetation. This data set contains measurements of assimilation, stomatal conductance, transpiration, internal CO2 concentration, and <span class="hlt">water</span> use efficiency conducted in the Southern Study Area (SSA) during the growing seasons of 1994 and 1996 using a portable <span class="hlt">gas</span> <span class="hlt">exchange</span> system. The data are stored in tabular ASCII files. The data files are available on a CD-ROM (see document number 20010000884), or from the Oak Ridge National Laboratory (ORNL) Distributed Active Archive Center (DAAC).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23844085','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23844085"><span>Plant <span class="hlt">water</span> use efficiency over geological time--evolution of leaf stomata configurations affecting plant <span class="hlt">gas</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Assouline, Shmuel; Or, Dani</p> <p>2013-01-01</p> <p>Plant <span class="hlt">gas</span> <span class="hlt">exchange</span> is a key process shaping global hydrological and carbon cycles and is often characterized by plant <span class="hlt">water</span> use efficiency (WUE - the ratio of CO2 gain to <span class="hlt">water</span> vapor loss). Plant fossil record suggests that plant adaptation to changing atmospheric CO2 involved correlated evolution of stomata density (d) and size (s), and related maximal aperture, amax . We interpreted the fossil record of s and d correlated evolution during the Phanerozoic to quantify impacts on <span class="hlt">gas</span> conductance affecting plant transpiration, E, and CO2 uptake, A, independently, and consequently, on plant WUE. A shift in stomata configuration from large s-low d to small s-high d in response to decreasing atmospheric CO2 resulted in large changes in plant <span class="hlt">gas</span> <span class="hlt">exchange</span> characteristics. The relationships between <span class="hlt">gas</span> conductance, gws , A and E and maximal relative transpiring leaf area, (amax ⋅d), exhibited hysteretic-like behavior. The new WUE trend derived from independent estimates of A and E differs from established WUE-CO2 trends for atmospheric CO2 concentrations exceeding 1,200 ppm. In contrast with a nearly-linear decrease in WUE with decreasing CO2 obtained by standard methods, the newly estimated WUE trend exhibits remarkably stable values for an extended geologic period during which atmospheric CO2 dropped from 3,500 to 1,200 ppm. Pending additional tests, the findings may affect projected impacts of increased atmospheric CO2 on components of the global hydrological cycle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA282842','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA282842"><span>Oceanic Whitecaps and Associated, Bubble-Mediated, <span class="hlt">Air</span>-Sea <span class="hlt">Exchange</span> Processes</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1992-10-01</p> <p>experiments performed in laboratory conditions using <span class="hlt">Air</span>-Sea <span class="hlt">Exchange</span> Monitoring System (A-SEMS). EXPERIMENTAL SET-UP In a first look, the <span class="hlt">Air</span>-Sea <span class="hlt">Exchange</span>...Model 225, equipped with a Model 519 plug-in module. Other complementary information on A-SEMS along with results from first tests and calibration...between 9.50C and 22.40C within the first 24 hours after transferring the <span class="hlt">water</span> sample into laboratory conditions. The results show an enhancement of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25232199','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25232199"><span>Nesting behaviour influences species-specific <span class="hlt">gas</span> <span class="hlt">exchange</span> across avian eggshells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Portugal, Steven J; Maurer, Golo; Thomas, Gavin H; Hauber, Mark E; Grim, Tomáš; Cassey, Phillip</p> <p>2014-09-15</p> <p> achieve optimal <span class="hlt">water</span> loss during incubation. We also suggest that eggs laid in cup nests and burrows may require a higher G(H2O) to overcome the increased humidity as a result from the confined nest microclimate lacking <span class="hlt">air</span> movements through the nest. Taken together, these comparative data imply that species-specific levels of <span class="hlt">gas</span> <span class="hlt">exchange</span> across avian eggshells are variable and evolve in response to ecological and physical variation resulting from parental and nesting behaviours. © 2014. Published by The Company of Biologists Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=286462','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=286462"><span>Effects of permethrin and amitraz on <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> loss in unfed adult females of Amblyomma americanum (Acari: Ixodidae)</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>Effects of permethrin and amitraz on metabolism of the lone star tick, Amblyomma americanum, were examined using a flow-through carbon dioxide (CO2) and <span class="hlt">water</span> vapor analyzer. Untreated adult female ticks exhibited a distinct discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> pattern (DGEP) with no measurable <span class="hlt">water</span> loss. Si...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17938120','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17938120"><span>Leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> of understory spruce-fir saplings in relict cloud forests, southern Appalachian Mountains, USA.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Reinhardt, Keith; Smith, William K</p> <p>2008-01-01</p> <p>The southern Appalachian spruce-fir (Picea rubens Sarg. and Abies fraseri (Pursh) Poir.) forest is found only on high altitude mountain tops that receive copious precipitation ( > 2000 mm year(-1)) and experience frequent cloud immersion. These high-elevation, temperate rain forests are immersed in clouds on approximately 65% of the total growth season days and for 30-40% of a typical summer day, and cloud deposition accounts for up to 50% of their annual <span class="hlt">water</span> budget. We investigated environmental influences on understory leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> relations at two sites: Mt. Mitchell, NC (MM; 35 degrees 45'53'' N, 82 degrees 15'53'' W, 2028 m elevation) and Whitetop Mtn., VA (WT; 36 degrees 38'19'' N, 81 degrees 36'19'' W, 1685 m elevation). We hypothesized that the cool, moist and cloudy conditions at these sites exert a strong influence on leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>. Maximum photosynthesis (A(max)) varied between 1.6 and 4.0 micromol CO(2) m(-2) s(-1) for both spruce and fir and saturated at irradiances between approximately 200 and 400 micromol m(-2) s(-1) at both sites. Leaf conductance (g) ranged between 0.05 and 0.25 mol m(-2) s(-1) at MM and between 0.15 and 0.40 mol m(-2) s(-1) at WT and was strongly associated with leaf-to-<span class="hlt">air</span> vapor pressure difference (LAVD). At both sites, g decreased exponentially as LAVD increased, with an 80-90% reduction in g between 0 and 0.5 kPa. Predawn leaf <span class="hlt">water</span> potentials remained between -0.25 and -0.5 MPa for the entire summer, whereas late afternoon values declined to between -1.25 and -1.75 MPa by late summer. Thus, leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> appeared tightly coupled to the response of g to LAVD, which maintained high <span class="hlt">water</span> status, even at the relatively low LAVD of these cloud forests. Moreover, the cloudy, humid environment of these refugial forests appears to exert a strong influence on tree leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> relations. Because global climate change is predicted to increase regional cloud ceiling levels, more research on</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AtmEn.147..200O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AtmEn.147..200O"><span>Determination of temperature dependent Henry's law constants of polychlorinated naphthalenes: Application to <span class="hlt">air</span>-sea <span class="hlt">exchange</span> in Izmir Bay, Turkey</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Odabasi, Mustafa; Adali, Mutlu</p> <p>2016-12-01</p> <p>The Henry's law constant (H) is a crucial variable to investigate the <span class="hlt">air-water</span> <span class="hlt">exchange</span> of persistent organic pollutants. H values for 32 polychlorinated naphthalene (PCN) congeners were measured using an inert <span class="hlt">gas</span>-stripping technique at five temperatures ranging between 5 and 35 °C. H values in deionized <span class="hlt">water</span> (at 25 °C) varied between 0.28 ± 0.08 Pa m3 mol-1 (PCN-73) and 18.01 ± 0.69 Pa m3 mol-1 (PCN-42). The agreement between the measured and estimated H values from the octanol-<span class="hlt">water</span> and octanol-<span class="hlt">air</span> partition coefficients was good (measured/estimated ratio = 1.00 ± 0.41, average ± SD). The calculated phase change enthalpies (ΔHH) were within the interval previously determined for other several semivolatile organic compounds (42.0-106.4 kJ mol-1). Measured H values, paired atmospheric and aqueous concentrations and meteorological variables were also used to reveal the level and direction of <span class="hlt">air</span>-sea <span class="hlt">exchange</span> fluxes of PCNs at the coast of Izmir Bay, Turkey. The net PCN <span class="hlt">air</span>-sea <span class="hlt">exchange</span> flux varied from -0.55 (volatilization, PCN-24/14) to 2.05 (deposition, PCN-23) ng m-2 day-1. PCN-19, PCN-24/14, PCN-42, and PCN-33/34/37 were mainly volatilized from seawater while the remaining congeners were mainly deposited. The overall number of the cases showing deposition was higher (67.9%) compared to volatilization (21.4%) and near equilibrium (10.7%).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?direntryid=315010&keyword=climate%20change&subject=climate%20change%20research&showcriteria=2&fed_org_id=111&datebeginpublishedpresented=03/10/2012&dateendpublishedpresented=03/10/2017&sortby=pubdateyear','PESTICIDES'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?direntryid=315010&keyword=climate%20change&subject=climate%20change%20research&showcriteria=2&fed_org_id=111&datebeginpublishedpresented=03/10/2012&dateendpublishedpresented=03/10/2017&sortby=pubdateyear"><span>A dynamic leaf <span class="hlt">gas-exchange</span> strategy is conserved in woody ...</span></a></p> <p><a target="_blank" href="http://www.epa.gov/pesticides/search.htm">EPA Pesticide Factsheets</a></p> <p></p> <p></p> <p>Rising atmospheric [CO2], ca, is expected to affect stomatal regulation of leaf <span class="hlt">gas-exchange</span> of woody plants, thus influencing energy fluxes as well as carbon (C), <span class="hlt">water</span> and nutrient cycling of forests. Researchers have reported that stomata regulate leaf <span class="hlt">gas-exchange</span> around “set points” that include a constant leaf internal [CO2], ci, a constant drawdown in CO2 (ca - ci), and a constant ci/ca. Because these set points can result in drastically different consequences for leaf <span class="hlt">gas-exchange</span>, it will be essential for the accuracy of Earth systems models that generalizable patterns in leaf <span class="hlt">gas-exchange</span> responses to ca be identified if any do exist. We hypothesized that the concept of optimal stomatal behavior, exemplified by woody plants shifting along a continuum of these set point strategies, would provide a unifying framework for understanding leaf <span class="hlt">gas-exchange</span> responses to ca. We analyzed studies reporting C stable isotope ratio (δ13C) or photosynthetic discrimination (∆13C) from woody plant taxa that grew across ca spanning at least 100 ppm for each species investigated. From these data we calculated ci, and in combination with known or estimated ca, leaf <span class="hlt">gas-exchange</span> regulation strategies were assessed. Overall, our analyses does not support the hypothesis that trees are canalized towards any of the proposed set points, particularly so for a constant ci. Rather, the results are consistent with the hypothesis that stomatal optimization regulates leaf <span class="hlt">gas</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/7785756','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/7785756"><span>[Heat and moisture <span class="hlt">exchangers</span> for conditioning of inspired <span class="hlt">air</span> of intubated patients in intensive care. The humidification properties of passive <span class="hlt">air</span> <span class="hlt">exchangers</span> under clinical conditions].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rathgeber, J; Züchner, K; Kietzmann, D; Weyland, W</p> <p>1995-04-01</p> <p>Heat and moisture <span class="hlt">exchangers</span> (HME) are used as artificial noses for intubated patients to prevent tracheo-bronchial or pulmonary damage resulting from dry and cold inspired gases. HME are mounted directly on the tracheal tube, where they collect a large fraction of the heat and moisture of the expired <span class="hlt">air</span>, adding this to the subsequent inspired breath. The effective performance depends on the <span class="hlt">water</span>-retention capacity of the HME: the amount of <span class="hlt">water</span> added to the inspired <span class="hlt">gas</span> cannot exceed the stored <span class="hlt">water</span> uptake of the previous breath. This study evaluates the efficiency of four different HME under laboratory and clinical conditions using a new moisture-measuring device. METHODS. In a first step, the absolute efficiency of four different HME (DAR Hygrobac, Gibeck Humid-Vent 2P, Pall BB 22-15 T, and Pall BB 100) was evaluated using a lung model simulating physiological heat and humidity conditions of the upper airways. The model was ventilated with tidal volumes of 500, 1,000, and 1,500 ml and different flow rates. The <span class="hlt">water</span> content of the ventilated <span class="hlt">air</span> was determined between tracheal tube and HME using a new high-resolution humidity meter and compared with the absolute <span class="hlt">water</span> loss of the exhaled <span class="hlt">air</span> at the <span class="hlt">gas</span> outlet of a Siemens Servo C ventilator measured with a dew-point hygrometer. Secondly, the moisturizing efficiency was evaluated under clinical conditions in an intensive care unit with 25 intubated patients. Maintaining the ventilatory conditions for each patient, the HME were randomly changed. The humidity data were determined as described above and compared with the laboratory findings. RESULTS AND DISCUSSION. The <span class="hlt">water</span> content at the respirator outlet is inversely equivalent to the humidity of the inspired gases and represents the <span class="hlt">water</span> loss from the respiratory tract if the patient is ventilated with dry gases. Moisture retention and heating capacity decreased with higher volumes and higher flow rates. These data are simple to obtain without affecting the</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('https://ntrs.nasa.gov/search.jsp?R=19850045092&hterms=water+gas+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dwater%2Bgas%2Bexchange','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850045092&hterms=water+gas+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dwater%2Bgas%2Bexchange"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span>-wind speed relation measured with sulfur hexafluoride on a lake</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wanninkhof, R.; Broecker, W. S.; Ledwell, J. R.</p> <p>1985-01-01</p> <p><span class="hlt">Gas-exchange</span> processes control the uptake and release of various gases in natural systems such as oceans, rivers, and lakes. Not much is known about the effect of wind speed on <span class="hlt">gas</span> <span class="hlt">exchange</span> in such systems. In the experiment described here, sulfur hexafluoride was dissolved in lake <span class="hlt">water</span>, and the rate of escape of the <span class="hlt">gas</span> with wind speed (at wind speeds up to 6 meters per second) was determined over a 1-month period. A sharp change in the wind speed dependence of the <span class="hlt">gas-exchange</span> coefficient was found at wind speeds of about 2.4 meters per second, in agreement with the results of wind-tunnel studies. However the <span class="hlt">gas-exchange</span> coefficients at wind speeds above 3 meters per second were smaller than those observed in wind tunnels and are in agreement with earlier lake and ocean results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ACP....17.9019B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ACP....17.9019B"><span>Estimation of bubble-mediated <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> from concurrent DMS and CO2 transfer velocities at intermediate-high wind speeds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bell, Thomas G.; Landwehr, Sebastian; Miller, Scott D.; de Bruyn, Warren J.; Callaghan, Adrian H.; Scanlon, Brian; Ward, Brian; Yang, Mingxi; Saltzman, Eric S.</p> <p>2017-07-01</p> <p>Simultaneous <span class="hlt">air</span>-sea fluxes and concentration differences of dimethylsulfide (DMS) and carbon dioxide (CO2) were measured during a summertime North Atlantic cruise in 2011. This data set reveals significant differences between the <span class="hlt">gas</span> transfer velocities of these two gases (Δkw) over a range of wind speeds up to 21 m s-1. These differences occur at and above the approximate wind speed threshold when waves begin breaking. Whitecap fraction (a proxy for bubbles) was also measured and has a positive relationship with Δkw, consistent with enhanced bubble-mediated transfer of the less soluble CO2 relative to that of the more soluble DMS. However, the correlation of Δkw with whitecap fraction is no stronger than with wind speed. Models used to estimate bubble-mediated transfer from in situ whitecap fraction underpredict the observations, particularly at intermediate wind speeds. Examining the differences between <span class="hlt">gas</span> transfer velocities of gases with different solubilities is a useful way to detect the impact of bubble-mediated <span class="hlt">exchange</span>. More simultaneous <span class="hlt">gas</span> transfer measurements of different solubility gases across a wide range of oceanic conditions are needed to understand the factors controlling the magnitude and scaling of bubble-mediated <span class="hlt">gas</span> <span class="hlt">exchange</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998WRR....34.3245B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998WRR....34.3245B"><span><span class="hlt">Air</span> sparging: <span class="hlt">Air-water</span> mass transfer coefficients</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Braida, Washington J.; Ong, Say Kee</p> <p>1998-12-01</p> <p>Experiments investigating the mass transfer of several dissolved volatile organic compounds (VOCs) across the <span class="hlt">air-water</span> interface were conducted using a single-<span class="hlt">air</span>- channel <span class="hlt">air</span>-sparging system. Three different porous media were used in the study. <span class="hlt">Air</span> velocities ranged from 0.2 cm s-1 to 2.5 cm s-1. The tortuosity factor for each porous medium and the <span class="hlt">air-water</span> mass transfer coefficients were estimated by fitting experimental data to a one-dimensional diffusion model. The estimated mass transfer coefficients KG ranged from 1.79 × 10-3 cm min-1 to 3.85 × 10-2 cm min-1. The estimated lumped <span class="hlt">gas</span> phase mass transfer coefficients KGa were found to be directly related to the <span class="hlt">air</span> diffusivity of the VOC, <span class="hlt">air</span> velocity, and particle size, and inversely related to the Henry's law constant of the VOCs. Of the four parameters investigated, the parameter that controlled or had a dominant effect on the lumped <span class="hlt">gas</span> phase mass transfer coefficient was the <span class="hlt">air</span> diffusivity of the VOC. Two empirical models were developed by correlating the Damkohler and the modified <span class="hlt">air</span> phase Sherwood numbers with the <span class="hlt">air</span> phase Peclet number, Henry's law constant, and the reduced mean particle size of porous media. The correlation developed in this study may be used to obtain better predictions of mass transfer fluxes for field conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DFDH36008C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDH36008C"><span>Surface nanobubble nucleation dynamics during <span class="hlt">water</span>-ethanol <span class="hlt">exchange</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chan, Chon U.; Ohl, Claus-Dieter</p> <p>2015-11-01</p> <p><span class="hlt">Water</span>-ethanol <span class="hlt">exchange</span> has been a promising nucleation method for surface attached nanobubbles since their discovery. In this process, <span class="hlt">water</span> and ethanol displace each other sequentially on a substrate. As the <span class="hlt">gas</span> solubility is 36 times higher in ethanol than <span class="hlt">water</span>, it was suggested that the <span class="hlt">exchange</span> process leads to transient supersaturation and is responsible for the nanobubble nucleation. In this work, we visualize the nucleation dynamics by controllably mixing <span class="hlt">water</span> and ethanol. It depicts the temporal evolution of the conventional <span class="hlt">exchange</span> in a single field of view, detailing the conditions for surface nanobubble nucleation and the flow field that influences their spatial organization. This technique can also pattern surface nanobubbles with variable size distribution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.1015G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.1015G"><span>Evaluation of the swell effect on the <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer in the coastal zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gutiérrez-Loza, Lucía; Ocampo-Torres, Francisco J.</p> <p>2016-04-01</p> <p><span class="hlt">Air</span>-sea <span class="hlt">gas</span> transfer processes are one of the most important factors regarding global climate and long-term global climate changes. Despite its importance, there is still a huge uncertainty on how to better parametrize these processes in order to include them on the global climate models. This uncertainty exposes the need to increase our knowledge on <span class="hlt">gas</span> transfer controlling mechanisms. In the coastal regions, breaking waves become a key factor to take into account when estimating <span class="hlt">gas</span> fluxes, however, there is still a lack of information and the influence of the ocean surface waves on the <span class="hlt">air</span>-sea interaction and <span class="hlt">gas</span> flux behavior must be validated. In this study, as part of the "Sea Surface Roughness as <span class="hlt">Air</span>-Sea Interaction Control" project, we evaluate the effect of the ocean surface waves on the <span class="hlt">gas</span> <span class="hlt">exchange</span> in the coastal zone. Direct estimates of the flux of CO2 (FCO2) and <span class="hlt">water</span> vapor (FH2O) through eddy covariance, were carried out from May 2014 to April 2015 in a coastal station located at the Northwest of Todos Santos Bay, Baja California, México. For the same period, ocean surface waves are recorded using an Acoustic Doppler Current Profiler (Workhorse Sentinel, Teledyne RD Instruments) with a sampling rate of 2 Hz and located at 10 m depth about 350 m away from the tower. We found the study area to be a weak sink of CO2 under moderate wind and wave conditions with a mean flux of -1.32 μmol/m2s. The correlation between the wind speed and FCO2 was found to be weak, suggesting that other physical processes besides wind may be important factors for the <span class="hlt">gas</span> <span class="hlt">exchange</span> modulation at coastal <span class="hlt">waters</span>. The results of the quantile regression analysis computed between FCO2 and (1) wind speed, (2) significant wave height, (3) wave steepness and (4) <span class="hlt">water</span> temperature, show that the significant wave height is the most correlated parameter with FCO2; Nevertheless, the behavior of their relation varies along the probability distribution of FCO2, with the linear regression</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70004693','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70004693"><span><span class="hlt">Exchange</span> of Groundwater and Surface-<span class="hlt">Water</span> Mediated by Permafrost Response to Seasonal and Long Term <span class="hlt">Air</span> Temperature Variation</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ge, Shemin; McKenzie, Jeffrey; Voss, Clifford; Wu, Qingbai</p> <p>2011-01-01</p> <p>Permafrost dynamics impact hydrologic cycle processes by promoting or impeding groundwater and surface <span class="hlt">water</span> <span class="hlt">exchange</span>. Under seasonal and decadal <span class="hlt">air</span> temperature variations, permafrost temperature changes control the <span class="hlt">exchanges</span> between groundwater and surface <span class="hlt">water</span>. A coupled heat transport and groundwater flow model, SUTRA, was modified to simulate groundwater flow and heat transport in the subsurface containing permafrost. The northern central Tibet Plateau was used as an example of model application. Modeling results show that in a yearly cycle, groundwater flow occurs in the active layer from May to October. Maximum groundwater discharge to the surface lags the maximum subsurface temperature by two months. Under an increasing <span class="hlt">air</span> temperature scenario of 3?C per 100 years, over the initial 40-year period, the active layer thickness can increase by three-fold. Annual groundwater discharge to the surface can experience a similar three-fold increase in the same period. An implication of these modeling results is that with increased warming there will be more groundwater flow in the active layer and therefore increased groundwater discharge to rivers. However, this finding only holds if sufficient upgradient <span class="hlt">water</span> is available to replenish the increased discharge. Otherwise, there will be an overall lowering of the <span class="hlt">water</span> table in the recharge portion of the catchment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70035273','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70035273"><span><span class="hlt">Exchange</span> of groundwater and surface-<span class="hlt">water</span> mediated by permafrost response to seasonal and long term <span class="hlt">air</span> temperature variation</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ge, S.; McKenzie, J.; Voss, C.; Wu, Q.</p> <p>2011-01-01</p> <p>Permafrost dynamics impact hydrologic cycle processes by promoting or impeding groundwater and surface <span class="hlt">water</span> <span class="hlt">exchange</span>. Under seasonal and decadal <span class="hlt">air</span> temperature variations, permafrost temperature changes control the <span class="hlt">exchanges</span> between groundwater and surface <span class="hlt">water</span>. A coupled heat transport and groundwater flow model, SUTRA, was modified to simulate groundwater flow and heat transport in the subsurface containing permafrost. The northern central Tibet Plateau was used as an example of model application. Modeling results show that in a yearly cycle, groundwater flow occurs in the active layer from May to October. Maximum groundwater discharge to the surface lags the maximum subsurface temperature by two months. Under an increasing <span class="hlt">air</span> temperature scenario of 3C per 100 years, over the initial 40-year period, the active layer thickness can increase by three-fold. Annual groundwater discharge to the surface can experience a similar three-fold increase in the same period. An implication of these modeling results is that with increased warming there will be more groundwater flow in the active layer and therefore increased groundwater discharge to rivers. However, this finding only holds if sufficient upgradient <span class="hlt">water</span> is available to replenish the increased discharge. Otherwise, there will be an overall lowering of the <span class="hlt">water</span> table in the recharge portion of the catchment. Copyright 2011 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22766042','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22766042"><span>Drought effect on growth, <span class="hlt">gas</span> <span class="hlt">exchange</span> and yield, in two strains of local barley Ardhaoui, under <span class="hlt">water</span> deficit conditions in southern Tunisia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thameur, Afwa; Lachiheb, Belgacem; Ferchichi, Ali</p> <p>2012-12-30</p> <p>Two local barley strains cv. Ardhaoui originated from Tlalit and Switir, sourthern Tunisia were grown in pots in a glasshouse assay, under well-<span class="hlt">watered</span> conditions for a month. Plants were then either subjected to <span class="hlt">water</span> deficit (treatment) or continually well-<span class="hlt">watered</span> (control). Control pots were irrigated several times each week to maintain soil moisture near field capacity (FC), while stress pots experienced soil drying by withholding irrigation until they reached 50% of FC. Variation in relative <span class="hlt">water</span> content, leaf area, leaf appearance rate and leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> (i.e. net CO(2) assimilation rate (A), transpiration (E), and stomatal conductance (gs)) in response to <span class="hlt">water</span> deficit was investigated. High leaf relative <span class="hlt">water</span> content (RWC) was maintained in Tlalit by stomatal closure and a reduction of leaf area. Reduction in leaf area was due to decline in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> during <span class="hlt">water</span> deficit. Tlalit was found to be drought tolerant and able to maintain higher leaf RWC under drought conditions. <span class="hlt">Water</span> deficit treatment reduced stomatal conductance by 43% at anthesis. High net CO(2) assimilation rate under <span class="hlt">water</span> deficit was associated with high RWC (r = 0.998; P < 0.01). Decline in net CO(2) assimilation rate was due mainly to stomatal closure. Significant differences between studied strains in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters were found, which can give some indications on the degree of drought tolerance. Thus, the ability of the low leaf area plants to maintain higher RWC could explain the differences in drought tolerance in studied barley strains. Results showed that Tlalit showed to be more efficient and more productive than Switir. Copyright © 2012 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.B32C..01H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.B32C..01H"><span>Small ponds play big role in greenhouse <span class="hlt">gas</span> emissions from inland <span class="hlt">waters</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Holgerson, M.; Raymond, P. A.</p> <p>2017-12-01</p> <p>Inland <span class="hlt">waters</span> are an important part of the global carbon cycle, but there is uncertainty in estimating their greenhouse <span class="hlt">gas</span> emissions. Uncertainty stems from different models and variable estimates of surface <span class="hlt">water</span> <span class="hlt">gas</span> concentrations, <span class="hlt">gas</span> <span class="hlt">exchange</span> rates, and the global size distribution of <span class="hlt">water</span> bodies. Emissions from small <span class="hlt">water</span> bodies are especially difficult to estimate because they are not globally mapped and few studies have assessed their greenhouse <span class="hlt">gas</span> concentrations and <span class="hlt">gas</span> <span class="hlt">exchange</span> rates. To overcome these limitations, we studied greenhouse gases and <span class="hlt">gas</span> <span class="hlt">exchange</span> rates in small ponds in temperate forests of the northeastern United States. We then compiled our data with direct measurements of CO2 and CH4 concentrations from 427 ponds and lakes worldwide, and upscaled to estimate greenhouse <span class="hlt">gas</span> emissions using estimates of <span class="hlt">gas</span> <span class="hlt">exchange</span> rates and the size distribution of lakes. We found that small ponds play a disproportionately large role in greenhouse <span class="hlt">gas</span> emissions. While small ponds only account for about 9% of global lakes and ponds by area, they contribute 15% of CO2 and 41% of diffusive CH4 emissions from inland freshwaters. Secondly, we measured <span class="hlt">gas</span> <span class="hlt">exchange</span> velocities (k) in small ponds and compiled direct measurements of k from 67 global <span class="hlt">water</span> bodies. We found that k is low but highly variable in small ponds, and increases and becomes even more variable with lake size, a finding that is not currently included in global carbon models. In a third study, we found that <span class="hlt">gas</span> <span class="hlt">exchange</span> in small ponds is highly sensitive to overnight cooling, which can lead to short bursts of increased k at night, with implications for greenhouse <span class="hlt">gas</span> emissions. Overall, these studies show that small ponds are a critical part of the global carbon cycle, and also highlight many knowledge gaps. Therefore, understanding small pond carbon cycling is an important research priority.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED259926.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED259926.pdf"><span>Heat Recovery Ventilation for Housing: <span class="hlt">Air-to-Air</span> Heat <span class="hlt">Exchangers</span>.</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>Corbett, Robert J.; Miller, Barbara</p> <p></p> <p>The <span class="hlt">air-to-air</span> heat <span class="hlt">exchanger</span> (a fan powered ventilation device that recovers heat from stale outgoing <span class="hlt">air</span>) is explained in this six-part publication. Topic areas addressed are: (1) the nature of <span class="hlt">air-to-air</span> heat <span class="hlt">exchangers</span> and how they work; (2) choosing and sizing the system; (3) installation, control, and maintenance of the system; (4) heat…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26020102','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26020102"><span><span class="hlt">Gas</span> Transfer in Cellularized Collagen-Membrane <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Devices.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lo, Justin H; Bassett, Erik K; Penson, Elliot J N; Hoganson, David M; Vacanti, Joseph P</p> <p>2015-08-01</p> <p>Chronic lower respiratory disease is highly prevalent in the United States, and there remains a need for alternatives to lung transplant for patients who progress to end-stage lung disease. Portable or implantable <span class="hlt">gas</span> oxygenators based on microfluidic technologies can address this need, provided they operate both efficiently and biocompatibly. Incorporating biomimetic materials into such devices can help replicate native <span class="hlt">gas</span> <span class="hlt">exchange</span> function and additionally support cellular components. In this work, we have developed microfluidic devices that enable blood <span class="hlt">gas</span> <span class="hlt">exchange</span> across ultra-thin collagen membranes (as thin as 2 μm). Endothelial, stromal, and parenchymal cells readily adhere to these membranes, and long-term culture with cellular components results in remodeling, reflected by reduced membrane thickness. Functionally, acellular collagen-membrane lung devices can mediate effective <span class="hlt">gas</span> <span class="hlt">exchange</span> up to ∼288 mL/min/m(2) of oxygen and ∼685 mL/min/m(2) of carbon dioxide, approaching the <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency noted in the native lung. Testing several configurations of lung devices to explore various physical parameters of the device design, we concluded that thinner membranes and longer <span class="hlt">gas</span> <span class="hlt">exchange</span> distances result in improved hemoglobin saturation and increases in pO2. However, in the design space tested, these effects are relatively small compared to the improvement in overall oxygen and carbon dioxide transfer by increasing the blood flow rate. Finally, devices cultured with endothelial and parenchymal cells achieved similar <span class="hlt">gas</span> <span class="hlt">exchange</span> rates compared with acellular devices. Biomimetic blood oxygenator design opens the possibility of creating portable or implantable microfluidic devices that achieve efficient <span class="hlt">gas</span> transfer while also maintaining physiologic conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AIPC.1738K0008N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AIPC.1738K0008N"><span>The predictive protective control of the heat <span class="hlt">exchanger</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nevriva, Pavel; Filipova, Blanka; Vilimec, Ladislav</p> <p>2016-06-01</p> <p>The paper deals with the predictive control applied to flexible cogeneration energy system FES. FES was designed and developed by the VITKOVICE POWER ENGINEERING joint-stock company and represents a new solution of decentralized cogeneration energy sources. In FES, the heating medium is flue <span class="hlt">gas</span> generated by combustion of a solid fuel. The heated medium is power <span class="hlt">gas</span>, which is a <span class="hlt">gas</span> mixture of <span class="hlt">air</span> and <span class="hlt">water</span> steam. Power <span class="hlt">gas</span> is superheated in the main heat <span class="hlt">exchanger</span> and led to <span class="hlt">gas</span> turbines. To protect the main heat <span class="hlt">exchanger</span> against damage by overheating, the novel predictive protective control based on the mathematical model of <span class="hlt">exchanger</span> was developed. The paper describes the principle, the design and the simulation of the predictive protective method applied to main heat <span class="hlt">exchanger</span> of FES.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/511713','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/511713"><span>Incorporating the <span class="hlt">gas</span> analyzer response time in <span class="hlt">gas</span> <span class="hlt">exchange</span> computations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mitchell, R R</p> <p>1979-11-01</p> <p>A simple method for including the <span class="hlt">gas</span> analyzer response time in the breath-by-breath computation of <span class="hlt">gas</span> <span class="hlt">exchange</span> rates is described. The method uses a difference equation form of a model for the <span class="hlt">gas</span> analyzer in the computation of oxygen uptake and carbon dioxide production and avoids a numerical differentiation required to correct the <span class="hlt">gas</span> fraction wave forms. The effect of not accounting for analyzer response time is shown to be a 20% underestimation in <span class="hlt">gas</span> <span class="hlt">exchange</span> rate. The present method accurately measures <span class="hlt">gas</span> <span class="hlt">exchange</span> rate, is relatively insensitive to measurement errors in the analyzer time constant, and does not significantly increase the computation time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24970854','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24970854"><span>Pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency during exercise breathing normoxic and hypoxic <span class="hlt">gas</span> in adults born very preterm with low diffusion capacity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Duke, Joseph W; Elliott, Jonathan E; Laurie, Steven S; Beasley, Kara M; Mangum, Tyler S; Hawn, Jerold A; Gladstone, Igor M; Lovering, Andrew T</p> <p>2014-09-01</p> <p>Adults with a history of very preterm birth (<32 wk gestational age; PRET) have reduced lung function and significantly lower lung diffusion capacity for carbon monoxide (DLCO) relative to individuals born at term (CONT). Low DLCO may predispose PRET to diffusion limitation during exercise, particularly while breathing hypoxic <span class="hlt">gas</span> because of a reduced O2 driving gradient and pulmonary capillary transit time. We hypothesized that PRET would have significantly worse pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency [i.e., increased alveolar-to-arterial Po2 difference (AaDO2)] during exercise breathing room <span class="hlt">air</span> or hypoxic <span class="hlt">gas</span> (FiO2 = 0.12) compared with CONT. To test this hypothesis, we compared the AaDO2 in PRET (n = 13) with a clinically mild reduction in DLCO (72 ± 7% of predicted) and CONT (n = 14) with normal DLCO (105 ± 10% of predicted) pre- and during exercise breathing room <span class="hlt">air</span> and hypoxic <span class="hlt">gas</span>. Measurements of temperature-corrected arterial blood gases, and direct measure of O2 saturation (SaO2), were made prior to and during exercise at 25, 50, and 75% of peak oxygen consumption (V̇o2peak) while breathing room <span class="hlt">air</span> and hypoxic <span class="hlt">gas</span>. In addition to DLCO, pulmonary function and exercise capacity were significantly less in PRET. Despite PRET having low DLCO, no differences were observed in the AaDO2 or SaO2 pre- or during exercise breathing room <span class="hlt">air</span> or hypoxic <span class="hlt">gas</span> compared with CONT. Although our findings were unexpected, we conclude that reduced pulmonary function and low DLCO resulting from very preterm birth does not cause a measureable reduction in pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency. Copyright © 2014 the American Physiological Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20709922','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20709922"><span>Phenotypic plasticity of <span class="hlt">gas</span> <span class="hlt">exchange</span> pattern and <span class="hlt">water</span> loss in Scarabaeus spretus (Coleoptera: Scarabaeidae): deconstructing the basis for metabolic rate variation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Terblanche, John S; Clusella-Trullas, Susana; Chown, Steven L</p> <p>2010-09-01</p> <p>Investigation of <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns and modulation of metabolism provide insight into metabolic control systems and evolution in diverse terrestrial environments. Variation in metabolic rate in response to environmental conditions has been explained largely in the context of two contrasting hypotheses, namely metabolic depression in response to stressful or resource-(e.g. <span class="hlt">water</span>) limited conditions, or elevation of metabolism at low temperatures to sustain life in extreme conditions. To deconstruct the basis for metabolic rate changes in response to temperature variation, here we undertake a full factorial study investigating the longer- and short-term effects of temperature exposure on <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns. We examined responses of traits of <span class="hlt">gas</span> <span class="hlt">exchange</span> [standard metabolic rate (SMR); discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> (DGE) cycle frequency; cuticular, respiratory and total <span class="hlt">water</span> loss rate (WLR)] to elucidate the magnitude and form of plastic responses in the dung beetle, Scarabaeus spretus. Results showed that short- and longer-term temperature variation generally have significant effects on SMR and WLR. Overall, acclimation to increased temperature led to a decline in SMR (from 0.071+/-0.004 ml CO(2) h(-1) in 15 degrees C-acclimated beetles to 0.039+/-0.004 ml CO(2) h(-1) in 25 degrees C-acclimated beetles measured at 20 degrees C) modulated by reduced DGE frequency (15 degrees C acclimation: 0.554+/-0.027 mHz, 20 degrees C acclimation: 0.257+/-0.030 mHz, 25 degrees C acclimation: 0.208+/-0.027 mHz recorded at 20 degrees C), reduced cuticular WLRs (from 1.058+/-0.537 mg h(-1) in 15 degrees C-acclimated beetles to 0.900+/-0.400 mg h(-1) in 25 degrees C-acclimated beetles measured at 20 degrees C) and reduced total WLR (from 4.2+/-0.5 mg h(-1) in 15 degrees C-acclimated beetles to 3.1+/-0.5 mg h(-1) in 25 degrees C-acclimated beetles measured at 25 degrees C). Respiratory WLR was reduced from 2.25+/-0.40 mg h(-1) in 15 degrees C-acclimated beetles to 1.60+/-0.40 mg h</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3046576','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3046576"><span>A Three-Dimensional Multiscale Model for <span class="hlt">Gas</span> <span class="hlt">Exchange</span> in Fruit1[C][W][OA</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ho, Quang Tri; Verboven, Pieter; Verlinden, Bert E.; Herremans, Els; Wevers, Martine; Carmeliet, Jan; Nicolaï, Bart M.</p> <p>2011-01-01</p> <p>Respiration of bulky plant organs such as roots, tubers, stems, seeds, and fruit depends very much on oxygen (O2) availability and often follows a Michaelis-Menten-like response. A multiscale model is presented to calculate <span class="hlt">gas</span> <span class="hlt">exchange</span> in plants using the microscale geometry of the tissue, or vice versa, local concentrations in the cells from macroscopic <span class="hlt">gas</span> concentration profiles. This approach provides a computationally feasible and accurate analysis of cell metabolism in any plant organ during hypoxia and anoxia. The predicted O2 and carbon dioxide (CO2) partial pressure profiles compared very well with experimental data, thereby validating the multiscale model. The important microscale geometrical features are the shape, size, and three-dimensional connectivity of cells and <span class="hlt">air</span> spaces. It was demonstrated that the <span class="hlt">gas-exchange</span> properties of the cell wall and cell membrane have little effect on the cellular <span class="hlt">gas</span> <span class="hlt">exchange</span> of apple (Malus × domestica) parenchyma tissue. The analysis clearly confirmed that cells are an additional route for CO2 transport, while for O2 the intercellular spaces are the main diffusion route. The simulation results also showed that the local <span class="hlt">gas</span> concentration gradients were steeper in the cells than in the surrounding <span class="hlt">air</span> spaces. Therefore, to analyze the cellular metabolism under hypoxic and anoxic conditions, the microscale model is required to calculate the correct intracellular concentrations. Understanding the O2 response of plants and plant organs thus not only requires knowledge of external conditions, dimensions, <span class="hlt">gas-exchange</span> properties of the tissues, and cellular respiration kinetics but also of microstructure. PMID:21224337</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRC..122.3696L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRC..122.3696L"><span>How well does wind speed predict <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer in the sea ice zone? A synthesis of radon deficit profiles in the upper <span class="hlt">water</span> column of the Arctic Ocean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Loose, B.; Kelly, R. P.; Bigdeli, A.; Williams, W.; Krishfield, R.; Rutgers van der Loeff, M.; Moran, S. B.</p> <p>2017-05-01</p> <p>We present 34 profiles of radon-deficit from the ice-ocean boundary layer of the Beaufort Sea. Including these 34, there are presently 58 published radon-deficit estimates of <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer velocity (k) in the Arctic Ocean; 52 of these estimates were derived from <span class="hlt">water</span> covered by 10% sea ice or more. The average value of k collected since 2011 is 4.0 ± 1.2 m d-1. This exceeds the quadratic wind speed prediction of weighted kws = 2.85 m d-1 with mean-weighted wind speed of 6.4 m s-1. We show how ice cover changes the mixed-layer radon budget, and yields an "effective <span class="hlt">gas</span> transfer velocity." We use these 58 estimates to statistically evaluate the suitability of a wind speed parameterization for k, when the ocean surface is ice covered. Whereas the six profiles taken from the open ocean indicate a statistically good fit to wind speed parameterizations, the same parameterizations could not reproduce k from the sea ice zone. We conclude that techniques for estimating k in the open ocean cannot be similarly applied to determine k in the presence of sea ice. The magnitude of k through gaps in the ice may reach high values as ice cover increases, possibly as a result of focused turbulence dissipation at openings in the free surface. These 58 profiles are presently the most complete set of estimates of k across seasons and variable ice cover; as dissolved tracer budgets they reflect <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> with no impact from <span class="hlt">air</span>-ice <span class="hlt">gas</span> <span class="hlt">exchange</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18245633','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18245633"><span>The hygric hypothesis does not hold <span class="hlt">water</span>: abolition of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> cycles does not affect <span class="hlt">water</span> loss in the ant Camponotus vicinus.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lighton, John R B; Turner, Robbin J</p> <p>2008-02-01</p> <p>The discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> cycle (DGC) of insects and other tracheate arthropods temporally decouples oxygen uptake and carbon dioxide emission and generates powerful concentration gradients for both <span class="hlt">gas</span> species between the outside world and the tracheal system. Although the DGC is considered an adaptation to reduce respiratory <span class="hlt">water</span> loss (RWL) - the "hygric hypothesis" - it is absent from many taxa, including xeric ones. The "chthonic hypothesis" states that the DGC originated as an adaptation to <span class="hlt">gas</span> <span class="hlt">exchange</span> in hypoxic and hypercapnic, i.e. underground, environments. If that is the case then the DGC is not the ancestral condition, and its expression is not necessarily a requirement for reducing RWL. Here we report a study of <span class="hlt">water</span> loss rate in the ant Camponotus vicinus, measured while its DGC was slowly eliminated by gradual hypoxia (hypoxic ramp de-DGCing). Metabolic rate remained constant. The DGC ceased at a mean P(O2) of 8.4 kPa. RWL in the absence of DGCs was not affected until P(O2) declined below 3.9 kPa. Below that value, non-DGC spiracular regulation failed, accompanied by a large increase in RWL. Thus, the spiracular control strategy of the DGC is not required for low RWL, even in animals that normally express the DGC.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23069190','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23069190"><span>Estimating oxygen diffusive conductances of <span class="hlt">gas-exchange</span> systems: A stereological approach illustrated with the human placenta.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mayhew, Terry M</p> <p>2014-01-01</p> <p>For many organisms, respiratory <span class="hlt">gas</span> <span class="hlt">exchange</span> is a vital activity and different types of <span class="hlt">gas-exchange</span> apparatus have evolved to meet individual needs. They include not only skin, gills, tracheal systems and lungs but also transient structures such as the chorioallantois of avian eggs and the placenta of eutherian mammals. The ability of these structures to allow passage of oxygen by passive diffusion can be expressed as a diffusive conductance (units: cm(3) O2 min(-1) kPa(-1)). Occasionally, the ability to estimate diffusive conductance by physiological techniques is compromised by the difficulty of obtaining O2 partial pressures on opposite sides of the tissue interface between the delivery medium (<span class="hlt">air</span>, <span class="hlt">water</span>, blood) and uptake medium (usually blood). An alternative strategy is to estimate a morphometric diffusive conductance by combining stereological estimates of key structural quantities (volumes, surface areas, membrane thicknesses) with complementary physicochemical data (O2-haemoglobin chemical reaction rates and Krogh's permeability coefficients). This approach has proved valuable in a variety of comparative studies on respiratory organs from diverse species. The underlying principles were formulated in pioneering studies on the pulmonary lung but are illustrated here by taking the human placenta as the <span class="hlt">gas</span> <span class="hlt">exchanger</span>. Copyright © 2012 Elsevier GmbH. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1084201','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1084201"><span>Method for controlling exhaust <span class="hlt">gas</span> heat recovery systems in vehicles</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Spohn, Brian L.; Claypole, George M.; Starr, Richard D</p> <p>2013-06-11</p> <p>A method of operating a vehicle including an engine, a transmission, an exhaust <span class="hlt">gas</span> heat recovery (EGHR) heat <span class="hlt">exchanger</span>, and an oil-to-<span class="hlt">water</span> heat <span class="hlt">exchanger</span> providing selective heat-<span class="hlt">exchange</span> communication between the engine and transmission. The method includes controlling a two-way valve, which is configured to be set to one of an engine position and a transmission position. The engine position allows heat-<span class="hlt">exchange</span> communication between the EGHR heat <span class="hlt">exchanger</span> and the engine, but does not allow heat-<span class="hlt">exchange</span> communication between the EGHR heat <span class="hlt">exchanger</span> and the oil-to-<span class="hlt">water</span> heat <span class="hlt">exchanger</span>. The transmission position allows heat-<span class="hlt">exchange</span> communication between the EGHR heat <span class="hlt">exchanger</span>, the oil-to-<span class="hlt">water</span> heat <span class="hlt">exchanger</span>, and the engine. The method also includes monitoring an ambient <span class="hlt">air</span> temperature and comparing the monitored ambient <span class="hlt">air</span> temperature to a predetermined cold ambient temperature. If the monitored ambient <span class="hlt">air</span> temperature is greater than the predetermined cold ambient temperature, the two-way valve is set to the transmission position.</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('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=307735&keyword=environmental+AND+assessment+AND+natural+AND+environment&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=307735&keyword=environmental+AND+assessment+AND+natural+AND+environment&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>Review of <span class="hlt">Air</span> <span class="hlt">Exchange</span> Rate Models for <span class="hlt">Air</span> Pollution Exposure Assessments</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>A critical aspect of <span class="hlt">air</span> pollution exposure assessments is estimation of the <span class="hlt">air</span> <span class="hlt">exchange</span> rate (AER) for various buildings, where people spend their time. The AER, which is rate the <span class="hlt">exchange</span> of indoor <span class="hlt">air</span> with outdoor <span class="hlt">air</span>, is an important determinant for entry of outdoor <span class="hlt">air</span> pol...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15852967','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15852967"><span>[Effects of soil <span class="hlt">water</span> status on <span class="hlt">gas</span> <span class="hlt">exchange</span> of peanut and early rice leaves].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chen, Jiazhou; Lü, Guoan; He, Yuanqiu</p> <p>2005-01-01</p> <p>The <span class="hlt">gas</span> <span class="hlt">exchange</span> characteristics of peanut and early rice leaves were investigated in experimental plots under different soil <span class="hlt">water</span> conditions over a long growth period. The results showed that at the branching stage of peanut, the stomatal conductance (Gs) and transpiration rate (Tr) decreased slightly under mild and moderate soil <span class="hlt">water</span> stress, while the net photosynthetic rate (Pn) and leaf <span class="hlt">water</span> use efficiency (WUE) increased. The Gs/Tr ratio also increased under mild <span class="hlt">water</span> stress, but decreased under moderate <span class="hlt">water</span> stress. At podding stage, the Gs, Tr, Gs/Tr ratio and Pn decreased, while WUE increased significantly under mild and moderate <span class="hlt">water</span> stress. The peanut was suffered from <span class="hlt">water</span> stress at its pod setting stage. At the grain filling stage of early rice, the Gs, Tr and Gs/Tr ratio fluctuated insignificantly under mild and moderate <span class="hlt">water</span> stress, while Pn and WUE increased significantly, with an increase in grain yield under mild <span class="hlt">water</span> stress. It's suggested that the combination of Gs and Gs/Tr ratio could be a reference index for crop <span class="hlt">water</span> stress, namely, crops could be hazarded by <span class="hlt">water</span> stress when Gs and Gs/Tr decreased synchronously.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5512930','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5512930"><span>Effect of <span class="hlt">Water</span> Vapor and Surface Morphology on the Low Temperature Response of Metal Oxide Semiconductor <span class="hlt">Gas</span> Sensors</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Maier, Konrad; Helwig, Andreas; Müller, Gerhard; Hille, Pascal; Eickhoff, Martin</p> <p>2015-01-01</p> <p>In this work the low temperature response of metal oxide semiconductor <span class="hlt">gas</span> sensors is analyzed. Important characteristics of this low-temperature response are a pronounced selectivity to acid- and base-forming gases and a large disparity of response and recovery time constants which often leads to an integrator-type of <span class="hlt">gas</span> response. We show that this kind of sensor performance is related to the trend of semiconductor <span class="hlt">gas</span> sensors to adsorb <span class="hlt">water</span> vapor in multi-layer form and that this ability is sensitively influenced by the surface morphology. In particular we show that surface roughness in the nanometer range enhances desorption of <span class="hlt">water</span> from multi-layer adsorbates, enabling them to respond more swiftly to changes in the ambient humidity. Further experiments reveal that reactive gases, such as NO2 and NH3, which are easily absorbed in the <span class="hlt">water</span> adsorbate layers, are more easily <span class="hlt">exchanged</span> across the liquid/<span class="hlt">air</span> interface when the humidity in the ambient <span class="hlt">air</span> is high. PMID:28793583</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10998035','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10998035"><span>Contamination of piped medical <span class="hlt">gas</span> supply with <span class="hlt">water</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hay, H</p> <p>2000-08-01</p> <p>The failure of anaesthetic equipment as a result of maintenance is extremely rare. The ingress of <span class="hlt">water</span> into the flowmeters of an anaesthetic machine from the piped medical <span class="hlt">air</span> supply is reported and is possibly unique. The piped medical <span class="hlt">air</span> supply was open to the atmosphere during maintenance. <span class="hlt">Water</span> condensed in the <span class="hlt">gas</span> pipeline and this was not noticed during subsequent testing. <span class="hlt">Water</span> was seen leaking from the orthopaedic <span class="hlt">air</span> tools used for surgery but was assumed to be from the autoclaving process. Later the same day, when medical <span class="hlt">air</span> from the piped source was used as part of the <span class="hlt">gas</span> mixture for a general anaesthetic, <span class="hlt">water</span> was seen filling the barrel of the flowmeter <span class="hlt">air</span> control valve. This could have had far-reaching and dangerous consequences for the patient, which were fortunately averted.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=65360&keyword=day+AND+night&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=65360&keyword=day+AND+night&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>MONITORING CYCLICAL <span class="hlt">AIR-WATER</span> ELEMENTAL MERCURY <span class="hlt">EXCHANGE</span></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>Previous experimental work has demonstrated that elemental mercury evasion from natural <span class="hlt">water</span> displays a diel cycle; evasion rates during the day can be two to three times evasion rates observed at night. A study with polychlorinated biphenyls (PCBS) found that diurnal PCB <span class="hlt">air</span>/wa...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5026132','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5026132"><span>Lung Structure and the Intrinsic Challenges of <span class="hlt">Gas</span> <span class="hlt">Exchange</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>Hsia, Connie C.W.; Hyde, Dallas M.; Weibel, Ewald R.</p> <p>2016-01-01</p> <p>Structural and functional complexities of the mammalian lung evolved to meet a unique set of challenges, namely, the provision of efficient delivery of inspired <span class="hlt">air</span> to all lung units within a confined thoracic space, to build a large <span class="hlt">gas</span> <span class="hlt">exchange</span> surface associated with minimal barrier thickness and a microvascular network to accommodate the entire right ventricular cardiac output while withstanding cyclic mechanical stresses that increase several folds from rest to exercise. Intricate regulatory mechanisms at every level ensure that the dynamic capacities of ventilation, perfusion, diffusion, and chemical binding to hemoglobin are commensurate with usual metabolic demands and periodic extreme needs for activity and survival. This article reviews the structural design of mammalian and human lung, its functional challenges, limitations, and potential for adaptation. We discuss (i) the evolutionary origin of alveolar lungs and its advantages and compromises, (ii) structural determinants of alveolar <span class="hlt">gas</span> <span class="hlt">exchange</span>, including architecture of conducting bronchovascular trees that converge in <span class="hlt">gas</span> <span class="hlt">exchange</span> units, (iii) the challenges of matching ventilation, perfusion, and diffusion and tissue-erythrocyte and thoracopulmonary interactions. The notion of erythrocytes as an integral component of the <span class="hlt">gas</span> <span class="hlt">exchanger</span> is emphasized. We further discuss the signals, sources, and limits of structural plasticity of the lung in alveolar hypoxia and following a loss of lung units, and the promise and caveats of interventions aimed at augmenting endogenous adaptive responses. Our objective is to understand how individual components are matched at multiple levels to optimize organ function in the face of physiological demands or pathological constraints. PMID:27065169</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27065169','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27065169"><span>Lung Structure and the Intrinsic Challenges of <span class="hlt">Gas</span> <span class="hlt">Exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hsia, Connie C W; Hyde, Dallas M; Weibel, Ewald R</p> <p>2016-03-15</p> <p>Structural and functional complexities of the mammalian lung evolved to meet a unique set of challenges, namely, the provision of efficient delivery of inspired <span class="hlt">air</span> to all lung units within a confined thoracic space, to build a large <span class="hlt">gas</span> <span class="hlt">exchange</span> surface associated with minimal barrier thickness and a microvascular network to accommodate the entire right ventricular cardiac output while withstanding cyclic mechanical stresses that increase several folds from rest to exercise. Intricate regulatory mechanisms at every level ensure that the dynamic capacities of ventilation, perfusion, diffusion, and chemical binding to hemoglobin are commensurate with usual metabolic demands and periodic extreme needs for activity and survival. This article reviews the structural design of mammalian and human lung, its functional challenges, limitations, and potential for adaptation. We discuss (i) the evolutionary origin of alveolar lungs and its advantages and compromises, (ii) structural determinants of alveolar <span class="hlt">gas</span> <span class="hlt">exchange</span>, including architecture of conducting bronchovascular trees that converge in <span class="hlt">gas</span> <span class="hlt">exchange</span> units, (iii) the challenges of matching ventilation, perfusion, and diffusion and tissue-erythrocyte and thoracopulmonary interactions. The notion of erythrocytes as an integral component of the <span class="hlt">gas</span> <span class="hlt">exchanger</span> is emphasized. We further discuss the signals, sources, and limits of structural plasticity of the lung in alveolar hypoxia and following a loss of lung units, and the promise and caveats of interventions aimed at augmenting endogenous adaptive responses. Our objective is to understand how individual components are matched at multiple levels to optimize organ function in the face of physiological demands or pathological constraints. Copyright © 2016 John Wiley & Sons, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AdSpR..51..465W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AdSpR..51..465W"><span>Plant mineral nutrition, <span class="hlt">gas</span> <span class="hlt">exchange</span> and photosynthesis in space: A review</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wolff, S. A.; Coelho, L. H.; Zabrodina, M.; Brinckmann, E.; Kittang, A.-I.</p> <p>2013-02-01</p> <p>Successful growth and development of higher plants in space rely on adequate availability and uptake of <span class="hlt">water</span> and nutrients, and efficient energy distribution through photosynthesis and <span class="hlt">gas</span> <span class="hlt">exchange</span>. In the present review, literature has been reviewed to assemble the relevant knowledge within space plant research for future planetary missions. Focus has been on fractional gravity, space radiation, magnetic fields and ultimately a combined effect of these factors on <span class="hlt">gas</span> <span class="hlt">exchange</span>, photosynthesis and transport of <span class="hlt">water</span> and solutes. Reduced gravity prevents buoyancy driven thermal convection in the physical environment around the plant and alters transport and <span class="hlt">exchange</span> of gases and liquids between the plant and its surroundings. In space experiments, indications of root zone hypoxia have frequently been reported, but studies on the influences of the space environment on plant nutrition and <span class="hlt">water</span> transport are limited or inconclusive. Some studies indicate that uptake of potassium is elevated when plants are grown under microgravity conditions. Based on the current knowledge, <span class="hlt">gas</span> <span class="hlt">exchange</span>, metabolism and photosynthesis seem to work properly in space when plants are provided with a well stirred atmosphere and grown at moderate light levels. Effects of space radiation on plant metabolism, however, have not been studied so far in orbit. Ground experiments indicated that shielding from the Earth's magnetic field alters plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and metabolism, though more studies are required to understand the effects of magnetic fields on plant growth. It has been shown that plants can grow and reproduce in the space environment and adapt to space conditions. However, the influences of the space environment may result in a long term effect over multiple generations or have an impact on the plants' role as food and part of a regenerative life support system. Suggestions for future plant biology research in space are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23720333','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23720333"><span>Evolution of <span class="hlt">air</span> breathing: oxygen homeostasis and the transitions from <span class="hlt">water</span> to land and sky.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hsia, Connie C W; Schmitz, Anke; Lambertz, Markus; Perry, Steven F; Maina, John N</p> <p>2013-04-01</p> <p>Life originated in anoxia, but many organisms came to depend upon oxygen for survival, independently evolving diverse respiratory systems for acquiring oxygen from the environment. Ambient oxygen tension (PO2) fluctuated through the ages in correlation with biodiversity and body size, enabling organisms to migrate from <span class="hlt">water</span> to land and <span class="hlt">air</span> and sometimes in the opposite direction. Habitat expansion compels the use of different <span class="hlt">gas</span> <span class="hlt">exchangers</span>, for example, skin, gills, tracheae, lungs, and their intermediate stages, that may coexist within the same species; coexistence may be temporally disjunct (e.g., larval gills vs. adult lungs) or simultaneous (e.g., skin, gills, and lungs in some salamanders). Disparate systems exhibit similar directions of adaptation: toward larger diffusion interfaces, thinner barriers, finer dynamic regulation, and reduced cost of breathing. Efficient respiratory <span class="hlt">gas</span> <span class="hlt">exchange</span>, coupled to downstream convective and diffusive resistances, comprise the "oxygen cascade"-step-down of PO2 that balances supply against toxicity. Here, we review the origin of oxygen homeostasis, a primal selection factor for all respiratory systems, which in turn function as gatekeepers of the cascade. Within an organism's lifespan, the respiratory apparatus adapts in various ways to upregulate oxygen uptake in hypoxia and restrict uptake in hyperoxia. In an evolutionary context, certain species also become adapted to environmental conditions or habitual organismic demands. We, therefore, survey the comparative anatomy and physiology of respiratory systems from invertebrates to vertebrates, <span class="hlt">water</span> to <span class="hlt">air</span> breathers, and terrestrial to aerial inhabitants. Through the evolutionary directions and variety of <span class="hlt">gas</span> <span class="hlt">exchangers</span>, their shared features and individual compromises may be appreciated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3926130','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3926130"><span>Evolution of <span class="hlt">Air</span> Breathing: Oxygen Homeostasis and the Transitions from <span class="hlt">Water</span> to Land and Sky</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hsia, Connie C. W.; Schmitz, Anke; Lambertz, Markus; Perry, Steven F.; Maina, John N.</p> <p>2014-01-01</p> <p>Life originated in anoxia, but many organisms came to depend upon oxygen for survival, independently evolving diverse respiratory systems for acquiring oxygen from the environment. Ambient oxygen tension (PO2) fluctuated through the ages in correlation with biodiversity and body size, enabling organisms to migrate from <span class="hlt">water</span> to land and <span class="hlt">air</span> and sometimes in the opposite direction. Habitat expansion compels the use of different <span class="hlt">gas</span> <span class="hlt">exchangers</span>, for example, skin, gills, tracheae, lungs, and their intermediate stages, that may coexist within the same species; coexistence may be temporally disjunct (e.g., larval gills vs. adult lungs) or simultaneous (e.g., skin, gills, and lungs in some salamanders). Disparate systems exhibit similar directions of adaptation: toward larger diffusion interfaces, thinner barriers, finer dynamic regulation, and reduced cost of breathing. Efficient respiratory <span class="hlt">gas</span> <span class="hlt">exchange</span>, coupled to downstream convective and diffusive resistances, comprise the “oxygen cascade”—step-down of PO2 that balances supply against toxicity. Here, we review the origin of oxygen homeostasis, a primal selection factor for all respiratory systems, which in turn function as gatekeepers of the cascade. Within an organism's lifespan, the respiratory apparatus adapts in various ways to upregulate oxygen uptake in hypoxia and restrict uptake in hyperoxia. In an evolutionary context, certain species also become adapted to environmental conditions or habitual organismic demands. We, therefore, survey the comparative anatomy and physiology of respiratory systems from invertebrates to vertebrates, <span class="hlt">water</span> to <span class="hlt">air</span> breathers, and terrestrial to aerial inhabitants. Through the evolutionary directions and variety of <span class="hlt">gas</span> <span class="hlt">exchangers</span>, their shared features and individual compromises may be appreciated. PMID:23720333</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21073780-fouling-reduction-characteristics-distributor-fluidized-bed-heat-exchanger-flue-gas-heat-recovery','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21073780-fouling-reduction-characteristics-distributor-fluidized-bed-heat-exchanger-flue-gas-heat-recovery"><span>Fouling reduction characteristics of a no-distributor-fluidized-bed heat <span class="hlt">exchanger</span> for flue <span class="hlt">gas</span> heat recovery</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>Jun, Y.D.; Lee, K.B.; Islam, S.Z.</p> <p>2008-07-01</p> <p>In conventional flue <span class="hlt">gas</span> heat recovery systems, the fouling by fly ashes and the related problems such as corrosion and cleaning are known to be major drawbacks. To overcome these problems, a single-riser no-distributor-fluidized-bed heat <span class="hlt">exchanger</span> is devised and studied. Fouling and cleaning tests are performed for a uniquely designed fluidized bed-type heat <span class="hlt">exchanger</span> to demonstrate the effect of particles on the fouling reduction and heat transfer enhancement. The tested heat <span class="hlt">exchanger</span> model (1 m high and 54 mm internal diameter) is a <span class="hlt">gas-to-water</span> type and composed of a main vertical tube and four auxiliary tubes through which particles circulatemore » and transfer heat. Through the present study, the fouling on the heat transfer surface could successfully be simulated by controlling <span class="hlt">air</span>-to-fuel ratios rather than introducing particles through an external feeder, which produced soft deposit layers with 1 to 1.5 mm thickness on the inside pipe wall. Flue <span class="hlt">gas</span> temperature at the inlet of heat <span class="hlt">exchanger</span> was maintained at 450{sup o}C at the <span class="hlt">gas</span> volume rate of 0.738 to 0.768 CMM (0.0123 to 0.0128 m{sup 3}/sec). From the analyses of the measured data, heat transfer performances of the heat <span class="hlt">exchanger</span> before and after fouling and with and without particles were evaluated. Results showed that soft deposits were easily removed by introducing glass bead particles, and also heat transfer performance increased two times by the particle circulation. In addition, it was found that this type of heat <span class="hlt">exchanger</span> had high potential to recover heat of waste gases from furnaces, boilers, and incinerators effectively and to reduce fouling related problems.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28282689','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28282689"><span><span class="hlt">Water</span> <span class="hlt">exchange</span> for screening colonoscopy increases adenoma detection rate: a multicenter, double-blinded, randomized controlled trial.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cadoni, Sergio; Falt, Přemysl; Rondonotti, Emanuele; Radaelli, Franco; Fojtik, Petr; Gallittu, Paolo; Liggi, Mauro; Amato, Arnaldo; Paggi, Silvia; Smajstrla, Vit; Urban, Ondřej; Erriu, Matteo; Koo, Malcolm; Leung, Felix W</p> <p>2017-05-01</p> <p>Background and study aims  Single-center studies, which were retrospective and/or involved unblinded colonoscopists, have suggested that <span class="hlt">water</span> <span class="hlt">exchange</span>, but not <span class="hlt">water</span> immersion, compared with <span class="hlt">air</span> insufflation significantly increases the adenoma detection rate (ADR), particularly in the proximal and right colon. Head-to-head comparison of the three techniques with ADR as primary outcome and blinded colonoscopists has not been reported to date. In a randomized controlled trial with blinded colonoscopists, we aimed to evaluate the impact of the three insertion techniques on ADR. Patients and methods  A total of 1224 patients aged 50 - 70 years (672 males) and undergoing screening colonoscopy were randomized 1:1:1 to <span class="hlt">water</span> <span class="hlt">exchange</span>, <span class="hlt">water</span> immersion, or <span class="hlt">air</span> insufflation. Split-dose bowel preparation was adopted to optimize colon cleansing. After the cecum had been reached, a second colonoscopist who was blinded to the insertion technique performed the withdrawal. The primary outcome was overall ADR according to the three insertion techniques (<span class="hlt">water</span> <span class="hlt">exchange</span>, <span class="hlt">water</span> immersion, and <span class="hlt">air</span> insufflation). Secondary outcomes were other pertinent overall and right colon procedure-related measures. Results  Baseline characteristics of the three groups were comparable. Compared with <span class="hlt">air</span> insufflation, <span class="hlt">water</span> <span class="hlt">exchange</span> achieved a significantly higher overall ADR (49.3 %, 95 % confidence interval [CI] 44.3 % - 54.2 % vs. 40.4 % 95 %CI 35.6 % - 45.3 %; P  = 0.03); <span class="hlt">water</span> <span class="hlt">exchange</span> showed comparable overall ADR vs. <span class="hlt">water</span> immersion (43.4 %, 95 %CI 38.5 % - 48.3 %; P  = 0.28). In the right colon, <span class="hlt">water</span> <span class="hlt">exchange</span> achieved a higher ADR than <span class="hlt">air</span> insufflation (24.0 %, 95 %CI 20.0 % - 28.5 % vs. 16.9 %, 95 %CI 13.4 % - 20.9 %; P  = 0.04) and a higher advanced ADR (6.1 %, 95 %CI 4.0 % - 9.0 % vs. 2.5 %, 95 %CI 1.2 % - 4.6 %; P  = 0.03). Compared with <span class="hlt">air</span> insufflation, the mean number of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?direntryid=20677','PESTICIDES'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?direntryid=20677"><span>VOLATILIZATION RATES FROM <span class="hlt">WATER</span> TO INDOOR <span class="hlt">AIR</span> ...</span></a></p> <p><a target="_blank" href="http://www.epa.gov/pesticides/search.htm">EPA Pesticide Factsheets</a></p> <p></p> <p></p> <p>Contaminated <span class="hlt">water</span> can lead to volatilization of chemicals to residential indoor <span class="hlt">air</span>. Previous research has focused on only one source (shower stalls) and has been limited to chemicals in which <span class="hlt">gas</span>-phase resistance to mass transfer is of marginal significance. As a result, attempts to extrapolate chemical emissions from high-volatility chemicals to lower volatility chemicals, or to sources other than showers, have been difficult or impossible. This study involved the development of two-phase, dynamic mass balance models for estimating chemical emissions from washing machines, dishwashers, and bathtubs. An existing model was adopted for showers only. Each model required the use of source- and chemical-specific mass transfer coefficients. <span class="hlt">Air</span> <span class="hlt">exchange</span> (ventilation) rates were required for dishwashers and washing machines as well. These parameters were estimated based on a series of 113 experiments involving 5 tracer chemicals (acetone, ethyl acetate, toluene, ethylbenzene, and cyclohexane) and 4 sources (showers, bathtubs, washing machines, and dishwashers). Each set of experiments led to the determination of chemical stripping efficiencies and mass transfer coefficients (overall, liquid-phase, <span class="hlt">gas</span>-phase), and to an assessment of the importance of <span class="hlt">gas</span>- phase resistance to mass transfer. Stripping efficiencies ranged from 6.3% to 80% for showers, 2.6% to 69% for bathtubs, 18% to 100% for dishwashers, and 3.8% to 100% for washing machines. Acetone and cyclohexane al</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=332572','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=332572"><span>Scaling leaf measurements to estimate cotton canopy <span class="hlt">gas</span> <span class="hlt">exchange</span></span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>Diurnal leaf and canopy <span class="hlt">gas</span> <span class="hlt">exchange</span> of well <span class="hlt">watered</span> field grown cotton were measured. Leaf measurements were made with a portable photosynthesis system and canopy measurements with open Canopy Evapo-Transpiration and Assimilation (CETA) systems. Leaf level measurements were arithmetically scaled to...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PrOce.138...18D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PrOce.138...18D"><span>Nitrous oxide and methane in Atlantic and Mediterranean <span class="hlt">waters</span> in the Strait of Gibraltar: <span class="hlt">Air</span>-sea fluxes and inter-basin <span class="hlt">exchange</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>de la Paz, M.; Huertas, I. E.; Flecha, S.; Ríos, A. F.; Pérez, F. F.</p> <p>2015-11-01</p> <p>The global ocean plays an important role in the overall budget of nitrous oxide (N2O) and methane (CH4), as both gases are produced within the ocean and released to the atmosphere. However, for large parts of the open and coastal oceans there is little or no spatial data coverage for N2O and CH4. Hence, a better assessment of marine emissions estimates is necessary. As a contribution to remedying the scarcity of data on marine regions, N2O and CH4 concentrations have been determined in the Strait of Gibraltar at the ocean Fixed Time series (GIFT). During six cruises performed between July 2011 and November 2014 samples were collected at the surface and various depths in the <span class="hlt">water</span> column, and subsequently measured using <span class="hlt">gas</span> chromatography. From this we were able to quantify the temporal variability of the <span class="hlt">gas</span> <span class="hlt">air</span>-sea <span class="hlt">exchange</span> in the area and examine the vertical distribution of N2O and CH4 in Atlantic and Mediterranean <span class="hlt">waters</span>. Results show that surface Atlantic <span class="hlt">waters</span> are nearly in equilibrium with the atmosphere whereas deeper Mediterranean <span class="hlt">waters</span> are oversaturated in N2O, and a gradient that gradually increases with depth was detected in the <span class="hlt">water</span> column. Temperature was found to be the main factor responsible for the seasonal variability of N2O in the surface layer. Furthermore, although CH4 levels did not reveal any feature clearly associated with the circulation of <span class="hlt">water</span> masses, vertical distributions showed that higher concentrations are generally observed in the Atlantic layer, and that the deeper Mediterranean <span class="hlt">waters</span> are considerably undersaturated (by up to 50%). Even though surface <span class="hlt">waters</span> act as a source of atmospheric N2O during certain periods, on an annual basis the net N2O flux in the Strait of Gibraltar is only 0.35 ± 0.27 μmol m-2 d-1, meaning that these <span class="hlt">waters</span> are almost in a neutral status with respect to the atmosphere. Seasonally, the region behaves as a slight sink for atmospheric CH4 in winter and as a source in spring and fall. Approximating</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28667337','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28667337"><span>Contrasting dynamics of leaf potential and <span class="hlt">gas</span> <span class="hlt">exchange</span> during progressive drought cycles and recovery in Amorpha fruticosa and Robinia pseudoacacia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yan, Weiming; Zheng, Shuxia; Zhong, Yangquanwei; Shangguan, Zhouping</p> <p>2017-06-30</p> <p>Leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> is closely associated with <span class="hlt">water</span> relations; however, less attention has been given to this relationship over successive drought events. Dynamic changes in <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> potential in the seedlings of two woody species, Amorpha fruticosa and Robinia pseudoacacia, were monitored during recurrent drought. The pre-dawn leaf <span class="hlt">water</span> potential declined in parallel with <span class="hlt">gas</span> <span class="hlt">exchange</span> in both species, and sharp declines in <span class="hlt">gas</span> <span class="hlt">exchange</span> occurred with decreasing <span class="hlt">water</span> potential. A significant correlation between pre-dawn <span class="hlt">water</span> potential and <span class="hlt">gas</span> <span class="hlt">exchange</span> was observed in both species and showed a right shift in R. pseudoacacia in the second drought. The results suggested that stomatal closure in early drought was mediated mainly by elevated foliar abscisic acid (ABA) in R. pseudoacacia, while a shift from ABA-regulated to leaf-<span class="hlt">water</span>-potential-driven stomatal closure was observed in A. fruticosa. After re-<span class="hlt">watering</span>, the pre-dawn <span class="hlt">water</span> potential recovered quickly, whereas stomatal conductance did not fully recover from drought in R. pseudoacacia, which affected the ability to tightly control transpiration post-drought. The dynamics of recovery from drought suggest that stomatal behavior post-drought may be restricted mainly by hydraulic factors, but non-hydraulic factors may also be involved in R. pseudoacacia.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880001086','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880001086"><span>The effect of wind and currents on <span class="hlt">gas</span> <span class="hlt">exchange</span> in an estuarine system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Broecker, W. S.; Ledwell, J. R.; Bopp, R.</p> <p>1987-01-01</p> <p>The objectives were to develop a non-volatile tracer to use in <span class="hlt">gas</span> <span class="hlt">exchange</span> experiments in laterally unconfined systems and to study applications of deliberate tracers in limnology and oceanography. Progress was made on both fronts but work on the development of the non-volatile tracer proved to be more difficult and labor intensive that anticipated so no field experiments using non-volatile tracers was performed as yet. In the search for a suitable non-volatile tracer for an ocean scale <span class="hlt">gas</span> <span class="hlt">exchange</span> experiment a tracer was discovered which does not have the required sensitivity for a large scale experiment, but is very easy to analyze and will be well suited for smaller experiments such as <span class="hlt">gas</span> <span class="hlt">exchange</span> determinations on rivers and streams. Sulfur hexafluoride, SF6, was used successfully as a volatile tracer along with tritium as a non-volatile tracer to study <span class="hlt">gas</span> <span class="hlt">exchange</span> rates from a primary stream. This is the first <span class="hlt">gas</span> <span class="hlt">exchange</span> experiment in which <span class="hlt">gas</span> <span class="hlt">exchange</span> rates were determined on a head <span class="hlt">water</span> stream where significant groundwater input occurs along the reach. In conjunction with SF6, Radon-222 measurements were performed on the groundwater and in the stream. The feasibility of using a combination of SF6 and radon is being studied to determine groundwater inputs and <span class="hlt">gas</span> <span class="hlt">exchange</span> of rates in streams with significant groundwater input without using a non-volatile tracer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23305981','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23305981"><span>Airway <span class="hlt">exchange</span> of highly soluble gases.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hlastala, Michael P; Powell, Frank L; Anderson, Joseph C</p> <p>2013-03-01</p> <p>Highly blood soluble gases <span class="hlt">exchange</span> with the bronchial circulation in the airways. On inhalation, <span class="hlt">air</span> absorbs highly soluble gases from the airway mucosa and equilibrates with the blood before reaching the alveoli. Highly soluble <span class="hlt">gas</span> partial pressure is identical throughout all alveoli. At the end of exhalation the partial pressure of a highly soluble <span class="hlt">gas</span> decreases from the alveolar level in the terminal bronchioles to the end-exhaled partial pressure at the mouth. A mathematical model simulated the airway <span class="hlt">exchange</span> of four gases (methyl isobutyl ketone, acetone, ethanol, and propylene glycol monomethyl ether) that have high <span class="hlt">water</span> and blood solubility. The impact of solubility on the relative distribution of airway <span class="hlt">exchange</span> was studied. We conclude that an increase in <span class="hlt">water</span> solubility shifts the distribution of <span class="hlt">gas</span> <span class="hlt">exchange</span> toward the mouth. Of the four gases studied, ethanol had the greatest decrease in partial pressure from the alveolus to the mouth at end exhalation. Single exhalation breath tests are inappropriate for estimating alveolar levels of highly soluble gases, particularly for ethanol.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4888954','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4888954"><span>Airway <span class="hlt">exchange</span> of highly soluble gases</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Powell, Frank L.; Anderson, Joseph C.</p> <p>2013-01-01</p> <p>Highly blood soluble gases <span class="hlt">exchange</span> with the bronchial circulation in the airways. On inhalation, <span class="hlt">air</span> absorbs highly soluble gases from the airway mucosa and equilibrates with the blood before reaching the alveoli. Highly soluble <span class="hlt">gas</span> partial pressure is identical throughout all alveoli. At the end of exhalation the partial pressure of a highly soluble <span class="hlt">gas</span> decreases from the alveolar level in the terminal bronchioles to the end-exhaled partial pressure at the mouth. A mathematical model simulated the airway <span class="hlt">exchange</span> of four gases (methyl isobutyl ketone, acetone, ethanol, and propylene glycol monomethyl ether) that have high <span class="hlt">water</span> and blood solubility. The impact of solubility on the relative distribution of airway <span class="hlt">exchange</span> was studied. We conclude that an increase in <span class="hlt">water</span> solubility shifts the distribution of <span class="hlt">gas</span> <span class="hlt">exchange</span> toward the mouth. Of the four gases studied, ethanol had the greatest decrease in partial pressure from the alveolus to the mouth at end exhalation. Single exhalation breath tests are inappropriate for estimating alveolar levels of highly soluble gases, particularly for ethanol. PMID:23305981</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhFl...29d5107S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhFl...29d5107S"><span>Surface velocity divergence model of <span class="hlt">air/water</span> interfacial <span class="hlt">gas</span> transfer in open-channel flows</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sanjou, M.; Nezu, I.; Okamoto, T.</p> <p>2017-04-01</p> <p><span class="hlt">Air/water</span> interfacial <span class="hlt">gas</span> transfer through a free surface plays a significant role in preserving and restoring <span class="hlt">water</span> quality in creeks and rivers. However, direct measurements of the <span class="hlt">gas</span> transfer velocity and reaeration coefficient are still difficult, and therefore a reliable prediction model needs to be developed. Varying systematically the bulk-mean velocity and <span class="hlt">water</span> depth, laboratory flume experiments were conducted and we measured surface velocities and dissolved oxygen (DO) concentrations in open-channel flows to reveal the relationship between DO transfer velocity and surface divergence (SD). Horizontal particle image velocimetry measurements provide the time-variations of surface velocity divergence. Positive and negative regions of surface velocity divergence are transferred downstream in time, as occurs in boil phenomenon on natural river free-surfaces. The result implies that interfacial <span class="hlt">gas</span> transfer is related to bottom-situated turbulence motion and vertical mass transfer. The original SD model focuses mainly on small-scale viscous motion, and this model strongly depends on the <span class="hlt">water</span> depth. Therefore, we modify the SD model theoretically to accommodate the effects of the <span class="hlt">water</span> depth on <span class="hlt">gas</span> transfer, introducing a non-dimensional parameter that includes contributions of depth-scale large-vortex motion, such as secondary currents, to surface renewal events related to DO transport. The modified SD model proved effective and reasonable without any dependence on the bulk mean velocity and <span class="hlt">water</span> depth, and has a larger coefficient of determination than the original SD model. Furthermore, modeling of friction velocity with the Reynolds number improves the practicality of a new formula that is expected to be used in studies of natural rivers.</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('https://www.osti.gov/servlets/purl/1411972','SCIGOV-DOEDE'); return false;" href="https://www.osti.gov/servlets/purl/1411972"><span>Diurnal leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> survey, Feb2016-May2016, PA-SLZ, PA-PNM: Panama</span></a></p> <p><a target="_blank" href="http://www.osti.gov/dataexplorer">DOE Data Explorer</a></p> <p>Rogers, Alistair [Brookhaven National Lab; Serbin, Shawn [Brookhaven National Lab; Ely, Kim [Brookhaven National Lab; Wu, Jin [BNL; Wolfe, Brett [Smithsonian; Dickman, Turin [Los Alamos National Lab; Collins, Adam [Los Alamos National Lab; Detto, Matteo [Princeton; Grossiord, Charlotte [Los Alamos National Lab; McDowell, Nate [Los Alamos National Lab; Michaletz, Sean</p> <p>2017-01-01</p> <p>Diurnal leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> survey measured on sunlit canopy trees on a monthly basis from Feb to May 2016 at SLZ and PNM. This data was collected as part of the 2016 ENSO campaign. See related datasets (existing and future) for further sample details, leaf <span class="hlt">water</span> potential, LMA, leaf spectra, other <span class="hlt">gas</span> <span class="hlt">exchange</span> and leaf chemistry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1917686R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1917686R"><span>Uncertainty of the global oceanic CO2 <span class="hlt">exchange</span> at the <span class="hlt">air-water</span> interface induced by the choice of the <span class="hlt">gas</span> <span class="hlt">exchange</span> velocity formulation and the wind product: quantification and spatial analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Roobaert, Alizee; Laruelle, Goulven; Landschützer, Peter; Regnier, Pierre</p> <p>2017-04-01</p> <p>In lakes, rivers, estuaries and the ocean, the quantification of <span class="hlt">air-water</span> CO2 <span class="hlt">exchange</span> (FCO2) is still characterized by large uncertainties partly due to the lack of agreement over the parameterization of the <span class="hlt">gas</span> <span class="hlt">exchange</span> velocity (k). Although the ocean is generally regarded as the best constrained system because k is only controlled by the wind speed, numerous formulations are still currently used, leading to potentially large differences in FCO2. Here, a quantitative global spatial analysis of FCO2 is presented using several k-wind speed formulations in order to compare the effect of the choice of parameterization of k on FCO2. This analysis is performed at a 1 degree resolution using a sea surface pCO2 product generated using a two-step artificial neuronal network by Landschützer et al. (2015) over the 1991-2011 period. Four different global wind speed datasets (CCMP, ERA, NCEP 1 and NCEP 2) are also used to assess the effect of the choice of one wind speed product over the other when calculating the global and regional oceanic FCO2. Results indicate that this choice of wind speed product only leads to small discrepancies globally (6 %) except with NCEP 2 which produces a more intense global FCO2 compared to the other wind products. Regionally, theses differences are even more pronounced. For a given wind speed product, the choice of parametrization of k yields global FCO2 differences ranging from 7 % to 16 % depending on the wind product used. We also provide latitudinal profiles of FCO2 and its uncertainty calculated combining all combinations between the different k-relationships and the four wind speed products. Wind speeds >14 m s-1, which only account for 7 % of all observations, contributes disproportionately to the global oceanic FCO2 and, for this range of wind speeds, the uncertainty induced by the choice of formulation for k is maximum ( 50 %).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.agu.org/journals/jc/v096/iC04/90JC02642/','USGSPUBS'); return false;" href="http://www.agu.org/journals/jc/v096/iC04/90JC02642/"><span>Atmospheric organochlorine pollutants and <span class="hlt">air</span>-sea <span class="hlt">exchange</span> of hexachlorocyclohexane in the Bering and Chukchi Seas</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hinckley, D.A.; Bidleman, T.F.; Rice, C.P.</p> <p>1991-01-01</p> <p>Organochlorine pesticides have been found in Arctic fish, marine mammals, birds, and plankton for some time. The lack of local sources and remoteness of the region imply long-range transport and deposition of contaminants into the Arctic from sources to the south. While on the third Soviet-American Joint Ecological Expedition to the Bering and Chukchi Seas (August 1988), high-volume <span class="hlt">air</span> samples were taken and analyzed for organochlorine pesticides. Hexachlorocyclohexane (HCH), hexachlorobenzene, polychlorinated camphenes, and chlordane (listed in order of abundance, highest to lowest) were quantified. The <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> of HCH was estimated at 18 stations during the cruise. Average alpha-HCH concentrations in concurrent atmosphere and surface <span class="hlt">water</span> samples were 250 pg m-3 and 2.4 ng L-1, respectively, and average gamma-HCH concentrations were 68 pg m-3 in the atmosphere and 0.6 ng L-1 in surface <span class="hlt">water</span>. Calculations based on experimentally derived Henry's law constants showed that the surface <span class="hlt">water</span> was undersaturated with respect to the atmosphere at most stations (alpha-HCH, average 79% saturation; gamma-HCH, average 28% saturation). The flux for alpha-HCH ranged from -47 ng m-2 day-1 (sea to <span class="hlt">air</span>) to 122 ng m-2 d-1 (<span class="hlt">air</span> to sea) and averaged 25 ng m-2 d-1 <span class="hlt">air</span> to sea. All fluxes of gamma-HCH were from <span class="hlt">air</span> to sea, ranged from 17 to 54 ng m-2 d-1, and averaged 31 ng m-2 d-1.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=oxygen&pg=3&id=EJ940769','ERIC'); return false;" href="https://eric.ed.gov/?q=oxygen&pg=3&id=EJ940769"><span><span class="hlt">Gas</span> Property Demonstrations Using Plastic <span class="hlt">Water</span> Bottles</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>Campbell, Dean J.; Bannon, Stephen J.; Gunter, Molly M.</p> <p>2011-01-01</p> <p>Plastic <span class="hlt">water</span> bottles are convenient containers for demonstrations of <span class="hlt">gas</span> properties illustrating Boyle's law, Charles's law, and Avogadro's law. The contents of iron-based disposable hand warmer packets can be used to remove oxygen <span class="hlt">gas</span> from the <span class="hlt">air</span> within an unfilled plastic <span class="hlt">water</span> bottle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1711535S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1711535S"><span>Estimating <span class="hlt">gas</span> <span class="hlt">exchange</span> of CO2 and CH4 between headwater systems and the atmosphere in Southwest Sweden</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Somlai, Celia; Natchimuthu, Sivakiruthika; Bastviken, David; Lorke, Andreas</p> <p>2015-04-01</p> <p>Quantifying the role of inland <span class="hlt">water</span> systems in terms of carbon sinks and sources and their connection to the terrestrial ecosystems and landscapes is fundamental for improving the balance approach of regional and global carbon budgets. Recent research showed that freshwater bodies emit significant amounts of CO2 and CH4 into the atmosphere. The extent of the emissions from small streams and headwaters, however, remains uncertain due to a limited availability of data. Studies have shown that headwater systems receive most of the terrestrial organic carbon, have the highest dissolved CO2 concentration and the highest <span class="hlt">gas</span> <span class="hlt">exchange</span> velocities and cover the largest fractional surface area within fluvial networks. The <span class="hlt">gas</span> <span class="hlt">exchange</span> between inland <span class="hlt">waters</span> and the atmosphere is controlled by two factors: the difference between the dissolved <span class="hlt">gas</span> concentration and its atmospheric equilibrium concentration, and the <span class="hlt">gas</span> <span class="hlt">exchange</span> velocity. The direct measurement of the dissolved <span class="hlt">gas</span> concentration of greenhouse gases can be measured straightforwardly, for example, by <span class="hlt">gas</span> chromatography from headspace extraction of <span class="hlt">water</span> sample. In contrast, direct measurement of <span class="hlt">gas</span> <span class="hlt">exchange</span> velocity is more complex and time consuming, as simultaneous measurements with a volatile and nonvolatile inert tracer <span class="hlt">gas</span> are needed. Here we analyze measurements of <span class="hlt">gas</span> <span class="hlt">exchange</span> velocities, concentrations and fluxes of dissolved CO2 and CH4, as well as loads of total organic and inorganic carbon in 10 reaches in headwater streams in Southwest Sweden. We compare the <span class="hlt">gas</span> <span class="hlt">exchange</span> velocities measured directly through tracer injections with those estimated through various empirical approaches, which are based on modelled and measured current velocity, stream depth and slope. Furthermore, we estimate the resulting uncertainties of the flux estimates. We also present different time series of dissolved CO2, CH4 and O2 concentration, <span class="hlt">water</span> temperature, barometric pressure, electro conductivity, and pH values</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2954549','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2954549"><span>Modified Perfluorocarbon Tracer Method for Measuring Effective Multizone <span class="hlt">Air</span> <span class="hlt">Exchange</span> Rates</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Shinohara, Naohide; Kataoka, Toshiyuki; Takamine, Koichi; Butsugan, Michio; Nishijima, Hirokazu; Gamo, Masashi</p> <p>2010-01-01</p> <p>A modified procedure was developed for the measurement of the effective <span class="hlt">air</span> <span class="hlt">exchange</span> rate, which represents the relationship between the pollutants emitted from indoor sources and the residents’ level of exposure, by placing the dosers of tracer <span class="hlt">gas</span> at locations that resemble indoor emission sources. To measure the 24-h-average effective <span class="hlt">air</span> <span class="hlt">exchange</span> rates in future surveys based on this procedure, a low-cost, easy-to-use perfluorocarbon tracer (PFT) doser with a stable dosing rate was developed by using double glass vials, a needle, a polyethylene-sintered filter, and a diffusion tube. Carbon molecular sieve cartridges and carbon disulfide (CS2) were used for passive sampling and extraction of the tracer <span class="hlt">gas</span>, respectively. Recovery efficiencies, sampling rates, and lower detection limits for 24-h sampling of hexafluorobenzene, octafluorotoluene, and perfluoroallylbenzene were 40% ± 3%, 72% ± 5%, and 84% ± 6%; 10.5 ± 1.1, 14.4 ± 1.4, and 12.2 ± 0.49 mL min−1; and 0.20, 0.17, and 0.26 μg m−3, respectively. PMID:20948928</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20948928','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20948928"><span>Modified perfluorocarbon tracer method for measuring effective multizone <span class="hlt">air</span> <span class="hlt">exchange</span> rates.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shinohara, Naohide; Kataoka, Toshiyuki; Takamine, Koichi; Butsugan, Michio; Nishijima, Hirokazu; Gamo, Masashi</p> <p>2010-09-01</p> <p>A modified procedure was developed for the measurement of the effective <span class="hlt">air</span> <span class="hlt">exchange</span> rate, which represents the relationship between the pollutants emitted from indoor sources and the residents' level of exposure, by placing the dosers of tracer <span class="hlt">gas</span> at locations that resemble indoor emission sources. To measure the 24-h-average effective <span class="hlt">air</span> <span class="hlt">exchange</span> rates in future surveys based on this procedure, a low-cost, easy-to-use perfluorocarbon tracer (PFT) doser with a stable dosing rate was developed by using double glass vials, a needle, a polyethylene-sintered filter, and a diffusion tube. Carbon molecular sieve cartridges and carbon disulfide (CS₂) were used for passive sampling and extraction of the tracer <span class="hlt">gas</span>, respectively. Recovery efficiencies, sampling rates, and lower detection limits for 24-h sampling of hexafluorobenzene, octafluorotoluene, and perfluoroallylbenzene were 40% ± 3%, 72% ± 5%, and 84% ± 6%; 10.5 ± 1.1, 14.4 ± 1.4, and 12.2 ± 0.49 mL min⁻¹; and 0.20, 0.17, and 0.26 μg m⁻³, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28313925','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28313925"><span>Effects of plant size and <span class="hlt">water</span> relations on <span class="hlt">gas</span> <span class="hlt">exchange</span> and growth of the desert shrub Larrea tridentata.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Franco, A C; de Soyza, A G; Virginia, R A; Reynolds, J F; Whitford, W G</p> <p>1994-03-01</p> <p>Larrea tridentata is a xerophytic evergreen shrub, dominant in the arid regions of the southwestern United States. We examined relationships between gasexchange characteristics, plant and soil <span class="hlt">water</span> relations, and growth responses of large versus small shrubs of L. tridentata over the course of a summer growing season in the Chihuahuan Desert of southern New Mexico, USA. The soil wetting front did not reach 0.6 m, and soils at depths of 0.6 and 0.9 m remained dry throughout the summer, suggesting that L. tridentata extracts <span class="hlt">water</span> largely from soil near the surface. Surface soil layers (<0.3 m) were drier under large plants, but predawn xylem <span class="hlt">water</span> potentials were similar for both plant sizes suggesting some access to deeper soil moisture reserves by large plants. Stem elongation rates were about 40% less in large, reproductively active shrubs than in small, reproductively inactive shrubs. Maximal net photosynthetic rates (P max ) occurred in early summer (21.3 μ mol m -2 s -1 ), when pre-dawn xylem <span class="hlt">water</span> potential (XWP) reached ca. -1 MPa. Although both shrub sizes exhibited similar responses to environmental factors, small shrubs recovered faster from short-term drought, when pre-dawn XWP reached about -4.5 MPa and P max decreased to only ca. 20% of unstressed levels. <span class="hlt">Gas</span> <span class="hlt">exchange</span> measurements yielded a strong relationship between stomatal conductance and photosynthesis, and the relationship between leaf-to-<span class="hlt">air</span> vapor pressure deficit and stomatal conductance was found to be influenced by pre-dawn XWP. Our results indicate that stomatal responses to <span class="hlt">water</span> stress and vapor pressure deficit are important in determining rates of carbon gain and <span class="hlt">water</span> loss in L. tridentata.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28653741','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28653741"><span><span class="hlt">Gas</span> <span class="hlt">exchanges</span> and <span class="hlt">water</span> use efficiency in the selection of tomato genotypes tolerant to <span class="hlt">water</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>Borba, M E A; Maciel, G M; Fraga Júnior, E F; Machado Júnior, C S; Marquez, G R; Silva, I G; Almeida, R S</p> <p>2017-06-20</p> <p><span class="hlt">Water</span> stress can affect the yield in tomato crops and, despite this, there are few types of research aiming to select tomato genotypes resistant to the <span class="hlt">water</span> stress using physiological parameters. This experiment aimed to study the variables that are related to the <span class="hlt">gas</span> <span class="hlt">exchanges</span> and the efficiency in <span class="hlt">water</span> use, in the selection of tomato genotypes tolerant to <span class="hlt">water</span> stress. It was done in a greenhouse, measuring 7 x 21 m, in a randomized complete block design, with four replications (blocks), being five genotypes in the F 2 BC 1 generation, which were previously obtained from an interspecific cross between Solanum pennellii versus S. lycopersicum and three check treatments, two susceptible [UFU-22 (pre-commercial line) and cultivar Santa Clara] and one resistant (S. pennellii). At the beginning of flowering, the plants were submitted to a <span class="hlt">water</span> stress condition, through irrigation suspension. After that CO 2 assimilation, internal CO 2 , stomatal conductance, transpiration, leaf temperature, instantaneous <span class="hlt">water</span> use efficiency, intrinsic efficiency of <span class="hlt">water</span> use, instantaneous carboxylation efficiency, chlorophyll a and b, and the potential leaf <span class="hlt">water</span> (Ψf) were observed. Almost all variables that were analyzed, except CO 2 assimilation and instantaneous carboxylation efficiency, demonstrated the superiority of the wild accession, S. pennellii, concerning the susceptible check treatments. The high photosynthetic rate and the low stomatal conductance and transpiration, presented by the UFU22/F 2 BC 1 #2 population, allowed a better <span class="hlt">water</span> use efficiency. Because of that, these physiological characteristics are promising in the selection of tomato genotypes tolerant to <span class="hlt">water</span> stress.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21827644','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21827644"><span>Performance evaluation on an <span class="hlt">air</span>-cooled heat <span class="hlt">exchanger</span> for alumina nanofluid under laminar flow.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Teng, Tun-Ping; Hung, Yi-Hsuan; Teng, Tun-Chien; Chen, Jyun-Hong</p> <p>2011-08-09</p> <p>This study analyzes the characteristics of alumina (Al2O3)/<span class="hlt">water</span> nanofluid to determine the feasibility of its application in an <span class="hlt">air</span>-cooled heat <span class="hlt">exchanger</span> for heat dissipation for PEMFC or electronic chip cooling. The experimental sample was Al2O3/<span class="hlt">water</span> nanofluid produced by the direct synthesis method at three different concentrations (0.5, 1.0, and 1.5 wt.%). The experiments in this study measured the thermal conductivity and viscosity of nanofluid with weight fractions and sample temperatures (20-60°C), and then used the nanofluid in an actual <span class="hlt">air</span>-cooled heat <span class="hlt">exchanger</span> to assess its heat <span class="hlt">exchange</span> capacity and pressure drop under laminar flow. Experimental results show that the nanofluid has a higher heat <span class="hlt">exchange</span> capacity than <span class="hlt">water</span>, and a higher concentration of nanoparticles provides an even better ratio of the heat <span class="hlt">exchange</span>. The maximum enhanced ratio of heat <span class="hlt">exchange</span> and pressure drop for all the experimental parameters in this study was about 39% and 5.6%, respectively. In addition to nanoparticle concentration, the temperature and mass flow rates of the working fluid can affect the enhanced ratio of heat <span class="hlt">exchange</span> and pressure drop of nanofluid. The cross-section aspect ratio of tube in the heat <span class="hlt">exchanger</span> is another important factor to be taken into consideration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3212002','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3212002"><span>Performance evaluation on an <span class="hlt">air</span>-cooled heat <span class="hlt">exchanger</span> for alumina nanofluid under laminar flow</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2011-01-01</p> <p>This study analyzes the characteristics of alumina (Al2O3)/<span class="hlt">water</span> nanofluid to determine the feasibility of its application in an <span class="hlt">air</span>-cooled heat <span class="hlt">exchanger</span> for heat dissipation for PEMFC or electronic chip cooling. The experimental sample was Al2O3/<span class="hlt">water</span> nanofluid produced by the direct synthesis method at three different concentrations (0.5, 1.0, and 1.5 wt.%). The experiments in this study measured the thermal conductivity and viscosity of nanofluid with weight fractions and sample temperatures (20-60°C), and then used the nanofluid in an actual <span class="hlt">air</span>-cooled heat <span class="hlt">exchanger</span> to assess its heat <span class="hlt">exchange</span> capacity and pressure drop under laminar flow. Experimental results show that the nanofluid has a higher heat <span class="hlt">exchange</span> capacity than <span class="hlt">water</span>, and a higher concentration of nanoparticles provides an even better ratio of the heat <span class="hlt">exchange</span>. The maximum enhanced ratio of heat <span class="hlt">exchange</span> and pressure drop for all the experimental parameters in this study was about 39% and 5.6%, respectively. In addition to nanoparticle concentration, the temperature and mass flow rates of the working fluid can affect the enhanced ratio of heat <span class="hlt">exchange</span> and pressure drop of nanofluid. The cross-section aspect ratio of tube in the heat <span class="hlt">exchanger</span> is another important factor to be taken into consideration. PMID:21827644</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15859112','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15859112"><span>Measurement of the oxygen mass transfer through the <span class="hlt">air-water</span> interface.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mölder, Erik; Mashirin, Alelxei; Tenno, Toomas</p> <p>2005-01-01</p> <p><span class="hlt">Gas</span> mass transfer through the liquid-<span class="hlt">gas</span> interface has enormous importance in various natural and industrial processes. Surfactants or insoluble compounds adsorbed onto an interface will inhibit the <span class="hlt">gas</span> mass transfer through the liquid-<span class="hlt">gas</span> surface. This study presents a technique for measuring the oxygen mass transfer through the <span class="hlt">air-water</span> interface. Experimental data obtained with the measuring device were incorporated into a novel mathematical model, which allowed one to calculate diffusion conduction of liquid surface layer and oxygen mass transfer coefficient in the liquid surface layer. A special measurement cell was constructed. The most important part of the measurement cell is a chamber containing the electrochemical oxygen sensor inside it. <span class="hlt">Gas</span> <span class="hlt">exchange</span> between the volume of the chamber and the external environment takes place only through the investigated surface layer. Investigated liquid was deoxygenated, which triggers the oxygen mass transfer from the chamber through the liquid-<span class="hlt">air</span> interface into the liquid phase. The decrease of oxygen concentration in the cell during time was measured. By using this data it is possible to calculate diffusional parameters of the <span class="hlt">water</span> surface layer. Diffusion conduction of oxygen through the <span class="hlt">air-water</span> surface layer of selected wastewaters was measured. The diffusion conduction of different wastewaters was about 3 to 6 times less than in the unpolluted <span class="hlt">water</span> surface. It was observed that the dilution of wastewater does not have a significant impact on the oxygen diffusion conduction through the wastewater surface layer. This fact can be explained with the presence of the compounds with high surface activity in the wastewater. Surfactants achieved a maximum adsorption and, accordingly, the maximum decrease of oxygen permeability already at a very low concentration of surfactants in the solution. Oxygen mass transfer coefficient of the surface layer of the <span class="hlt">water</span> is found to be Ds/ls = 0.13 x 10(-3) x cm/s. A simple</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ECSS..176....1M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ECSS..176....1M"><span>Temporal variability of <span class="hlt">air</span>-sea CO2 <span class="hlt">exchange</span> in a low-emission estuary</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mørk, Eva Thorborg; Sejr, Mikael Kristian; Stæhr, Peter Anton; Sørensen, Lise Lotte</p> <p>2016-07-01</p> <p>There is the need for further study of whether global estimates of <span class="hlt">air</span>-sea CO2 <span class="hlt">exchange</span> in estuarine systems capture the relevant temporal variability and, as such, the temporal variability of bulk parameterized and directly measured CO2 fluxes was investigated in the Danish estuary, Roskilde Fjord. The <span class="hlt">air</span>-sea CO2 fluxes showed large temporal variability across seasons and between days and that more than 30% of the net CO2 emission in 2013 was a result of two large fall and winter storms. The diurnal variability of ΔpCO2 was up to 400 during summer changing the estuary from a source to a sink of CO2 within the day. Across seasons the system was suggested to change from a sink of atmospheric CO2 during spring to near neutral during summer and later to a source of atmospheric CO2 during fall. Results indicated that Roskilde Fjord was an annual low-emission estuary, with an estimated bulk parameterized release of 3.9 ± 8.7 mol CO2 m-2 y-1 during 2012-2013. It was suggested that the production-respiration balance leading to the low annual emission in Roskilde Fjord, was caused by the shallow depth, long residence time and high <span class="hlt">water</span> quality in the estuary. In the data analysis the eddy covariance CO2 flux samples were filtered according to the H2Osbnd CO2 cross-sensitivity assessment suggested by Landwehr et al. (2014). This filtering reduced episodes of contradicting directions between measured and bulk parameterized <span class="hlt">air</span>-sea CO2 <span class="hlt">exchanges</span> and changed the net <span class="hlt">air</span>-sea CO2 <span class="hlt">exchange</span> from an uptake to a release. The CO2 <span class="hlt">gas</span> transfer velocity was calculated from directly measured CO2 fluxes and ΔpCO2 and agreed to previous observations and parameterizations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.B11K..06A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.B11K..06A"><span>Limitations on <span class="hlt">gas</span> <span class="hlt">exchange</span> recovery following natural drought in Californian oak woodlands.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ackerly, D.; Skelton, R. P.; Dawson, T.; Thompson, S.; Feng, X.; Weitz, A.; McLaughlin, B.</p> <p>2017-12-01</p> <p>Abstract Background/Question/Methods Drought can cause major damage to plant communities, but species damage thresholds and post-drought recovery of forest productivity are not yet predictable. We asked the question how should forest net primary productivity recover following exposure to severe drought? We used a natural drought period to investigate whether drought responses and post-drought recovery of canopy health could be predicted by properties of the <span class="hlt">water</span> transport system. We aimed to test the hypothesis that recovery of <span class="hlt">gas</span> <span class="hlt">exchange</span> and canopy health would be most severely limited by xylem embolism in stems. To do this we monitored leaf level <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> status for multiple individuals of two deciduous and two evergreen species for four years spanning a severe drought event and following subsequent rehydration. Results/Discussion Severe drought caused major declines in leaf <span class="hlt">water</span> potential, reduced stomatal conductance and assimilation rates and increased canopy bareness in our four canopy species. <span class="hlt">Water</span> potential surpassed levels associated with incipient embolism in leaves of most trees. In contrast, due to hydraulic segmentation, <span class="hlt">water</span> potential only rarely surpassed critical thresholds in the stems of the study trees. Individuals that surpassed critical thresholds of embolism in the stem displayed significant canopy dieback and mortality. Thus, recovery of plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and canopy health was predicted by xylem safety margin in stems, but not leaves, providing strong support for stem cavitation vulnerability as an index of damage under natural drought conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28250584','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28250584"><span>Brassinosteroids improve photosystem II efficiency, <span class="hlt">gas</span> <span class="hlt">exchange</span>, antioxidant enzymes and growth of cowpea plants exposed to <span class="hlt">water</span> deficit.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lima, J V; Lobato, A K S</p> <p>2017-01-01</p> <p><span class="hlt">Water</span> deficit is considered the main abiotic stress that limits agricultural production worldwide. Brassinosteroids (BRs) are natural substances that play roles in plant tolerance against abiotic stresses, including <span class="hlt">water</span> deficit. This research aims to determine whether BRs can mitigate the negative effects caused by <span class="hlt">water</span> deficiency, revealing how BRs act and their possible contribution to increased tolerance of cowpea plants to <span class="hlt">water</span> deficit. The experiment was a factorial design with the factors completely randomised, with two <span class="hlt">water</span> conditions (control and <span class="hlt">water</span> deficit) and three levels of brassinosteroids (0, 50 and 100 nM 24-epibrassinolide; EBR is an active BRs). Plants sprayed with 100 nM EBR under the <span class="hlt">water</span> deficit presented significant increases in Φ PSII , q P and ETR compared with plants subjected to the <span class="hlt">water</span> deficit without EBR. With respect to <span class="hlt">gas</span> <span class="hlt">exchange</span>, P N , E and g s exhibited significant reductions after <span class="hlt">water</span> deficit, but application of 100 nM EBR caused increases in these variables of 96, 24 and 33%, respectively, compared to the <span class="hlt">water</span> deficit + 0 nM EBR treatment. To antioxidant enzymes, EBR resulted in increases in SOD, CAT, APX and POX, indicating that EBR acts on the antioxidant system, reducing cell damage. The <span class="hlt">water</span> deficit caused significant reductions in Chl a , Chl b and total Chl, while plants sprayed with 100 nM EBR showed significant increases of 26, 58 and 33% in Chl a , Chl b and total Chl, respectively. This study revealed that EBR improves photosystem II efficiency, inducing increases in Φ PSII , q P and ETR. This substance also mitigated the negative effects on <span class="hlt">gas</span> <span class="hlt">exchange</span> and growth induced by the <span class="hlt">water</span> deficit. Increases in SOD, CAT, APX and POX of plants treated with EBR indicate that this steroid clearly increased the tolerance to the <span class="hlt">water</span> deficit, reducing reactive oxygen species, cell damage, and maintaining the photosynthetic pigments. Additionally, 100 nM EBR resulted in a better dose-response of cowpea</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3234696','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3234696"><span>Removal of infused <span class="hlt">water</span> predominantly during insertion (<span class="hlt">water</span> <span class="hlt">exchange</span>) is consistently associated with an increase in adenoma detection rate - review of data in randomized controlled trials (RCTs) of <span class="hlt">water</span>-related methods</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Harker, JO; Leung, JW; Siao-Salera, RM; Mann, SK; Ramirez, FC; Friedland, S; Amato, A; Radaelli, F; Paggi, S; Terruzzi, V; Hsieh, YH</p> <p>2011-01-01</p> <p>Introduction Variation in outcomes in RcTs comparing <span class="hlt">water</span>-related methods and <span class="hlt">air</span> insufflation raises challenging questions regarding the new approach. This report reviews impact of <span class="hlt">water</span> <span class="hlt">exchange</span> - simultaneous infusion and removal of infused <span class="hlt">water</span> during insertion on adenoma detection rate (ADR) defined as proportion of patients with a least one adenoma of any size. Methods Medline (2008–2011) searches, abstract of 2011 Digestive Disease Week (DDW) meeting and personal communications were considered to identify RcTs that compared <span class="hlt">water</span>-related methods and <span class="hlt">air</span> insufflation to aid insertion of colonoscope. Results Since 2008, eleven reports of RcTs (6 published, 1 submitted and 4 abstracts, n=1728) described ADR in patients randomized to be examined by <span class="hlt">air</span> and <span class="hlt">water</span>-related methods. The <span class="hlt">water</span>-related methods differed in timing of removal of the infused <span class="hlt">water</span> -predominantly during insertion (<span class="hlt">water</span> <span class="hlt">exchange</span>) (n=825) or predominantly during withdrawal (<span class="hlt">water</span> immersion) (n=903). <span class="hlt">Water</span> immersion was associated with both increases and decreases in ADR compared to respective <span class="hlt">air</span> method patients and the net overall change (-7%) was significant. On the other hand <span class="hlt">water</span> <span class="hlt">exchange</span> was associated with increases in ADR consistently and the net changes (overall, 8%; proximal overall, 11%; and proximal <10 mm, 12%) were all significant. Conclusion Comparative data generated the hypothesis that significantly larger increases in overall and proximal colon ADRs were associated with <span class="hlt">water</span> <span class="hlt">exchange</span> than <span class="hlt">water</span> immersion or <span class="hlt">air</span> insufflation during insertion. The hypothesis should be evaluated by RCTs to elucidate the mechanism of <span class="hlt">water</span> <span class="hlt">exchange</span> on adenoma detection. PMID:22163082</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23461476','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23461476"><span>Regulation and acclimation of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> in a piñon-juniper woodland exposed to three different precipitation regimes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Limousin, Jean-Marc; Bickford, Christopher P; Dickman, Lee T; Pangle, Robert E; Hudson, Patrick J; Boutz, Amanda L; Gehres, Nathan; Osuna, Jessica L; Pockman, William T; McDowell, Nate G</p> <p>2013-10-01</p> <p>Leaf <span class="hlt">gas-exchange</span> regulation plays a central role in the ability of trees to survive drought, but forecasting the future response of <span class="hlt">gas</span> <span class="hlt">exchange</span> to prolonged drought is hampered by our lack of knowledge regarding potential acclimation. To investigate whether leaf <span class="hlt">gas-exchange</span> rates and sensitivity to drought acclimate to precipitation regimes, we measured the seasonal variations of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> in a mature piñon-juniper Pinus edulis-Juniperus monosperma woodland after 3 years of precipitation manipulation. We compared trees receiving ambient precipitation with those in an irrigated treatment (+30% of ambient precipitation) and a partial rainfall exclusion (-45%). Treatments significantly affected leaf <span class="hlt">water</span> potential, stomatal conductance and photosynthesis for both isohydric piñon and anisohydric juniper. Leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> acclimated to the precipitation regimes in both species. Maximum <span class="hlt">gas-exchange</span> rates under well-<span class="hlt">watered</span> conditions, leaf-specific hydraulic conductance and leaf <span class="hlt">water</span> potential at zero photosynthetic assimilation all decreased with decreasing precipitation. Despite their distinct drought resistance and stomatal regulation strategies, both species experienced hydraulic limitation on leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> when precipitation decreased, leading to an intraspecific trade-off between maximum photosynthetic assimilation and resistance of photosynthesis to drought. This response will be most detrimental to the carbon balance of piñon under predicted increases in aridity in the southwestern USA. © 2013 John Wiley & Sons Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12857848','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12857848"><span>Hydraulic properties of rice and the response of <span class="hlt">gas</span> <span class="hlt">exchange</span> to <span class="hlt">water</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>Stiller, Volker; Lafitte, H Renee; Sperry, John S</p> <p>2003-07-01</p> <p>We investigated the role of xylem cavitation, plant hydraulic conductance, and root pressure in the response of rice (Oryza sativa) <span class="hlt">gas</span> <span class="hlt">exchange</span> to <span class="hlt">water</span> stress. In the field (Philippines), the percentage loss of xylem conductivity (PLC) from cavitation exceeded 60% in leaves even in <span class="hlt">watered</span> controls. The PLC versus leaf <span class="hlt">water</span> potential relationship indicated diurnal refilling of cavitated xylem. The leaf <span class="hlt">water</span> potential causing 50 PLC (P(50)) was -1.6 MPa and did not differ between upland versus lowland rice varieties. Greenhouse-grown varieties (Utah) were more resistant to cavitation with a 50 PLC of -1.9 MPa but also showed no difference between varieties. Six-day droughts caused concomitant reductions in leaf-specific photosynthetic rate, leaf diffusive conductance, and soil-leaf hydraulic conductance that were associated with cavitation-inducing <span class="hlt">water</span> potentials and the disappearance of nightly root pressure. The return of root pressure after drought was associated with the complete recovery of leaf diffusive conductance, leaf-specific photosynthetic rate, and soil-leaf hydraulic conductance. Root pressure after the 6-d drought (61.2 +/- 8.8 kPa) was stimulated 7-fold compared with well-<span class="hlt">watered</span> plants before drought (8.5 +/- 3.8 kPa). The results indicate: (a) that xylem cavitation plays a major role in the reduction of plant hydraulic conductance during drought, and (b) that rice can readily reverse cavitation, possibly aided by nocturnal root pressure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26710629','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26710629"><span>[Effects of different <span class="hlt">water</span> potentials on leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and chlorophyll fluorescence parameters of cucumber during post-flowering growth stage].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lin, Lu; Tang, Yun; Zhang, Ji-tao; Yan, Wan-li; Xiao, Jian-hong; Ding, Chao; Dong, Chuan; Ji, Zeng-shun</p> <p>2015-07-01</p> <p>Impacts of different substrate <span class="hlt">water</span> potentials (SWP) on leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and chlorophyll fluorescence parameters of greenhouse cucumber during its post-flowering growth stage were analyzed in this study. The results demonstrated that -10 and -30 kPa were the critical values for initiating stomatal and non-stomatal limitation of drought stress, respectively. During the stage of no drought stress (-10 kPa < SWP ≤ 0 kPa), <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters and chlorophyll fluorescence parameters were not different significantly among treatments. During the stage of stomatal limitation of drought stress (-30 kPa<SWP ≤ -10 kPa), with the decrease of SWP, the stomatal conductance (gs), intercellular carbon dioxide concentration (Ci), net photosynthetic rate (Pn) , apparent quantum efficiency (ε), transpiration rate (Tr), carboxylation efficiency (CE), maximum Rubisco-limited rate of carboxylation (Vc max), maximum rate of electron transport (Jmax), rate of triosephosphate utilization (VTPU), maximum and actual quantum efficiency of PSII (ΦPSII, and Fv/Fm) and photochemical quenching (qP) decreased, but the light compensation point (LCP), dark respiration rate (Rd), carbon dioxide compensation point (CCP), stomatal limitation value (LS), instantaneous <span class="hlt">water</span> use efficiency (WUEi) and non-photochemical quenching (qN) increased. In this stage, <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters changed faster than chlorophyll fluorescence parameters and differed significantly among treatments. During the stage of non-stomatal limitation of drought stress (-45 kPa≤SWP ≤ -30 kPa), with the decrease of SWP, light saturation point (LSP), Rd, CE, Vcmax, VTPU, LS, WUEi, ΦpPSII, Fv/Fm and qp decreased, while CCP, Ci and qN increased. In this stage, chlorophyll fluorescence parameters changed faster than <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters and differed significantly among treatments. In production of greenhouse cucumber, -10 and -5 kPa should be the lower and upper limit value of irrigation, respectively. The stomatal</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=546951','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=546951"><span><span class="hlt">Gas</span> <span class="hlt">Exchange</span> of Algae</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ammann, Elizabeth C. B.; Lynch, Victoria H.</p> <p>1967-01-01</p> <p>The oxygen production of a photosynthetic <span class="hlt">gas</span> <span class="hlt">exchanger</span> containing Chlorella pyrenoidosa (1% packed cell volume) was measured when various concentrations of carbon dioxide were present within the culture unit. The internal carbon dioxide concentrations were obtained by manipulating the entrance <span class="hlt">gas</span> concentration and the flow rate. Carbon dioxide percentages were monitored by means of electrodes placed directly in the nutrient medium. The concentration of carbon dioxide in the nutrient medium which produced maximal photosynthesis was in the range of 1.5 to 2.5% by volume. Results were unaffected by either the level of carbon dioxide in the entrance <span class="hlt">gas</span> or the rate of <span class="hlt">gas</span> flow. Entrance gases containing 2% carbon dioxide flowing at 320 ml/min, 3% carbon dioxide at 135 ml/min, and 4% carbon dioxide at 55 ml/min yielded optimal carbon dioxide concentrations in the particular unit studied. By using carbon dioxide electrodes implanted directly in the <span class="hlt">gas</span> <span class="hlt">exchanger</span> to optimize the carbon dioxide concentration throughout the culture medium, it should be possible to design more efficient large-scale units. PMID:4382391</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/2013JPS...243..946K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JPS...243..946K"><span>Numerical investigation of interfacial transport resistance due to <span class="hlt">water</span> droplets in proton <span class="hlt">exchange</span> membrane fuel cell <span class="hlt">air</span> channels</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Koz, Mustafa; Kandlikar, Satish G.</p> <p>2013-12-01</p> <p>Oxygen transport resistance at the <span class="hlt">air</span> flow channel and <span class="hlt">gas</span> diffusion layer (GDL) interface is needed in modelling the performance of a proton <span class="hlt">exchange</span> membrane fuel cell (PEMFC). This resistance is expressed through the non-dimensional Sherwood number (Sh). The effect of the presence of a droplet on Sh is studied numerically in an isolated <span class="hlt">air</span> flow channel using a commercially available package, COMSOL Multiphysics®. A droplet is represented as a solid obstruction placed on the GDL-channel interface and centred along the channel width. The effect of a single droplet is first studied for a range of superficial mean <span class="hlt">air</span> velocities and droplet sizes. Secondly, the effect of droplet spacing on Sh is studied through simulations of two consecutive droplets. Lastly, multiple droplets in a row are studied as a more representative case of a PEMFC <span class="hlt">air</span> flow channel. The results show that the droplets significantly increase Sh above the fully developed value in the wake region. This enhancement increases with the number of droplets, droplet size, and superficial mean <span class="hlt">air</span> velocity. Moreover, the analogy between mass and heat transfer is investigated by comparing Sh to the equivalent Nusselt number.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25311102','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25311102"><span>Nondestructive natural <span class="hlt">gas</span> hydrate recovery driven by <span class="hlt">air</span> and carbon dioxide.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kang, Hyery; Koh, Dong-Yeun; Lee, Huen</p> <p>2014-10-14</p> <p>Current technologies for production of natural <span class="hlt">gas</span> hydrates (NGH), which include thermal stimulation, depressurization and inhibitor injection, have raised concerns over unintended consequences. The possibility of catastrophic slope failure and marine ecosystem damage remain serious challenges to safe NGH production. As a potential approach, this paper presents <span class="hlt">air</span>-driven NGH recovery from permeable marine sediments induced by simultaneous mechanisms for methane liberation (NGH decomposition) and CH₄-<span class="hlt">air</span> or CH₄-CO₂/<span class="hlt">air</span> replacement. <span class="hlt">Air</span> is diffused into and penetrates NGH and, on its surface, forms a boundary between the <span class="hlt">gas</span> and solid phases. Then spontaneous melting proceeds until the chemical potentials become equal in both phases as NGH depletion continues and self-regulated CH4-<span class="hlt">air</span> replacement occurs over an arbitrary point. We observed the existence of critical methane concentration forming the boundary between decomposition and replacement mechanisms in the NGH reservoirs. Furthermore, when CO₂ was added, we observed a very strong, stable, self-regulating process of <span class="hlt">exchange</span> (CH₄ replaced by CO₂/<span class="hlt">air</span>; hereafter CH₄-CO₂/<span class="hlt">air</span>) occurring in the NGH. The proposed process will work well for most global <span class="hlt">gas</span> hydrate reservoirs, regardless of the injection conditions or geothermal gradient.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JaJAP..57a02BCI','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JaJAP..57a02BCI"><span>Simultaneous generation of acidic and alkaline <span class="hlt">water</span> using atmospheric <span class="hlt">air</span> plasma formed in <span class="hlt">water</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Imai, Shin-ichi; Sakaguchi, Yoshihiro; Shirafuji, Tatsuru</p> <p>2018-01-01</p> <p>Plasmas on <span class="hlt">water</span> surfaces and in <span class="hlt">water</span> can be generated at atmosphere pressure using several kinds of gases, including helium, argon, oxygen, and <span class="hlt">air</span>. Nitrates are generated in <span class="hlt">water</span> through the interaction between <span class="hlt">water</span> and atmospheric plasma that uses ambient <span class="hlt">air</span>. <span class="hlt">Water</span> that has been made acidic by the generation of nitric acid and the acidic <span class="hlt">water</span> can be used for the sterilization of medical instruments, toilet bowls, and washing machines. Dishwashers are another potential application, as alkaline <span class="hlt">water</span> is needed to remove grease from tableware. To investigate the production of alkaline <span class="hlt">water</span> and its mechanism, <span class="hlt">gas</span> component analysis was performed using an atmospheric quadrupole mass spectrometer. It was found that hydrogen <span class="hlt">gas</span> evolves from the <span class="hlt">water</span> surrounding both the positive and negative electrodes. The <span class="hlt">gas</span> and <span class="hlt">water</span> analyses carried out in this study revealed that acidic <span class="hlt">water</span> of pH 2.5 and alkaline <span class="hlt">water</span> of pH 10 can be simultaneously generated by our ambient <span class="hlt">air</span> plasma device, which has been altered from our original model. The alterative plasma device has a partition wall, which is made of conductive resin, between the positive and negative electrodes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24222707','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24222707"><span>Ex vivo assessment and validation of <span class="hlt">water</span> <span class="hlt">exchange</span> performance of 23 heat and moisture <span class="hlt">exchangers</span> for laryngectomized patients.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>van den Boer, Cindy; Muller, Sara H; Vincent, Andrew D; van den Brekel, Michiel W M; Hilgers, Frans J M</p> <p>2014-08-01</p> <p>Breathing through a tracheostoma results in insufficient warming and humidification of the inspired <span class="hlt">air</span>. This loss of <span class="hlt">air</span> conditioning, especially humidification, can be partially restored with the application of a heat and moisture <span class="hlt">exchanger</span> (HME) over the tracheostoma. For medical professionals, it is not easy to judge differences in <span class="hlt">water</span> <span class="hlt">exchange</span> performance of various HMEs owing to the lack of universal outcome measures. This study has three aims: assessment of the <span class="hlt">water</span> <span class="hlt">exchange</span> performance of commercially available HMEs for laryngectomized patients, validation of these results with absolute humidity outcomes, and assessment of the role of hygroscopic salt present in some of the tested HMEs. Measurements of weight and absolute humidity at end inspiration and end expiration at different breathing volumes of a healthy volunteer were performed using a microbalance and humidity sensor. Twenty-three HMEs from 6 different manufacturers were tested. Associations were determined between core weight, weight change, breathing volume, and absolute humidity, using both linear and nonlinear mixed effects models. <span class="hlt">Water</span> <span class="hlt">exchange</span> of the 23 HMEs at a breathing volume of 0.5 L varies between 0.5 and 3.6 mg. Both <span class="hlt">water</span> <span class="hlt">exchange</span> and wet core weight correlate strongly with the end-inspiratory absolute humidity values (r2 =0.89/0.87). Hygroscopic salt increases core weight. The 23 tested HMEs for laryngectomized patients show wide variation in <span class="hlt">water</span> <span class="hlt">exchange</span> performance. <span class="hlt">Water</span> <span class="hlt">exchange</span> correlates well with the end-inspiratory absolute humidity outcome, which validates the ex vivo weight change method. Wet core weight is a predictor of HME performance. Hygroscopic salt increases the weight of the core material. The results of this study can help medical professionals to obtain a more founded opinion about the performance of available HMEs for pulmonary rehabilitation in laryngectomized patients, and allow them to make an informed decision about which HME type to use.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.B43D2148B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.B43D2148B"><span>Assessing Near-surface Heat, <span class="hlt">Water</span> Vapor and Carbon Dioxide <span class="hlt">Exchange</span> Over a Coastal Salt-marsh</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bogoev, I.; O'Halloran, T. L.; LeMoine, J.</p> <p>2017-12-01</p> <p>Coastal ecosystems play an important role in mitigating the effects of climate change by storing significant quantities of carbon. A growing number of studies suggest that vegetated estuarine habitats, specifically salt marshes, have high long-term rates of carbon sequestration, perhaps even higher than mature tropical and temperate forests. Large amounts of carbon, accumulated over thousands of years, are stored in the plant materials and sediment. Improved understanding of the factors that control energy and carbon <span class="hlt">exchange</span> is needed to better guide restoration and conservation management practices. To that end, we recently established an observation system to study marsh-atmosphere interactions within the North Inlet-Winyah Bay National Estuarine Research Reserve. Near-surface fluxes of heat, <span class="hlt">water</span> vapor (H2O) and carbon dioxide (CO2) were measured by an eddy-covariance system consisting of an aerodynamic open-path H2O / CO2 <span class="hlt">gas</span> analyzer with a spatially integrated 3D sonic anemometer/thermometer (IRGASON). The IRGASON instrument provides co-located and highly synchronized, fast response H2O, CO2 and <span class="hlt">air</span>- temperature measurements, which eliminates the need for spectral corrections associated with the separation between the sonic anemometer and the <span class="hlt">gas</span> analyzer. This facilitates calculating the instantaneous CO2 molar mixing ratio relative to dry <span class="hlt">air</span>. Fluxes computed from CO2 and H2O mixing ratios, which are conserved quantities, do not require post-processing corrections for <span class="hlt">air</span>-density changes associated with temperature and <span class="hlt">water</span> vapor fluctuations. These corrections are particularly important for CO2, because they could be even larger than the measured flux. Here we present the normalized frequency spectra of <span class="hlt">air</span> temperature, <span class="hlt">water</span> vapor and CO2, as well as their co-spectra with the co-located vertical wind. We also show mean daily cycles of sensible, latent and CO2 fluxes and analyze correlations with <span class="hlt">air/water</span> temperature, wind speed and light availability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/874299','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/874299"><span>Method and apparatus for extracting <span class="hlt">water</span> from <span class="hlt">air</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Spletzer, Barry L.; Callow, Diane Schafer; Marron, Lisa C.; Salton, Jonathan R.</p> <p>2002-01-01</p> <p>The present invention provides a method and apparatus for extracting liquid <span class="hlt">water</span> from moist <span class="hlt">air</span> using minimal energy input. The method comprises compressing moist <span class="hlt">air</span> under conditions that foster the condensation of liquid <span class="hlt">water</span>. The <span class="hlt">air</span> can be decompressed under conditions that do not foster the vaporization of the condensate. The decompressed, dried <span class="hlt">air</span> can be <span class="hlt">exchanged</span> for a fresh charge of moist <span class="hlt">air</span> and the process repeated. The liquid condensate can be removed for use. The apparatus can comprise a compression chamber having a variable internal volume. An intake port allows moist <span class="hlt">air</span> into the compression chamber. An exhaust port allows dried <span class="hlt">air</span> out of the compression chamber. A condensation device fosters condensation at the desired conditions. A condensate removal port allows liquid <span class="hlt">water</span> to be removed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21350390-two-phase-gas-liquid-flow-characteristics-inside-plate-heat-exchanger','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21350390-two-phase-gas-liquid-flow-characteristics-inside-plate-heat-exchanger"><span>Two-phase <span class="hlt">gas</span>-liquid flow characteristics inside a plate heat <span class="hlt">exchanger</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>Nilpueng, Kitti; Wongwises, Somchai</p> <p></p> <p>In the present study, the <span class="hlt">air-water</span> two-phase flow characteristics including flow pattern and pressure drop inside a plate heat <span class="hlt">exchanger</span> are experimentally investigated. A plate heat <span class="hlt">exchanger</span> with single pass under the condition of counter flow is operated for the experiment. Three stainless steel commercial plates with a corrugated sinusoidal shape of unsymmetrical chevron angles of 55 and 10 are utilized for the pressure drop measurement. A transparent plate having the same configuration as the stainless steel plates is cast and used as a cover plate in order to observe the flow pattern inside the plate heat <span class="hlt">exchanger</span>. The <span class="hlt">air</span>-watermore » mixture flow which is used as a cold stream is tested in vertical downward and upward flow. The results from the present experiment show that the annular-liquid bridge flow pattern appeared in both upward and downward flows. However, the bubbly flow pattern and the slug flow pattern are only found in upward flow and downward flow, respectively. The variation of the <span class="hlt">water</span> and <span class="hlt">air</span> velocity has a significant effect on the two-phase pressure drop. Based on the present data, a two-phase multiplier correlation is proposed for practical application. (author)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17874769','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17874769"><span><span class="hlt">Air</span>-sea <span class="hlt">exchange</span> fluxes of synthetic polycyclic musks in the North Sea and the Arctic.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xie, Zhiyong; Ebinghaus, Ralf; Temme, Christian; Heemken, Olaf; Ruck, Wolfgang</p> <p>2007-08-15</p> <p>Synthetic polycyclic musk fragrances Galaxolide (HHCB) and Tonalide (AHTN) were measured simultaneously in <span class="hlt">air</span> and seawater in the Arctic and the North Sea and in the rural <span class="hlt">air</span> of northern Germany. Median concentrations of <span class="hlt">gas</span>-phase HHCB and AHTN were 4 and 18 pg m(-3) in the Arctic, 28 and 18 pg m(-3) in the North Sea, and 71 and 21 pg m(-3) in northern Germany, respectively. Various ratios of HHCB/AHTN implied that HHCB is quickly removed by atmospheric degradation, while AHTN is relatively persistent in the atmosphere. Dissolved concentrations ranged from 12 to 2030 pg L(-1) for HHCB and from below the method detection limit (3 pg L(-1)) to 965 pg L(-1) for AHTN with median values of 59 and 23 pg L(-1), respectively. The medians of volatilization fluxes for HHCB and AHTN were 27.2 and 14.2 ng m(-2) day(-1) and the depositional fluxes were 5.9 and 3.3 ng m(-2) day(-1), respectively, indicating <span class="hlt">water-to-air</span> volatilization is a significant process to eliminate HHCB and AHTN from the North Sea. In the Arctic, deposition fluxes dominated the <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> of HHCB and AHTN, suggesting atmospheric input controls the levels of HHCB and AHTN in the polar region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRC..123.2293B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRC..123.2293B"><span>Wave Attenuation and <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Velocity in Marginal Sea Ice Zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bigdeli, A.; Hara, T.; Loose, B.; Nguyen, A. T.</p> <p>2018-03-01</p> <p>The <span class="hlt">gas</span> transfer velocity in marginal sea ice zones exerts a strong control on the input of anthropogenic gases into the ocean interior. In this study, a sea state-dependent <span class="hlt">gas</span> <span class="hlt">exchange</span> parametric model is developed based on the turbulent kinetic energy dissipation rate. The model is tuned to match the conventional <span class="hlt">gas</span> <span class="hlt">exchange</span> parametrization in fetch-unlimited, fully developed seas. Next, fetch limitation is introduced in the model and results are compared to fetch limited experiments in lakes, showing that the model captures the effects of finite fetch on <span class="hlt">gas</span> <span class="hlt">exchange</span> with good fidelity. Having validated the results in fetch limited <span class="hlt">waters</span> such as lakes, the model is next applied in sea ice zones using an empirical relation between the sea ice cover and the effective fetch, while accounting for the sea ice motion effect that is unique to sea ice zones. The model results compare favorably with the available field measurements. Applying this parametric model to a regional Arctic numerical model, it is shown that, under the present conditions, <span class="hlt">gas</span> flux into the Arctic Ocean may be overestimated by 10% if a conventional parameterization is used.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15959823','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15959823"><span>Environmental sensitivity of <span class="hlt">gas</span> <span class="hlt">exchange</span> in different-sized trees.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>McDowell, Nate G; Licata, Julian; Bond, Barbara J</p> <p>2005-08-01</p> <p>The carbon isotope signature (delta13C) of foliar cellulose from sunlit tops of trees typically becomes enriched as trees of the same species in similar environments grow taller, indicative of size-related changes in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>. However, direct measurements of <span class="hlt">gas</span> <span class="hlt">exchange</span> in common environmental conditions do not always reveal size-related differences, even when there is a distinct size-related trend in delta13C of the very foliage used for the <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements. Since delta13C of foliage predominately reflects <span class="hlt">gas</span> <span class="hlt">exchange</span> during spring when carbon is incorporated into leaf cellulose, this implies that <span class="hlt">gas</span> <span class="hlt">exchange</span> differences in different-sized trees are most likely to occur in favorable environmental conditions during spring. If <span class="hlt">gas</span> <span class="hlt">exchange</span> differs with tree size during wet but not dry conditions, then this further implies that environmental sensitivity of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> varies as a function of tree size. These implications are consistent with theoretical relationships among height, hydraulic conductance and <span class="hlt">gas</span> <span class="hlt">exchange</span>. We investigated the environmental sensitivity of <span class="hlt">gas</span> <span class="hlt">exchange</span> in different-sized Douglas-fir (Pseudotsuga menziesii) via a detailed process model that specifically incorporates size-related hydraulic conductance [soil-plant-atmosphere (SPA)], and empirical measurements from both wet and dry periods. SPA predicted, and the empirical measurements verified, that differences in <span class="hlt">gas</span> <span class="hlt">exchange</span> associated with tree size are greatest in wet and mild environmental conditions and minimal during drought. The results support the hypothesis that annual net carbon assimilation and transpiration of trees are limited by hydraulic capacity as tree size increases, even though at particular points in time there may be no difference in <span class="hlt">gas</span> <span class="hlt">exchange</span> between different-sized trees. Maximum net ecosystem <span class="hlt">exchange</span> occurs in spring in Pacific Northwest forests; therefore, the presence of hydraulic limitations during this period may play a large role</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRG..122.1011H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRG..122.1011H"><span><span class="hlt">Gas</span> transfer velocities in small forested ponds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Holgerson, Meredith A.; Farr, Emily R.; Raymond, Peter A.</p> <p>2017-05-01</p> <p>Inland <span class="hlt">waters</span> actively <span class="hlt">exchange</span> gases with the atmosphere, and the <span class="hlt">gas</span> <span class="hlt">exchange</span> rate informs system biogeochemistry, ecology, and global carbon budgets. <span class="hlt">Gas</span> <span class="hlt">exchange</span> in medium- to large-sized lakes is largely regulated by wind; yet less is known about processes regulating <span class="hlt">gas</span> transfer in small ponds where wind speeds are low. In this study, we determined the <span class="hlt">gas</span> transfer velocity, k600, in four small (<250 m2) ponds by using a propane (C3H8) <span class="hlt">gas</span> injection. When estimated across 12 h periods, the average k600 ranged from 0.19 to 0.72 m d-1 across the ponds. We also estimated k600 at 2 to 3 h intervals during the day and evaluated the relationship with environmental conditions. The average daytime k600 ranged from 0.33 to 1.83 m d-1 across the ponds and was best predicted by wind speed and <span class="hlt">air</span> or <span class="hlt">air-water</span> temperature; however, the explanatory power was weak (R2 < 0.27) with high variability within and among ponds. To compare our results to larger <span class="hlt">water</span> bodies, we compiled direct measurements of k600 from 67 ponds and lakes worldwide. Our k600 estimates were within the range of estimates for other small ponds, and variability in k600 increased with lake size. However, the majority of studies were conducted on medium-sized lakes (0.01 to 1 km2), leaving small ponds and large lakes understudied. Overall, this study adds four small ponds to the existing body of research on <span class="hlt">gas</span> transfer velocities from inland <span class="hlt">waters</span> and highlights uncertainty in k600, with implications for calculating metabolism and carbon emissions in inland <span class="hlt">waters</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4052457','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4052457"><span>Responses of sap flow, leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and growth of hybrid aspen to elevated atmospheric humidity under field conditions</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Niglas, Aigar; Kupper, Priit; Tullus, Arvo; Sellin, Arne</p> <p>2014-01-01</p> <p>An increase in average <span class="hlt">air</span> temperature and frequency of rain events is predicted for higher latitudes by the end of the 21st century, accompanied by a probable rise in <span class="hlt">air</span> humidity. We currently lack knowledge on how forest trees acclimate to rising <span class="hlt">air</span> humidity in temperate climates. We analysed the leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>, sap flow and growth characteristics of hybrid aspen (Populus tremula × P. tremuloides) trees growing at ambient and artificially elevated <span class="hlt">air</span> humidity in an experimental forest plantation situated in the hemiboreal vegetation zone. Humidification manipulation did not affect the photosynthetic capacity of plants, but did affect stomatal responses: trees growing at elevated <span class="hlt">air</span> humidity had higher stomatal conductance at saturating photosynthetically active radiation (gs sat) and lower intrinsic <span class="hlt">water</span>-use efficiency (IWUE). Reduced stomatal limitation of photosynthesis in trees grown at elevated <span class="hlt">air</span> humidity allowed slightly higher net photosynthesis and relative current-year height increments than in trees at ambient <span class="hlt">air</span> humidity. Tree responses suggest a mitigating effect of higher <span class="hlt">air</span> humidity on trees under mild <span class="hlt">water</span> stress. At the same time, trees at higher <span class="hlt">air</span> humidity demonstrated a reduced sensitivity of IWUE to factors inducing stomatal closure and a steeper decline in canopy conductance in response to <span class="hlt">water</span> deficit, implying higher dehydration risk. Despite the mitigating impact of increased <span class="hlt">air</span> humidity under moderate drought, a future rise in atmospheric humidity at high latitudes may be disadvantageous for trees during weather extremes and represents a potential threat in hemiboreal forest ecosystems. PMID:24887000</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/7836191','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/7836191"><span>Analysis of factors affecting <span class="hlt">gas</span> <span class="hlt">exchange</span> in intravascular blood <span class="hlt">gas</span> <span class="hlt">exchanger</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Niranjan, S C; Clark, J W; San, K Y; Zwischenberger, J B; Bidani, A</p> <p>1994-10-01</p> <p>A mathematical model of an intravascular hollow-fiber <span class="hlt">gas-exchange</span> device, called IVOX, has been developed using a Krogh cylinder-like approach with a repeating unit structure comprised of a single fiber with <span class="hlt">gas</span> flowing through its lumen surrounded by a coaxial cylinder of blood flowing in the opposite direction. Species mass balances on O2 and CO2 result in a nonlinear coupled set of convective-diffusion parabolic partial differential equations that are solved numerically using an alternating-direction implicit finite-difference method. Computed results indicated the presence of a large resistance to <span class="hlt">gas</span> transport on the external (blood) side of the hollow-fiber <span class="hlt">exchanger</span>. Increasing <span class="hlt">gas</span> flow through the device favored CO2 removal from but not O2 addition to blood. Increasing blood flow over the device favored both CO2 removal as well as O2 addition. The rate of CO2 removal increased linearly with the transmural PCO2 gradient imposed across the device. The effect of fiber crimping on blood phase mass transfer resistance was evaluated indirectly by varying species blood diffusivity. Computed results indicated that CO2 excretion by IVOX can be significantly enhanced with improved bulk mixing of vena caval blood around the IVOX fibers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=240307&keyword=water+AND+gas+AND+exchange&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=240307&keyword=water+AND+gas+AND+exchange&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>Controls on <span class="hlt">gas</span> transfer velocities in a large river</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>The emission of biogenic gases from large rivers can be an important component of regional greenhouse <span class="hlt">gas</span> budgets. However, emission rate estimates are often poorly constrained due to uncertainties in the <span class="hlt">air-water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> rate. We used the floating chamber method to estim...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.H53A1376G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.H53A1376G"><span>The Effect of Thermal Convection on Earth-Atmosphere CO2 <span class="hlt">Gas</span> <span class="hlt">Exchange</span> in Aggregated Soil</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ganot, Y.; Weisbrod, N.; Dragila, M. I.</p> <p>2011-12-01</p> <p><span class="hlt">Gas</span> transport in soils and surface-atmosphere <span class="hlt">gas</span> <span class="hlt">exchange</span> are important processes that affect different aspects of soil science such as soil aeration, nutrient bio-availability, sorption kinetics, soil and groundwater pollution and soil remediation. Diffusion and convection are the two main mechanisms that affect <span class="hlt">gas</span> transport, fate and emissions in the soils and in the upper vadose zone. In this work we studied CO2 soil-atmosphere <span class="hlt">gas</span> <span class="hlt">exchange</span> under both day-time and night-time conditions, focusing on the impact of thermal convection (TCV) during the night. Experiments were performed in a climate-controlled laboratory. One meter long columns were packed with matrix of different grain size (sand, gravel and soil aggregates). <span class="hlt">Air</span> with 2000 ppm CO2 was injected into the bottom of the columns and CO2 concentration within the columns was continuously monitored by an Infra Red <span class="hlt">Gas</span> Analyzer. Two scenarios were compared for each soil: (1) isothermal conditions, representing day time conditions; and (2) thermal gradient conditions, i.e., atmosphere colder than the soil, representing night time conditions. Our results show that under isothermal conditions, diffusion is the major mechanism for surface-atmosphere <span class="hlt">gas</span> <span class="hlt">exchange</span> for all grain sizes; while under night time conditions the prevailing mechanism is dependent on the <span class="hlt">air</span> permeability of the matrix: for sand and gravel it is diffusion, and for soil aggregates it is TCV. Calculated CO2 flux for the soil aggregates column shows that the TCV flux was three orders of magnitude higher than the diffusive flux.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19285150','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19285150"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> in avian embryos and hatchlings.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mortola, Jacopo P</p> <p>2009-08-01</p> <p>The avian egg has been proven to be an excellent model for the study of the physical principles and the physiological characteristics of embryonic <span class="hlt">gas</span> <span class="hlt">exchange</span>. In recent years, it has become a model for the studies of the prenatal development of pulmonary ventilation, its chemical control and its interaction with extra-pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span>. Differently from mammals, in birds the initiation of pulmonary ventilation and the transition from diffusive to convective <span class="hlt">gas</span> <span class="hlt">exchange</span> are gradual and slow-occurring events amenable to detailed investigations. The absence of the placenta and of the mother permits the study of the mechanisms of embryonic adaptation to prenatal perturbations in a way that would be impossible with mammalian preparations. First, this review summarises the general aspects of the natural history of the avian egg that are pertinent to embryonic metabolism, growth and <span class="hlt">gas</span> <span class="hlt">exchange</span> and the characteristics of the structures participating in <span class="hlt">gas</span> <span class="hlt">exchange</span>. Then, the review focuses on the embryonic development of pulmonary ventilation, its regulation in relation to the embryo's environment and metabolic state, the effects that acute or sustained changes in embryonic temperature or oxygenation can have on growth, metabolism and ventilatory control.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27609764','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27609764"><span>The opening-closing rhythms of the subelytral cavity associated with <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns in diapausing Colorado potato beetle, Leptinotarsa decemlineata.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kuusik, Aare; Jõgar, Katrin; Metspalu, Luule; Ploomi, Angela; Merivee, Enno; Must, Anne; Williams, Ingrid H; Hiiesaar, Külli; Sibul, Ivar; Mänd, Marika</p> <p>2016-11-01</p> <p>The opening-closing rhythms of the subelytral cavity and associated <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns were monitored in diapausing Leptinotarsa decemlineata beetles. Measurements were made by means of a flow-through CO 2 analyser and a coulometric respirometer. Under the elytra of these beetles there is a more or less tightly enclosed space, the subelytral cavity (SEC). When the cavity was tightly closed, <span class="hlt">air</span> pressure inside was sub-atmospheric, as a result of oxygen uptake into the tracheae by the beetle. In about half of the beetles, regular opening-closing rhythms of the SEC were observed visually and also recorded; these beetles displayed a discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> pattern. The SEC opened at the start of the CO 2 burst and was immediately closed. On opening, a rapid passive suction inflow of atmospheric <span class="hlt">air</span> into the SEC occurred, recorded coulometrically as a sharp upward peak. As the CO 2 burst lasted beyond the closure of the SEC, we suggest that most of the CO 2 was expelled through the mesothoracic spiracles. In the remaining beetles, the SEC was continually semi-open, and cyclic <span class="hlt">gas</span> <span class="hlt">exchange</span> was exhibited. The locking mechanisms and structures between the elytra and between the elytra and the body were examined under a stereomicroscope and by means of microphotography. We conclude that at least some of the L. decemlineata diapausing beetles were able to close their subelytral cavity tightly, and that the cavity then served as a <span class="hlt">water</span>-saving device. © 2016. Published by The Company of Biologists Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=62405&keyword=FAN&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=62405&keyword=FAN&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>FACTORS AFFECTING <span class="hlt">AIR</span> <span class="hlt">EXCHANGE</span> IN TWO HOUSES</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><span class="hlt">Air</span> <span class="hlt">exchange</span> rate is critical to determining the relationship between indoor and outdoor concentrations of hazardous pollutants. Approximately 150 <span class="hlt">air</span> <span class="hlt">exchange</span> experiments were completed in two residences: a two-story detached house located in Redwood City, CA and a three-story...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/21355','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/21355"><span>Radiation-use efficiency and <span class="hlt">gas</span> <span class="hlt">exchange</span> responses to <span class="hlt">water</span> and nutrient availability in irrigated and fertilized stands of sweetgum and sycamore</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Christopher B. Allen; Rodney E. Will; Robert C. McGravey; David R. Coyle; Mark D. Coleman</p> <p>2005-01-01</p> <p>We investigated how <span class="hlt">water</span> and nutrient availability affect radiation-use effeciency (e) and assessed leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> as a possible mechanism for shifts in e. We measured aboveground net primary production (ANPP) and annual photosynthetically active radiation (PAR) capture to calculate e as well as leaf-level physiological variables (light-saturated net photosynthesis...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010005246','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010005246"><span>BOREAS TE-12 Leaf <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hall, Forrest G. (Editor); Curd, Shelaine (Editor); Arkebauer, Timothy J.; Yang, Litao</p> <p>2000-01-01</p> <p>The BOREAS TE-12 team collected several data sets in support of its efforts to characterize and interpret information on the reflectance, transmittance, and <span class="hlt">gas</span> <span class="hlt">exchange</span> of boreal vegetation. This data set contains measurements of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> conducted in the SSA during the growing seasons of 1994 and 1995 using a portable <span class="hlt">gas</span> <span class="hlt">exchange</span> system. The data are stored in tabular ASCII files. The data files are available on a CD-ROM (see document number 20010000884), or from the Oak Ridge National Laboratory (ORNL) Distributed Active Center (DAAC).</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('https://www.osti.gov/servlets/purl/873724','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/873724"><span>Method and apparatus for extracting <span class="hlt">water</span> from <span class="hlt">air</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Spletzer, Barry L.</p> <p>2001-01-01</p> <p>The present invention provides a method and apparatus for extracting liquid <span class="hlt">water</span> from moist <span class="hlt">air</span> using minimal energy input. The method comprises compressing moist <span class="hlt">air</span> under conditions that foster the condensation of liquid <span class="hlt">water</span> (ideally isothermal to a humidity of 1.0, then adiabatic thereafter). The <span class="hlt">air</span> can be decompressed under conditions that do not foster the vaporization of the condensate. The decompressed, dried <span class="hlt">air</span> can be <span class="hlt">exchanged</span> for a fresh charge of moist <span class="hlt">air</span> and the process repeated. The liquid condensate can be removed for use. The apparatus can comprise a compression chamber having a variable internal volume. An intake port allows moist <span class="hlt">air</span> into the compression chamber. An exhaust port allows dried <span class="hlt">air</span> out of the compression chamber. A condensation device fosters condensation at the desired conditions. A condensate removal port allows liquid <span class="hlt">water</span> to be removed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16271812','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16271812"><span>Atmospheric concentrations and <span class="hlt">air</span>-sea <span class="hlt">exchanges</span> of nonylphenol, tertiary octylphenol and nonylphenol monoethoxylate in the North Sea.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xie, Zhiyong; Lakaschus, Soenke; Ebinghaus, Ralf; Caba, Armando; Ruck, Wolfgang</p> <p>2006-07-01</p> <p>Concentrations of nonylphenol isomers (NP), tertiary octylphenol (t-OP) and nonylphenol monoethoxylate isomers (NP1EO) have been simultaneously determined in the sea <span class="hlt">water</span> and atmosphere of the North Sea. A decreasing concentration profile appeared following the distance increasing from the coast to the central part of the North Sea. <span class="hlt">Air</span>-sea <span class="hlt">exchanges</span> of t-OP and NP were estimated using the two-film resistance model based upon relative <span class="hlt">air-water</span> concentrations and experimentally derived Henry's law constant. The average of <span class="hlt">air</span>-sea <span class="hlt">exchange</span> fluxes was -12+/-6 ng m(-2)day(-1) for t-OP and -39+/-19 ng m(-2)day(-1) for NP, which indicates a net deposition is occurring. These results suggest that the <span class="hlt">air</span>-sea vapour <span class="hlt">exchange</span> is an important process that intervenes in the mass balance of alkylphenols in the North Sea.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25944919','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25944919"><span>Interruption to cutaneous <span class="hlt">gas</span> <span class="hlt">exchange</span> is not a likely mechanism of WNS-associated death in bats.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Carey, Charleve S; Boyles, Justin G</p> <p>2015-07-01</p> <p>Pseudogymnoascus destructans is the causative fungal agent of white-nose syndrome (WNS), an emerging fungal-borne epizootic. WNS is responsible for a catastrophic decline of hibernating bats in North America, yet we have limited understanding of the physiological interactions between pathogen and host. Pseudogymnoascus destructans severely damages wings and tail membranes, by causing dryness that leads to whole sections crumbling off. Four possible mechanisms have been proposed by which infection could lead to dehydration; in this study, we tested one: P. destructans infection could cause disruption to passive <span class="hlt">gas-exchange</span> pathways across the wing membranes, thereby causing a compensatory increase in <span class="hlt">water</span>-intensive pulmonary respiration. We hypothesized that total evaporative <span class="hlt">water</span> loss would be greater when passive <span class="hlt">gas</span> <span class="hlt">exchange</span> was inhibited. We found that bats did not lose more <span class="hlt">water</span> when passive pathways were blocked. This study provides evidence against the proposed proximal mechanism that disruption to passive <span class="hlt">gas</span> <span class="hlt">exchange</span> causes dehydration and death to WNS-infected bats. © 2015. Published by The Company of Biologists Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70040729','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70040729"><span>The impact of lower sea-ice extent on Arctic greenhouse-<span class="hlt">gas</span> <span class="hlt">exchange</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Parmentier, Frans-Jan W.; Christensen, Torben R.; Sørensen, Lise Lotte; Rysgaard, Søren; McGuire, A. David; Miller, Paul A.; Walker, Donald A.</p> <p>2013-01-01</p> <p>In September 2012, Arctic sea-ice extent plummeted to a new record low: two times lower than the 1979–2000 average. Often, record lows in sea-ice cover are hailed as an example of climate change impacts in the Arctic. Less apparent, however, are the implications of reduced sea-ice cover in the Arctic Ocean for marine–atmosphere CO2 <span class="hlt">exchange</span>. Sea-ice decline has been connected to increasing <span class="hlt">air</span> temperatures at high latitudes. Temperature is a key controlling factor in the terrestrial <span class="hlt">exchange</span> of CO2 and methane, and therefore the greenhouse-<span class="hlt">gas</span> balance of the Arctic. Despite the large potential for feedbacks, many studies do not connect the diminishing sea-ice extent with changes in the interaction of the marine and terrestrial Arctic with the atmosphere. In this Review, we assess how current understanding of the Arctic Ocean and high-latitude ecosystems can be used to predict the impact of a lower sea-ice cover on Arctic greenhouse-<span class="hlt">gas</span> <span class="hlt">exchange</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29267801','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29267801"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> and antioxidant activity in seedlings of C opaifera langsdorffii Desf. under different <span class="hlt">water</span> conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rosa, Derek B C J; Scalon, Silvana P Q; Cremon, Thais; Ceccon, Felipe; Dresch, Daiane M</p> <p>2017-01-01</p> <p>The aim of this study was to evaluate <span class="hlt">gas</span> <span class="hlt">exchange</span>, efficiency of the photosynthetic apparatus, and antioxidant activity in Copaifera langsdorffii Desf. The seedlings were cultivated under different conditions of <span class="hlt">water</span> availability, in order to improve the utilization efficiency of available <span class="hlt">water</span> resources. The seedlings were cultivated in four different <span class="hlt">water</span> retention capacities (WRC- 25%, 50%, 75%, and 100%), and evaluated at four different time (T- 30, 60, 90, and 120 days). During the experimental period, seedlings presented the highest values for carboxylation efficiency of Rubisco (A/Ci), intrinsic <span class="hlt">water</span> use efficiency (IWUE = A/gs), chlorophyll index, and stomatal opening, when grown in the substrate with 75% WRC, but the stomatal index (SI) was less the 25% WRC. The efficiency of photosystem II was not significantly altered by the treatments. Comparison between the extreme treatments in terms of <span class="hlt">water</span> availability, represented by 25% and 100% WRC, represent stress conditions for the species. <span class="hlt">Water</span> availability causes a high activity of antioxidant enzymes (superoxide dismutase, peroxidase, and catalase) in the plant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.H23D1302T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.H23D1302T"><span>Direct measurements of wind-<span class="hlt">water</span> momentum coupling in a marsh with emergent vegetation and implications for <span class="hlt">gas</span> transfer estimates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tse, I.; Poindexter, C.; Variano, E. A.</p> <p>2013-12-01</p> <p>Among the numerous ecological benefits of restoring wetlands is carbon sequestration. As emergent vegetation thrive, atmospheric CO2 is removed and converted into biomass that gradually become additional soil. Forecasts and management for these systems rely on accurate knowledge of <span class="hlt">gas</span> <span class="hlt">exchange</span> between the atmosphere and the wetland surface <span class="hlt">waters</span>. Our previous work showed that the rate of <span class="hlt">gas</span> transfer across the <span class="hlt">air-water</span> interface is affected by the amount of <span class="hlt">water</span> column mixing caused by winds penetrating through the plant canopy. Here, we present the first direct measurements of wind-<span class="hlt">water</span> momentum coupling made within a tule marsh. This work in Twitchell Island in the California Delta shows how momentum is imparted into the <span class="hlt">water</span> from wind stress and that this wind stress interacts with the surface <span class="hlt">waters</span> in an interesting way. By correlating three-component velocity signals from a sonic anemometer placed within the plant canopy with data from a novel Volumetric Particle Imager (VoPI) placed in the <span class="hlt">water</span>, we measure the flux of kinetic energy through the plant canopy and the time-scale of the response. We also use this unique dataset to estimate the <span class="hlt">air-water</span> drag coefficient using an adjoint method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4196106','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4196106"><span>Nondestructive natural <span class="hlt">gas</span> hydrate recovery driven by <span class="hlt">air</span> and carbon dioxide</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kang, Hyery; Koh, Dong-Yeun; Lee, Huen</p> <p>2014-01-01</p> <p>Current technologies for production of natural <span class="hlt">gas</span> hydrates (NGH), which include thermal stimulation, depressurization and inhibitor injection, have raised concerns over unintended consequences. The possibility of catastrophic slope failure and marine ecosystem damage remain serious challenges to safe NGH production. As a potential approach, this paper presents <span class="hlt">air</span>-driven NGH recovery from permeable marine sediments induced by simultaneous mechanisms for methane liberation (NGH decomposition) and CH4-<span class="hlt">air</span> or CH4-CO2/<span class="hlt">air</span> replacement. <span class="hlt">Air</span> is diffused into and penetrates NGH and, on its surface, forms a boundary between the <span class="hlt">gas</span> and solid phases. Then spontaneous melting proceeds until the chemical potentials become equal in both phases as NGH depletion continues and self-regulated CH4-<span class="hlt">air</span> replacement occurs over an arbitrary point. We observed the existence of critical methane concentration forming the boundary between decomposition and replacement mechanisms in the NGH reservoirs. Furthermore, when CO2 was added, we observed a very strong, stable, self-regulating process of <span class="hlt">exchange</span> (CH4 replaced by CO2/<span class="hlt">air</span>; hereafter CH4-CO2/<span class="hlt">air</span>) occurring in the NGH. The proposed process will work well for most global <span class="hlt">gas</span> hydrate reservoirs, regardless of the injection conditions or geothermal gradient. PMID:25311102</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25827140','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25827140"><span><span class="hlt">Air</span>-Seawater <span class="hlt">Exchange</span> of Organochlorine Pesticides along the Sediment Plume of a Large Contaminated River.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lin, Tian; Guo, Zhigang; Li, Yuanyuan; Nizzetto, Luca; Ma, Chuanliang; Chen, Yingjun</p> <p>2015-05-05</p> <p>Gaseous <span class="hlt">exchange</span> fluxes of organochlorine pesticides (OCPs) across the <span class="hlt">air-water</span> interface of the coastal East China Sea were determined in order to assess whether the contaminated plume of the Yangtze River could be an important regional source of OCPs to the atmosphere. Hexachlorocyclohexanes (HCHs), chlordane compounds (CHLs), and dichlorodiphenyltrichloroethanes (DDTs) were the most frequently detected OCPs in <span class="hlt">air</span> and <span class="hlt">water</span>. <span class="hlt">Air-water</span> <span class="hlt">exchange</span> was mainly characterized by net volatilization for all measured OCPs. The net gaseous <span class="hlt">exchange</span> flux ranged 10-240 ng/(m2·day) for γ-HCH, 60-370 ng/(m2·day) for trans-CHL, 97-410 ng/(m2·day) for cis-CHL, and ∼0 (e.g., equilibrium) to 490 ng/(m2·day) for p,p'-DDE. We found that the plume of the large contaminated river can serve as a significant regional secondary atmospheric source of legacy contaminants released in the catchment. In particular, the sediment plume represented the relevant source of DDT compounds (especially p,p'-DDE) sustaining net degassing when clean <span class="hlt">air</span> masses from the open ocean reached the plume area. In contrast, a mass balance showed that, for HCHs, contaminated river discharge (<span class="hlt">water</span> and sediment) plumes were capable of sustaining volatilization throughout the year. These results demonstrate the inconsistencies in the fate of HCHs and DDTs in this large estuarine system with declining primary sources.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhDT.........9G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhDT.........9G"><span>Investigation of <span class="hlt">water</span> droplet dynamics in PEM fuel cell <span class="hlt">gas</span> channels</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gopalan, Preethi</p> <p></p> <p><span class="hlt">Water</span> management in Proton <span class="hlt">Exchange</span> Membrane Fuel Cell (PEMFC) has remained one of the most important issues that need to be addressed before its commercialization in automotive applications. Accumulation of <span class="hlt">water</span> on the <span class="hlt">gas</span> diffusion layer (GDL) surface in a PEMFC introduces a barrier for transport of reactant gases through the GDL to the catalyst layer. Despite the fact that the channel geometry is one of the key design parameters of a fluidic system, very limited research is available to study the effect of microchannel geometry on the two-phase flow structure. In this study, the droplet-wall dynamics and two-phase pressure drop across the <span class="hlt">water</span> droplet present in a typical PEMFC channel, were examined in auto-competitive <span class="hlt">gas</span> channel designs (0.4 x 0.7 mm channel cross section). The liquid <span class="hlt">water</span> flow pattern inside the <span class="hlt">gas</span> channel was analyzed for different <span class="hlt">air</span> velocities. Experimental data was analyzed using the Concus-Finn condition to determine the wettability characteristics in the corner region. It was confirmed that the channel angle along with the <span class="hlt">air</span> velocity and the channel material influences the <span class="hlt">water</span> distribution and holdup within the channel. Dynamic contact angle emerged as an important parameter in controlling the droplet-wall interaction. Experiments were also performed to understand how the inlet location of the liquid droplet on the GDL surface affects the droplet dynamic behavior in the system. It was found that droplets emerging near the channel wall or under the land lead to corner filling of the channel. Improvements in the channel design has been proposed based on the artificial channel roughness created to act as capillary grooves to transport the liquid <span class="hlt">water</span> away from the land area. For droplets emerging near the center of the channel, beside the filling and no-filling behavior reported in the literature, a new droplet jumping behavior was observed. As droplets grew and touched the sidewalls, they jumped off to the sidewall leaving the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H51L..04B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H51L..04B"><span>Noble <span class="hlt">Gas</span> signatures of Enhanced Oil Recovery</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barry, P. H.; Kulongoski, J. T.; Tyne, R. L.; Hillegonds, D.; Byrne, D. J.; Landon, M. K.; Ballentine, C. J.</p> <p>2017-12-01</p> <p>Noble gases are powerful tracers of fluids from various oil and <span class="hlt">gas</span> production activities in hydrocarbon reservoirs and nearby groundwater. Non-radiogenic noble gases are introduced into undisturbed oil and natural <span class="hlt">gas</span> reservoirs through <span class="hlt">exchange</span> with formation <span class="hlt">waters</span> [1-3]. Reservoirs with extensive hydraulic fracturing, injection for enhanced oil recovery (EOR), and/or waste disposal also show evidence for a component of noble gases introduced from <span class="hlt">air</span> [4]. Isotopic and elemental ratios of noble gases can be used to 1) assess the migration history of the injected and formation fluids, and 2) determine the extent of <span class="hlt">exchange</span> between multiphase fluids in different reservoirs. We present noble <span class="hlt">gas</span> isotope and abundance data from casing, separator and injectate gases of the Lost Hills and Fruitvale oil fields in the San Joaquin basin, California. Samples were collected as part of the California State <span class="hlt">Water</span> Resource Control Board's Oil and <span class="hlt">Gas</span> Regional Groundwater Monitoring Program. Lost Hills (n=7) and Fruitvale (n=2) gases are geochemically distinct and duplicate samples are highly reproducible. Lost Hills casing <span class="hlt">gas</span> samples were collected from areas where EOR and hydraulic fracturing has occurred in the past several years, and from areas where EOR is absent. The Fruitvale samples were collected from a re-injection port. All samples are radiogenic in their He isotopes, typical of a crustal environment, and show enrichments in heavy noble gases, resulting from preferential adsorption on sediments. Fruitvale samples reflect <span class="hlt">air</span>-like surface conditions, with higher <span class="hlt">air</span>-derived noble <span class="hlt">gas</span> concentrations. Lost Hills gases show a gradation from pristine crustal signatures - indicative of closed-system <span class="hlt">exchange</span> with formation fluids - to strongly <span class="hlt">air</span>-contaminated signatures in the EOR region. Pristine samples can be used to determine the extent of hydrocarbon <span class="hlt">exchange</span> with fluids, whereas samples with excess <span class="hlt">air</span> can be used to quantify the extent of EOR. Determining noble</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26214174','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26214174"><span>Effect of Sediment <span class="hlt">Gas</span> Voids and Ebullition on Benthic Solute <span class="hlt">Exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Flury, Sabine; Glud, Ronnie N; Premke, Katrin; McGinnis, Daniel F</p> <p>2015-09-01</p> <p>The presence of free <span class="hlt">gas</span> in sediments and ebullition events can enhance the pore <span class="hlt">water</span> transport and solute <span class="hlt">exchange</span> across the sediment-<span class="hlt">water</span> interface. However, we experimentally and theoretically document that the presence of free <span class="hlt">gas</span> in sediments can counteract this enhancement effect. The apparent diffusivities (Da) of Rhodamine WT and bromide in sediments containing 8-18% <span class="hlt">gas</span> (Da,YE) were suppressed by 7-39% compared to the control (no <span class="hlt">gas</span>) sediments (Da,C). The measured ratios of Da,YE:Da,C were well within the range of ratios predicted by a theoretical soil model for <span class="hlt">gas</span>-bearing soils. Whereas <span class="hlt">gas</span> voids in sediments reduce the Da for soluble species, they represent a shortcut for low-soluble species such as methane and oxygen. Therefore, the presence of even minor amounts of <span class="hlt">gas</span> can increase the fluxes of low-soluble species (i.e., gases) by several factors, while simultaneously suppressing fluxes of dissolved species.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA615405','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA615405"><span>Forecasting Foreign Currency <span class="hlt">Exchange</span> Rates for <span class="hlt">Air</span> Force Budgeting</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2015-03-26</p> <p>FORECASTING FOREIGN CURRENCY <span class="hlt">EXCHANGE</span> RATES FOR <span class="hlt">AIR</span> FORCE BUDGETING THESIS MARCH 2015...States. AFIT-ENV-MS-15-M-178 FORECASTING FOREIGN CURRENCY <span class="hlt">EXCHANGE</span> RATES FOR <span class="hlt">AIR</span> FORCE BUDGETING THESIS Presented to the Faculty...FORECASTING FOREIGN CURRENCY <span class="hlt">EXCHANGE</span> RATES FOR <span class="hlt">AIR</span> FORCE BUDGETING Nicholas R. Gardner, BS Captain, USAF Committee Membership: Lt Col Jonathan</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDL20011S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDL20011S"><span>Boundary layers at a dynamic interface: <span class="hlt">air</span>-sea <span class="hlt">exchange</span> of heat and mass</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Szeri, Andrew</p> <p>2017-11-01</p> <p><span class="hlt">Exchange</span> of mass or heat across a turbulent liquid-<span class="hlt">gas</span> interface is a problem of critical interest, especially in <span class="hlt">air</span>-sea transfer of natural and man-made gases involved in climate change. The goal in this research area is to determine the <span class="hlt">gas</span> flux from <span class="hlt">air</span> to sea or vice versa. For sparingly soluble non-reactive gases, this is controlled by liquid phase turbulent velocity fluctuations that act on the thin species concentration boundary layer on the liquid side of the interface. If the fluctuations in surface-normal velocity and <span class="hlt">gas</span> concentration differences are known, then it is possible to determine the turbulent contribution to the <span class="hlt">gas</span> flux. However, there is no suitable fundamental direct approach in the general case where neither of these quantities can be easily measured. A new approach is presented to deduce key aspects about the near-surface turbulent motions from remote measurements, which allows one to determine the <span class="hlt">gas</span> transfer velocity, or <span class="hlt">gas</span> flux per unit area if overall concentration differences are known. The approach is illustrated with conceptual examples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040090386&hterms=water+gas+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dwater%2Bgas%2Bexchange','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040090386&hterms=water+gas+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dwater%2Bgas%2Bexchange"><span>Steady-state canopy <span class="hlt">gas</span> <span class="hlt">exchange</span>: system design and operation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bugbee, B.</p> <p>1992-01-01</p> <p>This paper describes the use of a commercial growth chamber for canopy photosynthesis, respiration, and transpiration measurements. The system was designed to measure transpiration via <span class="hlt">water</span> vapor fluxes, and the importance of this measurement is discussed. Procedures for continuous measurement of root-zone respiration are described, and new data is presented to dispel myths about sources of <span class="hlt">water</span> vapor interference in photosynthesis and in the measurement of CO2 by infrared <span class="hlt">gas</span> analysis. Mitchell (1992) has described the fundamentals of various approaches to measuring photosynthesis. Because our system evolved from experience with other types of single-leaf and canopy <span class="hlt">gas-exchange</span> systems, it is useful to review advantages and disadvantages of different systems as they apply to various research objectives.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19800000519&hterms=Water+turbine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DWater%2Bturbine','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19800000519&hterms=Water+turbine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DWater%2Bturbine"><span>Thermodynamic and transport properties of <span class="hlt">air/water</span> mixtures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fessler, T. E.</p> <p>1981-01-01</p> <p>Subroutine WETAIR calculates properties at nearly 1,500 K and 4,500 atmospheres. Necessary inputs are assigned values of combinations of density, pressure, temperature, and entropy. Interpolation of property tables obtains dry <span class="hlt">air</span> and <span class="hlt">water</span> (steam) properties, and simple mixing laws calculate properties of <span class="hlt">air/water</span> mixture. WETAIR is used to test <span class="hlt">gas</span> turbine engines and components operating in relatively humid <span class="hlt">air</span>. Program is written in SFTRAN and FORTRAN.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28577386','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28577386"><span>Hydraulics and <span class="hlt">gas</span> <span class="hlt">exchange</span> recover more rapidly from severe drought stress in small pot-grown grapevines than in field-grown plants.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Romero, Pascual; Botía, Pablo; Keller, Markus</p> <p>2017-09-01</p> <p>Modifications of plant hydraulics and shoot resistances (R shoot ) induced by <span class="hlt">water</span> withholding followed by rewatering, and their relationships with plant <span class="hlt">water</span> status, leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> use efficiency at the leaf level, were investigated in pot-grown and field-grown, own-rooted Syrah grapevines in an arid climate. <span class="hlt">Water</span> stress induced anisohydric behavior, gradually reducing stomatal conductance (g s ) and leaf photosynthesis (A) in response to decreasing midday stem <span class="hlt">water</span> potential (Ψ s ). <span class="hlt">Water</span> stress also rapidly increased intrinsic <span class="hlt">water</span>-use efficiency (A/g s ); this effect persisted for many days after rewatering. Whole-plant (K plant ), canopy (K canopy ), shoot (K shoot ) and leaf (K leaf ) hydraulic conductances decreased during <span class="hlt">water</span> stress, in tune with the gradual decrease in Ψ s , leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and whole plant <span class="hlt">water</span> use. <span class="hlt">Water</span>-stressed vines also had a lower Ψ gradient between stem and leaf (ΔΨ l ), which was correlated with lower leaf transpiration rate (E). E and ΔΨ l increased with increasing vapour pressure deficit (VPD) in non-stressed control vines but not in stressed vines. Perfusion of xylem-mobile dye showed that <span class="hlt">water</span> flow to petioles and leaves was substantially reduced or even stopped under moderate and severe drought stress. Leaf blade hydraulic resistance accounted for most of the total shoot resistance. However, hydraulic conductance of the whole root system (K root ) was not significantly reduced until <span class="hlt">water</span> stress became very severe in pot-grown vines. Significant correlations between K plant , K canopy and Ψ s , K canopy and leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>, K leaf and Ψ s , and K leaf and A support a link between <span class="hlt">water</span> supply, leaf <span class="hlt">water</span> status and <span class="hlt">gas</span> <span class="hlt">exchange</span>. Upon re-<span class="hlt">watering</span>, Ψ s recovered faster than <span class="hlt">gas</span> <span class="hlt">exchange</span> and leaf-shoot hydraulics. A gradual recovery of hydraulic functionality of plant organs was also observed, the leaves being the last to recover after rewatering. In pot-grown vines, K canopy recovered rather</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/868548','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/868548"><span><span class="hlt">Water</span> augmented indirectly-fired <span class="hlt">gas</span> turbine systems and method</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Bechtel, Thomas F.; Parsons, Jr., Edward J.</p> <p>1992-01-01</p> <p>An indirectly-fired <span class="hlt">gas</span> turbine system utilizing <span class="hlt">water</span> augmentation for increasing the net efficiency and power output of the system is described. <span class="hlt">Water</span> injected into the compressor discharge stream evaporatively cools the <span class="hlt">air</span> to provide a higher driving temperature difference across a high temperature <span class="hlt">air</span> heater which is used to indirectly heat the <span class="hlt">water</span>-containing <span class="hlt">air</span> to a turbine inlet temperature of greater than about 1,000.degree. C. By providing a lower <span class="hlt">air</span> heater hot side outlet temperature, heat rejection in the <span class="hlt">air</span> heater is reduced to increase the heat recovery in the <span class="hlt">air</span> heater and thereby increase the overall cycle efficiency.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22103582','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22103582"><span>Distribution and <span class="hlt">air</span>-sea <span class="hlt">exchange</span> of current-use pesticides (CUPs) from East Asia to the high Arctic Ocean.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhong, Guangcai; Xie, Zhiyong; Cai, Minghong; Möller, Axel; Sturm, Renate; Tang, Jianhui; Zhang, Gan; He, Jianfeng; Ebinghaus, Ralf</p> <p>2012-01-03</p> <p>Surface seawater and marine boundary layer <span class="hlt">air</span> samples were collected on the ice-breaker R/V Xuelong (Snow Dragon) from the East China Sea to the high Arctic (33.23-84.5° N) in July to September 2010 and have been analyzed for six current-use pesticides (CUPs): trifluralin, endosulfan, chlorothalonil, chlorpyrifos, dacthal, and dicofol. In all oceanic <span class="hlt">air</span> samples, the six CUPs were detected, showing highest level (>100 pg/m(3)) in the Sea of Japan. Gaseous CUPs basically decreased from East Asia (between 36.6 and 45.1° N) toward Bering and Chukchi Seas. The dissolved CUPs in ocean <span class="hlt">water</span> ranged widely from <MDL to 111 pg/L. Latitudinal trends of α-endosulfan, chlorpyrifos, and dicofol in seawater were roughly consistent with their latitudinal trends in <span class="hlt">air</span>. Trifluralin in seawater was relatively high in the Sea of Japan (35.2° N) and evenly distributed between 36.9 and 72.5° N, but it remained below the detection limit at the highest northern latitudes in Chukchi Sea. In contrast with other CUPs, concentrations of chlorothalonil and dacthal were more abundant in Chukchi Sea and in East Asia. The <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> of CUPs was generally dominated by net deposition. Latitudinal trends of fugacity ratios of α-endosulfan, chlorothalonil, and dacthal showed stronger deposition of these compounds in East Asia than in Chukchi Sea, while trifluralin showed stronger deposition in Chukchi Sea (-455 ± 245 pg/m(2)/day) than in the North Pacific (-241 ± 158 pg/m(2)/day). <span class="hlt">Air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> of chlorpyrifos varied from net volatilizaiton in East Asia (<40° N) to equilibrium or net deposition in the North Pacific and the Arctic.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1178521-guest-molecule-exchange-kinetics-ignik-sikumi-gas-hydrate-field-trial','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1178521-guest-molecule-exchange-kinetics-ignik-sikumi-gas-hydrate-field-trial"><span>Guest Molecule <span class="hlt">Exchange</span> Kinetics for the 2012 Ignik Sikumi <span class="hlt">Gas</span> Hydrate Field Trial</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>White, Mark D.; Lee, Won Suk</p> <p></p> <p>A commercially viable technology for producing methane from natural <span class="hlt">gas</span> hydrate reservoirs remains elusive. Short-term depressurization field tests have demonstrated the potential for producing natural <span class="hlt">gas</span> via dissociation of the clathrate structure, but the long-term performance of the depressurization technology ultimately requires a heat source to sustain the dissociation. A decade of laboratory experiments and theoretical studies have demonstrated the <span class="hlt">exchange</span> of pure CO2 and N2-CO2 mixtures with CH4 in sI <span class="hlt">gas</span> hydrates, yielding critical information about molecular mechanisms, recoveries, and <span class="hlt">exchange</span> kinetics. Findings indicated the potential for producing natural <span class="hlt">gas</span> with little to no production of <span class="hlt">water</span> and rapidmore » <span class="hlt">exchange</span> kinetics, generating sufficient interest in the guest-molecule <span class="hlt">exchange</span> technology for a field test. In 2012 the U.S. DOE/NETL, ConocoPhillips Company, and Japan Oil, <span class="hlt">Gas</span> and Metals National Corporation jointly sponsored the first field trial of injecting a mixture of N2-CO2 into a CH4-hydrate bearing formation beneath the permafrost on the Alaska North Slope. Known as the Ignik Sikumi #1 <span class="hlt">Gas</span> Hydrate Field Trial, this experiment involved three stages: 1) the injection of a N2-CO2 mixture into a targeted hydrate-bearing layer, 2) a 4-day pressurized soaking period, and 3) a sustained depressurization and fluid production period. Data collected during the three stages of the field trial were made available after an extensive quality check. These data included continuous temperature and pressure logs, injected and recovered fluid compositions and volumes. The Ignik Sikumi #1 data set is extensive, but contains no direct evidence of the guest-molecule <span class="hlt">exchange</span> process. This investigation is directed at using numerical simulation to provide an interpretation of the collected data. A numerical simulator, STOMP-HYDT-KE, was recently completed that solves conservation equations for energy, <span class="hlt">water</span>, mobile fluid guest molecules, and hydrate</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24438580','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24438580"><span>Argon used as dry suit insulation <span class="hlt">gas</span> for cold-<span class="hlt">water</span> diving.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vrijdag, Xavier Ce; van Ooij, Pieter-Jan Am; van Hulst, Robert A</p> <p>2013-06-03</p> <p>Cold-<span class="hlt">water</span> diving requires good thermal insulation because hypothermia is a serious risk. <span class="hlt">Water</span> conducts heat more efficiently compared to <span class="hlt">air</span>. To stay warm during a dive, the choice of thermal protection should be based on physical activity, the temperature of the <span class="hlt">water</span>, and the duration of exposure. A dry suit, a diving suit filled with <span class="hlt">gas</span>, is the most common diving suit in cold <span class="hlt">water</span>. <span class="hlt">Air</span> is the traditional dry suit inflation <span class="hlt">gas</span>, whereas the thermal conductivity of argon is approximately 32% lower compared to that of <span class="hlt">air</span>. This study evaluates the benefits of argon, compared to <span class="hlt">air</span>, as a thermal insulation <span class="hlt">gas</span> for a dry suit during a 1-h cold-<span class="hlt">water</span> dive by divers of the Royal Netherlands Navy. Seven male Special Forces divers made (in total) 19 dives in a diving basin with <span class="hlt">water</span> at 13 degrees C at a depth of 3 m for 1 h in upright position. A rubber dry suit and woollen undergarment were used with either argon (n = 13) or <span class="hlt">air</span> (n = 6) (blinded to the divers) as suit inflation <span class="hlt">gas</span>. Core temperature was measured with a radio pill during the dive. Before, halfway, and after the dive, subjective thermal comfort was recorded using a thermal comfort score. No diver had to abort the test due to cold. No differences in core temperature and thermal comfort score were found between the two groups. Core temperature remained unchanged during the dives. Thermal comfort score showed a significant decrease in both groups after a 60-min dive compared to baseline. In these tests the combination of the dry suit and undergarment was sufficient to maintain core temperature and thermal comfort for a dive of 1h in <span class="hlt">water</span> at 13 degrees C. The use of argon as a suit inflation <span class="hlt">gas</span> had no added value for thermal insulation compared to <span class="hlt">air</span> for these dives.</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://hdl.handle.net/2060/20010004659','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010004659"><span>BOREAS TE-9 In Situ Diurnal <span class="hlt">Gas</span> <span class="hlt">Exchange</span> of NAS Boreal Forest Stands</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hall, Forrest G. (Editor); Curd, Shelaine (Editor); Margolis, Hank; Coyea, Marie; Dang, Qinglai</p> <p>2000-01-01</p> <p>The BOREAS TE-9 team collected several data sets related to chemical and photosynthetic properties of leaves in boreal forest tree species. The purpose of the BOREAS TE-09 study was threefold: 1) to provide in situ <span class="hlt">gas</span> <span class="hlt">exchange</span> data that will be used to validate models of photosynthetic responses to light, temperature, and carbon dioxide (CO2); 2) to compare the photosynthetic responses of different tree crown levels (upper and lower); and 3) to characterize the diurnal <span class="hlt">water</span> potential curves for these sites to get an indication of the extent to which soil moisture supply to leaves might be limiting photosynthesis. The <span class="hlt">gas</span> <span class="hlt">exchange</span> data of the BOREAS NSA were collected to characterize diurnal <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> potential of two canopy levels of five boreal canopy cover types: young jack pine, old jack pine, old aspen, lowland old black spruce, and upland black spruce. These data were collected between 27-May-1994 and 17-Sep-1994. The data are provided in tabular ASCII files. The data files are available on a CD-ROM (see document number 20010000884), or from the Oak Ridge National Laboratory (ORNL) Distributed Active Archive Center (DAAC).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1060285','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1060285"><span>Development and Validation of a <span class="hlt">Gas</span>-Fired Residential Heat Pump <span class="hlt">Water</span> Heater - Final Report</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>Michael Garrabrant; Roger Stout; Paul Glanville</p> <p>2013-01-21</p> <p>For <span class="hlt">gas</span>-fired residential <span class="hlt">water</span> heating, the U.S. and Canada is predominantly supplied by minimum efficiency storage <span class="hlt">water</span> heaters with Energy Factors (EF) in the range of 0.59 to 0.62. Higher efficiency and higher cost ($700 - $2,000) options serve about 15% of the market, but still have EFs below 1.0, ranging from 0.65 to 0.95. To develop a new class of <span class="hlt">water</span> heating products that exceeds the traditional limit of thermal efficiency, the project team designed and demonstrated a packaged <span class="hlt">water</span> heater driven by a <span class="hlt">gas</span>-fired ammonia-<span class="hlt">water</span> absorption heat pump. This <span class="hlt">gas</span>-fired heat pump <span class="hlt">water</span> heater can achieve EFs ofmore » 1.3 or higher, at a consumer cost of $2,000 or less. Led by Stone Mountain Technologies Inc. (SMTI), with support from A.O. Smith, the <span class="hlt">Gas</span> Technology Institute (GTI), and Georgia Tech, the cross-functional team completed research and development tasks including cycle modeling, breadboard evaluation of two cycles and two heat <span class="hlt">exchanger</span> classes, heat pump/storage tank integration, compact solution pump development, combustion system specification, and evaluation of packaged prototype GHPWHs. The heat pump system extracts low grade heat from the ambient <span class="hlt">air</span> and produces high grade heat suitable for heating <span class="hlt">water</span> in a storage tank for domestic use. Product features that include conventional installation practices, standard footprint and reasonable economic payback, position the technology to gain significant market penetration, resulting in a large reduction of energy use and greenhouse <span class="hlt">gas</span> emissions from domestic hot <span class="hlt">water</span> production.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.B52B..02D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.B52B..02D"><span>High Resolution CH4 Emissions and Dissolved CH4 Measurements Elucidate Surface <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Processes in Toolik Lake, Arctic Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Del Sontro, T.; Sollberger, S.; Kling, G. W.; Shaver, G. R.; Eugster, W.</p> <p>2013-12-01</p> <p>Approximately 14% of the Alaskan North Slope is covered in lakes of various sizes and depths. Diffusive carbon emissions (CH4 and CO2) from these lakes offset the tundra sink by ~20 %, but the offset would substantially increase if ebullitive CH4 emissions were also considered. Ultimately, arctic lake CH4 emissions are not insignificant in the global CH4 budget and their contribution is bound to increase due to impacts from climate change. Here we present high resolution CH4 emission data as measured via eddy covariance and a Los Gatos <span class="hlt">gas</span> analyzer during the ice free period from Toolik Lake, a deep (20 m) Arctic lake located on the Alaskan North Slope, over the last few summers. Emissions are relatively low (< 25 mg CH4 m-2 d-1) with little variation over the summer. Diurnal variations regularly occur, however, with up to 3 times higher fluxes at night. <span class="hlt">Gas</span> <span class="hlt">exchange</span> is a relatively difficult process to estimate, but is normally done so as the product of the CH4 gradient across the <span class="hlt">air-water</span> interface and the <span class="hlt">gas</span> transfer velocity, k. Typically, k is determined based on the turbulence on the <span class="hlt">water</span> side of the interface, which is most commonly approximated by wind speed; however, it has become increasingly apparent that this assumption does not remain valid across all <span class="hlt">water</span> bodies. Dissolved CH4 profiles in Toolik revealed a subsurface peak in CH4 at the thermocline of up to 3 times as much CH4 as in the surface <span class="hlt">water</span>. We hypothesize that convective mixing at night due to cooling surface <span class="hlt">waters</span> brings the subsurface CH4 to the surface and causes the higher night fluxes. In addition to high resolution flux emission estimates, we also acquired high resolution data for dissolved CH4 in surface <span class="hlt">waters</span> of Toolik Lake during the last two summers using a CH4 equilibrator system connected to a Los Gatos <span class="hlt">gas</span> analyzer. Thus, having both the flux and the CH4 gradient across the <span class="hlt">air-water</span> interface measured directly, we can calculate k and investigate the processes influencing</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19750027247&hterms=respiratory&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Drespiratory','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19750027247&hterms=respiratory&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Drespiratory"><span>Automated measurement of respiratory <span class="hlt">gas</span> <span class="hlt">exchange</span> by an inert <span class="hlt">gas</span> dilution technique</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sawin, C. F.; Rummel, J. A.; Michel, E. L.</p> <p>1974-01-01</p> <p>A respiratory <span class="hlt">gas</span> analyzer (RGA) has been developed wherein a mass spectrometer is the sole transducer required for measurement of respiratory <span class="hlt">gas</span> <span class="hlt">exchange</span>. The mass spectrometer maintains all signals in absolute phase relationships, precluding the need to synchronize flow and <span class="hlt">gas</span> composition as required in other systems. The RGA system was evaluated by comparison with the Douglas bag technique. The RGA system established the feasibility of the inert <span class="hlt">gas</span> dilution method for measuring breath-by-breath respiratory <span class="hlt">gas</span> <span class="hlt">exchange</span>. This breath-by-breath analytical capability permits detailed study of transient respiratory responses to exercise.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15562064','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15562064"><span>Conductive heat <span class="hlt">exchange</span> with a gel-coated circulating <span class="hlt">water</span> mattress.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bräuer, Anselm; Pacholik, Larissa; Perl, Thorsten; English, Michael John Murray; Weyland, Wolfgang; Braun, Ulrich</p> <p>2004-12-01</p> <p>The use of forced-<span class="hlt">air</span> warming is associated with costs for the disposable blankets. As an alternative method, we studied heat transfer with a reusable gel-coated circulating <span class="hlt">water</span> mattress placed under the back in eight healthy volunteers. Heat flux was measured with six calibrated heat flux transducers. Additionally, mattress temperature, skin temperature, and core temperature were measured. <span class="hlt">Water</span> temperature was set to 25 degrees C, 30 degrees C, 35 degrees C, and 41 degrees C. Heat transfer was calculated by multiplying heat flux by contact area. Mattress temperature, skin temperature, and heat flux were used to determine the heat <span class="hlt">exchange</span> coefficient for conduction. Heat flux and <span class="hlt">water</span> temperature were related by the following equation: heat flux = 10.3 x <span class="hlt">water</span> temperature - 374 (r(2) = 0.98). The heat <span class="hlt">exchange</span> coefficient for conduction was 121 W . m(-2) . degrees C(-1). The maximal heat transfer with the gel-coated circulating <span class="hlt">water</span> mattress was 18.4 +/- 3.3 W. Because of the small effect on the heat balance of the body, a gel-coated circulating <span class="hlt">water</span> mattress placed only on the back cannot replace a forced-<span class="hlt">air</span> warming system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910041723&hterms=water+gas+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dwater%2Bgas%2Bexchange','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910041723&hterms=water+gas+exchange&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dwater%2Bgas%2Bexchange"><span>Relationship between <span class="hlt">gas</span> <span class="hlt">exchange</span>, wind speed, and radar backscatter in a large wind-wave tank</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wanninkhof, Richard H.; Bliven, L. F.</p> <p>1991-01-01</p> <p>The relationships between the <span class="hlt">gas</span> <span class="hlt">exchange</span>, wind speed, friction velocity, and radar backscatter from the <span class="hlt">water</span> surface was investigated using data obtained in a large <span class="hlt">water</span> tank in the Delft (Netherlands) wind-wave tunnel, filled with <span class="hlt">water</span> supersaturated with SF6, N2O, and CH4. Results indicate that the <span class="hlt">gas</span>-transfer velocities of these substances were related to the wind speed with a power law dependence. Microwave backscatter from <span class="hlt">water</span> surface was found to be related to <span class="hlt">gas</span> transfer velocities by a relationship in the form k(<span class="hlt">gas</span>) = a 10 exp (b A0), where k is the <span class="hlt">gas</span> transfer velocity for the particular <span class="hlt">gas</span>, the values of a and b are obtained from a least squares fit of the average backscatter cross section and <span class="hlt">gas</span> transfer at 80 m, and A0 is the directional (azimuthal) averaged return.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12095811','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12095811"><span>Effects of humidified and dry <span class="hlt">air</span> on corneal endothelial cells during vitreal fluid-<span class="hlt">air</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cekiç, Osman; Ohji, Masahito; Hayashi, Atsushi; Fang, Xiao Y; Kusaka, Shunji; Tano, Yasuo</p> <p>2002-07-01</p> <p>To report the immediate anatomic and functional alterations in corneal endothelial cells following use of humidified <span class="hlt">air</span> and dry <span class="hlt">air</span> during vitreal fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> in rabbits. Experimental study. Rabbits undergoing pars plana vitrectomy and lensectomy were perfused with either dry or humidified <span class="hlt">air</span> during fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> for designated durations. Three different experiments were performed. First, control and experimental corneas were examined by scanning electron microscopy (SEM). Second, corneas were stained with Phalloidin-FITC and examined by fluorescein microscopy. Finally, third, transendothelial permeability for carboxyfluorescein was determined using a diffusion chamber. While different from the corneal endothelial cells, those cells exposed to humidified <span class="hlt">air</span> were less stressed than cells exposed to dry <span class="hlt">air</span> by SEM. Actin cytoskeleton was found highly disorganized with dry <span class="hlt">air</span> exposure. Humidified <span class="hlt">air</span> maintained the normal actin cytoskeleton throughout the 20 minutes of fluid-<span class="hlt">air</span> <span class="hlt">exchange</span>. Paracellular carboxyfluorescein leakage was significantly higher in dry <span class="hlt">air</span> insufflated eyes compared with that of the humidified <span class="hlt">air</span> after 5, 10, and 20 minutes of fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> (P =.002, P =.004, and P =.002, respectively). Dry <span class="hlt">air</span> stress during fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> causes significant immediate alterations in monolayer appearance, actin cytoskeleton, and barrier function of corneal endothelium in aphakic rabbit eyes. Use of humidified <span class="hlt">air</span> largely prevents the alterations in monolayer appearance, actin cytoskeleton, and barrier function of corneal endothelial cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3718390','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3718390"><span>Atmospheric concentrations and air–soil <span class="hlt">gas</span> <span class="hlt">exchange</span> of polycyclic aromatic hydrocarbons (PAHs) in remote, rural village and urban areas of Beijing–Tianjin region, North 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>Wang, Wentao; Simonich, Staci; Giri, Basant; Chang, Ying; Zhang, Yuguang; Jia, Yuling; Tao, Shu; Wang, Rong; Wang, Bin; Li, Wei; Cao, Jun; Lu, Xiaoxia</p> <p>2013-01-01</p> <p>Forty passive <span class="hlt">air</span> samplers were deployed to study the occurrence of <span class="hlt">gas</span> and particulate phase PAHs in remote, rural village and urban areas of Beijing–Tianjin region, North China for four seasons (spring, summer, fall and winter) from 2007 to 2008. The influence of emissions on the spatial distribution pattern of <span class="hlt">air</span> PAH concentrations was addressed. In addition, the air–soil <span class="hlt">gas</span> <span class="hlt">exchange</span> of PAHs was studied using fugacity calculations. The median gaseous and particulate phase PAH concentrations were 222 ng/m3 and 114 ng/m3, respectively, with a median total PAH concentration of 349 ng/m3. Higher PAH concentrations were measured in winter than in other seasons. <span class="hlt">Air</span> PAH concentrations measured at the rural villages and urban sites in the northern mountain region were significantly lower than those measured at sites in the southern plain during all seasons. However, there was no significant difference in PAH concentrations between the rural villages and urban sites in the northern and southern areas. This urban–rural PAH distribution pattern was related to the location of PAH emission sources and the population distribution. The location of PAH emission sources explained 56%–77% of the spatial variation in ambient <span class="hlt">air</span> PAH concentrations. The annual median air–soil <span class="hlt">gas</span> <span class="hlt">exchange</span> flux of PAHs was 42.2 ng/m2/day from soil to <span class="hlt">air</span>. Among the 15 PAHs measured, acenaphthylene (ACY) and acenaphthene (ACE) contributed to more than half of the total <span class="hlt">exchange</span> flux. Furthermore, the air–soil <span class="hlt">gas</span> <span class="hlt">exchange</span> fluxes of PAHs at the urban sites were higher than those at the remote and rural sites. In summer, more gaseous PAHs volatilized from soil to <span class="hlt">air</span> because of higher temperatures and increased rainfall. However, in winter, more gaseous PAHs deposited from <span class="hlt">air</span> to soil due to higher PAH emissions and lower temperatures. The soil TOC concentration had no significant influence on the air–soil <span class="hlt">gas</span> <span class="hlt">exchange</span> of PAHs. PMID:21669328</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8625637','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8625637"><span>Perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span> in normal and acid-injured large sheep.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hernan, L J; Fuhrman, B P; Kaiser, R E; Penfil, S; Foley, C; Papo, M C; Leach, C L</p> <p>1996-03-01</p> <p>We hypothesized that a) perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span> could be accomplished in normal large sheep; b) the determinants of <span class="hlt">gas</span> <span class="hlt">exchange</span> would be similar during perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span> and conventional <span class="hlt">gas</span> ventilation; c)in large animals with lung injury, perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span> could be used to enhance <span class="hlt">gas</span> <span class="hlt">exchange</span> without adverse effects on hemodynamics; and d) the large animal with lung injury could be supported with an FIO2 of <1.0 during perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span>. Prospective, observational animal study and prospective randomized, controlled animal study. An animal laboratory in a university setting. Thirty adult ewes. Five normal ewes (61.0 +/- 4.0 kg) underwent perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span> to ascertain the effects of tidal volume, end-inspiratory pressure, and positive end-expiratory pressure (PEEP) on oxygenation. Respiratory rate, tidal volume, and minute ventilation were studied to determine their effects on CO2 clearance. Sheep, weighing 58.9 +/- 8.3 kg, had lung injury induced by instilling 2 mL/kg of 0.05 Normal hydrochloric acid into the trachea. Five minutes after injury, PEEP was increased to 10 cm H2O. Ten minutes after injury, sheep with Pao2 values of <100 torr (<13.3 kPa) were randomized to continue <span class="hlt">gas</span> ventilation (control, n=9) or to institute perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span> (n=9) by instilling 1.6 L of unoxygenated perflubron into the trachea and resuming <span class="hlt">gas</span> ventilation. Blood <span class="hlt">gas</span> and hemodynamic measurements were obtained throughout the 4-hr study. Both tidal volume and end-inspiratory pressure influenced oxygenation in normal sheep during perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span>. Minute ventilation determined CO2 clearance during perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span> in normal sheep. After acid aspiration lung injury, perfluorocarbon-associated <span class="hlt">gas</span> <span class="hlt">exchange</span> increased PaO2 and reduced intrapulmonary shunt fraction. Hypoxia and intrapulmonary shunting were unabated</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17575285','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17575285"><span>Leaf <span class="hlt">water</span> relations and net <span class="hlt">gas</span> <span class="hlt">exchange</span> responses of salinized Carrizo citrange seedlings during drought stress and recovery.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pérez-Pérez, J G; Syvertsen, J P; Botía, P; García-Sánchez, F</p> <p>2007-08-01</p> <p>Since salinity and drought stress can occur together, an assessment was made of their interacting effects on leaf <span class="hlt">water</span> relations, osmotic adjustment and net <span class="hlt">gas</span> <span class="hlt">exchange</span> in seedlings of the relatively chloride-sensitive Carrizo citrange, Citrus sinensis x Poncirus trifoliata. Plants were fertilized with nutrient solution with or without additional 100 mm NaCl (salt and no-salt treatments). After 7 d, half of the plants were drought stressed by withholding irrigation <span class="hlt">water</span> for 10 d. Thus, there were four treatments: salinized and non-salinized plants under drought-stress or well-<span class="hlt">watered</span> conditions. After the drought period, plants from all stressed treatments were re-<span class="hlt">watered</span> with nutrient solution without salt for 8 d to study recovery. Leaf <span class="hlt">water</span> relations, <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters, chlorophyll fluorescence, proline, quaternary ammonium compounds and leaf and root concentrations of Cl(-) and Na(+) were measured. Salinity increased leaf Cl(-) and Na(+) concentrations and decreased osmotic potential (Psi(pi)) such that leaf relative <span class="hlt">water</span> content (RWC) was maintained during drought stress. However, in non-salinized drought-stressed plants, osmotic adjustment did not occur and RWC decreased. The salinity-induced osmotic adjustment was not related to any accumulation of proline, quaternary ammonium compounds or soluble sugars. Net CO(2) assimilation rate (A(CO2)) was reduced in leaves from all stressed treatments but the mechanisms were different. In non-salinized drought-stressed plants, lower A(CO2) was related to low RWC, whereas in salinized plants decreased A(CO2) was related to high levels of leaf Cl(-) and Na(+). A(CO2) recovered after irrigation in all the treatments except in previously salinized drought-stressed leaves which had lower RWC and less chlorophyll but maintained high levels of Cl(-), Na(+) and quaternary ammonium compounds after recovery. High leaf levels of Cl(-) and Na(+) after recovery apparently came from the roots. Plants preconditioned by</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010022736','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010022736"><span>BOREAS TF-11 SSA-Fen Leaf <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Arkebauer, Timothy J.; Hall, Forrest G. (Editor); Knapp, David E. (Editor)</p> <p>2000-01-01</p> <p>The BOREAS TF-11 team gathered a variety of data to complement its tower flux measurements collected at the SSA-Fen site. This data set contains single-leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> data from the SSA-Fen site during 1994 and 1995. These leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> properties were measured for the dominant vascular plants using portable <span class="hlt">gas</span> <span class="hlt">exchange</span> systems. The data are stored in tabular ASCII files.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19443613','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19443613"><span>Leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>, carbon isotope discrimination, and grain yield in contrasting rice genotypes subjected to <span class="hlt">water</span> deficits during the reproductive stage.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Centritto, Mauro; Lauteri, Marco; Monteverdi, Maria Cristina; Serraj, Rachid</p> <p>2009-01-01</p> <p>Genotypic variations in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and yield were analysed in five upland-adapted and three lowland rice cultivars subjected to a differential soil moisture gradient, varying from well-<span class="hlt">watered</span> to severely <span class="hlt">water</span>-stressed conditions. A reduction in the amount of <span class="hlt">water</span> applied resulted in a significant decrease in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and, subsequently, in above-ground dry mass and grain yield, that varied among genotypes and distance from the line source. The comparison between the variable J and the Delta values in recently synthesized sugars methods, yielded congruent estimations of mesophyll conductance (g(m)), confirming the reliability of these two techniques. Our data demonstrate that g(m) is a major determinant of photosynthesis (A), because rice genotypes with inherently higher g(m) were capable of keeping higher A in stressed conditions. Furthermore, A, g(s), and g(m) of <span class="hlt">water</span>-stressed genotypes rapidly recovered to the well-<span class="hlt">watered</span> values upon the relief of <span class="hlt">water</span> stress, indicating that drought did not cause any lasting metabolic limitation to photosynthesis. The comparisons between the A/C(i) and corresponding A/C(c) curves, measured in the genotypes that showed intrinsically higher and lower instantaneous A, confirmed this finding. Moreover, the effect of drought stress on grain yield was correlated with the effects on both A and total diffusional limitations to photosynthesis. Overall, these data indicate that genotypes which showed higher photosynthesis and conductances were also generally more productive across the entire soil moisture gradient. The analysis of Delta revealed a substantial variation of <span class="hlt">water</span> use efficiency among the genotypes, both on the long-term (leaf pellet analysis) and short-term scale (leaf soluble sugars analysis).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018BGeo...15.3085H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018BGeo...15.3085H"><span>Use of argon to measure <span class="hlt">gas</span> <span class="hlt">exchange</span> in turbulent mountain streams</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hall, Robert O., Jr.; Madinger, Hilary L.</p> <p>2018-05-01</p> <p><span class="hlt">Gas</span> <span class="hlt">exchange</span> is a parameter needed in stream metabolism and trace <span class="hlt">gas</span> emissions models. One way to estimate <span class="hlt">gas</span> <span class="hlt">exchange</span> is via measuring the decline of added tracer gases such as sulfur hexafluoride (SF6). Estimates of oxygen (O2) <span class="hlt">gas</span> <span class="hlt">exchange</span> derived from SF6 additions require scaling via Schmidt number (Sc) ratio, but this scaling is uncertain under conditions of high <span class="hlt">gas</span> <span class="hlt">exchange</span> via bubbles because scaling depends on <span class="hlt">gas</span> solubility as well as Sc. Because argon (Ar) and O2 have nearly identical Schmidt numbers and solubility, Ar may be a useful tracer <span class="hlt">gas</span> for estimating stream O2 <span class="hlt">exchange</span>. Here we compared rates of <span class="hlt">gas</span> <span class="hlt">exchange</span> measured via Ar and SF6 for turbulent mountain streams in Wyoming, USA. We measured Ar as the ratio of Ar : N2 using a membrane inlet mass spectrometer (MIMS). Normalizing to N2 confers higher precision than simply measuring [Ar] alone. We consistently enriched streams with Ar from 1 to 18 % of ambient Ar concentration and could estimate <span class="hlt">gas</span> <span class="hlt">exchange</span> rate using an exponential decline model. The mean ratio of <span class="hlt">gas</span> <span class="hlt">exchange</span> of Ar relative to SF6 was 1.8 (credible interval 1.1 to 2.5) compared to the theoretical estimate 1.35, showing that using SF6 would have underestimated <span class="hlt">exchange</span> of Ar. Steep streams (slopes 11-12 %) had high rates of <span class="hlt">gas</span> <span class="hlt">exchange</span> velocity normalized to Sc = 600 (k600, 57-210 m d-1), and slope strongly predicted variation in k600 among all streams. We suggest that Ar is a useful tracer because it is easily measured, requires no scaling assumptions to estimate rates of O2 <span class="hlt">exchange</span>, and is not an intense greenhouse <span class="hlt">gas</span> as is SF6. We caution that scaling from rates of either Ar or SF6 <span class="hlt">gas</span> <span class="hlt">exchange</span> to CO2 is uncertain due to solubility effects in conditions of bubble-mediated <span class="hlt">gas</span> transfer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GBioC..31.1579S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GBioC..31.1579S"><span>Oceanic Uptake of Oxygen During Deep Convection Events Through Diffusive and Bubble-Mediated <span class="hlt">Gas</span> <span class="hlt">Exchange</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sun, Daoxun; Ito, Takamitsu; Bracco, Annalisa</p> <p>2017-10-01</p> <p>The concentration of dissolved oxygen (O2) plays fundamental roles in diverse chemical and biological processes throughout the oceans. The balance between the physical supply and the biological consumption controls the O2 level of the interior ocean, and the O2 supply to the deep <span class="hlt">waters</span> can only occur through deep convection in the polar oceans. We develop a theoretical framework describing the oceanic O2 uptake during open-ocean deep convection events and test it against a suite of numerical sensitivity experiments. Our framework allows for two predictions, confirmed by the numerical simulations. First, both the duration and the intensity of the wintertime cooling contribute to the total O2 uptake for a given buoyancy loss. Stronger cooling leads to deeper convection and the oxygenation can reach down to deeper depths. Longer duration of the cooling period increases the total amount of O2 uptake over the convective season. Second, the bubble-mediated influx of O2 tends to weaken the diffusive influx by shifting the <span class="hlt">air</span>-sea disequilibrium of O2 toward supersaturation. The degree of compensation between the diffusive and bubble-mediated <span class="hlt">gas</span> <span class="hlt">exchange</span> depends on the dimensionless number measuring the relative strength of oceanic vertical mixing and the <span class="hlt">gas</span> transfer velocity. Strong convective mixing, which may occur under strong cooling, reduces the degree of compensation so that the two components of <span class="hlt">gas</span> <span class="hlt">exchange</span> together drive exceptionally strong oceanic O2 uptake.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1358252','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1358252"><span>Miniaturized <span class="hlt">Air</span>-to-Refrigerant Heat <span class="hlt">Exchangers</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>Radermacher, Reinhard; Bacellar, Daniel; Aute, Vikrant</p> <p></p> <p><span class="hlt">Air</span>-to-refrigerant Heat <span class="hlt">eXchangers</span> (HX) are an essential component of Heating, Ventilation, <span class="hlt">Air</span>-Conditioning, and Refrigeration (HVAC&R) systems, serving as the main heat transfer component. The major limiting factor to HX performance is the large airside thermal resistance. Recent literature aims at improving heat transfer performance by utilizing enhancement methods such as fins and small tube diameters; this has lead to almost exhaustive research on the microchannel HX (MCHX). The objective of this project is to develop a miniaturized <span class="hlt">air</span>-to-refrigerant HX with at least 20% reduction in volume, material volume, and approach temperature compared to current state-of-the-art multiport flat tube designs andmore » also be capable of production within five years. Moreover, the proposed HX’s are expected to have good <span class="hlt">water</span> drainage and should succeed in both evaporator and condenser applications. The project leveraged Parallel-Parametrized Computational Fluid Dynamics (PPCFD) and Approximation-Assisted Optimization (AAO) techniques to perform multi-scale analysis and shape optimization with the intent of developing novel HX designs whose thermal-hydraulic performance exceeds that of state-of-the-art MCHX. Nine heat <span class="hlt">exchanger</span> geometries were initially chosen for detailed analysis, selected from 35+ geometries which were identified in previous work at the University of Maryland, College Park. The newly developed optimization framework was exercised for three design optimization problems: (DP I) 1.0kW radiator, (DP II) 10kW radiator and (DP III) 10kW two-phase HX. DP I consisted of the design and optimization of 1.0kW <span class="hlt">air-to-water</span> HX’s which exceeded the project requirements of 20% volume/material reduction and 20% better performance. Two prototypes for the 1.0kW HX were prototyped, tested and validated using newly-designed airside and refrigerant side test facilities. DP II, a scaled version DP I for 10kW <span class="hlt">air-to-water</span> HX applications, also yielded optimized HX</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2013-08-14/pdf/2013-19756.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2013-08-14/pdf/2013-19756.pdf"><span>78 FR 49484 - <span class="hlt">Exchange</span> of <span class="hlt">Air</span> Force Real Property for Non-<span class="hlt">Air</span> Force Real Property</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2013-08-14</p> <p>... DEPARTMENT OF DEFENSE Department of <span class="hlt">Air</span> Force <span class="hlt">Exchange</span> of <span class="hlt">Air</span> Force Real Property for Non-<span class="hlt">Air</span> Force Real Property SUMMARY: Notice identifies excess Federal real property under administrative jurisdiction of the United States <span class="hlt">Air</span> Force it intends to <span class="hlt">exchange</span> for real property not currently owned by the...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28620915','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28620915"><span><span class="hlt">Gas</span> <span class="hlt">exchange</span> recovery following natural drought is rapid unless limited by loss of leaf hydraulic conductance: evidence from an evergreen woodland.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Skelton, Robert P; Brodribb, Timothy J; McAdam, Scott A M; Mitchell, Patrick J</p> <p>2017-09-01</p> <p>Drought can cause major damage to plant communities, but species damage thresholds and postdrought recovery of forest productivity are not yet predictable. We used an El Niño drought event as a natural experiment to test whether postdrought recovery of <span class="hlt">gas</span> <span class="hlt">exchange</span> could be predicted by properties of the <span class="hlt">water</span> transport system, or if metabolism, primarily high abscisic acid concentration, might delay recovery. We monitored detailed physiological responses, including shoot sapflow, leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>, leaf <span class="hlt">water</span> potential and foliar abscisic acid (ABA), during drought and through the subsequent rehydration period for a sample of eight canopy and understory species. Severe drought caused major declines in leaf <span class="hlt">water</span> potential, elevated foliar ABA concentrations and reduced stomatal conductance and assimilation rates in our eight sample species. Leaf <span class="hlt">water</span> potential surpassed levels associated with incipient loss of leaf hydraulic conductance in four species. Following heavy rainfall <span class="hlt">gas</span> <span class="hlt">exchange</span> in all species, except those trees predicted to have suffered hydraulic impairment, recovered to prestressed rates within 1 d. Recovery of plant <span class="hlt">gas</span> <span class="hlt">exchange</span> was rapid and could be predicted by the hydraulic safety margin, providing strong support for leaf vulnerability to <span class="hlt">water</span> deficit as an index of damage under natural drought conditions. © 2017 The Authors. New Phytologist © 2017 New Phytologist Trust.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10369593','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10369593"><span>Effect of humidity on posterior lens opacification during fluid-<span class="hlt">air</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Harlan, J B; Lee, E T; Jensen, P S; de Juan, E</p> <p>1999-06-01</p> <p>To study the relationship of humidity and the rate of lens opacity formation during fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> using an animal model. Vitrectomy and fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> was carried out using 16 eyes of 8 pigmented rabbits. One eye of each rabbit was exposed to dry <span class="hlt">air</span> and the fellow eye received humidified <span class="hlt">air</span> using an intraocular <span class="hlt">air</span> humidifier. In each case, the percent humidity of the intraocular <span class="hlt">air</span> was measured using an in-line hygrometer. Elapsed time from initial <span class="hlt">air</span> entry to lens feathering was recorded for each eye, with the surgeon-observer unaware of the percent humidity of the <span class="hlt">air</span> infusion. In each rabbit, use of humidified <span class="hlt">air</span> resulted in a delay in lens feathering (P<.02), with an overall increase in time to feathering of 80% for humidified <span class="hlt">air</span> vs room <span class="hlt">air</span>. Use of a humidifier during fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> prolongs intraoperative lens clarity in the rabbit model, suggesting that humidified <span class="hlt">air</span> should prolong lens clarity during phakic fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> in patients. Use of humidified <span class="hlt">air</span> during vitrectomy and fluid-<span class="hlt">air</span> <span class="hlt">exchange</span> may retard the intraoperative loss of lens clarity, promoting better visualization of the posterior segment and enhancing surgical performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930038655&hterms=Water+turbine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DWater%2Bturbine','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930038655&hterms=Water+turbine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DWater%2Bturbine"><span>Two and three-dimensional prediffuser combustor studies with <span class="hlt">air-water</span> mixture</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Laing, Peter; Ehresman, C. M.; Murthy, S. N. B.</p> <p>1993-01-01</p> <p>Two- and three-dimensional <span class="hlt">gas</span> turbine prediffuser-combustor sectors were experimentally studied under a number of mixture and flow conditions in a tunnel operating with a two-phase, <span class="hlt">air</span>-liquid film-droplet mixture. It is concluded that <span class="hlt">water</span> vaporization in the combustor causes changes in both local <span class="hlt">gas</span> temperature and state of vitiation and reduces reaction rates. Substantial accumulation of <span class="hlt">water</span> and <span class="hlt">water</span> vapor takes place in pocket over the combustor volume, even when the <span class="hlt">air-water</span> mixture is steady in time. The accuracy of determining combustor performance changes increases with a better knowledge of the state of the <span class="hlt">air-water</span> mixture in the primary zone. To establish flame-out conditions it is considered to be necessary to combine the prediction of detailed flowfield and chemical activity with that of flame stability and motion characteristics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.B11B0442B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.B11B0442B"><span>Regulation of leaf-<span class="hlt">gas</span> <span class="hlt">exchange</span> strategies of woody plants under elevated CO2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Belmecheri, S.; Guerrieri, R.; Voelker, S.</p> <p>2016-12-01</p> <p>Estimates of vegetation <span class="hlt">water</span> use efficiency (WUE) have increasingly been assessed using both eddy covariance and plant stable isotope techniques but these data have often lead to differing conclusions. Eddy covariance can provide forest ecosystem-level responses of coupled carbon and <span class="hlt">water</span> <span class="hlt">exchanges</span> to recent global change phenomena. These direct observations, however, are generally less than one or two decades, thus documenting ecosystem-level responses at elevated [CO2] concentrations (350-400 ppm). Therefore, eddy covariance data cannot directly address plant physiological mechanisms and adaptation to climate variability and anthropogenic factors, e.g., increasing atmospheric [CO2]. By contrast, tree based carbon isotope approaches can retrospectively assess intrinsic WUE over long periods and have documented physiological responses to ambient atmospheric [CO2] (ca), which have often been contextualized within generalized strategies for stomatal regulation of leaf <span class="hlt">gas-exchange</span>. These include maintenance of a constant leaf internal [CO2] (ci), a constant drawdown in [CO2] (ca - ci), and a constant ci/ca . Tree carbon isotope studies, however, cannot account for changes in leaf area of individual trees or canopies, which makes scaling up a difficult task. The limitations of these different approaches to understanding how forest <span class="hlt">water</span> use efficiency has been impacted by rising [CO2] has contributed to the uncertainty in global terrestrial carbon cycling and the "missing" terrestrial carbon sink. We examined stable C isotope ratios (d13C) from woody plants over a wide range of [CO2] (200-400 ppm) to test for patterns of ci-regulation in response to rising ca. The analyses are not consistent with any of the leaf <span class="hlt">gas-exchange</span> regulation strategies noted above. The data suggest that ca - ci is still recently increasing in most species but that the rate of increase is less than expected from paleo trees which grew at much lower [CO2]. This evidence demonstrates that a</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://www.dtic.mil/docs/citations/ADA603185','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA603185"><span>Elimination of Acid Cleaning of High Temperature Salt <span class="hlt">Water</span> Heat <span class="hlt">Exchangers</span>: Redesigned Pre-Production Full-Scale Heat Pipe Bleed <span class="hlt">Air</span> Cooler for Shipboard Evaluation</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2011-11-01</p> <p>Cleaning of High Temperature Salt <span class="hlt">Water</span> Heat <span class="hlt">Exchangers</span> ESTCP WP-200302 Subtitle: Redesigned Pre-production Full-Scale Heat Pipe Bleed <span class="hlt">Air</span> Cooler For...FINAL 3. DATES COVERED (From - To) 1-Jan-2003 – 1-Oct-2009 4. TITLE AND SUBTITLE Elimination of Acid Cleaning of High Temperature Salt <span class="hlt">Water</span> Heat...6-5 Figure 6- 6 HP-BAC Tube Sheet Being Immersed in Ultrasonic Cleaning Tank ..................................... 6-6 Figure 6- 7 Heat Pipe</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28634999','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28634999"><span>Plasticity in <span class="hlt">gas-exchange</span> physiology of mature Scots pine and European larch drive short- and long-term adjustments to changes in <span class="hlt">water</span> availability.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Feichtinger, Linda M; Siegwolf, Rolf T W; Gessler, Arthur; Buchmann, Nina; Lévesque, Mathieu; Rigling, Andreas</p> <p>2017-09-01</p> <p>Adjustment mechanisms of trees to changes in soil-<span class="hlt">water</span> availability over long periods are poorly understood, but crucial to improve estimates of forest development in a changing climate. We compared mature trees of Scots pine (Pinus sylvestris) and European larch (Larix decidua) growing along <span class="hlt">water</span>-permeable channels (irrigated) and under natural conditions (control) at three sites in inner-Alpine dry valleys. At two sites, the irrigation had been stopped in the 1980s. We combined measurements of basal area increment (BAI), tree height and <span class="hlt">gas-exchange</span> physiology (Δ 13 C) for the period 1970-2009. At one site, the Δ 13 C of irrigated pine trees was higher than that of the control in all years, while at the other sites, it differed in pine and larch only in years with dry climatic conditions. During the first decade after the sudden change in <span class="hlt">water</span> availability, the BAI and Δ 13 C of originally irrigated pine and larch trees decreased instantly, but subsequently reached higher levels than those of the control by 2009 (15 years afterwards). We found a high plasticity in the <span class="hlt">gas-exchange</span> physiology of pine and larch and site-specific responses to changes in <span class="hlt">water</span> availability. Our study highlights the ability of trees to adjust to new conditions, thus showing high resilience. © 2017 John Wiley & Sons Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3234695','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3234695"><span>Removal of infused <span class="hlt">water</span> predominantly during insertion (<span class="hlt">water</span> <span class="hlt">exchange</span>) is consistently associated with a greater reduction of pain score - review of randomized controlled trials (RCTs) of <span class="hlt">water</span> method colonoscopy</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Harker, JO; Leung, JW; Siao-Salera, RM; Mann, SK; Ramirez, FC; Friedland, S; Amato, A; Radaelli, F; Paggi, S; Terruzzi, V; Hsieh, YH</p> <p>2011-01-01</p> <p>Introduction Variation in the outcomes in RcTs comparing <span class="hlt">water</span>-related methods and <span class="hlt">air</span> insufflation during the insertion phase of colonoscopy raises challenging questions regarding the approach. This report reviews the impact of <span class="hlt">water</span> <span class="hlt">exchange</span> on the variation in attenuation of pain during colonoscopy by <span class="hlt">water</span>-related methods. Methods Medline (2008 to 2011) searches, abstracts of the 2011 Digestive Disease Week (DDW) and personal communications were considered to identify RcTs that compared <span class="hlt">water</span>-related methods and <span class="hlt">air</span> insufflation to aid insertion of the colonoscope. Results: Since 2008 nine published and one submitted RcTs and five abstracts of RcTs presented at the 2011 DDW have been identified. Thirteen RcTs (nine published, one submitted and one abstract, n=1850) described reduction of pain score during or after colonoscopy (eleven reported statistical significance); the remaining reports described lower doses of medication used, or lower proportion of patients experiencing severe pain in colonoscopy performed with <span class="hlt">water</span>-related methods compared with <span class="hlt">air</span> insufflation (Tables 1 and 2). The <span class="hlt">water</span>-related methods notably differ in the timing of removal of the infused <span class="hlt">water</span> - predominantly during insertion (<span class="hlt">water</span> <span class="hlt">exchange</span>) versus predominantly during withdrawal (<span class="hlt">water</span> immersion). Use of <span class="hlt">water</span> <span class="hlt">exchange</span> was consistently associated with a greater attenuation of pain score in patients who did not receive full sedation (Table 3). Conclusion The comparative data reveal that a greater attenuation of pain was associated with <span class="hlt">water</span> <span class="hlt">exchange</span> than <span class="hlt">water</span> immersion during insertion. The intriguing results should be subjected to further evaluation by additional RcTs to elucidate the mechanism of the pain-alleviating impact of the <span class="hlt">water</span> method. PMID:22163081</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19740006538','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19740006538"><span>Effect of <span class="hlt">water</span> injection on nitric oxide emissions of a <span class="hlt">gas</span> turbine combustor burning natural <span class="hlt">gas</span> fuel</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Marchionna, N. R.; Diehl, L. A.; Trout, A. M.</p> <p>1973-01-01</p> <p>The effect of direct <span class="hlt">water</span> injection on the exhaust <span class="hlt">gas</span> emissions of a turbojet combustor burning natural <span class="hlt">gas</span> fuel was investigated. The results are compared with the results from similar tests using ASTM Jet-A fuel. Increasing <span class="hlt">water</span> injection decreased the emissions of oxides of nitrogen (NOX) and increased the emissions of carbon monoxide and unburned hydrocarbons. The greatest percentage decrease in NOX with increasing <span class="hlt">water</span> injection was at the lowest inlet-<span class="hlt">air</span> temperature tested. The effect of increasing inlet-<span class="hlt">air</span> temperature was to decrease the effect of the <span class="hlt">water</span> injection. The reduction in NOX due to <span class="hlt">water</span> injection was almost identical to the results obtained with Jet-A fuel. However, the emission indices of unburned hydrocarbons, carbon monoxide, and percentage nitric oxide in NOX were not.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22675191','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22675191"><span>Respiratory dynamics of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> in the tracheal system of the desert locust, Schistocerca gregaria.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Groenewald, Berlizé; Hetz, Stefan K; Chown, Steven L; Terblanche, John S</p> <p>2012-07-01</p> <p><span class="hlt">Gas</span> <span class="hlt">exchange</span> dynamics in insects is of fundamental importance to understanding evolved variation in breathing patterns, such as discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> cycles (DGCs). Most insects do not rely solely on diffusion for the <span class="hlt">exchange</span> of respiratory gases but may also make use of respiratory movements (active ventilation) to supplement <span class="hlt">gas</span> <span class="hlt">exchange</span> at rest. However, their temporal dynamics have not been widely investigated. Here, intratracheal pressure, V(CO2) and body movements of the desert locust Schistocerca gregaria were measured simultaneously during the DGC and revealed several important aspects of <span class="hlt">gas</span> <span class="hlt">exchange</span> dynamics. First, S. gregaria employs two different ventilatory strategies, one involving dorso-ventral contractions and the other longitudinal telescoping movements. Second, although a true spiracular closed (C)-phase of the DGC could be identified by means of subatmospheric intratracheal pressure recordings, some CO(2) continued to be released. Third, strong pumping actions do not necessarily lead to CO(2) release and could be used to ensure mixing of gases in the closed tracheal system, or enhance <span class="hlt">water</span> vapour reabsorption into the haemolymph from fluid-filled tracheole tips by increasing the hydrostatic pressure or forcing fluid into the haemocoel. Finally, this work showed that the C-phase of the DGC can occur at any pressure. These results provide further insights into the mechanistic basis of insect <span class="hlt">gas</span> <span class="hlt">exchange</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930057487&hterms=water&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D70%26Ntt%3Dwater','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930057487&hterms=water&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D70%26Ntt%3Dwater"><span>A novel membrane device for the removal of <span class="hlt">water</span> vapor and <span class="hlt">water</span> droplets from <span class="hlt">air</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ray, Rod; Newbold, David D.; Mccray, Scott B.; Friesen, Dwayne T.; Kliss, Mark</p> <p>1992-01-01</p> <p>One of the key challenges facing NASA engineers is the development of systems for separating liquids and gases in microgravity environments. In this paper, a novel membrane-based phase separator is described. This device, known as a <span class="hlt">water</span> recovery heat <span class="hlt">exchanger</span> (WRHEX), overcomes the inherent deficiencies of current phase-separation technology. Specifically, the WRHEX cools and removes <span class="hlt">water</span> vapor or <span class="hlt">water</span> droplets from feed-<span class="hlt">air</span> streams without the use of a vacuum or centrifugal force. As is shown in this paper, only a low-power <span class="hlt">air</span> blower and a small stream of recirculated cool <span class="hlt">water</span> is required for WRHEX operation. This paper presents the results of tests using this novel membrane device over a wide range of operating conditions. The data show that the WRHEX produces a dry <span class="hlt">air</span> stream containing no entrained or liquid <span class="hlt">water</span> - even when the feed <span class="hlt">air</span> contains <span class="hlt">water</span> droplets or mist. An analysis of the operation of the WRHEX is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Th%26Ae..24..483K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Th%26Ae..24..483K"><span>Estimate for interstage <span class="hlt">water</span> injection in <span class="hlt">air</span> compressor incorporated into <span class="hlt">gas</span>-turbine cycles and combined power plants cycles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kler, A. M.; Zakharov, Yu. B.; Potanina, Yu. M.</p> <p>2017-05-01</p> <p>The objects of study are the <span class="hlt">gas</span> turbine (GT) plant and combined cycle power plant (CCPP) with opportunity for injection between the stages of <span class="hlt">air</span> compressor. The objective of this paper is technical and economy optimization calculations for these classes of plants with <span class="hlt">water</span> interstage injection. The integrated development environment "System of machine building program" was a tool for creating the mathematic models for these classes of power plants. Optimization calculations with the criterion of minimum for specific capital investment as a function of the unit efficiency have been carried out. For a <span class="hlt">gas</span>-turbine plant, the economic gain from <span class="hlt">water</span> injection exists for entire range of power efficiency. For the combined cycle plant, the economic benefit was observed only for a certain range of plant's power efficiency.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5592373-quality-site-seasonal-report-army-air-force-exchange-service-headquarters-building-sfbp-august-through-may','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5592373-quality-site-seasonal-report-army-air-force-exchange-service-headquarters-building-sfbp-august-through-may"><span>Quality site seasonal report: Army <span class="hlt">Air</span> Force <span class="hlt">Exchange</span> Service Headquarters Building, SFBP 1343, August 1984 through May 1985</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>Pollock, E.O. Jr.</p> <p>1987-10-15</p> <p>The active solar Domestic Hot <span class="hlt">Water</span> (DHW) system at the HQ Army-<span class="hlt">Air</span> Force <span class="hlt">Exchange</span> Service (AAFES) Building was designed and constructed as part of the Solar in Federal Buildings Programs (SFBP). This retrofitted system is one of eight of the systems in the SFBP selected for quality monitoring. The purpose of this monitoring effort is to document the performance of quality state-of-the-art solar systems in large federal building applications. The six-story HQ AAFES Building houses a cafeteria, officer's mess and club and office space for 2400 employees. The siphon-return drainback system uses 1147 ft/sup 2/ of Aircraftsman flat-plate collectors tomore » collect solar energy which is used to preheat domestic hot <span class="hlt">water</span>. Solar energy is stored in a 1329-gallon tank and transferred to the hot <span class="hlt">water</span> load through a heat <span class="hlt">exchanger</span> located in the 356-gallon DHW preheat tank. Auxiliary energy is supplied by two <span class="hlt">gas</span> fired boilers which boost the temperature to 130/sup 0/F before it is distributed to the load. Highlights of the performance of the HQ AAFES Building solar system during the monitoring period from August 1984 through May 1985 are presented in this report.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1225372','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1225372"><span>External reflection FTIR of peptide monolayer films in situ at the <span class="hlt">air/water</span> interface: experimental design, spectra-structure correlations, and effects of hydrogen-deuterium <span class="hlt">exchange</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>Flach, C R; Brauner, J W; Taylor, J W; Baldwin, R C; Mendelsohn, R</p> <p>1994-01-01</p> <p>A Fourier transform infrared spectrometer has been interfaced with a surface balance and a new external reflection infrared sampling accessory, which permits the acquisition of spectra from protein monolayers in situ at the <span class="hlt">air/water</span> interface. The accessory, a sample shuttle that permits the collection of spectra in alternating fashion from sample and background troughs, reduces interference from <span class="hlt">water</span> vapor rotation-vibration bands in the amide I and amide II regions of protein spectra (1520-1690 cm-1) by nearly an order of magnitude. Residual interference from <span class="hlt">water</span> vapor absorbance ranges from 50 to 200 microabsorbance units. The performance of the device is demonstrated through spectra of synthetic peptides designed to adopt alpha-helical, antiparallel beta-sheet, mixed beta-sheet/beta-turn, and unordered conformations at the <span class="hlt">air/water</span> interface. The extent of <span class="hlt">exchange</span> on the surface can be monitored from the relative intensities of the amide II and amide I modes. Hydrogen-deuterium <span class="hlt">exchange</span> may lower the amide I frequency by as much as 11-12 cm-1 for helical secondary structures. This shifts the vibrational mode into a region normally associated with unordered structures and leads to uncertainties in the application of algorithms commonly used for determination of secondary structure from amide I contours of proteins in D2O solution. PMID:7919013</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24749994','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24749994"><span>Effect of impeller design and spacing on <span class="hlt">gas</span> <span class="hlt">exchange</span> in a percutaneous respiratory assist catheter.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jeffries, R Garrett; Frankowski, Brian J; Burgreen, Greg W; Federspiel, William J</p> <p>2014-12-01</p> <p>Providing partial respiratory assistance by removing carbon dioxide (CO2 ) can improve clinical outcomes in patients suffering from acute exacerbations of chronic obstructive pulmonary disease and acute respiratory distress syndrome. An intravenous respiratory assist device with a small (25 Fr) insertion diameter eliminates the complexity and potential complications associated with external blood circuitry and can be inserted by nonspecialized surgeons. The impeller percutaneous respiratory assist catheter (IPRAC) is a highly efficient CO2 removal device for percutaneous insertion to the vena cava via the right jugular or right femoral vein that utilizes an array of impellers rotating within a hollow-fiber membrane bundle to enhance <span class="hlt">gas</span> <span class="hlt">exchange</span>. The objective of this study was to evaluate the effects of new impeller designs and impeller spacing on <span class="hlt">gas</span> <span class="hlt">exchange</span> in the IPRAC using computational fluid dynamics (CFD) and in vitro deionized <span class="hlt">water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> testing. A CFD <span class="hlt">gas</span> <span class="hlt">exchange</span> and flow model was developed to guide a progressive impeller design process. Six impeller blade geometries were designed and tested in vitro in an IPRAC device with 2- or 10-mm axial spacing and varying numbers of blades (2-5). The maximum CO2 removal efficiency (<span class="hlt">exchange</span> per unit surface area) achieved was 573 ± 8 mL/min/m(2) (40.1 mL/min absolute). The <span class="hlt">gas</span> <span class="hlt">exchange</span> rate was found to be largely independent of blade design and number of blades for the impellers tested but increased significantly (5-10%) with reduced axial spacing allowing for additional shaft impellers (23 vs. 14). CFD <span class="hlt">gas</span> <span class="hlt">exchange</span> predictions were within 2-13% of experimental values and accurately predicted the relative improvement with impellers at 2- versus 10-mm axial spacing. The ability of CFD simulation to accurately forecast the effects of influential design parameters suggests it can be used to identify impeller traits that profoundly affect facilitated <span class="hlt">gas</span> <span class="hlt">exchange</span>. Copyright © 2014 International Center for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1232686','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1232686"><span>Low Cost Polymer heat <span class="hlt">Exchangers</span> for Condensing Boilers</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>Butcher, Thomas; Trojanowski, Rebecca; Wei, George</p> <p>2015-09-30</p> <p>Work in this project sought to develop a suitable design for a low cost, corrosion resistant heat <span class="hlt">exchanger</span> as part of a high efficiency condensing boiler. Based upon the design parameters and cost analysis several geometries and material options were explored. The project also quantified and demonstrated the durability of the selected polymer/filler composite under expected operating conditions. The core material idea included a polymer matrix with fillers for thermal conductivity improvement. While the work focused on conventional heating oil, this concept could also be applicable to natural <span class="hlt">gas</span>, low sulfur heating oil, and biodiesel- although these are considered tomore » be less challenging environments. An extruded polymer composite heat <span class="hlt">exchanger</span> was designed, built, and tested during this project, demonstrating technical feasibility of this corrosion-resistant material approach. In such flue <span class="hlt">gas-to-air</span> heat <span class="hlt">exchangers</span>, the controlling resistance to heat transfer is in the <span class="hlt">gas</span>-side convective layer and not in the tube material. For this reason, the lower thermal conductivity polymer composite heat <span class="hlt">exchanger</span> can achieve overall heat transfer performance comparable to a metal heat <span class="hlt">exchanger</span>. However, with the polymer composite, the surface temperature on the <span class="hlt">gas</span> side will be higher, leading to a lower <span class="hlt">water</span> vapor condensation rate.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.B51F0483P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.B51F0483P"><span>Automatable Measurement of <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Rate in Streams: Oxygen-Carbon Method</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pennington, R.; Haggerty, R.; Argerich, A.; Wondzell, S. M.</p> <p>2015-12-01</p> <p><span class="hlt">Gas</span> <span class="hlt">exchange</span> rates between streams and the atmosphere are critically important to measurement of in-stream ecologic processes, as well as fate and transport of hazardous pollutants such as mercury and PCBs. Methods to estimate <span class="hlt">gas</span> <span class="hlt">exchange</span> rates include empirical relations to hydraulics, and direct injection of a tracer <span class="hlt">gas</span> such as propane or SF6. Empirical relations are inconsistent and inaccurate, particularly for lower order, high-roughness streams. <span class="hlt">Gas</span> injections are labor-intensive, and measured <span class="hlt">gas</span> <span class="hlt">exchange</span> rates are difficult to extrapolate in time since they change with discharge and stream geometry. We propose a novel method for calculation of <span class="hlt">gas</span> <span class="hlt">exchange</span> rates utilizing O2, pCO2, pH, and temperature data. Measurements, which can be automated using data loggers and probes, are made on the upstream and downstream end of the study reach. <span class="hlt">Gas</span> <span class="hlt">exchange</span> rates are then calculated from a solution to the transport equations for oxygen and dissolved inorganic carbon. Field tests in steep, low order, high roughness streams of the HJ Andrews Experimental Forest indicate the method to be viable along stream reaches with high downstream <span class="hlt">gas</span> concentration gradients and high rates of <span class="hlt">gas</span> transfer velocity. Automated and continuous collection of oxygen and carbonate chemistry data is increasingly common, thus the method may be used to estimate <span class="hlt">gas</span> <span class="hlt">exchange</span> rates through time, and is well suited for interactivity with databases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeCoA.194..291B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeCoA.194..291B"><span>Noble gases solubility models of hydrocarbon charge mechanism in the Sleipner Vest <span class="hlt">gas</span> field</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barry, P. H.; Lawson, M.; Meurer, W. P.; Warr, O.; Mabry, J. C.; Byrne, D. J.; Ballentine, C. J.</p> <p>2016-12-01</p> <p>Noble gases are chemically inert and variably soluble in crustal fluids. They are primarily introduced into hydrocarbon reservoirs through <span class="hlt">exchange</span> with formation <span class="hlt">waters</span>, and can be used to assess migration pathways and mechanisms, as well as reservoir storage conditions. Of particular interest is the role groundwater plays in hydrocarbon transport, which is reflected in hydrocarbon-<span class="hlt">water</span> volume ratios. Here, we present compositional, stable isotope and noble <span class="hlt">gas</span> isotope and abundance data from the Sleipner Vest field, in the Norwegian North Sea. Sleipner Vest gases are generated from primary cracking of kerogen and the thermal cracking of oil. <span class="hlt">Gas</span> was emplaced into the Sleipner Vest from the south and subsequently migrated to the east, filling and spilling into the Sleipner Ost fields. Gases principally consist of hydrocarbons (83-93%), CO2 (5.4-15.3%) and N2 (0.6-0.9%), as well as trace concentrations of noble gases. Helium isotopes (3He/4He) are predominantly radiogenic and range from 0.065 to 0.116 RA; reported relative to <span class="hlt">air</span> (RA = 1.4 × 10-6; Clarke et al., 1976; Sano et al., 1988), showing predominantly (>98%) crustal contributions, consistent with Ne (20Ne/22Ne from 9.70 to 9.91; 21Ne/22Ne from 0.0290 to 0.0344) and Ar isotopes (40Ar/36Ar from 315 to 489). <span class="hlt">Air</span>-derived noble <span class="hlt">gas</span> isotopes (20Ne, 36Ar, 84Kr, 132Xe) are introduced into the hydrocarbon system by direct <span class="hlt">exchange</span> with <span class="hlt">air</span>-saturated <span class="hlt">water</span> (ASW). The distribution of <span class="hlt">air</span>-derived noble <span class="hlt">gas</span> species are controlled by phase partitioning processes; in that they preferentially partition into the <span class="hlt">gas</span> (i.e., methane) phase, due to their low solubilities in fluids. Therefore, the extent of <span class="hlt">exchange</span> between hydrocarbon phases and formation <span class="hlt">waters</span> - that have previously equilibrated with the atmosphere - can be determined by investigating <span class="hlt">air</span>-derived noble <span class="hlt">gas</span> species. We utilize both elemental ratios to address process (i.e., open vs. closed system) and concentrations to quantify the extent of hydrocarbon-<span class="hlt">water</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.H23C0898R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.H23C0898R"><span>Rapid, Real-time Methane Detection in Ground <span class="hlt">Water</span> Using a New <span class="hlt">Gas-Water</span> Equilibrator Design</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ruybal, C. J.; DiGiulio, D. C.; Wilkin, R. T.; Hargrove, K. D.; McCray, J. E.</p> <p>2014-12-01</p> <p>Recent increases in unconventional <span class="hlt">gas</span> development have been accompanied by public concern for methane contamination in drinking <span class="hlt">water</span> wells near production areas. Although not a regulated pollutant, methane may be a marker contaminant for others that are less mobile in groundwater and thus may be detected later, or at a location closer to the source. In addition, methane poses an explosion hazard if exsolved concentrations reach 5 - 15% volume in <span class="hlt">air</span>. Methods for determining dissolved gases, such as methane, have evolved over 60 years. However, the response time of these methods is insufficient to monitor trends in methane concentration in real-time. To enable rapid, real-time monitoring of aqueous methane concentrations during ground <span class="hlt">water</span> purging, a new <span class="hlt">gas-water</span> equilibrator (GWE) was designed that increases <span class="hlt">gas-water</span> mass <span class="hlt">exchange</span> rates of methane for measurement. Monitoring of concentration trends allows a comparison of temporal trends between sampling events and comparison of baseline conditions with potential post-impact conditions. These trends may be a result of removal of stored casing <span class="hlt">water</span>, pre-purge ambient borehole flow, formation physical and chemical heterogeneity, or flow outside of well casing due to inadequate seals. Real-time information in the field can help focus an investigation, aid in determining when to collect a sample, save money by limiting costs (e.g. analytical, sample transport and storage), and provide an immediate assessment of local methane concentrations. Four domestic <span class="hlt">water</span> wells, one municipal <span class="hlt">water</span> well, and one agricultural <span class="hlt">water</span> well were sampled for traditional laboratory analysis and compared to the field GWE results. Aqueous concentrations measured on the GWE ranged from non-detect to 1,470 μg/L methane. Some trends in aqueous methane concentrations measured on the GWE were observed during purging. Applying a paired t-test comparing the new GWE method and traditional laboratory analysis yielded a p-value 0</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010APS..DFD.MX008H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010APS..DFD.MX008H"><span>Coaxial twin-fluid atomization with pattern <span class="hlt">air</span> <span class="hlt">gas</span> streams</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hei Ng, Chin; Aliseda, Alberto</p> <p>2010-11-01</p> <p>Coaxial twin-fluid atomization has numerous industrial applications, most notably fuel injection and spray coating. In the coating process of pharmaceutical tablets, the coaxial atomizing <span class="hlt">air</span> stream is accompanied by two diametrically opposed side jets that impinge on the liquid/<span class="hlt">gas</span> coaxial jets at an angle to produce an elliptical shape of the spray's cross section. Our study focuses on the influence of these side jets on the break up process and on the droplet velocity and diameter distribution along the cross section. The ultimate goal is to predict the size distribution and volume flux per unit area in the spray. With this predictive model, an optimal atomizing <span class="hlt">air</span>/pattern <span class="hlt">air</span> ratio can be found to achieve the desired coating result. This model is also crucial in scaling up the laboratory setup to production level. We have performed experiments with different atomized liquids, such as <span class="hlt">water</span> and glycerine-<span class="hlt">water</span> mixtures, that allow us to establish the effect of liquid viscosity, through the Ohnesorge number, in the spray characteristics. The <span class="hlt">gas</span> Reynolds number of our experiments ranges from 9000 to 18000 and the Weber number ranges from 400 to 1600. We will present the effect of pattern <span class="hlt">air</span> in terms of the resulting droplets size, droplet number density and velocity at various distances downstream of the nozzle where the effect of pattern <span class="hlt">air</span> is significant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1411867','SCIGOV-DOEDE'); return false;" href="https://www.osti.gov/servlets/purl/1411867"><span>CO2 response (ACi) <span class="hlt">gas</span> <span class="hlt">exchange</span>, calculated Vcmax & Jmax parameters, Feb2016-May2016, PA-SLZ, PA-PNM: Panama</span></a></p> <p><a target="_blank" href="http://www.osti.gov/dataexplorer">DOE Data Explorer</a></p> <p>Rogers, Alistair [Brookhaven National Lab; Serbin, Shawn [Brookhaven National Lab; Ely, Kim [Brookhaven National Lab; Wu, Jin [BNL; Wolfe, Brett [Smithsonian; Dickman, Turin [Los Alamos National Lab; Collins, Adam [Los Alamos National Lab; Detto, Matteo [Princeton; Grossiord, Charlotte [Los Alamos National Lab; McDowell, Nate [Los Alamos National Lab; Michaletz, Sean</p> <p>2017-01-01</p> <p>CO2 response (ACi) <span class="hlt">gas</span> <span class="hlt">exchange</span> measured on leaves collected from sunlit canopy trees on a monthly basis from Feb to May 2016 at SLZ and PNM. Dataset includes calculated Vcmax and Jmax parameters. This data was collected as part of the 2016 ENSO campaign. See related datasets (existing and future) for further sample details, leaf <span class="hlt">water</span> potential, LMA, leaf spectra, other <span class="hlt">gas</span> <span class="hlt">exchange</span> and leaf chemistry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110000601','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110000601"><span><span class="hlt">Air</span> Circulation and Heat <span class="hlt">Exchange</span> Under Reduced Pressures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rygalov, V.; Wheeler, R.; Dixon, M.; Fowler, P.; Hillhouse, L.</p> <p>2010-01-01</p> <p>Heat <span class="hlt">exchange</span> rates decrease non-linearly with reductions in atmospheric pressure. This decrease creates risk of thermal stress (elevated leaf temperatures) for plants under reduced pressures. Forced convection (fans) significantly increases heat <span class="hlt">exchange</span> rate under almost all pressures except below 10 kPa. Plant cultivation techniques under reduced pressures will require forced convection. The cooling curve technique is a reliable means of assessing the influence of environmental variables like pressure and gravity on <span class="hlt">gas</span> <span class="hlt">exchange</span> of plant. These results represent the extremes of <span class="hlt">gas</span> <span class="hlt">exchange</span> conditions for simple systems under variable pressures. In reality, dense plant canopies will exhibit responses in between these extremes. More research is needed to understand the dependence of forced convection on atmospheric pressure. The overall thermal balance model should include latent and radiative <span class="hlt">exchange</span> components.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24395505','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24395505"><span>Fungal oxygen <span class="hlt">exchange</span> between denitrification intermediates and <span class="hlt">water</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rohe, Lena; Anderson, Traute-Heidi; Braker, Gesche; Flessa, Heinz; Giesemann, Anette; Wrage-Mönnig, Nicole; Well, Reinhard</p> <p>2014-02-28</p> <p>Fungi can contribute greatly to N2O production from denitrification. Therefore, it is important to quantify the isotopic signature of fungal N2O. The isotopic composition of N2O can be used to identify and analyze the processes of N2O production and N2O reduction. In contrast to bacteria, information about the oxygen <span class="hlt">exchange</span> between denitrification intermediates and <span class="hlt">water</span> during fungal denitrification is lacking, impeding the explanatory power of stable isotope methods. Six fungal species were anaerobically incubated with the electron acceptors nitrate or nitrite and (18)O-labeled <span class="hlt">water</span> to determine the oxygen <span class="hlt">exchange</span> between denitrification intermediates and <span class="hlt">water</span>. After seven days of incubation, <span class="hlt">gas</span> samples were analyzed for N2O isotopologues by isotope ratio mass spectrometry. All the fungal species produced N2O. N2O production was greater when nitrite was the sole electron acceptor (129 to 6558 nmol N2O g dw(-1)  h(-1)) than when nitrate was the electron acceptor (6 to 47 nmol N2O g dw(-1)  h(-1)). Oxygen <span class="hlt">exchange</span> was complete with nitrate as electron acceptor in one of five fungi and with nitrite in two of six fungi. Oxygen <span class="hlt">exchange</span> of the other fungi varied (41 to 89% with nitrite and 11 to 61% with nitrate). This is the first report on oxygen <span class="hlt">exchange</span> with <span class="hlt">water</span> during fungal denitrification. The <span class="hlt">exchange</span> appears to be within the range previously reported for bacterial denitrification. This adds to the difficulty of differentiating N2O producing processes based on the origin of N2O-O. However, the large oxygen <span class="hlt">exchange</span> repeatedly observed for bacteria and now also fungi could lead to less variability in the δ(18)O values of N2O from soils, which could facilitate the assessment of the extent of N2O reduction. Copyright © 2013 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ACP....1611125W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ACP....1611125W"><span>Emission-dominated <span class="hlt">gas</span> <span class="hlt">exchange</span> of elemental mercury vapor over natural surfaces in 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, Xun; Lin, Che-Jen; Yuan, Wei; Sommar, Jonas; Zhu, Wei; Feng, Xinbin</p> <p>2016-09-01</p> <p>Mercury (Hg) emission from natural surfaces plays an important role in global Hg cycling. The present estimate of global natural emission has large uncertainty and remains unverified against field data, particularly for terrestrial surfaces. In this study, a mechanistic model is developed for estimating the emission of elemental mercury vapor (Hg0) from natural surfaces in China. The development implements recent advancements in the understanding of <span class="hlt">air</span>-soil and <span class="hlt">air</span>-foliage <span class="hlt">exchange</span> of Hg0 and redox chemistry in soil and on surfaces, incorporates the effects of soil characteristics and land use changes by agricultural activities, and is examined through a systematic set of sensitivity simulations. Using the model, the net <span class="hlt">exchange</span> of Hg0 between the atmosphere and natural surfaces of mainland China is estimated to be 465.1 Mg yr-1, including 565.5 Mg yr-1 from soil surfaces, 9.0 Mg yr-1 from <span class="hlt">water</span> bodies, and -100.4 Mg yr-1 from vegetation. The <span class="hlt">air</span>-surface <span class="hlt">exchange</span> is strongly dependent on the land use and meteorology, with 9 % of net emission from forest ecosystems; 50 % from shrubland, savanna, and grassland; 33 % from cropland; and 8 % from other land uses. Given the large agricultural land area in China, farming activities play an important role on the <span class="hlt">air</span>-surface <span class="hlt">exchange</span> over farmland. Particularly, rice field shift from a net sink (3.3 Mg uptake) during April-October (rice planting) to a net source when the farmland is not flooded (November-March). Summing up the emission from each land use, more than half of the total emission occurs in summer (51 %), followed by spring (28 %), autumn (13 %), and winter (8 %). Model verification is accomplished using observational data of <span class="hlt">air-soil/air-water</span> fluxes and Hg deposition through litterfall for forest ecosystems in China and Monte Carlo simulations. In contrast to the earlier estimate by Shetty et al. (2008) that reported large emission from vegetative surfaces using an evapotranspiration approach, the estimate in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880002873','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880002873"><span>Operation of an experimental algal <span class="hlt">gas</span> <span class="hlt">exchanger</span> for use in a CELSS</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smernoff, David T.; Wharton, Robert A., Jr.; Averner, Maurice M.</p> <p>1987-01-01</p> <p>Concepts of a Closed Ecological Life Support System (CELSS) anticipate the use of photosynthetic organisms (higher plants and algae) for <span class="hlt">air</span> revitalization. The rates of production and uptake of carbon dioxide and oxygen between the crew and the photosynthetic organisms are mismatched. An algal system used for <span class="hlt">gas</span> <span class="hlt">exchange</span> only will have the difficulty of an accumulation or depletion of these gases beyond physiologically tolerable limits (in a closed system the mismatch between assimilatory quotient (AQ) and respiratory quotient (RQ) is balanced by the operation of the waste processor). The results are given of a study designed to test the feasibility of using environmental manipulations to maintain physiologically appropriate atmospheres for algae and mice in a <span class="hlt">gas</span> closed system. Specifically, the atmosphere behavior of this system is considered with algae grown on nitrate or urea and at different light intensities and optical densities. Manipulation of both allow operation of the system in a <span class="hlt">gas</span> stable manner. Operation of such a system in a CELSS may be useful for reduction of buffer sizes, as a backup system for higher plant <span class="hlt">air</span> revitalization and to supply extra oxygen to the waste processor or during crew changes.</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/26436513','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26436513"><span>Concentrations, Trends, and <span class="hlt">Air-Water</span> <span class="hlt">Exchange</span> of PAHs and PBDEs Derived from Passive Samplers in Lake Superior in 2011.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ruge, Zoe; Muir, Derek; Helm, Paul; Lohmann, Rainer</p> <p>2015-12-01</p> <p>Polycyclic aromatic hydrocarbons (PAHs) and polybrominated diphenylethers (PBDEs) are both currently released into the environment from anthropogenic activity. Both are hence primarily associated with populated or industrial areas, although wildfires can be an important source of PAHs, as well. Polyethylene passive samplers (PEs) were simultaneously deployed in surface <span class="hlt">water</span> and near surface atmosphere to determine spatial trends and <span class="hlt">air-water</span> gaseous <span class="hlt">exchange</span> of 21 PAHs and 11 PBDEs at 19 sites across Lake Superior in 2011. Surface <span class="hlt">water</span> and atmospheric PAH concentrations were greatest at urban sites (up to 65 ng L(-1) and 140 ng m(-3), respectively, averaged from June to October). Near populated regions, PAHs displayed net <span class="hlt">air-to-water</span> deposition, but were near equilibrium off-shore. Retene, probably depositing following major wildfires in the region, dominated dissolved PAH concentrations at most Lake Superior sites. Atmospheric and dissolved PBDEs were greatest near urban and populated sites (up to 6.8 pg L(-1) and 15 pg m(-3), respectively, averaged from June to October), dominated by BDE-47. At most coastal sites, there was net gaseous deposition of BDE-47, with less brominated congeners contributing to Sault Ste. Marie and eastern open lake fluxes. Conversely, the central open lake and Eagle Harbor sites generally displayed volatilization of PBDEs into the atmosphere, mainly BDE-47.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22923680','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22923680"><span>The grapevine root-specific aquaporin VvPIP2;4N controls root hydraulic conductance and leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> under well-<span class="hlt">watered</span> conditions but not under <span class="hlt">water</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>Perrone, Irene; Gambino, Giorgio; Chitarra, Walter; Vitali, Marco; Pagliarani, Chiara; Riccomagno, Nadia; Balestrini, Raffaella; Kaldenhoff, Ralf; Uehlein, Norbert; Gribaudo, Ivana; Schubert, Andrea; Lovisolo, Claudio</p> <p>2012-10-01</p> <p>We functionally characterized the grape (Vitis vinifera) VvPIP2;4N (for Plasma membrane Intrinsic Protein) aquaporin gene. Expression of VvPIP2;4N in Xenopus laevis oocytes increased their swelling rate 54-fold. Northern blot and quantitative reverse transcription-polymerase chain reaction analyses showed that VvPIP2;4N is the most expressed PIP2 gene in root. In situ hybridization confirmed root localization in the cortical parenchyma and close to the endodermis. We then constitutively overexpressed VvPIP2;4N in grape 'Brachetto', and in the resulting transgenic plants we analyzed (1) the expression of endogenous and transgenic VvPIP2;4N and of four other aquaporins, (2) whole-plant, root, and leaf ecophysiological parameters, and (3) leaf abscisic acid content. Expression of transgenic VvPIP2;4N inhibited neither the expression of the endogenous gene nor that of other PIP aquaporins in both root and leaf. Under well-<span class="hlt">watered</span> conditions, transgenic plants showed higher stomatal conductance, <span class="hlt">gas</span> <span class="hlt">exchange</span>, and shoot growth. The expression level of VvPIP2;4N (endogenous + transgene) was inversely correlated to root hydraulic resistance. The leaf component of total plant hydraulic resistance was low and unaffected by overexpression of VvPIP2;4N. Upon <span class="hlt">water</span> stress, the overexpression of VvPIP2;4N induced a surge in leaf abscisic acid content and a decrease in stomatal conductance and leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>. Our results show that aquaporin-mediated modifications of root hydraulics play a substantial role in the regulation of <span class="hlt">water</span> flow in well-<span class="hlt">watered</span> grapevine plants, while they have a minor role upon drought, probably because other signals, such as abscisic acid, take over the control of <span class="hlt">water</span> flow.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23845983','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23845983"><span>Probing the regional distribution of pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> through single-breath <span class="hlt">gas</span>- and dissolved-phase 129Xe MR imaging.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kaushik, S Sivaram; Freeman, Matthew S; Cleveland, Zackary I; Davies, John; Stiles, Jane; Virgincar, Rohan S; Robertson, Scott H; He, Mu; Kelly, Kevin T; Foster, W Michael; McAdams, H Page; Driehuys, Bastiaan</p> <p>2013-09-01</p> <p>Although some central aspects of pulmonary function (ventilation and perfusion) are known to be heterogeneous, the distribution of diffusive <span class="hlt">gas</span> <span class="hlt">exchange</span> remains poorly characterized. A solution is offered by hyperpolarized 129Xe magnetic resonance (MR) imaging, because this <span class="hlt">gas</span> can be separately detected in the lung's <span class="hlt">air</span> spaces and dissolved in its tissues. Early dissolved-phase 129Xe images exhibited intensity gradients that favored the dependent lung. To quantitatively corroborate this finding, we developed an interleaved, three-dimensional radial sequence to image the gaseous and dissolved 129Xe distributions in the same breath. These images were normalized and divided to calculate "129Xe <span class="hlt">gas</span>-transfer" maps. We hypothesized that, for healthy volunteers, 129Xe <span class="hlt">gas</span>-transfer maps would retain the previously observed posture-dependent gradients. This was tested in nine subjects: when the subjects were supine, 129Xe <span class="hlt">gas</span> transfer exhibited a posterior-anterior gradient of -2.00 ± 0.74%/cm; when the subjects were prone, the gradient reversed to 1.94 ± 1.14%/cm (P < 0.001). The 129Xe <span class="hlt">gas</span>-transfer maps also exhibited significant heterogeneity, as measured by the coefficient of variation, that correlated with subject total lung capacity (r = 0.77, P = 0.015). <span class="hlt">Gas</span>-transfer intensity varied nonmonotonically with slice position and increased in slices proximal to the main pulmonary arteries. Despite substantial heterogeneity, the mean <span class="hlt">gas</span> transfer for all subjects was 1.00 ± 0.01 while supine and 1.01 ± 0.01 while prone (P = 0.25), indicating good "matching" between <span class="hlt">gas</span>- and dissolved-phase distributions. This study demonstrates that single-breath <span class="hlt">gas</span>- and dissolved-phase 129Xe MR imaging yields 129Xe <span class="hlt">gas</span>-transfer maps that are sensitive to altered <span class="hlt">gas</span> <span class="hlt">exchange</span> caused by differences in lung inflation and posture.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017WRR....53.9519L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017WRR....53.9519L"><span>The <span class="hlt">Gas</span>-Absorption/Chemical-Reaction Method for Measuring <span class="hlt">Air-Water</span> Interfacial Area in Natural Porous Media</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lyu, Ying; Brusseau, Mark L.; El Ouni, Asma; Araujo, Juliana B.; Su, Xiaosi</p> <p>2017-11-01</p> <p>The <span class="hlt">gas</span>-absorption/chemical-reaction (GACR) method used in chemical engineering to quantify <span class="hlt">gas</span>-liquid interfacial area in reactor systems is adapted for the first time to measure the effective <span class="hlt">air-water</span> interfacial area of natural porous media. Experiments were conducted with the GACR method, and two standard methods (X-ray microtomographic imaging and interfacial partitioning tracer tests) for comparison, using model glass beads and a natural sand. The results of a series of experiments conducted under identical conditions demonstrated that the GACR method exhibited excellent repeatability for measurement of interfacial area (Aia). Coefficients of variation for Aia were 3.5% for the glass beads and 11% for the sand. Extrapolated maximum interfacial areas (Am) obtained with the GACR method were statistically identical to independent measures of the specific solid surface areas of the media. For example, the Am for the glass beads is 29 (±1) cm-1, compared to 32 (±3), 30 (±2), and 31 (±2) cm-1 determined from geometric calculation, N2/BET measurement, and microtomographic measurement, respectively. This indicates that the method produced accurate measures of interfacial area. Interfacial areas determined with the GACR method were similar to those obtained with the standard methods. For example, Aias of 47 and 44 cm-1 were measured with the GACR and XMT methods, respectively, for the sand at a <span class="hlt">water</span> saturation of 0.57. The results of the study indicate that the GACR method is a viable alternative for measuring <span class="hlt">air-water</span> interfacial areas. The method is relatively quick, inexpensive, and requires no specialized instrumentation compared to the standard methods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22268690','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22268690"><span>Sorption-induced effects of humic substances on mass transfer of organic pollutants through aqueous diffusion boundary layers: the example of <span class="hlt">water/air</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ramus, Ksenia; Kopinke, Frank-Dieter; Georgi, Anett</p> <p>2012-02-21</p> <p>This study examines the effect of dissolved humic substances (DHS) on the rate of <span class="hlt">water-gas</span> <span class="hlt">exchange</span> of organic compounds under conditions where diffusion through the aqueous boundary layer is rate-determining. A synthetic surfactant was applied for comparison. Mass-transfer coefficients were determined from the rate of depletion of the model compounds by means of an apparatus containing a stirred aqueous solution with continuous purging of the headspace above the solution. In addition, experiments with continuous passive dosing of analytes into the <span class="hlt">water</span> phase were conducted to simulate a system where thermodynamic activity of the chemical in the aqueous phase is identical in the presence and absence of DHS. The experimental results show that DHS and surfactants can affect <span class="hlt">water-gas</span> <span class="hlt">exchange</span> rates by the superposition of two mechanisms: (1) hydrodynamic effects due to surface film formation ("surface smoothing"), and (2) sorption-induced effects. Whether sorption accelerates or retards mass transfer depends on its effect on the thermodynamic activity of the pollutant in the aqueous phase. Mass transfer will be retarded if the activity (or freely dissolved concentration) of the pollutant is decreased due to sorption. If it remains unchanged (e.g., due to fast equilibration with a sediment acting as a large source phase), then DHS and surfactant micelles can act as an additional shuttle for the pollutants, enhancing the flux through the boundary layer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28965822','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28965822"><span>A new method for noninvasive measurement of pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> using expired <span class="hlt">gas</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>West, John B; Prisk, G Kim</p> <p>2018-01-01</p> <p>Measurement of the <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency of the lung is often required in the practice of pulmonary medicine and in other settings. The traditional standard is the values of the PO2, PCO2, and pH of arterial blood. However arterial puncture requires technical expertise, is invasive, uncomfortable for the patient, and expensive. Here we describe how the composition of expired <span class="hlt">gas</span> can be used in conjunction with pulse oximetry to obtain useful measures of <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency. The new procedure is noninvasive, well tolerated by the patient, and takes only a few minutes. It could be particularly useful when repeated measurements of pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> are required. One product of the procedure is the difference between the PO2 of end-tidal alveolar <span class="hlt">gas</span> and the calculated PO2 of arterial blood. This measurement is related to the classical alveolar-arterial PO2 difference based on ideal alveolar <span class="hlt">gas</span>. However that traditional index is heavily influenced by lung units with low ventilation-perfusion ratios, whereas the new index has a broader physiological basis because it includes contributions from the whole lung. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19246596','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19246596"><span><span class="hlt">Water</span> and nitrogen conditions affect the relationships of Delta13C and Delta18O to <span class="hlt">gas</span> <span class="hlt">exchange</span> and growth in durum wheat.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cabrera-Bosquet, Llorenç; Molero, Gemma; Nogués, Salvador; Araus, José Luis</p> <p>2009-01-01</p> <p>Whereas the effects of <span class="hlt">water</span> and nitrogen (N) on plant Delta(13)C have been reported previously, these factors have scarcely been studied for Delta(18)O. Here the combined effect of different <span class="hlt">water</span> and N regimes on Delta(13)C, Delta(18)O, <span class="hlt">gas</span> <span class="hlt">exchange</span>, <span class="hlt">water</span>-use efficiency (WUE), and growth of four genotypes of durum wheat [Triticum turgidum L. ssp. durum (Desf.) Husn.] cultured in pots was studied. <span class="hlt">Water</span> and N supply significantly increased plant growth. However, a reduction in <span class="hlt">water</span> supply did not lead to a significant decrease in <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters, and consequently Delta(13)C was only slightly modified by <span class="hlt">water</span> input. Conversely, N fertilizer significantly decreased Delta(13)C. On the other hand, <span class="hlt">water</span> supply decreased Delta(18)O values, whereas N did not affect this parameter. Delta(18)O variation was mainly determined by the amount of transpired <span class="hlt">water</span> throughout plant growth (T(cum)), whereas Delta(13)C variation was explained in part by a combination of leaf N and stomatal conductance (g(s)). Even though the four genotypes showed significant differences in cumulative transpiration rates and biomass, this was not translated into significant differences in Delta(18)O(s). However, genotypic differences in Delta(13)C were observed. Moreover, approximately 80% of the variation in biomass across growing conditions and genotypes was explained by a combination of both isotopes, with Delta(18)O alone accounting for approximately 50%. This illustrates the usefulness of combining Delta(18)O and Delta(13)C in order to assess differences in plant growth and total transpiration, and also to provide a time-integrated record of the photosynthetic and evaporative performance of the plant during the course of crop growth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23715084','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23715084"><span>A review of <span class="hlt">air</span> <span class="hlt">exchange</span> rate models for <span class="hlt">air</span> pollution exposure assessments.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Breen, Michael S; Schultz, Bradley D; Sohn, Michael D; Long, Thomas; Langstaff, John; Williams, Ronald; Isaacs, Kristin; Meng, Qing Yu; Stallings, Casson; Smith, Luther</p> <p>2014-11-01</p> <p>A critical aspect of <span class="hlt">air</span> pollution exposure assessments is estimation of the <span class="hlt">air</span> <span class="hlt">exchange</span> rate (AER) for various buildings where people spend their time. The AER, which is the rate of <span class="hlt">exchange</span> of indoor <span class="hlt">air</span> with outdoor <span class="hlt">air</span>, is an important determinant for entry of outdoor <span class="hlt">air</span> pollutants and for removal of indoor-emitted <span class="hlt">air</span> pollutants. This paper presents an overview and critical analysis of the scientific literature on empirical and physically based AER models for residential and commercial buildings; the models highlighted here are feasible for exposure assessments as extensive inputs are not required. Models are included for the three types of airflows that can occur across building envelopes: leakage, natural ventilation, and mechanical ventilation. Guidance is provided to select the preferable AER model based on available data, desired temporal resolution, types of airflows, and types of buildings included in the exposure assessment. For exposure assessments with some limited building leakage or AER measurements, strategies are described to reduce AER model uncertainty. This review will facilitate the selection of AER models in support of <span class="hlt">air</span> pollution exposure assessments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880029838&hterms=Chlorella&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DChlorella','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880029838&hterms=Chlorella&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DChlorella"><span>Operation of an experimental algal <span class="hlt">gas</span> <span class="hlt">exchanger</span> for use in a CELSS</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smernoff, David T.; Wharton, Robert A., Jr.; Averner, Maurice M.</p> <p>1987-01-01</p> <p>Concepts of a CELSS anticipate the use of photosynthetic organisms for <span class="hlt">air</span> revitalization. The rates of production and uptake of carbon dioxide and oxygen between the crew and the photosynthetic organisms are mismatched. An algal system used for <span class="hlt">gas</span> <span class="hlt">exchange</span> only will have the difficulty of an accumulation or depletion of these gases beyond physiologically tolerable limits. The results of a study designed to test the feasibility of using environmental manipulations to maintain physiologically appropriate atmospheres for algae (Chlorella pyrenoidosa) and mice (Mus musculus strain DW/J) in a <span class="hlt">gas</span>-closed system is reported. Specifically, the atmosphere behavior of this system with Chlorella grown on nitrate or urea and at different light intensities and optical densities is considered. Manipulation of both the photosynthetic rate and the assimilatory quotient of the alga has been found to reduce the mismatch of <span class="hlt">gas</span> requirements and allow operation of the system in a <span class="hlt">gas</span>-stable manner.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19825869','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19825869"><span>Xylem anatomy correlates with <span class="hlt">gas</span> <span class="hlt">exchange</span>, <span class="hlt">water</span>-use efficiency and growth performance under contrasting <span class="hlt">water</span> regimes: evidence from Populus deltoides x Populus nigra hybrids.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fichot, Régis; Laurans, Françoise; Monclus, Romain; Moreau, Alain; Pilate, Gilles; Brignolas, Franck</p> <p>2009-12-01</p> <p>Six Populus deltoides Bartr. ex Marsh. x P. nigra L. genotypes were selected to investigate whether stem xylem anatomy correlated with <span class="hlt">gas</span> <span class="hlt">exchange</span> rates, <span class="hlt">water</span>-use efficiency (WUE) and growth performance. Clonal copies of the genotypes were grown in a two-plot common garden test under contrasting <span class="hlt">water</span> regimes, with one plot maintained irrigated and the other one subjected to moderate summer <span class="hlt">water</span> deficit. The six genotypes displayed a large range of xylem anatomy, mean vessel and fibre diameter varying from about 40 to 60 microm and from 7.5 to 10.5 microm, respectively. Decreased <span class="hlt">water</span> availability resulted in a reduced cell size and an important rise in vessel density, but the extent of xylem plasticity was both genotype and trait dependent. Vessel diameter and theoretical xylem-specific hydraulic conductivity correlated positively with stomatal conductance, carbon isotope discrimination and growth performance-related traits and negatively with intrinsic WUE, especially under <span class="hlt">water</span> deficit conditions. Vessel diameter and vessel density measured under <span class="hlt">water</span> deficit conditions correlated with the relative losses in biomass production in response to <span class="hlt">water</span> deprivation; this resulted from the fact that a more plastic xylem structure was generally accompanied by a larger loss in biomass production.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22209165','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22209165"><span>The effect of strobilurins on leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>, <span class="hlt">water</span> use efficiency and ABA content in grapevine under field conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Diaz-Espejo, Antonio; Cuevas, María Victoria; Ribas-Carbo, Miquel; Flexas, Jaume; Martorell, Sebastian; Fernández, José Enrique</p> <p>2012-03-01</p> <p>Strobilurins are one of the most important classes of agricultural fungicides. In addition to their anti-fungal effect, strobilurins have been reported to produce simultaneous effects in plant physiology. This study investigated whether the use of strobilurin fungicide improved <span class="hlt">water</span> use efficiency in leaves of grapevines grown under field conditions in a Mediterranean climate in southern Spain. Fungicide was applied three times in the vineyard and measurements of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>, plant <span class="hlt">water</span> status, abscisic acid concentration in sap ([ABA]), and carbon isotope composition in leaves were performed before and after applications. No clear effect on stomatal conductance, leaf <span class="hlt">water</span> potential and intrinsic <span class="hlt">water</span> use efficiency was found after three fungicide applications. ABA concentration was observed to increase after fungicide application on the first day, vanishing three days later. Despite this transient effect, evolution of [ABA] matched well with the evolution of leaf carbon isotope ratio, which can be used as a surrogate for plant <span class="hlt">water</span> use efficiency. Morning stomatal conductance was negatively correlated to [ABA]. Yield was enhanced in strobilurin treated plants, whereas fruit quality remained unaltered. Published by Elsevier GmbH.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010005746','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010005746"><span>BOREAS TE-5 Leaf <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hall, Forrest G. (Editor); Curd, Shelaine (Editor); Ehleriinger, Jim; Brooks, J. Renee; Flanagan, Larry</p> <p>2000-01-01</p> <p>The BOREAS TE-5 team collected measurements in the NSA and SSA on <span class="hlt">gas</span> <span class="hlt">exchange</span>, <span class="hlt">gas</span> composition, and tree growth. The leaf photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> data were collected in the BOREAS NSA and the SSA from 06-Jun- 1994 to 13-Sep- 1994 using a LI-COR 6200 portable photosynthesis system. The data were collected to compare the photosynthetic capacity, stomata] conductance, and leaf intercellular CO, concentrations among the major tree species at the BOREAS sites. The data are average values from diurnal measurements on the upper canopy foliage (sun leaves). The data are available in tabular ASCII files. The data files are available on a CD-ROM (see document number 20010000884), or from the Oak Ridge National Laboratory (ORNL) Distributed Activity Archive Center (DAAC).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1714022B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1714022B"><span>Atmospheric <span class="hlt">exchange</span> of carbon dioxide and methane of a small <span class="hlt">water</span> body and a floating mat in the Luther Marsh peatland, Ontario, Canada</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burger, Magdalena; Berger, Sina; Blodau, Christian</p> <p>2015-04-01</p> <p>Recent investigations have suggested that small <span class="hlt">water</span> bodies cover larger areas in northern peatlands than previously assumed. Their role in the carbon cycle and <span class="hlt">gas</span> <span class="hlt">exchange</span> rates are poorly constrained so far. To address this issue we measured CO2 and CH4 fluxes on a small <span class="hlt">water</span> body (ca. 700 m2) and the surrounding floating mat in the Luther Marsh peatland in Ontario, Canada from July to September 2014. To this end we used closed chambers combined with a portable Los Gatos high-resolution trace <span class="hlt">gas</span> analyzer at different <span class="hlt">water</span> depths and distances from the shore on the pond and with different dominating plant types on the floating mat surrounding the pond. In addition, CO2 concentrations were recorded in high temporal resolution using an infrared sensor system during selected periods. <span class="hlt">Air</span> and <span class="hlt">water</span> temperature, humidity and temperature of the floating mat, wind speed and direction, photosynthetically active radiation, <span class="hlt">air</span> pressure and relative humidity were also recorded as auxiliary data at the study site. The results show that pond and floating mat were sources of methane throughout the whole measuring period. Methane emissions via the ebullition pathway occurred predominantly near the shore and on the floating mat. During the daytime measurements the floating mat acted as a net sink and the pond as a net source of CO2. The dynamics of CO2 <span class="hlt">exchange</span> was also strongly time dependent, as CO2 emissions from the pond strongly increased after mid-August. This suggests that photosynthesis was more affected by seasonal decline than respiration process in the pond and that the allochthonous component of the CO2 flux increased in relative importance towards fall.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19077169','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19077169"><span>Surviving floods: leaf <span class="hlt">gas</span> films improve O₂ and CO₂ <span class="hlt">exchange</span>, root aeration, and growth of completely submerged rice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pedersen, Ole; Rich, Sarah Meghan; Colmer, Timothy David</p> <p>2009-04-01</p> <p>When completely submerged, the leaves of some species retain a surface <span class="hlt">gas</span> film. Leaf <span class="hlt">gas</span> films on submerged plants have recently been termed 'plant plastrons', analogous with the plastrons of aquatic insects. In aquatic insects, surface <span class="hlt">gas</span> layers (i.e. plastrons) enlarge the <span class="hlt">gas-water</span> interface to promote O₂ uptake when under <span class="hlt">water</span>; however, the function of leaf <span class="hlt">gas</span> films has rarely been considered. The present study demonstrates that <span class="hlt">gas</span> films on leaves of completely submerged rice facilitate entry of O₂ from floodwaters when in darkness and CO₂ entry when in light. O₂ microprofiles showed that the improved <span class="hlt">gas</span> <span class="hlt">exchange</span> was not caused by differences in diffusive boundary layers adjacent to submerged leaves with or without <span class="hlt">gas</span> films; instead, reduced resistance to <span class="hlt">gas</span> <span class="hlt">exchange</span> was probably due to the enlarged <span class="hlt">water-gas</span> interface (cf. aquatic insects). When <span class="hlt">gas</span> films were removed artificially, underwater net photosynthesis declined to only 20% of the rate with <span class="hlt">gas</span> films present, such that, after 7 days of complete submergence, tissue sugar levels declined, and both shoot and root growth were reduced. Internal aeration of roots in anoxic medium, when shoots were in aerobic floodwater in darkness or when in light, was improved considerably when leaf <span class="hlt">gas</span> films were present. Thus, leaf <span class="hlt">gas</span> films contribute to the submergence tolerance of rice, in addition to those traits already recognized, such as the shoot-elongation response, aerenchyma and metabolic adjustments to O₂ deficiency and oxidative stress. © 2009 The Authors. Journal compilation © 2009 Blackwell Publishing Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23480170','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23480170"><span>Degradation and rearrangement of a lung surfactant lipid at the <span class="hlt">air-water</span> interface during exposure to the pollutant <span class="hlt">gas</span> ozone.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thompson, Katherine C; Jones, Stephanie H; Rennie, Adrian R; King, Martin D; Ward, Andrew D; Hughes, Brian R; Lucas, Claire O M; Campbell, Richard A; Hughes, Arwel V</p> <p>2013-04-09</p> <p>The presence of unsaturated lipids in lung surfactant is important for proper respiratory function. In this work, we have used neutron reflection and surface pressure measurements to study the reaction of the ubiquitous pollutant <span class="hlt">gas</span>-phase ozone, O3, with pure and mixed phospholipid monolayers at the <span class="hlt">air-water</span> interface. The results reveal that the reaction of the unsaturated lipid 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine, POPC, with ozone leads to the rapid loss of the terminal C9 portion of the oleoyl strand of POPC from the <span class="hlt">air-water</span> interface. The loss of the C9 portion from the interface is accompanied by an increase in the surface pressure (decrease in surface tension) of the film at the <span class="hlt">air-water</span> interface. The results suggest that the portion of the oxidized oleoyl strand that is still attached to the lipid headgroup rapidly reverses its orientation and penetrates the <span class="hlt">air-water</span> interface alongside the original headgroup, thus increasing the surface pressure. The reaction of POPC with ozone also leads to a loss of material from the palmitoyl strand, but the loss of palmitoyl material occurs after the loss of the terminal C9 portion from the oleoyl strand of the molecule, suggesting that the palmitoyl material is lost in a secondary reaction step. Further experiments studying the reaction of mixed monolayers composed of unsaturated lipid POPC and saturated lipid dipalmitoyl-sn-glycero-3-phosphocholine, DPPC, revealed that no loss of DPPC from the <span class="hlt">air-water</span> interface occurs, eliminating the possibility that a reactive species such as an OH radical is formed and is able to attack nearby lipid chains. The reaction of ozone with the mixed films does cause a significant change in the surface pressure of the <span class="hlt">air-water</span> interface. Thus, the reaction of unsaturated lipids in lung surfactant changes and impairs the physical properties of the film at the <span class="hlt">air-water</span> interface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26614785','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26614785"><span>Leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> performance and the lethal <span class="hlt">water</span> potential of five European species during drought.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Shan; Feifel, Marion; Karimi, Zohreh; Schuldt, Bernhard; Choat, Brendan; Jansen, Steven</p> <p>2016-02-01</p> <p>Establishing physiological thresholds to drought-induced mortality in a range of plant species is crucial in understanding how plants respond to severe drought. Here, five common European tree species were selected (Acer campestre L., Acer pseudoplatanus L., Carpinus betulus L., Corylus avellana L. and Fraxinus excelsior L.) to study their hydraulic thresholds to mortality. Photosynthetic parameters during desiccation and the recovery of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> after rewatering were measured. Stem vulnerability curves and leaf pressure-volume curves were investigated to understand the hydraulic coordination of stem and leaf tissue traits. Stem and root samples from well-<span class="hlt">watered</span> and severely drought-stressed plants of two species were observed using transmission electron microscopy to visualize mortality of cambial cells. The lethal <span class="hlt">water</span> potential (ψlethal) correlated with stem P99 (i.e., the xylem <span class="hlt">water</span> potential at 99% loss of hydraulic conductivity, PLC). However, several plants that were stressed beyond the <span class="hlt">water</span> potential at 100% PLC showed complete recovery during the next spring, which suggests that the ψlethal values were underestimated. Moreover, we observed a 1 : 1 relationship between the xylem <span class="hlt">water</span> potential at the onset of embolism and stomatal closure, confirming hydraulic coordination between leaf and stem tissues. Finally, ultrastructural changes in the cytoplasm of cambium tissue and mortality of cambial cells are proposed to provide an alternative approach to investigate the point of no return associated with plant death. © The Author 2015. Published by Oxford University Press. All rights reserved. For Permissions, please email: journals.permissions@oup.com.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29375728','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29375728"><span>Steel reinforced composite silicone membranes and its integration to microfluidic oxygenators for high performance <span class="hlt">gas</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matharoo, Harpreet; Dabaghi, Mohammadhossein; Rochow, Niels; Fusch, Gerhard; Saraei, Neda; Tauhiduzzaman, Mohammed; Veldhuis, Stephen; Brash, John; Fusch, Christoph; Selvaganapathy, P Ravi</p> <p>2018-01-01</p> <p>Respiratory distress syndrome (RDS) is one of the main causes of fatality in newborn infants, particularly in neonates with low birth-weight. Commercial extracorporeal oxygenators have been used for low-birth-weight neonates in neonatal intensive care units. However, these oxygenators require high blood volumes to prime. In the last decade, microfluidics oxygenators using enriched oxygen have been developed for this purpose. Some of these oxygenators use thin polydimethylsiloxane (PDMS) membranes to facilitate <span class="hlt">gas</span> <span class="hlt">exchange</span> between the blood flowing in the microchannels and the ambient <span class="hlt">air</span> outside. However, PDMS is elastic and the thin membranes exhibit significant deformation and delamination under pressure which alters the architecture of the devices causing poor oxygenation or device failure. Therefore, an alternate membrane with high stability, low deformation under pressure, and high <span class="hlt">gas</span> <span class="hlt">exchange</span> was desired. In this paper, we present a novel composite membrane consisting of an ultra-thin stainless-steel mesh embedded in PDMS, designed specifically for a microfluidic single oxygenator unit (SOU). In comparison to homogeneous PDMS membranes, this composite membrane demonstrated high stability, low deformation under pressure, and high <span class="hlt">gas</span> <span class="hlt">exchange</span>. In addition, a new design for oxygenator with sloping profile and tapered inlet configuration has been introduced to achieve the same <span class="hlt">gas</span> <span class="hlt">exchange</span> at lower pressure drops. SOUs were tested by bovine blood to evaluate <span class="hlt">gas</span> <span class="hlt">exchange</span> properties. Among all tested SOUs, the flat design SOU with composite membrane has the highest oxygen <span class="hlt">exchange</span> of 40.32 ml/min m 2 . The superior performance of the new device with composite membrane was demonstrated by constructing a lung assist device (LAD) with a low priming volume of 10 ml. The LAD was achieved by the oxygen uptake of 0.48-0.90 ml/min and the CO 2 release of 1.05-2.27 ml/min at blood flow rates ranging between 8 and 48 ml/min. This LAD was shown to increase the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title40-vol34/pdf/CFR-2012-title40-vol34-sec1065-645.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title40-vol34/pdf/CFR-2012-title40-vol34-sec1065-645.pdf"><span>40 CFR 1065.645 - Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>.</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>... 40 Protection of Environment 34 2012-07-01 2012-07-01 false Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>. 1065.645 Section 1065.645 Protection of Environment ENVIRONMENTAL PROTECTION AGENCY (CONTINUED) <span class="hlt">AIR</span> POLLUTION CONTROLS ENGINE-TESTING PROCEDURES Calculations and Data Requirements § 1065.645 Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>. This section describes how to...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title40-vol34/pdf/CFR-2013-title40-vol34-sec1065-645.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title40-vol34/pdf/CFR-2013-title40-vol34-sec1065-645.pdf"><span>40 CFR 1065.645 - Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>.</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>... 40 Protection of Environment 34 2013-07-01 2013-07-01 false Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>. 1065.645 Section 1065.645 Protection of Environment ENVIRONMENTAL PROTECTION AGENCY (CONTINUED) <span class="hlt">AIR</span> POLLUTION CONTROLS ENGINE-TESTING PROCEDURES Calculations and Data Requirements § 1065.645 Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>. This section describes how to...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol33/pdf/CFR-2011-title40-vol33-sec1065-645.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol33/pdf/CFR-2011-title40-vol33-sec1065-645.pdf"><span>40 CFR 1065.645 - Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>.</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>... 40 Protection of Environment 33 2011-07-01 2011-07-01 false Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>. 1065.645 Section 1065.645 Protection of Environment ENVIRONMENTAL PROTECTION AGENCY (CONTINUED) <span class="hlt">AIR</span> POLLUTION CONTROLS ENGINE-TESTING PROCEDURES Calculations and Data Requirements § 1065.645 Amount of <span class="hlt">water</span> in an ideal <span class="hlt">gas</span>. This section describes how to...</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/23791347','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23791347"><span>Atmospheric partitioning and the <span class="hlt">air-water</span> <span class="hlt">exchange</span> of polycyclic aromatic hydrocarbons in a large shallow Chinese lake (Lake Chaohu).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Qin, Ning; He, Wei; Kong, Xiang-Zhen; Liu, Wen-Xiu; He, Qi-Shuang; Yang, Bin; Ouyang, Hui-Ling; Wang, Qing-Mei; Xu, Fu-Liu</p> <p>2013-11-01</p> <p>The residual levels of polycyclic aromatic hydrocarbons (PAHs) in the atmosphere and in dissolved phase from Lake Chaohu were measured by (GC-MS). The composition and seasonal variation were investigated. The diffusive <span class="hlt">air-water</span> <span class="hlt">exchange</span> flux was estimated by a two-film model, and the uncertainty in the flux calculations and the sensitivity of the parameters were evaluated. The following results were obtained: (1) the average residual levels of all PAHs (PAH16) in the atmosphere from Lake Chaohu were 60.85±46.17 ng m(-3) in the gaseous phase and 14.32±23.82 ng m(-3) in the particulate phase. The dissolved PAH16 level was 173.46±132.89 ng L(-1). (2) The seasonal variation of average PAH16 contents ranged from 43.09±33.20 ng m(-3) (summer) to 137.47±41.69 ng m(-3) (winter) in gaseous phase, from 6.62±2.72 ng m(-3) (summer) to 56.13±22.99 ng m(-3) (winter) in particulate phase, and 142.68±74.68 ng L(-1) (winter) to 360.00±176.60 ng L(-1) (summer) in <span class="hlt">water</span> samples. Obvious seasonal trends of PAH16 concentrations were found in the atmosphere and <span class="hlt">water</span>. The values of PAH16 for both the atmosphere and the <span class="hlt">water</span> were significantly correlated with temperature. (3) The monthly diffusive <span class="hlt">air-water</span> <span class="hlt">exchange</span> flux of total PAH16 ranged from -1.77×10(4) ng m(-2) d(-1) to 1.11×10(5) ng m(-2) d(-1), with an average value of 3.45×10(4) ng m(-2) d(-1). (4) The results of a Monte Carlo simulation showed that the monthly average PAH fluxes ranged from -3.4×10(3) ng m(-2) d(-1) to 1.6×10(4) ng m(-2) d(-1) throughout the year, and the uncertainties for individual PAHs were compared. (5) According to the sensitivity analysis, the concentrations of dissolved and gaseous phase PAHs were the two most important factors affecting the results of the flux calculations. Copyright © 2013 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29593081','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29593081"><span>Cuticular <span class="hlt">gas</span> <span class="hlt">exchange</span> by Antarctic sea spiders.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lane, Steven J; Moran, Amy L; Shishido, Caitlin M; Tobalske, Bret W; Woods, H Arthur</p> <p>2018-04-25</p> <p>Many marine organisms and life stages lack specialized respiratory structures, like gills, and rely instead on cutaneous respiration, which they facilitate by having thin integuments. This respiratory mode may limit body size, especially if the integument also functions in support or locomotion. Pycnogonids, or sea spiders, are marine arthropods that lack gills and rely on cutaneous respiration but still grow to large sizes. Their cuticle contains pores, which may play a role in <span class="hlt">gas</span> <span class="hlt">exchange</span>. Here, we examined alternative paths of <span class="hlt">gas</span> <span class="hlt">exchange</span> in sea spiders: (1) oxygen diffuses across pores in the cuticle, a common mechanism in terrestrial eggshells, (2) oxygen diffuses directly across the cuticle, a common mechanism in small aquatic insects, or (3) oxygen diffuses across both pores and cuticle. We examined these possibilities by modeling diffusive oxygen fluxes across all pores in the body of sea spiders and asking whether those fluxes differed from measured metabolic rates. We estimated fluxes across pores using Fick's law parameterized with measurements of pore morphology and oxygen gradients. Modeled oxygen fluxes through pores closely matched oxygen consumption across a range of body sizes, which means the pores facilitate oxygen diffusion. Furthermore, pore volume scaled hypermetrically with body size, which helps larger species facilitate greater diffusive oxygen fluxes across their cuticle. This likely presents a functional trade-off between <span class="hlt">gas</span> <span class="hlt">exchange</span> and structural support, in which the cuticle must be thick enough to prevent buckling due to external forces but porous enough to allow sufficient <span class="hlt">gas</span> <span class="hlt">exchange</span>. © 2018. Published by The Company of Biologists Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21619278','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21619278"><span>An analytical method for trifluoroacetic Acid in <span class="hlt">water</span> and <span class="hlt">air</span> samples using headspace <span class="hlt">gas</span> chromatographic determination of the methyl ester.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zehavi, D; Seiber, J N</p> <p>1996-10-01</p> <p>An analytical method has been developed for the determination of trace levels of trifluoroacetic acid (TFA), an atmospheric breakdown product of several of the hydrofluorocarbon (HFC) and hydrochlorofluorocarbon (HCFC) replacements for the chlorofluorocarbon (CFC) refrigerants, in <span class="hlt">water</span> and <span class="hlt">air</span>. TFA is derivatized to the volatile methyl trifluoroacetate (MTFA) and determined by automated headspace <span class="hlt">gas</span> chromatography (HSGC) with electron-capture detection or manual HSGC using GC/MS in the selected ion monitoring (SIM) mode. The method is based on the reaction of an aqueous sample containing TFA with dimethyl sulfate (DMS) in concentrated sulfuric acid in a sealed headspace vial under conditions favoring distribution of MTFA to the vapor phase. <span class="hlt">Water</span> samples are prepared by evaporative concentration, during which TFA is retained as the anion, followed by extraction with diethyl ether of the acidified sample and then back-extraction of TFA (as the anion) in aqueous bicarbonate solution. The extraction step is required for samples with a relatively high background of other salts and organic materials. <span class="hlt">Air</span> samples are collected in sodium bicarbonate-glycerin-coated glass denuder tubes and prepared by rinsing the denuder contents with <span class="hlt">water</span> to form an aqueous sample for derivatization and analysis. Recoveries of TFA from spiked <span class="hlt">water</span>, with and without evaporative concentration, and from spiked <span class="hlt">air</span> were quantitative, with estimated detection limits of 10 ng/mL (unconcentrated) and 25 pg/mL (concentrated 250 mL:1 mL) for <span class="hlt">water</span> and 1 ng/m(3) (72 h at 5 L/min) for <span class="hlt">air</span>. Several environmental <span class="hlt">air</span>, fogwater, rainwater, and surface <span class="hlt">water</span> samples were successfully analyzed; many showed the presence of TFA.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24097404','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24097404"><span>A membrane inlet mass spectrometry system for noble gases at natural abundances in <span class="hlt">gas</span> and <span class="hlt">water</span> samples.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Visser, Ate; Singleton, Michael J; Hillegonds, Darren J; Velsko, Carol A; Moran, Jean E; Esser, Bradley K</p> <p>2013-11-15</p> <p>Noble gases dissolved in groundwater can reveal paleotemperatures, recharge conditions, and precise travel times. The collection and analysis of noble <span class="hlt">gas</span> samples are cumbersome, involving noble <span class="hlt">gas</span> purification, cryogenic separation and static mass spectrometry. A quicker and more efficient sample analysis method is required for introduced tracer studies and laboratory experiments. A Noble <span class="hlt">Gas</span> Membrane Inlet Mass Spectrometry (NG-MIMS) system was developed to measure noble gases at natural abundances in <span class="hlt">gas</span> and <span class="hlt">water</span> samples. The NG-MIMS system consists of a membrane inlet, a dry-ice <span class="hlt">water</span> trap, a carbon-dioxide trap, two getters, a gate valve, a turbomolecular pump and a quadrupole mass spectrometer equipped with an electron multiplier. Noble gases isotopes (4)He, (22)Ne, (38)Ar, (84)Kr and (132)Xe are measured every 10 s. The NG-MIMS system can reproduce measurements made on a traditional noble <span class="hlt">gas</span> mass spectrometer system with precisions of 2%, 8%, 1%, 1% and 3% for He, Ne, Ar, Kr and Xe, respectively. Noble <span class="hlt">gas</span> concentrations measured in an artificial recharge pond were used to monitor an introduced xenon tracer and to reconstruct temperature variations to within 2 °C. Additional experiments demonstrated the capability to measure noble gases in <span class="hlt">gas</span> and in <span class="hlt">water</span> samples, in real time. The NG-MIMS system is capable of providing analyses sufficiently accurate and precise for introduced noble <span class="hlt">gas</span> tracers at managed aquifer recharge facilities, groundwater fingerprinting based on excess <span class="hlt">air</span> and noble <span class="hlt">gas</span> recharge temperature, and field and laboratory studies investigating ebullition and diffusive <span class="hlt">exchange</span>. Copyright © 2013 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1024521','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1024521"><span>Kitchen Appliance Upgrades Improve <span class="hlt">Water</span> Efficiency at DOD <span class="hlt">Exchange</span> Facilities: Best Management Practice Case Study #11: Commercial Kitchen Equipment (Brochure)</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>Not Available</p> <p>2011-09-01</p> <p>The <span class="hlt">Exchange</span>, formerly the Army and <span class="hlt">Air</span> Force <span class="hlt">Exchange</span> Service (AAFES), is a joint military activity and the U.S. Department of Defense?s (DOD) oldest and largest retailer. The <span class="hlt">Exchange</span> is taking a leadership role in <span class="hlt">water</span> efficiency improvements in their commercial kitchens by integrating <span class="hlt">water</span> efficiency concepts into the organization?s overall sustainability plan and objectives.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GBioC..31..961W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GBioC..31..961W"><span>The impact of changing wind speeds on <span class="hlt">gas</span> transfer and its effect on global <span class="hlt">air</span>-sea CO2 fluxes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wanninkhof, R.; Triñanes, J.</p> <p>2017-06-01</p> <p>An increase in global wind speeds over time is affecting the global uptake of CO2 by the ocean. We determine the impact of changing winds on <span class="hlt">gas</span> transfer and CO2 uptake by using the recently updated, global high-resolution, cross-calibrated multiplatform wind product (CCMP-V2) and a fixed monthly pCO2 climatology. In particular, we assess global changes in the context of regional wind speed changes that are attributed to large-scale climate reorganizations. The impact of wind on global CO2 <span class="hlt">gas</span> fluxes as determined by the bulk formula is dependent on several factors, including the functionality of the <span class="hlt">gas</span> <span class="hlt">exchange</span>-wind speed relationship and the regional and seasonal differences in the <span class="hlt">air-water</span> partial pressure of CO2 gradient (ΔpCO2). The latter also controls the direction of the flux. Fluxes out of the ocean are influenced more by changes in the low-to-intermediate wind speed range, while ingassing is impacted more by changes in higher winds because of the regional correlations between wind and ΔpCO2. <span class="hlt">Gas</span> <span class="hlt">exchange</span>-wind speed parameterizations with a quadratic and third-order polynomial dependency on wind, each of which meets global constraints, are compared. The changes in <span class="hlt">air</span>-sea CO2 fluxes resulting from wind speed trends are greatest in the equatorial Pacific and cause a 0.03-0.04 Pg C decade-1 increase in outgassing over the 27 year time span. This leads to a small overall decrease of 0.00 to 0.02 Pg C decade-1 in global net CO2 uptake, contrary to expectations that increasing winds increase net CO2 uptake.<abstract type="synopsis"><title type="main">Plain Language SummaryThe effects of changing winds are isolated from the total change in trends in global <span class="hlt">air</span>-sea CO2 fluxes over the last 27 years. The overall effect of increasing winds over time has a smaller impact than expected as the impact in regions of outgassing is greater than for the regions acting as a CO2 sink.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5818131-how-gas-cools-apples-can-fall-up','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5818131-how-gas-cools-apples-can-fall-up"><span>How <span class="hlt">gas</span> cools (or, apples can fall up)</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>Not Available</p> <p>1987-01-01</p> <p>This primer on <span class="hlt">gas</span> cooling systems explains the basics of heat <span class="hlt">exchange</span> within a refrigeration system, the principle of reverse-cycle refrigeration, and how a <span class="hlt">gas</span>-engine-driven heat pump can provide cooling, additional winter heating capacity, and hot <span class="hlt">water</span> year-round. <span class="hlt">Gas</span> cooling equipment available or under development include natural <span class="hlt">gas</span> chillers, engine-driven chillers, and absorption chillers. In cogeneration systems, heat recovered from an engine's exhaust and coolant may be used in an absorption chiller to provide <span class="hlt">air</span>-conditioning. <span class="hlt">Gas</span> desiccant cooling systems may be used in buildings and businesses that are sensitive to high humidity levels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..1611486M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..1611486M"><span>Wind variability and sheltering effects on measurements and modeling of <span class="hlt">air-water</span> <span class="hlt">exchange</span> for a small lake</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Markfort, Corey D.; Resseger, Emily; Porté-Agel, Fernando; Stefan, Heinz</p> <p>2014-05-01</p> <p>Lakes with a surface area of less than 10 km2 account for over 50% of the global cumulative lake surface <span class="hlt">water</span> area, and make up more than 99% of the total number of global lakes, ponds, and wetlands. Within the boreal regions as well as some temperate and tropical areas, a significant proportion of land cover is characterized by lakes or wetlands, which can have a dramatic effect on land-atmosphere fluxes as well as the local and regional energy budget. Many of these small <span class="hlt">water</span> bodies are surrounded by complex terrain and forest, which cause the wind blowing over a small lake or wetland to be highly variable. Wind mixing of the lake surface layer affects thermal stratification, surface temperature and <span class="hlt">air-water</span> <span class="hlt">gas</span> transfer, e.g. O2, CO2, and CH4. As the wind blows from the land to the lake, wake turbulence behind trees and other shoreline obstacles leads to a recirculation zone and enhanced turbulence. This wake flow results in the delay of the development of wind shear stress on the lake surface, and the fetch required for surface shear stress to fully develop may be ~O(1 km). Interpretation of wind measurements made on the lake is hampered by the unknown effect of wake turbulence. We present field measurements designed to quantify wind variability over a sheltered lake. The wind data and <span class="hlt">water</span> column temperature profiles are used to evaluate a new method to quantify wind sheltering of lakes that takes into account lake size, shape and the surrounding landscape features. The model is validated against field data for 36 Minnesota lakes. Effects of non-uniform sheltering and lake shape are also demonstrated. The effects of wind sheltering must be included in lake models to determine the effect of wind-derived energy inputs on lake stratification, surface <span class="hlt">gas</span> transfer, lake <span class="hlt">water</span> quality, and fish habitat. These effects are also important for correctly modeling momentum, heat, moisture and trace <span class="hlt">gas</span> flux to the atmosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6242231-water-splitting-titanium-exchanged-zeolite-technical-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6242231-water-splitting-titanium-exchanged-zeolite-technical-report"><span>'<span class="hlt">water</span> splitting' by titanium <span class="hlt">exchanged</span> zeolite A. Technical report</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>Kuznicki, S.M.; Eyring, E.M.</p> <p>1978-09-01</p> <p>Visually detectable and chromatographically and mass spectrally identified hydrogen <span class="hlt">gas</span> evolves from titanium (III) <span class="hlt">exchanged</span> zeolite A immersed in <span class="hlt">water</span> and illuminated with visible light. Titanium(III) <span class="hlt">exchanged</span> zeolite X and zeolite Y do not produce this reaction. A photochemically produced, oxygenated titanium free radical (detected by electron spin resonance) not previously described is the species in the zeolite that reduces protons to molecular hydrogen. The other product of this reduction step is a nonradical, oxygenated titanium species of probable empirical formula TiO4. Heating the spent oxygenated titanium containing zeolite A under vacuum at 375 C restores over fifty percent ofmore » the free radical. Unlike previously reported systems, heating does not restore the original aquotitanium(III) species in the zeolite. Thus a means other than heating must be found to achieve a closed photochemical cycle that harnesses visible solar energy in the production of molecular hydrogen. The titanium <span class="hlt">exchanged</span> zeolite A does, however, lend itself to a thermolysis of <span class="hlt">water</span> previously described by Kasai and Bishop. (Author)« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/7203850-enhance-gas-processing-reflux-heat-exchangers','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/7203850-enhance-gas-processing-reflux-heat-exchangers"><span>Enhance <span class="hlt">gas</span> processing with reflux heat-<span class="hlt">exchangers</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>Finn, A.J.</p> <p>1994-05-01</p> <p>Despite recent successes of membrane-based separations in low-throughput applications, cryogenic processing remains the best route for separating and purifying <span class="hlt">gas</span> mixtures, especially when high recoveries are required. Now conventional units are being modified to yield even higher recoveries at lower costs. Throughout the chemical process industries (CPI), this is being accomplished with reflux or plate-fin <span class="hlt">exchangers</span>, especially for processing of natural <span class="hlt">gas</span>, and offgases from refineries and petrochemical facilities. The concept of utilizing a heat <span class="hlt">exchanger</span> as a multi stage rectification device is not new. However, only in the last fifteen years or so has accurate design of reflux exchangersmore » become feasible. Also helpful have been the availability of prediction techniques for high-quality thermodynamic data, and process simulators that can rapidly solve the complex material, equilibrium and enthalpy relationships involved in simulating the performance of reflux <span class="hlt">exchangers</span>. Four projects that show the value and effectiveness of reflux <span class="hlt">exchangers</span> are discussed below in more detail. The first example considers hydrogen recovery from demethanizer overheads; the second highlights a low energy process for NGL and LPG recovery from natural <span class="hlt">gas</span>. The third is a simple process for recovery of ethylene from fluid-catalytic cracker (FCC) offgas; and the fourth is a similar process for olefin recovery from dehydrogenation-reactor offgas.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=65336&keyword=Austria&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=65336&keyword=Austria&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>CO2 AND O3 ALTER PHOTOSYNTHESIS AND <span class="hlt">WATER</span> VAPOR <span class="hlt">EXCHANGE</span> FOR PINUS PONDEROSA NEEDLES</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>1. Effects of CO2 and O3 were determined for a key component of ecosystem carbon and <span class="hlt">water</span> cycling: needle <span class="hlt">gas</span> <span class="hlt">exchange</span> (photosynthesis, conductance, transpiration and <span class="hlt">water</span> use efficiency). The measurements were made on Pinus ponderosa seedlings grown in outdoor, sunlit, mesoc...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000119049&hterms=water+purification&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dwater%2Bpurification','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000119049&hterms=water+purification&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dwater%2Bpurification"><span>Regenerable <span class="hlt">Air</span> Purification System for <span class="hlt">Gas</span>-Phase Contaminant Control</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Constantinescu, Ileana C.; Finn, John E.; LeVan, M. Douglas; Lung, Bernadette (Technical Monitor)</p> <p>2000-01-01</p> <p>Tests of a pre-prototype regenerable <span class="hlt">air</span> purification system (RAPS) that uses <span class="hlt">water</span> vapor to displace adsorbed contaminants from an adsorbent column have been performed at NASA Ames Research Center. A unit based on this design can be used for removing trace <span class="hlt">gas</span>-phase contaminants from spacecraft cabin <span class="hlt">air</span> or from polluted process streams including incinerator exhaust. During the normal operation mode, contaminants are removed from the <span class="hlt">air</span> on the column. Regeneration of the column is performed on-line. During regeneration, contaminants are displaced and destroyed inside the closed oxidation loop. In this presentation we discuss initial experimental results for the performance of RAPS in the removal and treatment of several important spacecraft contaminant species from <span class="hlt">air</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16731054','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16731054"><span>Ammonia as a respiratory <span class="hlt">gas</span> in <span class="hlt">water</span> and <span class="hlt">air</span>-breathing fishes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Randall, David J; Ip, Yuen K</p> <p>2006-11-01</p> <p>Ammonia is produced in the liver and excreted as NH(3) by diffusion across the gills. Elevated ammonia results in an increase in gill ventilation, perhaps via stimulation of gill oxygen chemo-receptors. Acidification of the <span class="hlt">water</span> around the fish by carbon dioxide and acid excretion enhances ammonia excretion and constitutes "environmental ammonia detoxification". Fish have difficulties in excreting ammonia in alkaline <span class="hlt">water</span> or high concentrations of environmental ammonia, or when out of <span class="hlt">water</span>. The mudskipper, Periphthalmodon schlosseri, is capable of active NH(4)(+) transport, maintaining low internal levels of ammonia. To prevent a back flux of NH(3), these <span class="hlt">air</span>-breathing fish can increase gill acid excretion and reduce the membrane NH(3) permeability by modifying the phospholipid and cholesterol compositions of their skin. Several <span class="hlt">air</span>-breathing fish species can excrete ammonia into <span class="hlt">air</span> through NH(3) volatilization. Some fish detoxify ammonia to glutamine or urea. The brains of some fish can tolerate much higher levels of ammonia than other animals. Studies of these fish may offer insights into the nature of ammonia toxicity in general.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20018864-economic-analysis-condensers-water-recovery-steam-injected-gas-turbines','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20018864-economic-analysis-condensers-water-recovery-steam-injected-gas-turbines"><span>Economic analysis of condensers for <span class="hlt">water</span> recovery in steam injected <span class="hlt">gas</span> turbines</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>De Paepe, M.; Huvenne, P.; Dick, E.</p> <p>1998-07-01</p> <p>Steam injection cycles are interesting for small power ranges because of the high efficiency and the relatively low investment costs. A big disadvantage is the consumption of <span class="hlt">water</span> by the cycle. <span class="hlt">Water</span> recovery is seldom realized in industrial practice. In this paper an analysis of the technical and economical possibilities of <span class="hlt">water</span> recovery by condensation of <span class="hlt">water</span> out of the exhaust gases is made. Three <span class="hlt">gas</span> turbines are considered : the Kawasaki M1A-13CC (2.3 MWe), the Allison 501KH (6.8 MWe) and the General Electric LM1600 (17 MWe). For every <span class="hlt">gas</span> turbine two types of condensers are designed. In the watermore » cooled condenser finned tubes are used to cool the exhaust gases, flowing at the outside of the tubes. The <span class="hlt">water</span> itself flows at the inside of the tubes and is cooled by a <span class="hlt">water</span> to <span class="hlt">air</span> cooler. In the <span class="hlt">air</span> cooled condenser the exhaust gases flow at the inside of the tubes and the cooling <span class="hlt">air</span> at the outside. The investment costs of the condensers is compared to the costs of the total installation. The investment costs are relatively smaller if the produced power goes up. The <span class="hlt">water</span> cooled condenser with <span class="hlt">water</span> to <span class="hlt">air</span> cooler is cheaper than the <span class="hlt">air</span> cooled condenser. Using a condenser results in higher exploitation costs due to the fans and pumps. It is shown that the <span class="hlt">air</span> cooled condenser has lower exploitation costs than the <span class="hlt">water</span> cooled one. Pay back time of the total installation does not significantly vary compared to the installation without recovery. <span class="hlt">Water</span> prices are determined for which <span class="hlt">water</span> recovery is profitable. For the <span class="hlt">water</span> cooled condenser the turning point lies at 2.2 Euro/m; for the <span class="hlt">air</span> cooled condenser this is 0.6 Euro/m.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..1611838B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..1611838B"><span>The influence of <span class="hlt">water</span> vapor on atmospheric <span class="hlt">exchange</span> measurements with an ICOS* based Laser absorption analyzer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bunk, Rüdiger; Quan, Zhi; Wandel, Matthias; Yi, Zhigang; Bozem, Heiko; Kesselmeier, Jürgen</p> <p>2014-05-01</p> <p>Carbonyl sulfide and carbon monoxide are both atmospheric trace gases of high interest. Recent advances in the field of spectroscopy have enabled instruments that measure the concentration of the above and other trace gases very fast and with good precision. Increasing the effective path length by reflecting the light between two mirrors in a cavity, these instruments reach impressive sensitivities. Often it is possible to measure the concentration of more than one trace <span class="hlt">gas</span> at the same time. The OCS/CO2 Analyzer by LGR (Los Gatos Research, Inc.) measures the concentration of <span class="hlt">water</span> vapor [H2O], carbonyl sulfide [COS], carbon dioxide [CO2] and carbon monoxide [CO] simultaneously. For that the cavity is saturated with light, than the attenuation of light is measured as in standard absorption spectroscopy. The instrument proved to be very fast with good precision and to be able to detect even very low concentrations, especially for COS (as low as 30ppt in the case of COS). However, we observed a rather strong cross sensitivity to <span class="hlt">water</span> vapor. Altering the <span class="hlt">water</span> vapor content of the sampled <span class="hlt">air</span> with two different methods led to a change in the perceived concentration of COS, CO and CO2. This proved especially problematic for enclosure (cuvette) measurements, where the concentrations of one of the above species in an empty cuvette are compared to the concentration of another cuvette containing a plant whose <span class="hlt">exchange</span> of trace gases with the atmosphere is of interest. There, the plants transpiration leads to a large difference in <span class="hlt">water</span> vapor content between the cuvettes and that in turn produces artifacts in the concentration differences between the cuvettes for the other above mentioned trace gases. For CO, simultaneous measurement with a UV-Emission Analyzer (AL 5002, Aerolaser) and the COS/CO Analyzer showed good agreement of perceived concentrations as long as the sample <span class="hlt">gas</span> was dry and an increasing difference in perceived concentration when the sample <span class="hlt">gas</span> was</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28819793','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28819793"><span>The mechanisms underlying the production of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> cycles in insects.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matthews, Philip G D</p> <p>2018-03-01</p> <p>This review examines the control of <span class="hlt">gas</span> <span class="hlt">exchange</span> in insects, specifically examining what mechanisms could explain the emergence of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> cycles (DGCs). DGCs are <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns consisting of alternating breath-hold periods and bouts of <span class="hlt">gas</span> <span class="hlt">exchange</span>. While all insects are capable of displaying a continuous pattern of <span class="hlt">gas</span> <span class="hlt">exchange</span>, this episodic pattern is known to occur within only some groups of insects and then only sporadically or during certain phases of their life cycle. Investigations into DGCs have tended to emphasise the role of chemosensory thresholds in triggering spiracle opening as critical for producing these <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns. However, a chemosensory basis for episodic breathing also requires an as-of-yet unidentified hysteresis between internal respiratory stimuli, chemoreceptors, and the spiracles. What has been less appreciated is the role that the insect's central nervous system (CNS) might play in generating episodic patterns of ventilation. The active ventilation displayed by many insects during DGCs suggests that this pattern could be the product of directed control by the CNS rather than arising passively as a result of self-sustaining oscillations in internal oxygen and carbon dioxide levels. This paper attempts to summarise what is currently known about insect <span class="hlt">gas</span> <span class="hlt">exchange</span> regulation, examining the location and control of ventilatory pattern generators in the CNS, the influence of chemoreceptor feedback in the form of O 2 and CO 2 /pH fluctuations in the haemolymph, and the role of state-dependent changes in CNS activity on ventilatory control. This information is placed in the context of what is currently known regarding the production of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> patterns.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ACP....17.8177K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ACP....17.8177K"><span>The boundary condition for vertical velocity and its interdependence with surface <span class="hlt">gas</span> <span class="hlt">exchange</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kowalski, Andrew S.</p> <p>2017-07-01</p> <p>The law of conservation of linear momentum is applied to surface <span class="hlt">gas</span> <span class="hlt">exchanges</span>, employing scale analysis to diagnose the vertical velocity (w) in the boundary layer. Net upward momentum in the surface layer is forced by evaporation (E) and defines non-zero vertical motion, with a magnitude defined by the ratio of E to the <span class="hlt">air</span> density, as w = <mstyle displaystyle="false">E/ρ</mstyle>. This is true even right down at the surface where the boundary condition is w|0 = <mstyle displaystyle="false"><mfrac style="text">E/ρ|0</mfrac></mstyle> (where w|0 and ρ|0 represent the vertical velocity and density of <span class="hlt">air</span> at the surface). This Stefan flow velocity implies upward transport of a non-diffusive nature that is a general feature of the troposphere but is of particular importance at the surface, where it assists molecular diffusion with upward <span class="hlt">gas</span> migration (of H2O, for example) but opposes that of downward-diffusing species like CO2 during daytime. The definition of flux-gradient relationships (eddy diffusivities) requires rectification to exclude non-diffusive transport, which does not depend on scalar gradients. At the microscopic scale, the role of non-diffusive transport in the process of evaporation from inside a narrow tube - with vapour transport into an overlying, horizontal airstream - was described long ago in classical mechanics and is routinely accounted for by chemical engineers, but has been neglected by scientists studying stomatal conductance. Correctly accounting for non-diffusive transport through stomata, which can appreciably reduce net CO2 transport and marginally boost that of <span class="hlt">water</span> vapour, should improve characterisations of ecosystem and plant functioning.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5388163','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5388163"><span><span class="hlt">Exchange</span> Bias Optimization by Controlled Oxidation of Cobalt Nanoparticle Films Prepared by Sputter <span class="hlt">Gas</span> Aggregation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Antón, Ricardo López; González, Juan A.; Andrés, Juan P.; Normile, Peter S.; Canales-Vázquez, Jesús; Muñiz, Pablo; Riveiro, José M.; De Toro, José A.</p> <p>2017-01-01</p> <p>Porous films of cobalt nanoparticles have been obtained by sputter <span class="hlt">gas</span> aggregation and controllably oxidized by <span class="hlt">air</span> annealing at 100 °C for progressively longer times (up to more than 1400 h). The magnetic properties of the samples were monitored during the process, with a focus on the <span class="hlt">exchange</span> bias field. <span class="hlt">Air</span> annealing proves to be a convenient way to control the Co/CoO ratio in the samples, allowing the optimization of the <span class="hlt">exchange</span> bias field to a value above 6 kOe at 5 K. The occurrence of the maximum in the <span class="hlt">exchange</span> bias field is understood in terms of the density of CoO uncompensated spins and their degree of pinning, with the former reducing and the latter increasing upon the growth of a progressively thicker CoO shell. Vertical shifts exhibited in the magnetization loops are found to correlate qualitatively with the peak in the <span class="hlt">exchange</span> bias field, while an increase in vertical shift observed for longer oxidation times may be explained by a growing fraction of almost completely oxidized particles. The presence of a hummingbird-like form in magnetization loops can be understood in terms of a combination of hard (biased) and soft (unbiased) components; however, the precise origin of the soft phase is as yet unresolved. PMID:28336895</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28336895','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28336895"><span><span class="hlt">Exchange</span> Bias Optimization by Controlled Oxidation of Cobalt Nanoparticle Films Prepared by Sputter <span class="hlt">Gas</span> Aggregation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Antón, Ricardo López; González, Juan A; Andrés, Juan P; Normile, Peter S; Canales-Vázquez, Jesús; Muñiz, Pablo; Riveiro, José M; De Toro, José A</p> <p>2017-03-11</p> <p>Porous films of cobalt nanoparticles have been obtained by sputter <span class="hlt">gas</span> aggregation and controllably oxidized by <span class="hlt">air</span> annealing at 100 °C for progressively longer times (up to more than 1400 h). The magnetic properties of the samples were monitored during the process, with a focus on the <span class="hlt">exchange</span> bias field. <span class="hlt">Air</span> annealing proves to be a convenient way to control the Co/CoO ratio in the samples, allowing the optimization of the <span class="hlt">exchange</span> bias field to a value above 6 kOe at 5 K. The occurrence of the maximum in the <span class="hlt">exchange</span> bias field is understood in terms of the density of CoO uncompensated spins and their degree of pinning, with the former reducing and the latter increasing upon the growth of a progressively thicker CoO shell. Vertical shifts exhibited in the magnetization loops are found to correlate qualitatively with the peak in the <span class="hlt">exchange</span> bias field, while an increase in vertical shift observed for longer oxidation times may be explained by a growing fraction of almost completely oxidized particles. The presence of a hummingbird-like form in magnetization loops can be understood in terms of a combination of hard (biased) and soft (unbiased) components; however, the precise origin of the soft phase is as yet unresolved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6039646-process-gas-hear-recovery','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6039646-process-gas-hear-recovery"><span>Process <span class="hlt">gas</span> hear recovery</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>Anderson, W.M.; Thurner, R.P.</p> <p>1977-01-01</p> <p>In considering the use of regenerative and recuperative heat <span class="hlt">exchangers</span> for process-<span class="hlt">gas</span> heat recovery general information regarding heat-<span class="hlt">exchanger</span> effectiveness versus initial capital investment and operating costs is discussed. Specific examples for preheating combustion <span class="hlt">air</span> for process furnaces and for using primary and secondary heat <span class="hlt">exchangers</span> in conjunction with an <span class="hlt">air</span>-pollution-control system for drying and curing ovens cover basic heat-<span class="hlt">exchanger</span> design and application considerations as well as investment-payback factors.</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('https://www.osti.gov/servlets/purl/1175556','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1175556"><span>Co-flow anode/cathode supply heat <span class="hlt">exchanger</span> for a solid-oxide fuel cell assembly</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Haltiner, Jr., Karl J.; Kelly, Sean M.</p> <p>2005-11-22</p> <p>In a solid-oxide fuel cell assembly, a co-flow heat <span class="hlt">exchanger</span> is provided in the flow paths of the reformate <span class="hlt">gas</span> and the cathode <span class="hlt">air</span> ahead of the fuel cell stack, the reformate <span class="hlt">gas</span> being on one side of the <span class="hlt">exchanger</span> and the cathode <span class="hlt">air</span> being on the other. The reformate <span class="hlt">gas</span> is at a substantially higher temperature than is desired in the stack, and the cathode <span class="hlt">gas</span> is substantially cooler than desired. In the co-flow heat <span class="hlt">exchanger</span>, the temperatures of the reformate and cathode streams converge to nearly the same temperature at the outlet of the <span class="hlt">exchanger</span>. Preferably, the heat <span class="hlt">exchanger</span> is formed within an integrated component manifold (ICM) for a solid-oxide fuel cell assembly.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009JGRG..114.0C08W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009JGRG..114.0C08W"><span>Diminished mercury emission from <span class="hlt">waters</span> with duckweed cover</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wollenberg, Jennifer L.; Peters, Stephen C.</p> <p>2009-06-01</p> <p>Duckweeds (Lemnaceae) are a widely distributed type of floating vegetation in freshwater systems. Under suitable conditions, duckweeds form a dense vegetative mat on the <span class="hlt">water</span> surface, which reduces light penetration into the <span class="hlt">water</span> column and limits <span class="hlt">gas</span> <span class="hlt">exchange</span> at the <span class="hlt">water-air</span> interface by decreasing the area of open <span class="hlt">water</span> surface. Experiments were conducted to determine whether duckweed decreases mercury emission by limiting <span class="hlt">gas</span> diffusion across the <span class="hlt">water-air</span> interface and attenuating light, or, conversely, enhances emission via transpiration of mercury vapor. Microcosm flux chamber experiments indicate that duckweed decreases mercury emission from the <span class="hlt">water</span> surface compared to open <span class="hlt">water</span> controls. Fluxes under duckweed were 17-67% lower than in controls, with lower fluxes occurring at higher percent cover. The decrease in mercury emission suggests that duckweed may limit emission through one of several mechanisms, including limited <span class="hlt">gas</span> transport across the <span class="hlt">air-water</span> interface, decreased photoreactions due to light attenuation, and plant-mercury interactions. The results of this experiment were applied to a model lake system to illustrate the magnitude of potential effects on mercury cycling. The mercury retained in the lake as a result of hindered emission may increase bioaccumulation potential in lakes with duckweed cover.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29351442','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29351442"><span>Measurements of pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> efficiency using expired <span class="hlt">gas</span> and oximetry: results in normal subjects.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>West, John B; Wang, Daniel L; Prisk, G Kim</p> <p>2018-04-01</p> <p>We are developing a novel, noninvasive method for measuring the efficiency of pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> in patients with lung disease. The patient wears an oximeter, and we measure the partial pressures of oxygen and carbon dioxide in inspired and expired <span class="hlt">gas</span> using miniature analyzers. The arterial Po 2 is then calculated from the oximeter reading and the oxygen dissociation curve, using the end-tidal Pco 2 to allow for the Bohr effect. This calculation is only accurate when the oxygen saturation is <94%, and therefore, these normal subjects breathed 12.5% oxygen. When the procedure is used in patients with hypoxemia, they breathe <span class="hlt">air</span>. The Po 2 difference between the end-tidal and arterial values is called the "oxygen deficit." Preliminary data show that this index increases substantially in patients with lung disease. Here we report measurements of the oxygen deficit in 20 young normal subjects (age 19 to 31 yr) and 11 older normal subjects (47 to 88 yr). The mean value of the oxygen deficit in the young subjects was 2.02 ± 3.56 mmHg (means ± SD). This mean is remarkably small. The corresponding value in the older group was 7.53 ± 5.16 mmHg (means ± SD). The results are consistent with the age-related trend of the traditional alveolar-arterial difference, which is calculated from the calculated ideal alveolar Po 2 minus the measured arterial Po 2 . That measurement requires an arterial blood sample. The present study suggests that this noninvasive procedure will be valuable in assessing the degree of impaired <span class="hlt">gas</span> <span class="hlt">exchange</span> in patients with lung disease.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17706251','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17706251"><span><span class="hlt">Air</span>--sea gaseous <span class="hlt">exchange</span> of PCB at the Venice lagoon (Italy).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Manodori, L; Gambaro, A; Moret, I; Capodaglio, G; Cescon, P</p> <p>2007-10-01</p> <p><span class="hlt">Water</span> bodies are important storage media for persistent organic pollutants (POPs) such as polychlorinated biphenyls (PCBs) and this function is increased in coastal regions because their inputs are higher than those to the open sea. The <span class="hlt">air-water</span> interface is extensively involved with the global cycling of PCBs because it is the place where they accumulate due to depositional processes and where they may be emitted by gaseous <span class="hlt">exchange</span>. In this work the parallel collection of <span class="hlt">air</span>, microlayer and sub-superficial <span class="hlt">water</span> samples was performed in July 2005 at a site in the Venice lagoon to evaluate the summer gaseous flux of PCBs. The total concentration of PCBs (sum of 118 congeners) in <span class="hlt">air</span> varies from 87 to 273 pg m(-3), whereas in the operationally defined dissolved phase of microlayer and sub-superficial <span class="hlt">water</span> samples it varies from 159 to 391 pg L(-1). No significant enrichment of dissolved PCB into the microlayer has been observed, although a preferential accumulation of most hydrophobic congeners occurs. Due to this behaviour, we believe that the modified two-layer model was the most suitable approach for the evaluation of the flux at the <span class="hlt">air</span>-sea interface, because it takes into account the influence of the microlayer. From its application it appears that PCB volatilize from the lagoon <span class="hlt">waters</span> with a net flux varying from 58 to 195 ng m(-2)d(-1) (uncertainty: +/-50-64%) due to the strong influence of wind speed. This flux is greater than those reported in the literature for the atmospheric deposition and rivers input and reveals that PCB are actively emitted from the Venice lagoon in summer months.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18849091','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18849091"><span>Juvenile Rhus glabra leaves have higher temperatures and lower <span class="hlt">gas</span> <span class="hlt">exchange</span> rates than mature leaves when compared in the field during periods of high irradiance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Snider, John L; Choinski, John S; Wise, Robert R</p> <p>2009-05-01</p> <p>We sought to test the hypothesis that stomatal development determines the timing of <span class="hlt">gas</span> <span class="hlt">exchange</span> competency, which then influences leaf temperature through transpirationally driven leaf cooling. To test this idea, daily patterns of <span class="hlt">gas</span> <span class="hlt">exchange</span> and leaflet temperature were obtained from leaves of two distinctively different developmental stages of smooth sumac (Rhus glabra) grown in its native habitat. Juvenile and mature leaves were also sampled for ultrastructural studies of stomatal development. When plants were sampled in May-June, the hypothesis was supported: juvenile leaflets were (for part of the day) from 1.4 to 6.0 degrees C warmer than mature leaflets and as much as 2.0 degrees C above ambient <span class="hlt">air</span> temperature with lower stomatal conductance and photosynthetic rates than mature leaflets. When measurements were taken from July to October, no significant differences were observed, although mature leaflet <span class="hlt">gas</span> <span class="hlt">exchange</span> rates declined to the levels of the juvenile leaves. The <span class="hlt">gas</span> <span class="hlt">exchange</span> data were supported by the observations that juvenile leaves had approximately half the number of functional stomata on a leaf surface area basis as did mature leaves. It was concluded that leaf temperature and stage of leaf development in sumac are strongly linked with the higher surface temperatures observed in juvenile leaflets in the early spring possibly being involved in promoting photosynthesis and leaf expansion when <span class="hlt">air</span> temperatures are cooler.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28372820','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28372820"><span><span class="hlt">Air-water</span> <span class="hlt">exchange</span> of PAHs and OPAHs at a superfund mega-site.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tidwell, Lane G; Blair Paulik, L; Anderson, Kim A</p> <p>2017-12-15</p> <p>Chemical fate is a concern at environmentally contaminated sites, but characterizing that fate can be difficult. Identifying and quantifying the movement of chemicals at the <span class="hlt">air-water</span> interface are important steps in characterizing chemical fate. Superfund sites are often suspected sources of <span class="hlt">air</span> pollution due to legacy sediment and <span class="hlt">water</span> contamination. A quantitative assessment of polycyclic aromatic hydrocarbons (PAHs) and oxygenated PAH (OPAHs) diffusive flux in a river system that contains a Superfund Mega-site, and passes through residential, urban and agricultural land, has not been reported before. Here, passive sampling devices (PSDs) were used to measure 60 polycyclic aromatic hydrocarbons (PAHs) and 22 oxygenated PAH (OPAHs) in <span class="hlt">air</span> and <span class="hlt">water</span>. From these concentrations the magnitude and direction of contaminant flux between these two compartments was calculated. The magnitude of PAH flux was greater at sites near or within the Superfund Mega-site than outside of the Superfund Mega-site. The largest net individual PAH deposition at a single site was naphthalene at a rate of -14,200 (±5780) (ng/m 2 )/day. The estimated one-year total flux of phenanthrene was -7.9×10 5 (ng/m 2 )/year. Human health risk associated with inhalation of vapor phase PAHs and dermal exposure to PAHs in <span class="hlt">water</span> were assessed by calculating benzo[a]pyrene equivalent concentrations. Excess lifetime cancer risk estimates show potential increased risk associated with exposure to PAHs at sites within and in close proximity to the Superfund Mega-site. Specifically, estimated excess lifetime cancer risk associated with dermal exposure and inhalation of PAHs was above 1 in 1 million within the Superfund Mega-site. The predominant depositional flux profile observed in this study suggests that the river <span class="hlt">water</span> in this Superfund site is largely a sink for airborne PAHs, rather than a source. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/14466','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/14466"><span><span class="hlt">Water</span> vapor mass balance method for determining <span class="hlt">air</span> infiltration rates in houses</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>David R. DeWalle; Gordon M. Heisler</p> <p>1980-01-01</p> <p>A <span class="hlt">water</span> vapor mass balance technique that includes the use of common humidity-control equipment can be used to determine average <span class="hlt">air</span> infiltration rates in buildings. Only measurements of the humidity inside and outside the home, the mass of vapor <span class="hlt">exchanged</span> by a humidifier/dehumidifier, and the volume of interior <span class="hlt">air</span> space are needed. This method gives results that...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1985htcg.agarS....N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1985htcg.agarS....N"><span>Heat <span class="hlt">exchangers</span> in regenerative <span class="hlt">gas</span> turbine cycles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nina, M. N. R.; Aguas, M. P. N.</p> <p>1985-09-01</p> <p>Advances in compact heat <span class="hlt">exchanger</span> design and fabrication together with fuel cost rises continuously improve the attractability of regenerative <span class="hlt">gas</span> turbine helicopter engines. In this study cycle parameters aiming at reduced specific fuel consumption and increased payload or mission range, have been optimized together with heat <span class="hlt">exchanger</span> type and size. The discussion is based on a typical mission for an attack helicopter in the 900 kw power class. A range of heat <span class="hlt">exchangers</span> is studied to define the most favorable geometry in terms of lower fuel consumption and minimum engine plus fuel weight. Heat <span class="hlt">exchanger</span> volume, frontal area ratio and pressure drop effect on cycle efficiency are considered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28873626','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28873626"><span>Effectiveness of <span class="hlt">water-air</span> and octanol-<span class="hlt">air</span> partition coefficients to predict lipophilic flavor release behavior from O/W emulsions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tamaru, Shunji; Igura, Noriyuki; Shimoda, Mitsuya</p> <p>2018-01-15</p> <p>Flavor release from food matrices depends on the partition of volatile flavor compounds between the food matrix and the vapor phase. Thus, we herein investigated the relationship between released flavor concentrations and three different partition coefficients, namely octanol-<span class="hlt">water</span>, octanol-<span class="hlt">air</span>, and <span class="hlt">water-air</span>, which represented the oil, <span class="hlt">water</span>, and <span class="hlt">air</span> phases present in emulsions. Limonene, 2-methylpyrazine, nonanal, benzaldehyde, ethyl benzoate, α-terpineol, benzyl alcohol, and octanoic acid were employed. The released concentrations of these flavor compounds from oil-in-<span class="hlt">water</span> (O/W) emulsions were measured under equilibrium using static headspace <span class="hlt">gas</span> chromatography. The results indicated that <span class="hlt">water-air</span> and octanol-<span class="hlt">air</span> partition coefficients correlated with the logarithms of the released concentrations in the headspace for highly lipophilic flavor compounds. Moreover, the same tendency was observed over various oil volume ratios in the emulsions. Our findings therefore suggest that octanol-<span class="hlt">air</span> and <span class="hlt">water-air</span> partition coefficients can be used to predict the released concentration of lipophilic flavor compounds from O/W emulsions. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.C11A0352L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.C11A0352L"><span>Radon and radium in the ice-covered Arctic Ocean, and what they reveal about <span class="hlt">gas</span> <span class="hlt">exchange</span> in the sea ice zone.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Loose, B.; Kelly, R. P.; Bigdeli, A.; Moran, S. B.</p> <p>2014-12-01</p> <p>The polar sea ice zones are regions of high primary productivity and interior <span class="hlt">water</span> mass formation. Consequently, the seasonal sea ice cycle appears important to both the solubility and biological carbon pumps. To estimate net CO2 transfer in the sea ice zone, we require accurate estimates of the <span class="hlt">air</span>-sea <span class="hlt">gas</span> transfer velocity. In the open ocean, the <span class="hlt">gas</span> transfer velocity is driven by wind, waves and bubbles - all of which are strongly altered by the presence of sea ice, making it difficult to translate open ocean estimates of <span class="hlt">gas</span> transfer to the ice zone. In this study, we present profiles of 222Rn and 226Ra throughout the mixed-layer and euphotic zone. Profiles were collected spanning a range of sea ice cover conditions from 40 to 100%. The profiles of Rn/Ra can be used to estimate the <span class="hlt">gas</span> transfer velocity, but the 3.8 day half-life of 222Rn implies that mixed layer radon will have a memory of the past ~20 days of <span class="hlt">gas</span> <span class="hlt">exchange</span> forcing, which may include a range of sea ice cover conditions. Here, we compare individual estimates of the <span class="hlt">gas</span> transfer velocity to the turbulent forcing conditions constrained from shipboard and regional reanalysis data to more appropriately capture the time history upper ocean Rn/Ra.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRC..122.2781S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRC..122.2781S"><span>Boundary layers at a dynamic interface: <span class="hlt">Air</span>-sea <span class="hlt">exchange</span> of heat and mass</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Szeri, Andrew J.</p> <p>2017-04-01</p> <p><span class="hlt">Exchange</span> of mass or heat across a turbulent liquid-<span class="hlt">gas</span> interface is a problem of critical interest, especially in <span class="hlt">air</span>-sea transfer of natural and anthropogenic gases involved in the study of climate. The goal in this research area is to determine the <span class="hlt">gas</span> flux from <span class="hlt">air</span> to sea or vice versa. For sparingly soluble nonreactive gases, this is controlled by liquid phase turbulent velocity fluctuations that act on the thin species concentration boundary layer on the liquid side of the interface. If the fluctuations in surface-normal velocity w' and <span class="hlt">gas</span> concentration c' are known, then it is possible to determine the turbulent contribution to the <span class="hlt">gas</span> flux. However, there is no suitable fundamental direct approach in the general case where neither w' nor c' can be easily measured. A new approach is presented to deduce key aspects about the near-surface turbulent motions from measurements that can be taken by an infrared (IR) camera. An equation is derived with inputs being the surface temperature and heat flux, and a solution method developed for the surface-normal strain experienced over time by boundary layers at the interface. Because the thermal and concentration boundary layers experience the same near-surface fluid motions, the solution for the surface-normal strain determines the <span class="hlt">gas</span> flux or <span class="hlt">gas</span> transfer velocity. Examples illustrate the approach in the cases of complete surface renewal, partial surface renewal, and insolation. The prospects for use of the approach in flows characterized by sheared interfaces or rapid boundary layer straining are explored.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18465177','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18465177"><span>Preoperative gender differences in pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> in morbidly obese subjects.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zavorsky, Gerald S; Christou, Nicolas V; Kim, Do Jun; Carli, Franco; Mayo, Nancy E</p> <p>2008-12-01</p> <p>Morbidly obese men may have poorer pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> compared to morbidly obese women (see Zavorsky et al., Chest 131:362-367, 2007). The purpose was to compare pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> in morbidly obese men and women at rest and throughout exercise. Twenty-five women (age=38+/-10 years, 164+/-7 cm, body mass index or BMI = 51+/-7 kg/m(2), peak oxygen consumption or VO(2peak)=2.0+/-0.4 l/min) and 17 men (age=43+/-9 years, 178+/-7 cm, BMI=50+/-10 kg/m(2), VO(2peak)=2.6+/-0.8 l/min) were recruited to perform a graded exercise test on a cycle ergometer with temperature-corrected arterial blood-<span class="hlt">gas</span> samples taken at rest and every minute of exercise, including peak exercise. At rest, women were 98% predicted for pulmonary diffusion compared to 88% predicted in men. At rest, women had better pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> compared to the men which was related to women having a lower waist-to-hip ratio (WHR; p<0.01). Only 20% of the subjects had an excessive alveolar-to-arterial oxygen partial pressure difference (>or=25 mmHg) at peak exercise, but 75% of the subjects showed inadequate compensatory hyperventilation at peak exercise (arterial carbon dioxide pressure >35 mmHg), and both were not different between genders. At rest, morbidly obese men have poorer pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> and pulmonary diffusion compared to morbidly obese women. The better <span class="hlt">gas</span> <span class="hlt">exchange</span> in women is related to the lower WHR in the women. During exercise, few subjects showed disturbances in pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> despite demonstrating poor compensatory hyperventilation at peak exercise.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16615688','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16615688"><span>Testing of heat <span class="hlt">exchangers</span> in membrane oxygenators using <span class="hlt">air</span> pressure.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hamilton, Carole; Stein, Jutta; Seidler, Rainer; Kind, Robert; Beck, Karin; Tosok, Jürgen; Upterfofel, Jörg</p> <p>2006-03-01</p> <p>All heat <span class="hlt">exchangers</span> (HE) in membrane oxygenators are tested by the manufacturer for <span class="hlt">water</span> leaks during the production phase. However, for safety reasons, it is highly recommended that HEs be tested again before clinical use. The most common method is to attach the heater-cooler to the HE and allow the <span class="hlt">water</span> to recirculate for at least 10 min, during which time a <span class="hlt">water</span> leak should be evident. To improve the detection of <span class="hlt">water</span> leaks, a test was devised using a pressure manometer with an integrated bulb used to pressurize the HE with <span class="hlt">air</span>. The cardiopulmonary bypass system is set up as per protocol. A pressure manometer adapted to a 1/2" tubing is connected to the <span class="hlt">water</span> inlet side of the oxygenator. The <span class="hlt">water</span> outlet side is blocked with a short piece of 1/2" deadend tubing. The HE is pressurized with 250 mmHg for at least 30 sec and observed for any drop. Over the last 2 years, only one oxygenator has been detected with a <span class="hlt">water</span> leak in which the <span class="hlt">air</span>-method leaktest was performed. This unit was sent back to the manufacturer who confirmed the failure. Even though the incidence of <span class="hlt">water</span> leaks is very low, it does occur and it is, therefore, important that all HEs are tested before they are used clinically. This method of using a pressure manometer offers many advantages, as the HE can be tested outside of the operating room (OR), allowing earlier testing of the oxygenator, no <span class="hlt">water</span> contact is necessary, and it is simple, easy and quick to perform.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5735977','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5735977"><span>Assessing the Effects of <span class="hlt">Water</span> Deficit on Photosynthesis Using Parameters Derived from Measurements of Leaf <span class="hlt">Gas</span> <span class="hlt">Exchange</span> and of Chlorophyll a Fluorescence</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Urban, Laurent; Aarrouf, Jawad; Bidel, Luc P. R.</p> <p>2017-01-01</p> <p><span class="hlt">Water</span> deficit (WD) is expected to increase in intensity, frequency and duration in many parts of the world as a consequence of global change, with potential negative effects on plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and growth. We review here the parameters that can be derived from measurements made on leaves, in the field, and that can be used to assess the effects of WD on the components of plant photosynthetic rate, including stomatal conductance, mesophyll conductance, photosynthetic capacity, light absorbance, and efficiency of absorbed light conversion into photosynthetic electron transport. We also review some of the parameters related to dissipation of excess energy and to rerouting of electron fluxes. Our focus is mainly on the techniques of <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements and of measurements of chlorophyll a fluorescence (ChlF), either alone or combined. But we put also emphasis on some of the parameters derived from analysis of the induction phase of maximal ChlF, notably because they could be used to assess damage to photosystem II. Eventually we briefly present the non-destructive methods based on the ChlF excitation ratio method which can be used to evaluate non-destructively leaf contents in anthocyanins and flavonols. PMID:29312367</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29312367','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29312367"><span>Assessing the Effects of <span class="hlt">Water</span> Deficit on Photosynthesis Using Parameters Derived from Measurements of Leaf <span class="hlt">Gas</span> <span class="hlt">Exchange</span> and of Chlorophyll a Fluorescence.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Urban, Laurent; Aarrouf, Jawad; Bidel, Luc P R</p> <p>2017-01-01</p> <p><span class="hlt">Water</span> deficit (WD) is expected to increase in intensity, frequency and duration in many parts of the world as a consequence of global change, with potential negative effects on plant <span class="hlt">gas</span> <span class="hlt">exchange</span> and growth. We review here the parameters that can be derived from measurements made on leaves, in the field, and that can be used to assess the effects of WD on the components of plant photosynthetic rate, including stomatal conductance, mesophyll conductance, photosynthetic capacity, light absorbance, and efficiency of absorbed light conversion into photosynthetic electron transport. We also review some of the parameters related to dissipation of excess energy and to rerouting of electron fluxes. Our focus is mainly on the techniques of <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements and of measurements of chlorophyll a fluorescence (ChlF), either alone or combined. But we put also emphasis on some of the parameters derived from analysis of the induction phase of maximal ChlF, notably because they could be used to assess damage to photosystem II. Eventually we briefly present the non-destructive methods based on the ChlF excitation ratio method which can be used to evaluate non-destructively leaf contents in anthocyanins and flavonols.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16667848','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16667848"><span>Influence of vesicular arbuscular mycorrhizae and leaf age on net <span class="hlt">gas</span> <span class="hlt">exchange</span> of citrus leaves.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Syvertsen, J P; Graham, J H</p> <p>1990-11-01</p> <p>The purpose of this study was to test the hypothesis that vesicular arbuscular mycorrhizal (VAM) fungi affect net assimilation of CO(2) (A) of different-aged citrus leaves independent of mineral nutrition effects of mycorrhizae. Citrus aurantium L., sour orange plants were grown for 6 months in a sandy soil low in phosphorus that was either infested with the VAM fungus, Glomus intraradices Schenck & Smith, or fertilized with additional phosphorus and left nonmycorrhizal (NM). Net CO(2) assimilation, stomatal conductance, <span class="hlt">water</span> use efficiency, and mineral nutrient status for expanding, recently expanded, and mature leaves were evaluated as well as plant size and relative growth rate of leaves. Nutrient status and net <span class="hlt">gas</span> <span class="hlt">exchange</span> varied with leaf age. G. intraradices-inoculated plants had well-established colonization (79% of root length) and were comparable in relative growth rate and size at final harvest with NM plants. Leaf mineral concentrations were generally the same for VAM and NM plants except for nitrogen. Although leaf nitrogen was apparently sufficient for high rates of A, VAM plants did have higher nitrogen concentrations than NM at the time of <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements. G. intraradices had no effect on A, stomatal conductance, or <span class="hlt">water</span> use efficiency, irrespective of leaf age. These results show that well-established VAM colonization does not affect net <span class="hlt">gas</span> <span class="hlt">exchange</span> of citrus plants that are comparable in size, growth rate, and nutritional status with NM plants.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29238999','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29238999"><span>Physiological <span class="hlt">gas</span> <span class="hlt">exchange</span> mapping of hyperpolarized 129 Xe using spiral-IDEAL and MOXE in a model of regional radiation-induced lung injury.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zanette, Brandon; Stirrat, Elaine; Jelveh, Salomeh; Hope, Andrew; Santyr, Giles</p> <p>2018-02-01</p> <p>To map physiological <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters using dissolved hyperpolarized (HP) 129 Xe in a rat model of regional radiation-induced lung injury (RILI) with spiral-IDEAL and the model of xenon <span class="hlt">exchange</span> (MOXE). Results are compared to quantitative histology of pulmonary tissue and red blood cell (RBC) distribution. Two cohorts (n = 6 each) of age-matched rats were used. One was irradiated in the right-medial lung, producing regional injury. <span class="hlt">Gas</span> <span class="hlt">exchange</span> was mapped 4 weeks postirradiation by imaging dissolved-phase HP 129 Xe using spiral-IDEAL at five <span class="hlt">gas</span> <span class="hlt">exchange</span> timepoints using a clinical 1.5 T scanner. Physiological lung parameters were extracted regionally on a voxel-wise basis using MOXE. Mean <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters, specifically <span class="hlt">air</span>-capillary barrier thickness (δ) and hematocrit (HCT) in the right-medial lung were compared to the contralateral lung as well as nonirradiated control animals. Whole-lung spectroscopic analysis of <span class="hlt">gas</span> <span class="hlt">exchange</span> was also performed. δ was significantly increased (1.43 ± 0.12 μm from 1.07 ± 0.09 μm) and HCT was significantly decreased (17.2 ± 1.2% from 23.6 ± 1.9%) in the right-medial lung (i.e., irradiated region) compared to the contralateral lung of the irradiated rats. These changes were not observed in healthy controls. δ and HCT correlated with histologically measured increases in pulmonary tissue heterogeneity (r = 0.77) and decreases in RBC distribution (r = 0.91), respectively. No changes were observed using whole-lung analysis. This work demonstrates the feasibility of mapping <span class="hlt">gas</span> <span class="hlt">exchange</span> using HP 129 Xe in an animal model of RILI 4 weeks postirradiation. Spatially resolved <span class="hlt">gas</span> <span class="hlt">exchange</span> mapping is sensitive to regional injury between cohorts that was undetected with whole-lung <span class="hlt">gas</span> <span class="hlt">exchange</span> analysis, in agreement with histology. <span class="hlt">Gas</span> <span class="hlt">exchange</span> mapping holds promise for assessing regional lung function in RILI and other pulmonary diseases. © 2017 The Authors. Medical Physics published by Wiley</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhDT........23V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhDT........23V"><span>Development and Characterization of <span class="hlt">Gas</span> Diffusion Layer Using Carbon Slurry Dispersed by Ammonium Lauryl Sulfate for Proton <span class="hlt">Exchange</span> Member Fuel Cells</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Villacorta, Rashida</p> <p></p> <p><span class="hlt">Gas</span> diffusion layers (GDLs) are a critical and essential part of proton <span class="hlt">exchange</span> membrane fuel cells (PEMFCs). They carry out various important functions such as transportation of reactants to and from the reaction sites. The material properties and structural characteristics of the substrate and the microporous layer strongly influence fuel cell performance. The microporous layer of the GDLs was fabricated with the carbon slurry dispersed in <span class="hlt">water</span> containing ammonium lauryl sulfate (ALS) using the wire rod coating method. GDLs were fabricated with different materials to compose the microporous layer and evaluated the effects on PEMFC power output performance. The consistency of the carbon slurry was achieved by adding 25 wt. % of PTFE, a binding agent with a 75:25 ratio of carbon (Pureblack and vapor grown carbon fiber). The GDLs were investigated in PEMFC under various relative humidity (RH) conditions using H2/O2 and H2/<span class="hlt">Air</span>. GDLs were also fabricated with the carbon slurry dispersed in <span class="hlt">water</span> containing sodium dodecyl sulfate (SDS) and multiwalled carbon nanotubes (MWCNTs) with isopropyl alcohol (IPA) based for fuel cell performance comparison. MWCNTs and SDS exhibits the highest performance at 60% and 70% RH with a peak power density of 1100 mW.cm-2 and 850 mW.cm-2 using <span class="hlt">air</span> and oxygen as an oxidant. This means that the <span class="hlt">gas</span> diffusion characteristics of these two samples were optimum at 60 and 70 % RH with high limiting current density range. It was also found that the composition of the carbon slurry, specifically ALS concentration has the highest peak power density of 1300 and 500mW.cm-2 for both H2/O 2 and H2/<span class="hlt">Air</span> at 100% RH. However, SDS and MWCNTs demonstrates the lowest power density using <span class="hlt">air</span> and oxygen as an oxidants at 100% RH.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/875007','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/875007"><span>Method and apparatus for extracting <span class="hlt">water</span> from <span class="hlt">air</span> using a desiccant</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Spletzer, Barry L.; Callow, Diane Schafer</p> <p>2003-01-01</p> <p>The present invention provides a method and apparatus for extracting liquid <span class="hlt">water</span> from moist <span class="hlt">air</span> using minimal energy input. The method can be considered as four phases: (1) adsorbing <span class="hlt">water</span> from <span class="hlt">air</span> into a desiccant, (2) isolating the <span class="hlt">water</span>-laden desiccant from the <span class="hlt">air</span> source, (3) desorbing <span class="hlt">water</span> as vapor from the desiccant into a chamber, and (4) isolating the desiccant from the chamber, and compressing the vapor in the chamber to form liquid condensate. The liquid condensate can be removed for use. Careful design of the dead volumes and pressure balances can minimize the energy required. The dried <span class="hlt">air</span> can be <span class="hlt">exchanged</span> for fresh moist <span class="hlt">air</span> and the process repeated. An apparatus comprises a first chamber in fluid communication with a desiccant, and having ports to intake moist <span class="hlt">air</span> and exhaust dried <span class="hlt">air</span>. The apparatus also comprises a second chamber in fluid communication with the desiccant. The second chamber allows variable internal pressure, and has a port for removal of liquid condensate. Each chamber can be configured to be isolated or in communication with the desiccant. The first chamber can be configured to be isolated or in communication with a course of moist <span class="hlt">air</span>. Various arrangements of valves, pistons, and chambers are described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/7300088-operating-experiences-rotary-air-air-heat-exchangers-hospitals-schools-nursing-homes-swimming-pools','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/7300088-operating-experiences-rotary-air-air-heat-exchangers-hospitals-schools-nursing-homes-swimming-pools"><span>Operating experiences with rotary <span class="hlt">air-to-air</span> heat <span class="hlt">exchangers</span>: hospitals, schools, nursing homes, swimming pools</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>Pearson, R.J.</p> <p>1976-01-01</p> <p>Systems utilizing rotary <span class="hlt">air-to-air</span> heat <span class="hlt">exchangers</span> are discussed. Basic considerations of use (fresh <span class="hlt">air</span> requirements, system configurations, cost considerations), typical system layout/design considerations, and operating observations by engineers, staff and maintenance personnel are described.</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('https://www.ncbi.nlm.nih.gov/pubmed/14753695','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14753695"><span>Study of the influence of surfactants on the transfer of gases into liquids by inverse <span class="hlt">gas</span> chromatography.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Atta, Khan Rashid; Gavril, Dimitrios; Loukopoulos, Vassilios; Karaiskakis, George</p> <p>2004-01-16</p> <p>The experimental technique of the reversed-flow version of inverse <span class="hlt">gas</span> chromatography was applied for the study of effects of surfactants in reducing <span class="hlt">air-water</span> <span class="hlt">exchange</span> rates. The vinyl chloride (VC)-<span class="hlt">water</span> system was used as a model, which is of great importance in environmental chemistry. Using suitable mathematical analysis, various physicochemical quantities were calculated, among which the most significant are: Partition coefficients of the VC <span class="hlt">gas</span> between the surfactant interface and the carrier <span class="hlt">gas</span> nitrogen, as well as between the bulk of the <span class="hlt">water</span> + surfactant solution and the carrier <span class="hlt">gas</span> nitrogen, overall mass transfer coefficients of VC in the liquid (<span class="hlt">water</span> + surfactant) and the <span class="hlt">gas</span> (nitrogen) phases, <span class="hlt">water</span> and surfactant film transfer coefficients, nitrogen, <span class="hlt">water</span> and surfactant phase resistances for the transfer of VC into the <span class="hlt">water</span> solution, relative resistance of surfactant in the transfer of VC into the bulk of solution, <span class="hlt">exchange</span> velocity of VC between nitrogen and the liquid solution, and finally the thickness of the surfactant stagnant film in the liquid phase, according to the three phase resistance model. From the variation of the above parameters with the surfactant's concentration, important conclusions concerning the effects of surfactants on the transfer of a <span class="hlt">gas</span> at the <span class="hlt">air</span>-liquid interface, as well as to the bulk of the liquid were extracted. An interesting finding of this work was also that by successive addition of surfactant, the critical micelle concentration of surfactant was obtained, after which follows a steady-state for the transfer of the <span class="hlt">gas</span> into the <span class="hlt">water</span> body, which could be attributed to the transition from mono- to multi-layer state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28003523','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28003523"><span>An experimental evolution study confirms that discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> does not contribute to body <span class="hlt">water</span> conservation in locusts.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Talal, Stav; Ayali, Amir; Gefen, Eran</p> <p>2016-12-01</p> <p>The adaptive nature of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> (DGE) in insects is contentious. The classic 'hygric hypothesis', which posits that DGE serves to reduce respiratory <span class="hlt">water</span> loss (RWL), is still the best supported. We thus focused on the hygric hypothesis in this first-ever experimental evolution study of any of the competing adaptive hypotheses. We compared populations of the migratory locust (Locusta migratoria) that underwent 10 consecutive generations of selection for desiccation resistance with control populations. Selected locusts survived 36% longer under desiccation stress but DGE prevalence did not differ between these and control populations (approx. 75%). Evolved changes in DGE properties in the selected locusts included longer cycle and interburst durations. However, in contrast with predictions of the hygric hypothesis, these changes were not associated with reduced RWL rates. Other responses observed in the selected locusts were higher body <span class="hlt">water</span> content when hydrated and lower total evaporative <span class="hlt">water</span> loss rates. Hence, our data suggest that DGE cycle properties in selected locusts are a consequence of an evolved increased ability to store <span class="hlt">water</span>, and thus an improved capacity to buffer accumulated CO 2 , rather than an adaptive response to desiccation. We conclude that DGE is unlikely to be an evolutionary response to dehydration challenge in locusts. © 2016 The Author(s).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/33600','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/33600"><span>Leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> of mature bottomland oak trees</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Rico M. Gazal; Mark E. Kubiske; Kristina F. Connor</p> <p>2009-01-01</p> <p>We determined how changes in environmental moisture affected leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> in Nuttall (Quercus texana Buckley), overcup (Q. lyrata Walt.), and dominant and codominant swamp chestnut (Q. michauxii Nutt.) oak trees in Mississippi and Louisiana. We used canopy access towers to measure leaf level <span class="hlt">gas</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016E%26ES...49e2007B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016E%26ES...49e2007B"><span>Dynamic <span class="hlt">water</span> behaviour due to one trapped <span class="hlt">air</span> pocket in a laboratory pipeline apparatus</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bergant, A.; Karadžić, U.; Tijsseling, A.</p> <p>2016-11-01</p> <p>Trapped <span class="hlt">air</span> pockets may cause severe operational problems in hydropower and <span class="hlt">water</span> supply systems. A locally isolated <span class="hlt">air</span> pocket creates distinct amplitude, shape and timing of pressure pulses. This paper investigates dynamic behaviour of a single trapped <span class="hlt">air</span> pocket. The <span class="hlt">air</span> pocket is incorporated as a boundary condition into the discrete <span class="hlt">gas</span> cavity model (DGCM). DGCM allows small <span class="hlt">gas</span> cavities to form at computational sections in the method of characteristics (MOC). The growth of the pocket and <span class="hlt">gas</span> cavities is described by the <span class="hlt">water</span> hammer compatibility equation(s), the continuity equation for the cavity volume, and the equation of state of an ideal <span class="hlt">gas</span>. Isentropic behaviour is assumed for the trapped <span class="hlt">gas</span> pocket and an isothermal bath for small <span class="hlt">gas</span> cavities. Experimental investigations have been performed in a laboratory pipeline apparatus. The apparatus consists of an upstream end high-pressure tank, a horizontal steel pipeline (total length 55.37 m, inner diameter 18 mm), four valve units positioned along the pipeline including the end points, and a downstream end tank. A trapped <span class="hlt">air</span> pocket is captured between two ball valves at the downstream end of the pipeline. The transient event is initiated by rapid opening of the upstream end valve; the downstream end valve stays closed during the event. Predicted and measured results for a few typical cases are compared and discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/id0443.photos.220119p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/id0443.photos.220119p/"><span>ETR COMPRESSOR BUILDING, TRA643. CAMERA FACES NORTHEAST. <span class="hlt">WATER</span> HEAT <span class="hlt">EXCHANGER</span> ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>ETR COMPRESSOR BUILDING, TRA-643. CAMERA FACES NORTHEAST. <span class="hlt">WATER</span> HEAT <span class="hlt">EXCHANGER</span> IS IN LEFT FOREGROUND. A PARTIALLY ASSEMBLED PLANT <span class="hlt">AIR</span> CONDITIONER IS AT CENTER. WORKERS AT RIGHT ASSEMBLE 4000 HORSEPOWER COMPRESSOR DRIVE MOTOR AT RIGHT. INL NEGATIVE NO. 56-3714. R.G. Larsen, Photographer, 11/13/1956 - Idaho National Engineering Laboratory, Test Reactor Area, Materials & Engineering Test Reactors, Scoville, Butte County, ID</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JPS...297..202X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JPS...297..202X"><span>Template-directed fabrication of porous <span class="hlt">gas</span> diffusion layer for magnesium <span class="hlt">air</span> batteries</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xue, Yejian; Miao, He; Sun, Shanshan; Wang, Qin; Li, Shihua; Liu, Zhaoping</p> <p>2015-11-01</p> <p>The uniform micropore distribution in the <span class="hlt">gas</span> diffusion layers (GDLs) of the <span class="hlt">air</span>-breathing cathode is very important for the metal <span class="hlt">air</span> batteries. In this work, the super-hydrophobic GDL with the interconnected regular pores is prepared by a facile silica template method, and then the electrochemical properties of the Mg <span class="hlt">air</span> batteries containing these GDLs are investigated. The results indicate that the interconnected and uniform pore structure, the available <span class="hlt">water</span>-breakout pressure and the high <span class="hlt">gas</span> permeability coefficient of the GDL can be obtained by the application of 30% silica template. The maximum power density of the Mg <span class="hlt">air</span> battery containing the GDL with 30% regular pores reaches 88.9 mW cm-2 which is about 1.2 times that containing the pristine GDL. Furthermore, the GDL with 30% regular pores exhibits the improved the long term hydrophobic stability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26290590','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26290590"><span>The effect of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> on respiratory <span class="hlt">water</span> loss in grasshoppers (Orthoptera: Acrididae) varies across an aridity gradient.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huang, Shu-Ping; Talal, Stav; Ayali, Amir; Gefen, Eran</p> <p>2015-08-01</p> <p>The significance of discontinuous <span class="hlt">gas-exchange</span> cycles (DGC) in reducing respiratory <span class="hlt">water</span> loss (RWL) in insects is contentious. Results from single-species studies are equivocal in their support of the classic 'hygric hypothesis' for the evolution of DGC, whereas comparative analyses generally support a link between DGC and <span class="hlt">water</span> balance. In this study, we investigated DGC prevalence and characteristics and RWL in three grasshopper species (Acrididae, subfamily Pamphaginae) across an aridity gradient in Israel. In order to determine whether DGC contributes to a reduction in RWL, we compared the DGC characteristics and RWL associated with CO2 release (transpiration ratio, i.e. the molar ratio of RWL to CO2 emission rates) among these species. Transpiration ratios of DGC and continuous breathers were also compared intraspecifically. Our data show that DGC characteristics, DGC prevalence and the transpiration ratios correlate well with habitat aridity. The xeric-adapted Tmethis pulchripennis exhibited a significantly shorter burst period and lower transpiration ratio compared with the other two mesic species, Ocneropsis bethlemita and Ocneropsis lividipes. However, DGC resulted in significant <span class="hlt">water</span> savings compared with continuous <span class="hlt">exchange</span> in T. pulchripennis only. These unique DGC characteristics for T. pulchripennis were correlated with its significantly higher mass-specific tracheal volume. Our data suggest that the origin of DGC may not be adaptive, but rather that evolved modulation of cycle characteristics confers a fitness advantage under stressful conditions. This modulation may result from morphological and/or physiological modifications. © 2015. Published by The Company of Biologists Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AtmEn..75..348P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AtmEn..75..348P"><span>Intensive measurements of <span class="hlt">gas</span>, <span class="hlt">water</span>, and energy <span class="hlt">exchange</span> between vegetation and troposphere during the MONTES campaign in a vegetation gradient from short semi-desertic shrublands to tall wet temperate forests in the NW Mediterranean Basin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Peñuelas, J.; Guenther, A.; Rapparini, F.; Llusia, J.; Filella, I.; Seco, R.; Estiarte, M.; Mejia-Chang, M.; Ogaya, R.; Ibañez, J.; Sardans, J.; Castaño, L. M.; Turnipseed, A.; Duhl, T.; Harley, P.; Vila, J.; Estavillo, J. M.; Menéndez, S.; Facini, O.; Baraldi, R.; Geron, C.; Mak, J.; Patton, E. G.; Jiang, X.; Greenberg, J.</p> <p>2013-08-01</p> <p>MONTES (“Woodlands”) was a multidisciplinary international field campaign aimed at measuring energy, <span class="hlt">water</span> and especially <span class="hlt">gas</span> <span class="hlt">exchange</span> between vegetation and atmosphere in a gradient from short semi-desertic shrublands to tall wet temperate forests in NE Spain in the North Western Mediterranean Basin (WMB). The measurements were performed at a semidesertic area (Monegros), at a coastal Mediterranean shrubland area (Garraf), at a typical Mediterranean holm oak forest area (Prades) and at a wet temperate beech forest (Montseny) during spring (April 2010) under optimal plant physiological conditions in driest-warmest sites and during summer (July 2010) with drought and heat stresses in the driest-warmest sites and optimal conditions in the wettest-coolest site. The objective of this campaign was to study the differences in <span class="hlt">gas</span>, <span class="hlt">water</span> and energy <span class="hlt">exchange</span> occurring at different vegetation coverages and biomasses. Particular attention was devoted to quantitatively understand the <span class="hlt">exchange</span> of biogenic volatile organic compounds (BVOCs) because of their biological and environmental effects in the WMB. A wide range of instruments (GC-MS, PTR-MS, meteorological sensors, O3 monitors,…) and vertical platforms such as masts, tethered balloons and aircraft were used to characterize the <span class="hlt">gas</span>, <span class="hlt">water</span> and energy <span class="hlt">exchange</span> at increasing footprint areas by measuring vertical profiles. In this paper we provide an overview of the MONTES campaign: the objectives, the characterization of the biomass and <span class="hlt">gas</span>, <span class="hlt">water</span> and energy <span class="hlt">exchange</span> in the 4 sites-areas using satellite data, the estimation of isoprene and monoterpene emissions using MEGAN model, the measurements performed and the first results. The isoprene and monoterpene emission rates estimated with MEGAN and emission factors measured at the foliar level for the dominant species ranged from about 0 to 0.2 mg m-2 h-1 in April. The warmer temperature in July resulted in higher model estimates from about 0 to ca. 1.6 mg m-2 h-1 for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130000766','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130000766"><span>International Space Station Common Cabin <span class="hlt">Air</span> Assembly Condensing Heat <span class="hlt">Exchanger</span> Hydrophilic Coating Operation, Recovery, and Lessons Learned</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Balistreri, Steven F.; Steele, John W.; Caron, Mark E.; Laliberte, Yvon J.; Shaw, Laura A.</p> <p>2013-01-01</p> <p>The ability to control the temperature and humidity of an environment or habitat is critical for human survival. These factors are important to maintaining human health and comfort, as well as maintaining mechanical and electrical equipment in good working order to support the human and to accomplish mission objectives. The temperature and humidity of the International Space Station (ISS) United States On-orbit Segment (USOS) cabin <span class="hlt">air</span> is controlled by the Common Cabin <span class="hlt">Air</span> Assembly (CCAA). The CCAA consists of a fan, a condensing heat <span class="hlt">exchanger</span> (CHX), an <span class="hlt">air/water</span> separator, temperature and liquid sensors, and electrical controlling hardware and software. The CHX is the primary component responsible for control of temperature and humidity. The CCAA CHX contains a chemical coating that was developed to be hydrophilic and thus attract <span class="hlt">water</span> from the humid influent <span class="hlt">air</span>. This attraction forms the basis for <span class="hlt">water</span> removal and therefore cabin humidity control. However, there have been several instances of CHX coatings becoming hydrophobic and repelling <span class="hlt">water</span>. When this behavior is observed in an operational CHX in the ISS segments, the unit s ability to remove moisture from the <span class="hlt">air</span> is compromised and the result is liquid <span class="hlt">water</span> carryover into downstream ducting and systems. This <span class="hlt">water</span> carryover can have detrimental effects on the ISS cabin atmosphere quality and on the health of downstream hardware. If the <span class="hlt">water</span> carryover is severe and widespread, this behavior can result in an inability to maintain humidity levels in the USOS. This paper will describe the operation of the five CCAAs within the USOS, the potential causes of the hydrophobic condition, and the impacts of the resulting <span class="hlt">water</span> carryover to downstream systems. It will describe the history of this behavior and the actual observed impacts to the ISS USOS. Information on mitigation steps to protect the health of future CHX hydrophilic coatings as well as remediation and recovery of the full heat <span class="hlt">exchanger</span> will be</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=262549','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=262549"><span>Quantifying the measurement errors in a LI-6400 <span class="hlt">gas</span> <span class="hlt">exchange</span> system and their effects on the parameterization of Farquhar et al. model for C3 leaves</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>The LI-6400 <span class="hlt">gas</span> <span class="hlt">exchange</span> system (Li-Cor, Inc, Lincoln, NE, USA) has been widely used for the measurement of net <span class="hlt">gas</span> <span class="hlt">exchanges</span> and calibration/parameterization of leaf models. Measurement errors due to diffusive leakages of <span class="hlt">water</span> vapor and carbon dioxide between inside and outside of the leaf chamber...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1361339-european-regional-climate-zone-modeling-commercial-absorption-heat-pump-hot-water-heater','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1361339-european-regional-climate-zone-modeling-commercial-absorption-heat-pump-hot-water-heater"><span>European Regional Climate Zone Modeling of a Commercial Absorption Heat Pump Hot <span class="hlt">Water</span> Heater</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>Sharma, Vishaldeep; Shen, Bo; Keinath, Chris</p> <p>2017-01-01</p> <p>High efficiency <span class="hlt">gas</span>-burning hot <span class="hlt">water</span> heating takes advantage of a condensing heat <span class="hlt">exchanger</span> to deliver improved combustion efficiency over a standard non-condensing configuration. The <span class="hlt">water</span> heating is always lower than the <span class="hlt">gas</span> heating value. In contrast, <span class="hlt">Gas</span> Absorption Heat Pump (GAHP) hot <span class="hlt">water</span> heating combines the efficiency of <span class="hlt">gas</span> burning with the performance increase from a heat pump to offer significant <span class="hlt">gas</span> energy savings. An ammonia-<span class="hlt">water</span> system also has the advantage of zero Ozone Depletion Potential and low Global Warming Potential. In comparison with <span class="hlt">air</span> source electric heat pumps, the absorption system can maintain higher coefficients of performance in coldermore » climates. In this work, a GAHP commercial <span class="hlt">water</span> heating system was compared to a condensing <span class="hlt">gas</span> storage system for a range of locations and climate zones across Europe. The thermodynamic performance map of a single effect ammonia-<span class="hlt">water</span> absorption system was used in a building energy modeling software that could also incorporate the changing ambient <span class="hlt">air</span> temperature and <span class="hlt">water</span> mains temperature for a specific location, as well as a full-service restaurant <span class="hlt">water</span> draw pattern.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.891a2135C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.891a2135C"><span>Research on Heat <span class="hlt">Exchange</span> Process in Aircraft <span class="hlt">Air</span> Conditioning System</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chichindaev, A. V.</p> <p>2017-11-01</p> <p>Using of heat-<span class="hlt">exchanger</span>-condenser in the <span class="hlt">air</span> conditioning system of the airplane Tu-204 (Boeing, Airbus, Superjet 100, MS-21, etc.) for cooling the compressed <span class="hlt">air</span> by the cold <span class="hlt">air</span> with negative temperature exiting the turbine results in a number of operational problems. Mainly it’s frosting of the heat <span class="hlt">exchange</span> surface, which is the cause of live-section channels frosting, resistance increasing and airflow in the system decreasing. The purpose of this work is to analyse the known freeze-up-fighting methods for heat-<span class="hlt">exchanger</span>-condenser, description of the features of anti-icing protection and offering solutions to this problem. For the problem of optimizing the design of heat <span class="hlt">exchangers</span> in this work used generalized criterion that describes the ratio of thermal resistances of cold and hot sections, which include: the ratio of the initial values of heat transfer agents flow state; heat <span class="hlt">exchange</span> surface finning coefficients; factors which describes the ratio of operating parameters and finning area. By controlling the ratio of the thermal resistances can be obtained the desired temperature of the heat <span class="hlt">exchange</span> surface, which would prevent freezing. The work presents the results of a numerical study of the effect of different combinations of regime and geometrical factors changes on reduction of the heat-<span class="hlt">exchanger</span>-condenser freezing surface area, including using of variable ratio of thermal resistances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20160009653&hterms=water&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26Nf%3DPublication-Date%257CBTWN%2B20140101%2B20141231%26N%3D0%26No%3D30%26Ntt%3Dwater','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20160009653&hterms=water&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26Nf%3DPublication-Date%257CBTWN%2B20140101%2B20141231%26N%3D0%26No%3D30%26Ntt%3Dwater"><span>Recent Advances in <span class="hlt">Water</span> Analysis with <span class="hlt">Gas</span> Chromatograph Mass Spectrometers</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>MacAskill, John A.; Tsikata, Edem</p> <p>2014-01-01</p> <p>We report on progress made in developing a <span class="hlt">water</span> sampling system for detection and analysis of volatile organic compounds in <span class="hlt">water</span> with a <span class="hlt">gas</span> chromatograph mass spectrometer (GCMS). Two approaches are described herein. The first approach uses a custom <span class="hlt">water</span> pre-concentrator for performing trap and purge of VOCs from <span class="hlt">water</span>. The second approach uses a custom micro-volume, split-splitless injector that is compatible with <span class="hlt">air</span> and <span class="hlt">water</span>. These <span class="hlt">water</span> sampling systems will enable a single GC-based instrument to analyze <span class="hlt">air</span> and <span class="hlt">water</span> samples for VOC content. As reduced mass, volume, and power is crucial for long-duration, manned space-exploration, these <span class="hlt">water</span> sampling systems will demonstrate the ability of a GCMS to monitor both <span class="hlt">air</span> and <span class="hlt">water</span> quality of the astronaut environment, thereby reducing the amount of required instrumentation for long duration habitation. Laboratory prototypes of these <span class="hlt">water</span> sampling systems have been constructed and tested with a quadrupole ion trap mass spectrometer as well as a thermal conductivity detector. Presented herein are details of these <span class="hlt">water</span> sampling system with preliminary test results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28079171','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28079171"><span>Comfortable, high-efficiency heat pump with desiccant-coated, <span class="hlt">water</span>-sorbing heat <span class="hlt">exchangers</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tu, Y D; Wang, R Z; Ge, T S; Zheng, X</p> <p>2017-01-12</p> <p>Comfortable, efficient, and affordable heating, ventilation, and <span class="hlt">air</span> conditioning systems in buildings are highly desirable due to the demands of energy efficiency and environmental friendliness. Traditional vapor-compression <span class="hlt">air</span> conditioners exhibit a lower coefficient of performance (COP) (typically 2.8-3.8) owing to the cooling-based dehumidification methods that handle both sensible and latent loads together. Temperature- and humidity-independent control or desiccant systems have been proposed to overcome these challenges; however, the COP of current desiccant systems is quite small and additional heat sources are usually needed. Here, we report on a desiccant-enhanced, direct expansion heat pump based on a <span class="hlt">water</span>-sorbing heat <span class="hlt">exchanger</span> with a desiccant coating that exhibits an ultrahigh COP value of more than 7 without sacrificing any comfort or compactness. The pump's efficiency is doubled compared to that of pumps currently used in conventional room <span class="hlt">air</span> conditioners, which is a revolutionary HVAC breakthrough. Our proposed <span class="hlt">water</span>-sorbing heat <span class="hlt">exchanger</span> can independently handle sensible and latent loads at the same time. The desiccants adsorb moisture almost isothermally and can be regenerated by condensation heat. This new approach opens up the possibility of achieving ultrahigh efficiency for a broad range of temperature- and humidity-control applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5227918','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5227918"><span>Comfortable, high-efficiency heat pump with desiccant-coated, <span class="hlt">water</span>-sorbing heat <span class="hlt">exchangers</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>Tu, Y. D.; Wang, R. Z.; Ge, T. S.; Zheng, X.</p> <p>2017-01-01</p> <p>Comfortable, efficient, and affordable heating, ventilation, and <span class="hlt">air</span> conditioning systems in buildings are highly desirable due to the demands of energy efficiency and environmental friendliness. Traditional vapor-compression <span class="hlt">air</span> conditioners exhibit a lower coefficient of performance (COP) (typically 2.8–3.8) owing to the cooling-based dehumidification methods that handle both sensible and latent loads together. Temperature- and humidity-independent control or desiccant systems have been proposed to overcome these challenges; however, the COP of current desiccant systems is quite small and additional heat sources are usually needed. Here, we report on a desiccant-enhanced, direct expansion heat pump based on a <span class="hlt">water</span>-sorbing heat <span class="hlt">exchanger</span> with a desiccant coating that exhibits an ultrahigh COP value of more than 7 without sacrificing any comfort or compactness. The pump’s efficiency is doubled compared to that of pumps currently used in conventional room <span class="hlt">air</span> conditioners, which is a revolutionary HVAC breakthrough. Our proposed <span class="hlt">water</span>-sorbing heat <span class="hlt">exchanger</span> can independently handle sensible and latent loads at the same time. The desiccants adsorb moisture almost isothermally and can be regenerated by condensation heat. This new approach opens up the possibility of achieving ultrahigh efficiency for a broad range of temperature- and humidity-control applications. PMID:28079171</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NatSR...740437T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatSR...740437T"><span>Comfortable, high-efficiency heat pump with desiccant-coated, <span class="hlt">water</span>-sorbing heat <span class="hlt">exchangers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tu, Y. D.; Wang, R. Z.; Ge, T. S.; Zheng, X.</p> <p>2017-01-01</p> <p>Comfortable, efficient, and affordable heating, ventilation, and <span class="hlt">air</span> conditioning systems in buildings are highly desirable due to the demands of energy efficiency and environmental friendliness. Traditional vapor-compression <span class="hlt">air</span> conditioners exhibit a lower coefficient of performance (COP) (typically 2.8-3.8) owing to the cooling-based dehumidification methods that handle both sensible and latent loads together. Temperature- and humidity-independent control or desiccant systems have been proposed to overcome these challenges; however, the COP of current desiccant systems is quite small and additional heat sources are usually needed. Here, we report on a desiccant-enhanced, direct expansion heat pump based on a <span class="hlt">water</span>-sorbing heat <span class="hlt">exchanger</span> with a desiccant coating that exhibits an ultrahigh COP value of more than 7 without sacrificing any comfort or compactness. The pump’s efficiency is doubled compared to that of pumps currently used in conventional room <span class="hlt">air</span> conditioners, which is a revolutionary HVAC breakthrough. Our proposed <span class="hlt">water</span>-sorbing heat <span class="hlt">exchanger</span> can independently handle sensible and latent loads at the same time. The desiccants adsorb moisture almost isothermally and can be regenerated by condensation heat. This new approach opens up the possibility of achieving ultrahigh efficiency for a broad range of temperature- and humidity-control applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16827010','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16827010"><span>Homeostatic maintenance of ponderosa pine <span class="hlt">gas</span> <span class="hlt">exchange</span> in response to stand density changes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>McDowell, Nate G; Adams, Henry D; Bailey, John D; Hess, Marcey; Kolb, Thomas E</p> <p>2006-06-01</p> <p>Homeostatic maintenance of <span class="hlt">gas</span> <span class="hlt">exchange</span> optimizes carbon gain per <span class="hlt">water</span> loss. Homeostasis is regulated by short-term physiological and long-term structural mechanisms, both of which may respond to changes in resource availability associated with competition. Therefore, stand density regulation via silvicultural manipulations may facilitate growth and survival through mechanisms operating at both short and long timescales. We investigated the responses of ponderosa pine (Pinus ponderosa) to stand basal area manipulations in Arizona, USA. Stand basal area was manipulated to seven replicated levels in 1962 and was maintained for four decades by decadal thinning. We measured basal area increment (BAI) to assess the response and sustainability of wood growth, carbon isotope discrimination (A) inferred from annual rings to assess the response of crown <span class="hlt">gas</span> <span class="hlt">exchange</span>, and ratios of leaf area to sapwood area (A(l):A(s)) to assess longer term structural acclimation. Basal area treatments increased soil <span class="hlt">water</span> potential (r2 = 0.99) but did not affect photosynthetic capacity. BAI increased within two years of thinning, and the 40-year mean BAI was negatively correlated with stand basal area (r2 = 0.98). delta was negatively correlated with stand basal area for years 5 through 12 after thinning (r2 = 0.90). However, delta was relatively invariant with basal area for the period 13-40 years after initial thinning despite maintenance of treatment basal areas via repeated decadal thinnings. Independent <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements verified that the ratio of photosynthesis to stomatal conductance was invariant with basal area, but absolute values of both were elevated at lower basal areas. A(l):A(s) was negatively correlated with basal area (r2 = 0.93). We hypothesize that increased A(l):A(s) is a homeostatic response to increased <span class="hlt">water</span> availability that maximizes <span class="hlt">water</span>-use efficiency and whole-tree carbon uptake. Elevated A(l):A(s) of trees at low basal areas was associated with greater</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29341855','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29341855"><span>Probabilistic assessment of the potential indoor <span class="hlt">air</span> impacts of vent-free <span class="hlt">gas</span> heating appliances in energy-efficient homes in the United States.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Whitmyre, Gary K; Pandian, Muhilan D</p> <p>2018-06-01</p> <p>Use of vent-free <span class="hlt">gas</span> heating appliances for supplemental heating in U.S. homes is increasing. However, there is currently a lack of information on the potential impact of these appliances on indoor <span class="hlt">air</span> quality for homes constructed according to energy-efficient and green building standards. A probabilistic analysis was conducted to estimate the impact of vent-free <span class="hlt">gas</span> heating appliances on indoor <span class="hlt">air</span> concentrations of carbon monoxide (CO), nitrogen dioxide (NO 2 ), carbon dioxide (CO 2 ), <span class="hlt">water</span> vapor, and oxygen in "tight" energy-efficient homes in the United States. A total of 20,000 simulations were conducted for each Department of Energy (DOE) heating region to capture a wide range of home sizes, appliance features, and conditions, by varying a number of parameters, e.g., room volume, house volume, outdoor humidity, <span class="hlt">air</span> <span class="hlt">exchange</span> rates, appliance input rates (Btu/hr), and house heat loss factors. Predicted airborne levels of CO were below the U.S. Environmental Protection Agency (EPA) standard of 9 ppm for all modeled cases. The airborne concentrations of NO 2 were below the U.S. Consumer Product Safety Commission (CPSC) guideline of 0.3 ppm and the Health Canada benchmark of 0.25 ppm in all cases and were below the World Health Organization (WHO) standard of 0.11 ppm in 99-100% of all cases. Predicted levels of CO 2 were below the Health Canada standard of 3500 ppm for all simulated cases. Oxygen levels in the room of vent-free heating appliance use were not significantly reduced. The great majority of cases in all DOE regions were associated with relative humidity (RH) levels from all indoor <span class="hlt">water</span> vapor sources that were less than the EPA-recommended 70% RH maximum to avoid active mold and mildew growth. The conclusion of this investigation is that when installed in accordance with the manufacturer's instructions, vent-free <span class="hlt">gas</span> heating appliances maintain acceptable indoor <span class="hlt">air</span> quality in tight energy-efficient homes, as defined by the standards referenced in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25642383','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25642383"><span>Unlocking Chain <span class="hlt">Exchange</span> in Highly Amphiphilic Block Polymer Micellar Systems: Influence of Agitation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Murphy, Ryan P; Kelley, Elizabeth G; Rogers, Simon A; Sullivan, Millicent O; Epps, Thomas H</p> <p>2014-11-18</p> <p>Chain <span class="hlt">exchange</span> between block polymer micelles in highly selective solvents, such as <span class="hlt">water</span>, is well-known to be arrested under quiescent conditions, yet this work demonstrates that simple agitation methods can induce rapid chain <span class="hlt">exchange</span> in these solvents. Aqueous solutions containing either pure poly(butadiene- b -ethylene oxide) or pure poly(butadiene- b -ethylene oxide- d 4 ) micelles were combined and then subjected to agitation by vortex mixing, concentric cylinder Couette flow, or nitrogen <span class="hlt">gas</span> sparging. Subsequently, the extent of chain <span class="hlt">exchange</span> between micelles was quantified using small angle neutron scattering. Rapid vortex mixing induced chain <span class="hlt">exchange</span> within minutes, as evidenced by a monotonic decrease in scattered intensity, whereas Couette flow and sparging did not lead to measurable chain <span class="hlt">exchange</span> over the examined time scale of hours. The linear kinetics with respect to agitation time suggested a surface-limited <span class="hlt">exchange</span> process at the <span class="hlt">air-water</span> interface. These findings demonstrate the strong influence of processing conditions on block polymer solution assemblies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28833173','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28833173"><span>Disruption of stomatal lineage signaling or transcriptional regulators has differential effects on mesophyll development, but maintains coordination of <span class="hlt">gas</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dow, Graham J; Berry, Joseph A; Bergmann, Dominique C</p> <p>2017-10-01</p> <p>Stomata are simultaneously tasked with permitting the uptake of carbon dioxide for photosynthesis while limiting <span class="hlt">water</span> loss from the plant. This process is mainly regulated by guard cell control of the stomatal aperture, but recent advancements have highlighted the importance of several genes that control stomatal development. Using targeted genetic manipulations of the stomatal lineage and a combination of <span class="hlt">gas</span> <span class="hlt">exchange</span> and microscopy techniques, we show that changes in stomatal development of the epidermal layer lead to coupled changes in the underlying mesophyll tissues. This coordinated response tends to match leaf photosynthetic potential (V cmax ) with <span class="hlt">gas-exchange</span> capacity (g smax ), and hence the uptake of carbon dioxide for <span class="hlt">water</span> lost. We found that different genetic regulators systematically altered tissue coordination in separate ways: the transcription factor SPEECHLESS (SPCH) primarily affected leaf size and thickness, whereas peptides in the EPIDERMAL PATTERNING FACTOR (EPF) family altered cell density in the mesophyll. It was also determined that interlayer coordination required the cell-surface receptor TOO MANY MOUTHS (TMM). These results demonstrate that stomata-specific regulators can alter mesophyll properties, which provides insight into how molecular pathways can organize leaf tissues to coordinate <span class="hlt">gas</span> <span class="hlt">exchange</span> and suggests new strategies for improving plant <span class="hlt">water</span>-use efficiency. © 2017 The Authors. New Phytologist © 2017 New Phytologist Trust.</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('http://adsabs.harvard.edu/abs/2006GeoRL..3315404D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006GeoRL..3315404D"><span>Circadian rhythms constrain leaf and canopy <span class="hlt">gas</span> <span class="hlt">exchange</span> in an Amazonian forest</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Doughty, Christopher E.; Goulden, Michael L.; Miller, Scott D.; da Rocha, Humberto R.</p> <p>2006-08-01</p> <p>We used a controlled-environment leaf <span class="hlt">gas-exchange</span> system and the micrometeorological technique eddy covariance to determine whether circadian rhythms constrain the rates of leaf and canopy <span class="hlt">gas</span> <span class="hlt">exchange</span> in an Amazonian forest over a day. When exposed to continuous and constant light for 20 to 48 hours leaves of eleven of seventeen species reduced their photosynthetic rates and closed their stomata during the normally dark period and resumed active <span class="hlt">gas</span> <span class="hlt">exchange</span> during the normally light period. Similarly, the rate of whole-forest CO2 uptake at a predetermined irradiance declined during the late afternoon and early morning and increased during the middle of the day. We attribute these cycles to circadian rhythms that are analogous to ones that have been reported for herbaceous plants in the laboratory. The importance of endogenous <span class="hlt">gas</span> <span class="hlt">exchange</span> rhythms presents a previously unrecognized challenge for efforts to both interpret and model land-atmosphere energy and mass <span class="hlt">exchange</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26991124','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26991124"><span>Optimal allocation of leaf epidermal area for <span class="hlt">gas</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>de Boer, Hugo J; Price, Charles A; Wagner-Cremer, Friederike; Dekker, Stefan C; Franks, Peter J; Veneklaas, Erik J</p> <p>2016-06-01</p> <p>A long-standing research focus in phytology has been to understand how plants allocate leaf epidermal space to stomata in order to achieve an economic balance between the plant's carbon needs and <span class="hlt">water</span> use. Here, we present a quantitative theoretical framework to predict allometric relationships between morphological stomatal traits in relation to leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and the required allocation of epidermal area to stomata. Our theoretical framework was derived from first principles of diffusion and geometry based on the hypothesis that selection for higher anatomical maximum stomatal conductance (gsmax ) involves a trade-off to minimize the fraction of the epidermis that is allocated to stomata. Predicted allometric relationships between stomatal traits were tested with a comprehensive compilation of published and unpublished data on 1057 species from all major clades. In support of our theoretical framework, stomatal traits of this phylogenetically diverse sample reflect spatially optimal allometry that minimizes investment in the allocation of epidermal area when plants evolve towards higher gsmax . Our results specifically highlight that the stomatal morphology of angiosperms evolved along spatially optimal allometric relationships. We propose that the resulting wide range of viable stomatal trait combinations equips angiosperms with developmental and evolutionary flexibility in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> unrivalled by gymnosperms and pteridophytes. © 2016 The Authors New Phytologist © 2016 New Phytologist Trust.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16443233','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16443233"><span>Determination of ethephon residues in <span class="hlt">water</span> by <span class="hlt">gas</span> chromatography with cubic mass spectrometry after ion-<span class="hlt">exchange</span> purification and derivatisation with N-(tert-butyldimethylsilyl)-N-methyltrifluoroacetamide.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Royer, A; Laporte, F; Bouchonnet, S; Communal, P-Y</p> <p>2006-03-03</p> <p>An analytical method has been developed for the determination of residues of ethephon (2-chloroethyl phosphonic acid) in drinking and surface <span class="hlt">water</span>. The procedure is based on de-ionisation with an anion/cation-<span class="hlt">exchange</span> resin, solid phase extraction by means of anion-<span class="hlt">exchange</span> polystyrene-divinylbenzene extraction disks, elution with a mixture of methanol and 10 M hydrochloric acid (98/2, v/v), redisolution into acetonitrile after evaporation and silylation with N-(tert-butyldimethylsilyl)-N-methyltrifluoroacetamide (MTBSTFA). Quantification is performed by <span class="hlt">gas</span> chromatography with ion-trap cubic mass spectrometric detection in the electron impact mode (GC-EI-MS3). Method validation was conducted using samples of mineral, tap, and river <span class="hlt">water</span> that were fortified with ethephon at concentration levels ranging from 0.1 to 1.0 microg/L. The mean recovery from all the fortified samples (n = 36) amounted to 88% with a relative standard deviation of 17%. The method, therefore, was shown to allow accurate determination of ethephon residues in drinking and surface <span class="hlt">water</span> with a limit of quantification of 0.1 microg/L.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20030093589','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20030093589"><span>Membrane-Based <span class="hlt">Gas</span> Traps for Ammonia, Freon-21, and <span class="hlt">Water</span> Systems to Simplify Ground Processing</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ritchie, Stephen M. C.</p> <p>2003-01-01</p> <p><span class="hlt">Gas</span> traps are critical for the smooth operation of coolant loops because <span class="hlt">gas</span> bubbles can cause loss of centrifugal pump prime, interference with sensor readings, inhibition of heat transfer, and blockage of passages to remote systems. Coolant loops are ubiquitous in space flight hardware, and thus there is a great need for this technology. Conventional <span class="hlt">gas</span> traps will not function in micro-gravity due to the absence of buoyancy forces. Therefore, clever designs that make use of adhesion and momentum are required for adequate separation, preferable in a single pass. The <span class="hlt">gas</span> traps currently used in <span class="hlt">water</span> coolant loops on the International Space Station are composed of membrane tube sets in a shell. Each tube set is composed of a hydrophilic membrane (used for <span class="hlt">water</span> transport and capture of bubbles) and a hydrophobic membrane (used for venting of <span class="hlt">air</span> bubbles). For the hydrophilic membrane, there are two critical pressures, the pressure drop and the bubble pressure. The pressure drop is the decrease in system pressure across the <span class="hlt">gas</span> trap. The bubble pressure is the pressure required for <span class="hlt">air</span> bubbles to pass across the <span class="hlt">water</span> filled membrane. A significant difference between these pressures is needed to ensure complete capture of <span class="hlt">air</span> bubbles in a single pass. Bubbles trapped by the device adsorb on the hydrophobic membrane in the interior of the hydrophilic membrane tube. After adsorption, the <span class="hlt">air</span> is vented due to a pressure drop of approximately 1 atmosphere across the membrane. For <span class="hlt">water</span> systems, the <span class="hlt">air</span> is vented to the ambient (cabin). Because <span class="hlt">water</span> vapor can also transport across the hydrophobic membrane, it is critical that a minimum surface area is used to avoid excessive <span class="hlt">water</span> loss (would like to have a closed loop for the coolant). The currently used <span class="hlt">gas</span> traps only provide a difference in pressure drop and bubble pressure of 3-4 psid. This makes the <span class="hlt">gas</span> traps susceptible to failure at high bubble loading and if <span class="hlt">gas</span> venting is impaired. One mechanism for the latter</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1714679M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1714679M"><span>Carbon speciation at the <span class="hlt">air</span>-sea interface during rain</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McGillis, Wade; Hsueh, Diana; Takeshita, Yui; Donham, Emily; Markowitz, Michele; Turk, Daniela; Martz, Todd; Price, Nicole; Langdon, Chris; Najjar, Raymond; Herrmann, Maria; Sutton, Adrienne; Loose, Brice; Paine, Julia; Zappa, Christopher</p> <p>2015-04-01</p> <p>This investigation demonstrates the surface ocean dilution during rain events on the ocean and quantifies the lowering of surface pCO2 affecting the <span class="hlt">air</span>-sea <span class="hlt">exchange</span> of carbon dioxide. Surface salinity was measured during rain events in Puerto Rico, the Florida Keys, East Coast USA, Panama, and the Palmyra Atoll. End-member analysis is used to determine the subsequent surface ocean carbonate speciation. Surface ocean carbonate chemistry was measured during rain events to verify any approximations made. The physical processes during rain (cold, fresh <span class="hlt">water</span> intrusion and buoyancy, surface waves and shear, microscale mixing) are described. The role of rain on surface mixing, biogeochemistry, and <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15277564','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15277564"><span>Effect of <span class="hlt">water</span> depth and <span class="hlt">water</span> velocity upon the surfacing frequency of the bimodally respiring freshwater turtle, Rheodytes leukops.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gordos, Matthew A; Franklin, Craig E; Limpus, Colin J</p> <p>2004-08-01</p> <p>This study examines the effect of increasing <span class="hlt">water</span> depth and <span class="hlt">water</span> velocity upon the surfacing behaviour of the bimodally respiring turtle, Rheodytes leukops. Surfacing frequency was recorded for R. leukops at varying <span class="hlt">water</span> depths (50, 100, 150 cm) and <span class="hlt">water</span> velocities (5, 15, 30 cm s(-1)) during independent trials to provide an indirect cost-benefit analysis of aquatic versus pulmonary respiration. With increasing <span class="hlt">water</span> velocity, R. leukops decreased its surfacing frequency twentyfold, thus suggesting a heightened reliance upon aquatic <span class="hlt">gas</span> <span class="hlt">exchange</span>. An elevated reliance upon aquatic respiration, which presumably translates into a decreased <span class="hlt">air</span>-breathing frequency, may be metabolically more efficient for R. leukops compared to the expenditure (i.e. time and energy) associated with <span class="hlt">air</span>-breathing within fast-flowing riffle zones. Additionally, R. leukops at higher <span class="hlt">water</span> velocities preferentially selected low-velocity microhabitats, presumably to avoid the metabolic expenditure associated with high <span class="hlt">water</span> flow. Alternatively, increasing <span class="hlt">water</span> depth had no effect upon the surfacing frequency of R. leukops, suggesting little to no change in the respiratory partitioning of the species across treatment settings. Routinely long dives (>90 min) recorded for R. leukops indicate a high reliance upon aquatic O2 uptake regardless of <span class="hlt">water</span> depth. Moreover, metabolic and temporal costs attributed to pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span> within a pool-like environment were likely minimal for R. leukops, irrespective of <span class="hlt">water</span> depth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28406724','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28406724"><span>Fifty Years of Research in ARDS. <span class="hlt">Gas</span> <span class="hlt">Exchange</span> in Acute Respiratory Distress Syndrome.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Radermacher, Peter; Maggiore, Salvatore Maurizio; Mercat, Alain</p> <p>2017-10-15</p> <p>Acute respiratory distress syndrome (ARDS) is characterized by severe impairment of <span class="hlt">gas</span> <span class="hlt">exchange</span>. Hypoxemia is mainly due to intrapulmonary shunt, whereas increased alveolar dead space explains the alteration of CO 2 clearance. Assessment of the severity of <span class="hlt">gas</span> <span class="hlt">exchange</span> impairment is a requisite for the characterization of the syndrome and the evaluation of its severity. Confounding factors linked to hemodynamic status can greatly influence the relationship between the severity of lung injury and the degree of hypoxemia and/or the effects of ventilator settings on <span class="hlt">gas</span> <span class="hlt">exchange</span>. Apart from situations of rescue treatment, targeting optimal <span class="hlt">gas</span> <span class="hlt">exchange</span> in ARDS has become less of a priority compared with prevention of injury. A complex question for clinicians is to understand when improvement in oxygenation and alveolar ventilation is related to a lower degree or risk of injury for the lungs. In this regard, a full understanding of <span class="hlt">gas</span> <span class="hlt">exchange</span> mechanism in ARDS is imperative for individualized symptomatic support of patients with ARDS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.B51K..07P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.B51K..07P"><span>The role of hydrodynamic transport in greenhouse <span class="hlt">gas</span> fluxes at a wetland with emergent vegetation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Poindexter, C.; Gilson, E.; Knox, S. H.; Matthes, J. H.; Verfaillie, J. G.; Baldocchi, D. D.; Variano, E. A.</p> <p>2013-12-01</p> <p>In wetlands with emergent vegetation, the hydrodynamic transport of dissolved gases is often neglected because emergent plants transport gases directly and limit wind-driven <span class="hlt">air-water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span> by sheltering the <span class="hlt">water</span> surface. Nevertheless, wetland hydrodynamics, and thermally-driven stirring in particular, have the potential to impact <span class="hlt">gas</span> fluxes in these environments. We are evaluating the importance of hydrodynamic dissolved <span class="hlt">gas</span> transport at a re-established marsh on Twitchell Island in the Sacramento-San Joaquin Delta (California, USA). At this marsh, the U.S. Geological Survey has previously observed rapid accumulation of organic material (carbon sequestration) as well as very high methane emissions. To assess the role of hydrodynamics in the marsh's greenhouse <span class="hlt">gas</span> fluxes, we measured dissolved carbon dioxide and methane in the <span class="hlt">water</span> column on a bi-weekly basis beginning in July 2012. We employed a model for <span class="hlt">air-water</span> <span class="hlt">gas</span> fluxes in wetlands with emergent vegetation that predicts <span class="hlt">gas</span> transfer velocities from meteorological conditions. Modeled <span class="hlt">air-water</span> <span class="hlt">gas</span> fluxes were compared with net <span class="hlt">gas</span> fluxes measured at the marsh via the eddy covariance technique. This comparison revealed that hydrodynamic transport due to thermal convection was responsible for approximately one third of net carbon dioxide and methane fluxes. The cooling at the <span class="hlt">water</span> surface driving thermal convection occurred each night and was most pronounced during the warmest months of the year. These finding have implications for the prediction and management of greenhouse <span class="hlt">gas</span> fluxes at re-established marshes in the Sacramento-San Joaquin Delta and other similar wetlands.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=315010&Lab=NHEERL&keyword=smith&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=315010&Lab=NHEERL&keyword=smith&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>A dynamic leaf <span class="hlt">gas-exchange</span> strategy is conserved in woody plants under changing ambient CO2: evidence from carbon isotope discrimination in paleo and CO2 enrichment studies</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>Rising atmospheric [CO2], ca, is expected to affect stomatal regulation of leaf <span class="hlt">gas-exchange</span> of woody plants, thus influencing energy fluxes as well as carbon (C), <span class="hlt">water</span> and nutrient cycling of forests. Researchers have reported that stomata regulate leaf <span class="hlt">gas-exchange</span> around “set...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28337628','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28337628"><span>Study of plasma off-<span class="hlt">gas</span> treatment from spent ion <span class="hlt">exchange</span> resin pyrolysis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Castro, Hernán Ariel; Luca, Vittorio; Bianchi, Hugo Luis</p> <p>2017-03-23</p> <p>Polystyrene divinylbenzene-based ion <span class="hlt">exchange</span> resins are employed extensively within nuclear power plants (NPPs) and research reactors for purification and chemical control of the cooling <span class="hlt">water</span> system. To maintain the highest possible <span class="hlt">water</span> quality, the resins are regularly replaced as they become contaminated with a range of isotopes derived from compromised fuel elements as well as corrosion and activation products including 14 C, 60 Co, 90 Sr, 129 I, and 137 Cs. Such spent resins constitute a major proportion (in volume terms) of the solid radioactive waste generated by the nuclear industry. Several treatment and conditioning techniques have been developed with a view toward reducing the spent resin volume and generating a stable waste product suitable for long-term storage and disposal. Between them, pyrolysis emerges as an attractive option. Previous work of our group suggests that the pyrolysis treatment of the resins at low temperatures between 300 and 350 °C resulted in a stable waste product with a significant volume reduction (>50%) and characteristics suitable for long-term storage and/or disposal. However, another important issue to take into account is the complexity of the off-<span class="hlt">gas</span> generated during the process and the different technical alternatives for its conditioning. Ongoing work addresses the characterization of the ion <span class="hlt">exchange</span> resin treatment's off-<span class="hlt">gas</span>. Additionally, the application of plasma technology for the treatment of the off-<span class="hlt">gas</span> current was studied as an alternative to more conventional processes utilizing oil- or <span class="hlt">gas</span>-fired post-combustion chambers operating at temperatures in excess of 1000 °C. A laboratory-scale flow reactor, using inductively coupled plasma, operating under sub-atmospheric conditions was developed. Fundamental experiments using model compounds have been performed, demonstrating a high destruction and removal ratio (>99.99%) for different reaction media, at low reactor temperatures and moderate power consumption</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017DSRI..122...17M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017DSRI..122...17M"><span>The <span class="hlt">air</span>-sea <span class="hlt">exchange</span> of mercury in the low latitude Pacific and Atlantic Oceans</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mason, Robert P.; Hammerschmidt, Chad R.; Lamborg, Carl H.; Bowman, Katlin L.; Swarr, Gretchen J.; Shelley, Rachel U.</p> <p>2017-04-01</p> <p><span class="hlt">Air</span>-sea <span class="hlt">exchange</span> is an important component of the global mercury (Hg) cycle as it mediates the rate of increase in ocean Hg, and therefore the rate of change in levels of methylmercury (MeHg), the most toxic and bioaccumulative form of Hg in seafood and the driver of human health concerns. <span class="hlt">Gas</span> evasion of elemental Hg (Hg0) from the ocean is an important sink for ocean Hg with previous studies suggesting that evasion is not uniform across ocean basins. To understand further the factors controlling Hg0 evasion, and its relationship to atmospheric Hg deposition, we made measurements of dissolved Hg0 (DHg0) in surface <span class="hlt">waters</span>, along with measurements of Hg in precipitation and on aerosols, and Hg0 in marine <span class="hlt">air</span>, during two GEOTRACES cruises; GP16 in the equatorial South Pacific and GA03 in the North Atlantic. We contrast the concentrations and estimated evasion fluxes of Hg0 during these cruises, and the factors influencing this <span class="hlt">exchange</span>. Concentrations of DHg0 and fluxes were lower during the GP16 cruise than during the GA03 cruise, and likely reflect the lower atmospheric deposition in the South Pacific. An examination of Hg/Al ratios for aerosols from the cruises suggests that they were anthropogenically-enriched relative to crustal material, although to a lesser degree for the South Pacific than the aerosols over the North Atlantic. Both regions appear to be net sources of Hg0 to the atmosphere (evasion>deposition) and the reasons for this are discussed. Overall, the studies reported here provide further clarification on the factors controlling evasion of Hg0 from the ocean surface, and the role of anthropogenic inputs in influencing ocean Hg concentrations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997AdSpR..20.2045.','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997AdSpR..20.2045."><span>Element <span class="hlt">exchange</span> in a <span class="hlt">water</span>-and <span class="hlt">gas</span>-closed biological life support system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p></p> <p>1997-01-01</p> <p>Liquid human wastes and household <span class="hlt">water</span> used for nutrition of wheat made possible to realize 24% closure for the mineral <span class="hlt">exchange</span> in an experiment with a 2-component version of ``Bios-3'' life support system (LSS) Input-output balances of revealed, that elements (primarily trace elements) within the system. The structural materials (steel, titanium), expanded clay aggregate, and catalytic furnace catalysts. By the end of experiment, the permanent nutrient solution, plants, and the human diet gradually built up Ni, Cr, Al, Fe, V, Zn, Cu, and Mo. Thorough selection and pretreatment of materials can substantially reduce this accumulation. To enhance closure of the mineral <span class="hlt">exchange</span> involves processing of human- metabolic wastes and inedible biomes inside LSS. An efficient method to oxidize wastes by hydrogen peroxide in a quartz reactor at the temperature of 80°C controlled electromagnetic field is proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997AdSpR..20.2045G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997AdSpR..20.2045G"><span>Element <span class="hlt">exchange</span> in a <span class="hlt">water</span>-and <span class="hlt">gas</span>-closed biological life support system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gribovskaya, I. V.; Kudenko, Yu. A.; Gitelson, J. I.</p> <p>1997-01-01</p> <p>Liquid human wastes and household <span class="hlt">water</span> used for nutrition of wheat made possible to realize 24% closure for the mineral <span class="hlt">exchange</span> in an experiment with a 2-component version of ``Bios-3'' life support system (LSS) Input-output balances of revealed, that elements (primarily trace elements) within the system. The structural materials (steel, titanium), expanded clay aggregate, and catalytic furnace catalysts. By the end of experiment, the permanent nutrient solution, plants, and the human diet gradually built up Ni, Cr, Al, Fe, V, Zn, Cu, and Mo. Thorough selection and pretreatment of materials can substantially reduce this accumulation. To enhance closure of the mineral <span class="hlt">exchange</span> involves processing of human- metabolic wastes and inedible biomes inside LSS. An efficient method to oxidize wastes by hydrogen peroxide in a quartz reactor at the temperature of 80 degC controlled electromagnetic field is proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25412353','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25412353"><span>Polycyclic aromatic hydrocarbon (PAH) and oxygenated PAH (OPAH) <span class="hlt">air-water</span> <span class="hlt">exchange</span> during the deepwater horizon oil spill.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tidwell, Lane G; Allan, Sarah E; O'Connell, Steven G; Hobbie, Kevin A; Smith, Brian W; Anderson, Kim A</p> <p>2015-01-06</p> <p>Passive sampling devices were used to measure <span class="hlt">air</span> vapor and <span class="hlt">water</span> dissolved phase concentrations of 33 polycyclic aromatic hydrocarbons (PAHs) and 22 oxygenated PAHs (OPAHs) at four Gulf of Mexico coastal sites prior to, during, and after shoreline oiling from the Deepwater Horizon oil spill (DWH). Measurements were taken at each site over a 13 month period, and flux across the <span class="hlt">water-air</span> boundary was determined. This is the first report of vapor phase and flux of both PAHs and OPAHs during the DWH. Vapor phase sum PAH and OPAH concentrations ranged between 1 and 24 ng/m(3) and 0.3 and 27 ng/m(3), respectively. PAH and OPAH concentrations in <span class="hlt">air</span> exhibited different spatial and temporal trends than in <span class="hlt">water</span>, and <span class="hlt">air-water</span> flux of 13 individual PAHs were strongly associated with the DWH incident. The largest PAH volatilizations occurred at the sites in Alabama and Mississippi in the summer, each nominally 10,000 ng/m(2)/day. Acenaphthene was the PAH with the highest observed volatilization rate of 6800 ng/m(2)/day in September 2010. This work represents additional evidence of the DWH incident contributing to <span class="hlt">air</span> contamination, and provides one of the first quantitative <span class="hlt">air-water</span> chemical flux determinations with passive sampling technology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=20677&Lab=NCEA&keyword=gas+AND+liquid&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=20677&Lab=NCEA&keyword=gas+AND+liquid&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>Volatilization Rates from <span class="hlt">Water</span> to Indoor <span class="hlt">Air</span> Phase II</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>Contaminated <span class="hlt">water</span> can lead to volatilization of chemicals to residential indoor <span class="hlt">air</span>. Previous research has focused on only one source (shower stalls) and has been limited to chemicals in which <span class="hlt">gas</span>-phase resistance to mass transfer is of marginal significance. As a result, attemp...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17330473','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17330473"><span>[<span class="hlt">Gas</span> <span class="hlt">exchange</span> features of Ambrosia artemisiifolia leaves and fruits and their correlations with soil heavy metals].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zu, Yuangang; Wang, Wenjie; Chen, Huafeng; Yang, Fengjian; Zhang, Zhonghua</p> <p>2006-12-01</p> <p>Ambrosia artemisiifolia can survive well in the habitats of heavy human disturbance and partial soil pollution. Weather its photosynthetic features benefit their survival is worthwhile to concern. With a refuse dump in Changchun City (43 degrees 50'N, 125 degrees 23'E) as study site, this paper analyzed the contents of soil Cu, Pb, Zn, Mn, Cr, Co, Ni, Cd, As, Sb and Hg at ten plots, and measured in situ the <span class="hlt">gas</span> <span class="hlt">exchange</span> in A. artemisiifolia leaves and young fruits. The results showed that the study site was slightly contaminated by Ni, but the contents of other soil heavy metals were approached to or substantially lower than their threshold values. The net photosynthetic rate of leaves ranged from 1.88 to 9.41 micromol x m(-2) x s(-1), while that of young fruits could be up to 2. 81 micromol x m(-2) s(-1). Averagely, the respiration rate, stomatal conductance, photosynthetic rate, and <span class="hlt">water</span> utilization efficiency of leaves were 1.81 micromol x m(-2) x s(-1), 75.7 mmol x m(-2) x s(-1), 6.05 micromol x m(-2) x s(-1), and 4.72 micromol CO2 x mmol(-1) H2O, being 5.26, 0.64, 1.31 and 1.69 times as much as those of young fruits, respectively, indicating that the respiratory and photosynthetic capacities and <span class="hlt">water</span> use efficiency of A. artemisiifolia young fruits were equivalent to or higher than those of its leaves. Many test heavy metals, such as Cu, Pb, Zn, Cd, As, Sb and Hg, had no significant effects on the <span class="hlt">gas</span> <span class="hlt">exchange</span> features of leaves and fruits, but there were significant correlations of Ni and Cr with the stomatal conductance and <span class="hlt">water</span> use efficiency of leaves and young fruits, Cr with the gross photosynthesis of leaves, and As with the stomatal conductance of young fruits, suggesting that a majority of test soil heavy metals had no direct effects on the <span class="hlt">gas</span> <span class="hlt">exchange</span> in A. artemisiifolia leaves and fruits, but soil Ni, Cr and As with the contents approached to or substantially lower than the threshold values could affect the <span class="hlt">gas</span> <span class="hlt">exchange</span> features of A</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870051492&hterms=water+cycle&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dwater%2Bcycle','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870051492&hterms=water+cycle&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dwater%2Bcycle"><span><span class="hlt">Air</span> Evaporation closed cycle <span class="hlt">water</span> recovery technology - Advanced energy saving designs</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Morasko, Gwyndolyn; Putnam, David F.; Bagdigian, Robert</p> <p>1986-01-01</p> <p>The <span class="hlt">Air</span> Evaporation <span class="hlt">water</span> recovery system is a visible candidate for Space Station application. A four-man <span class="hlt">Air</span> Evaporation open cycle system has been successfully demonstrated for waste <span class="hlt">water</span> recovery in manned chamber tests. The design improvements described in this paper greatly enhance the system operation and energy efficiency of the <span class="hlt">air</span> evaporation process. A state-of-the-art wick feed design which results in reduced logistics requirements is presented. In addition, several design concepts that incorporate regenerative features to minimize the energy input to the system are discussed. These include a recuperative heat <span class="hlt">exchanger</span>, a heat pump for energy transfer to the <span class="hlt">air</span> heater, and solar collectors for evaporative heat. The addition of the energy recovery devices will result in an energy reduction of more than 80 percent over the systems used in earlier manned chamber tests.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1212391','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1212391"><span>Metal-<span class="hlt">air</span> cell with ion <span class="hlt">exchange</span> material</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Friesen, Cody A.; Wolfe, Derek; Johnson, Paul Bryan</p> <p>2015-08-25</p> <p>Embodiments of the invention are related to anion <span class="hlt">exchange</span> membranes used in electrochemical metal-<span class="hlt">air</span> cells in which the membranes function as the electrolyte material, or are used in conjunction with electrolytes such as ionic liquid electrolytes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880002876','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880002876"><span>Sunlight supply and <span class="hlt">gas</span> <span class="hlt">exchange</span> systems in microalgal bioreactor</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mori, K.; Ohya, H.; Matsumoto, K.; Furune, H.</p> <p>1987-01-01</p> <p>The bioreactor with sunlight supply system and <span class="hlt">gas</span> <span class="hlt">exchange</span> systems presented has proved feasible in ground tests and shows much promise for space use as a closed ecological life support system device. The chief conclusions concerning the specification of total system needed for a life support system for a man in a space station are the following: (1) Sunlight supply system - compactness and low electrical consumption; (2) Bioreactor system - high density and growth rate of chlorella; and (3) <span class="hlt">Gas</span> <span class="hlt">exchange</span> system - enough for O2 production and CO2 assimilation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015A%26A...584A..98F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015A%26A...584A..98F"><span>Hydrogen isotope <span class="hlt">exchanges</span> between <span class="hlt">water</span> and methanol in interstellar ices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Faure, A.; Faure, M.; Theulé, P.; Quirico, E.; Schmitt, B.</p> <p>2015-12-01</p> <p>The deuterium fractionation of <span class="hlt">gas</span>-phase molecules in hot cores is believed to reflect the composition of interstellar ices. The deuteration of methanol is a major puzzle, however, because the isotopologue ratio [CH2DOH]/[CH3OD], which is predicted to be equal to 3 by standard grain chemistry models, is much larger (~20) in low-mass hot corinos and significantly lower (~1) in high-mass hot cores. This dichotomy in methanol deuteration between low-mass and massive protostars is currently not understood. In this study, we report a simplified rate equation model of the deuterium chemistry occurring in the icy mantles of interstellar grains. We apply this model to the chemistry of hot corinos and hot cores, with IRAS 16293-2422 and the Orion KL Compact Ridge as prototypes, respectively. The chemistry is based on a statistical initial deuteration at low temperature followed by a warm-up phase during which thermal hydrogen/deuterium (H/D) <span class="hlt">exchanges</span> occur between <span class="hlt">water</span> and methanol. The <span class="hlt">exchange</span> kinetics is incorporated using laboratory data. The [CH2DOH]/[CH3OD] ratio is found to scale inversely with the D/H ratio of <span class="hlt">water</span>, owing to the H/D <span class="hlt">exchange</span> equilibrium between the hydroxyl (-OH) functional groups of methanol and <span class="hlt">water</span>. Our model is able to reproduce the observed [CH2DOH]/[CH3OD] ratios provided that the primitive fractionation of <span class="hlt">water</span> ice [HDO]/[H2O] is ~2% in IRAS 16293-2422 and ~0.6% in Orion KL. We conclude that the molecular D/H ratios measured in hot cores may not be representative of the original mantles because molecules with <span class="hlt">exchangeable</span> deuterium atoms can equilibrate with <span class="hlt">water</span> ice during the warm-up phase.</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://www.ncbi.nlm.nih.gov/pubmed/22145748','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22145748"><span>Selective permeation of moisture and VOCs through polymer membranes used in total heat <span class="hlt">exchangers</span> for indoor <span class="hlt">air</span> ventilation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, L-Z; Zhang, X-R; Miao, Q-Z; Pei, L-X</p> <p>2012-08-01</p> <p>Fresh <span class="hlt">air</span> ventilation is central to indoor environmental control. Total heat <span class="hlt">exchangers</span> can be key equipment for energy conservation in ventilation. Membranes have been used for total heat <span class="hlt">exchangers</span> for more than a decade. Much effort has been spent to achieve <span class="hlt">water</span> vapor permeability of various membranes; however, relatively little attention has been paid to the selectivity of moisture compared with volatile organic compounds (VOCs) through such membranes. In this investigation, the most commonly used membranes, both hydrophilic and hydrophobic ones, are tested for their permeability for moisture and five VOCs (acetic acid, formaldehyde, acetaldehyde, toluene, and ethane). The selectivity of moisture vs. VOCs in these membranes is then evaluated. With a solution-diffusion model, the solubility and diffusivity of moisture and VOCs in these membranes are calculated. The resulting data could provide some reference for future material selection. Total heat <span class="hlt">exchangers</span> are important equipment for fresh <span class="hlt">air</span> ventilation with energy conservation. However, their implications for indoor <span class="hlt">air</span> quality in terms of volatile organic compound permeation have not been known. The data in this article help us to clarify the impacts on indoor VOC levels of membrane-based heat <span class="hlt">exchangers</span>. Guidelines for material selection can be obtained for future use total heat <span class="hlt">exchangers</span> for building ventilation. © 2011 John Wiley & Sons A/S.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150020902','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150020902"><span>Technology Candidates for <span class="hlt">Air-to-Air</span> and <span class="hlt">Air</span>-to-Ground Data <span class="hlt">Exchange</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Haynes, Brian D.</p> <p>2015-01-01</p> <p>Technology Candidates for <span class="hlt">Air-to-Air</span> and <span class="hlt">Air</span>-to-Ground Data <span class="hlt">Exchange</span> is a two-year research effort to visualize the U. S. aviation industry at a point 50 years in the future, and to define potential communication solutions to meet those future data <span class="hlt">exchange</span> needs. The research team, led by XCELAR, was tasked with identifying future National Airspace System (NAS) scenarios, determining requirements and functions (including gaps), investigating technical and business issues for <span class="hlt">air</span>, ground, & <span class="hlt">air</span>-to-ground interactions, and reporting on the results. The project was conducted under technical direction from NASA and in collaboration with XCELAR's partner, National Institute of Aerospace, and NASA technical representatives. Parallel efforts were initiated to define the information <span class="hlt">exchange</span> functional needs of the future NAS, and specific communication link technologies to potentially serve those needs. Those efforts converged with the mapping of each identified future NAS function to potential enabling communication solutions; those solutions were then compared with, and ranked relative to, each other on a technical basis in a structured analysis process. The technical solutions emerging from that process were then assessed from a business case perspective to determine their viability from a real-world adoption and deployment standpoint. The results of that analysis produced a proposed set of future solutions and most promising candidate technologies. Gap analyses were conducted at two points in the process, the first examining technical factors, and the second as part of the business case analysis. In each case, no gaps or unmet needs were identified in applying the solutions evaluated to the requirements identified. The future communication solutions identified in the research comprise both specific link technologies and two enabling technologies that apply to most or all specific links. As a result, the research resulted in a new analysis approach, viewing the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890040446&hterms=quality+life&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dquality%2Blife','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890040446&hterms=quality+life&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dquality%2Blife"><span><span class="hlt">Air</span> and <span class="hlt">water</span> quality monitor assessment of life support subsystems</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Whitley, Ken; Carrasquillo, Robyn L.; Holder, D.; Humphries, R.</p> <p>1988-01-01</p> <p>Preprotype <span class="hlt">air</span> revitalization and <span class="hlt">water</span> reclamation subsystems (Mole Sieve, Sabatier, Static Feed Electrolyzer, Trace Contaminant Control, and Thermoelectric Integrated Membrane Evaporative Subsystem) were operated and tested independently and in an integrated arrangement. During each test, <span class="hlt">water</span> and/or <span class="hlt">gas</span> samples were taken from each subsystem so that overall subsystem performance could be determined. The overall test design and objectives for both subsystem and integrated subsystem tests were limited, and no effort was made to meet <span class="hlt">water</span> or <span class="hlt">gas</span> specifications. The results of chemical analyses for each of the participating subsystems are presented along with other selected samples which were analyzed for physical properties and microbiologicals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010BGeo....7..121A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010BGeo....7..121A"><span>From biota to chemistry and climate: towards a comprehensive description of trace <span class="hlt">gas</span> <span class="hlt">exchange</span> between the biosphere and atmosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arneth, A.; Sitch, S.; Bondeau, A.; Butterbach-Bahl, K.; Foster, P.; Gedney, N.; de Noblet-Ducoudré, N.; Prentice, I. C.; Sanderson, M.; Thonicke, K.; Wania, R.; Zaehle, S.</p> <p>2010-01-01</p> <p><span class="hlt">Exchange</span> of non-CO2 trace gases between the land surface and the atmosphere plays an important role in atmospheric chemistry and climate. Recent studies have highlighted its importance for interpretation of glacial-interglacial ice-core records, the simulation of the pre-industrial and present atmosphere, and the potential for large climate-chemistry and climate-aerosol feedbacks in the coming century. However, spatial and temporal variations in trace <span class="hlt">gas</span> emissions and the magnitude of future feedbacks are a major source of uncertainty in atmospheric chemistry, <span class="hlt">air</span> quality and climate science. To reduce such uncertainties Dynamic Global Vegetation Models (DGVMs) are currently being expanded to mechanistically represent processes relevant to non-CO2 trace <span class="hlt">gas</span> <span class="hlt">exchange</span> between land biota and the atmosphere. In this paper we present a review of important non-CO2 trace <span class="hlt">gas</span> emissions, the state-of-the-art in DGVM modelling of processes regulating these emissions, identify key uncertainties for global scale model applications, and discuss a methodology for model integration and evaluation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009BGD.....6.7717A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009BGD.....6.7717A"><span>From biota to chemistry and climate: towards a comprehensive description of trace <span class="hlt">gas</span> <span class="hlt">exchange</span> between the biosphere and atmosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arneth, A.; Sitch, S.; Bondeau, A.; Butterbach-Bahl, K.; Foster, P.; Gedney, N.; de Noblet-Ducoudré, N.; Prentice, I. C.; Sanderson, M.; Thonicke, K.; Wania, R.; Zaehle, S.</p> <p>2009-07-01</p> <p><span class="hlt">Exchange</span> of non-CO2 trace gases between the land surface and the atmosphere plays an important role in atmospheric chemistry and climate. Recent studies have highlighted its importance for interpretation of glacial-interglacial ice-core records, the simulation of the pre-industrial and present atmosphere, and the potential for large climate-chemistry and climate-aerosol feedbacks in the coming century. However, spatial and temporal variations in trace <span class="hlt">gas</span> emissions and the magnitude of future feedbacks are a major source of uncertainty in atmospheric chemistry, <span class="hlt">air</span> quality and climate science. To reduce such uncertainties Dynamic Global Vegetation Models (DGVMs) are currently being expanded to mechanistically represent processes relevant to non-CO2 trace <span class="hlt">gas</span> <span class="hlt">exchange</span> between land biota and the atmosphere. In this paper we present a review of important non-CO2 trace <span class="hlt">gas</span> emissions, the state-of-the-art in DGVM modelling of processes regulating these emissions, identify key uncertainties for global scale model applications, and discuss a methodology for model integration and evaluation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/19485','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/19485"><span>A two-cell chamber for measuring <span class="hlt">gas</span> <span class="hlt">exchange</span> in tree seedlings</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Keith F. Jensen; Frederick W. Bender; Roberta G. Masters</p> <p>1973-01-01</p> <p>A two-celled chamber for measuring <span class="hlt">gas</span> <span class="hlt">exchange</span> in tree seedlings is described. Temperature is controlled within ± 0.5º C by means of a copper coil. The two cells are independent of one another, and one cell can be used as a preconditioning cell while <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements are being made in the second cell.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/1208301','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/1208301"><span>[Comparative study of respiratory <span class="hlt">exchanging</span> surfaces in birds and mammals].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jammes, Y</p> <p>1975-01-01</p> <p>Anatomical studies of the respiratory apparatus of birds show evidences for a <span class="hlt">gas</span> <span class="hlt">exchanging</span> tubular system (parabronchi and <span class="hlt">air</span> capillaries); these <span class="hlt">exchanging</span> structures are entirely dissociated from the ventilatory drive acting on the <span class="hlt">air</span> sacs. A "cross-current" <span class="hlt">gas</span> <span class="hlt">exchanging</span> system (perpendicular disposition of <span class="hlt">air</span> and blood capillaries) allow a good wash-out of carbon dioxide (PaCO2 lower than PECO2). The great efficiency of this lung is allowed by its very large diffusive surface (ASa) and by the high values of lung specific oxygen diffusing capacity (DO2/ASa) and of O2 extraction coefficient in inspired <span class="hlt">air</span>. The ventilatory pattern of birds is characterized by a greater tidal volume and a smaller respiratory frequency than in mammals of same weight. Respiratory centers of birds receive afferences from lung stretch receptors, CO2-sensitive lung receptors and arterial chemoreceptors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28393872','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28393872"><span><span class="hlt">Water</span> availability drives <span class="hlt">gas</span> <span class="hlt">exchange</span> and growth of trees in northeastern US, not elevated CO2 and reduced acid deposition.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Levesque, Mathieu; Andreu-Hayles, Laia; Pederson, Neil</p> <p>2017-04-10</p> <p>Dynamic global vegetation models (DGVM) exhibit high uncertainty about how climate change, elevated atmospheric CO 2 (atm. CO 2 ) concentration, and atmospheric pollutants will impact carbon sequestration in forested ecosystems. Although the individual roles of these environmental factors on tree growth are understood, analyses examining their simultaneous effects are lacking. We used tree-ring isotopic data and structural equation modeling to examine the concurrent and interacting effects of <span class="hlt">water</span> availability, atm. CO 2 concentration, and SO 4 and nitrogen deposition on two broadleaf tree species in a temperate mesic forest in the northeastern US. <span class="hlt">Water</span> availability was the strongest driver of <span class="hlt">gas</span> <span class="hlt">exchange</span> and tree growth. Wetter conditions since the 1980s have enhanced stomatal conductance, photosynthetic assimilation rates and, to a lesser extent, tree radial growth. Increased <span class="hlt">water</span> availability seemingly overrides responses to reduced acid deposition, CO 2 fertilization, and nitrogen deposition. Our results indicate that <span class="hlt">water</span> availability as a driver of ecosystem productivity in mesic temperate forests is not adequately represented in DGVMs, while CO 2 fertilization is likely overrepresented. This study emphasizes the importance to simultaneously consider interacting climatic and biogeochemical drivers when assessing forest responses to global environmental changes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70185308','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70185308"><span>Numerical evaluation of static-chamber measurements of soil-atmospheric <span class="hlt">gas</span> <span class="hlt">exchange</span>--Identification of physical processes</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Healy, Richard W.; Striegl, Robert G.; Russell, Thomas F.; Hutchinson, Gordon L.; Livingston, Gerald P.</p> <p>1996-01-01</p> <p>The <span class="hlt">exchange</span> of gases between soil and atmosphere is an important process that affects atmospheric chemistry and therefore climate. The static-chamber method is the most commonly used technique for estimating the rate of that <span class="hlt">exchange</span>. We examined the method under hypothetical field conditions where diffusion was the only mechanism for <span class="hlt">gas</span> transport and the atmosphere outside the chamber was maintained at a fixed concentration. Analytical and numerical solutions to the soil <span class="hlt">gas</span> diffusion equation in one and three dimensions demonstrated that <span class="hlt">gas</span> flux density to a static chamber deployed on the soil surface was less in magnitude than the ambient <span class="hlt">exchange</span> rate in the absence of the chamber. This discrepancy, which increased with chamber deployment time and <span class="hlt">air</span>-filled porosity of soil, is attributed to two physical factors: distortion of the soil <span class="hlt">gas</span> concentration gradient (the magnitude was decreased in the vertical component and increased in the radial component) and the slow transport rate of diffusion relative to mixing within the chamber. Instantaneous flux density to a chamber decreased continuously with time; steepest decreases occurred so quickly following deployment and in response to such slight changes in mean chamber headspace concentration that they would likely go undetected by most field procedures. Adverse influences of these factors were reduced by mixing the chamber headspace, minimizing deployment time, maximizing the height and radius of the chamber, and pushing the rim of the chamber into the soil. Nonlinear models were superior to a linear regression model for estimating flux densities from mean headspace concentrations, suggesting that linearity of headspace concentration with time was not necessarily a good indicator of measurement accuracy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22788714','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22788714"><span>Statics and dynamics of free and hydrogen-bonded OH groups at the <span class="hlt">air/water</span> interface.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vila Verde, Ana; Bolhuis, Peter G; Campen, R Kramer</p> <p>2012-08-09</p> <p>We use classical atomistic molecular dynamics simulations of two <span class="hlt">water</span> models (SPC/E and TIP4P/2005) to investigate the orientation and reorientation dynamics of two subpopulations of OH groups belonging to <span class="hlt">water</span> molecules at the <span class="hlt">air/water</span> interface at 300 K: those OH groups that donate a hydrogen bond (called "bonded") and those that do not (called "free"). Free interfacial OH groups reorient in two distinct regimes: a fast regime from 0 to 1 ps and a slow regime thereafter. Qualitatively similar behavior was reported by others for free OH groups near extended hydrophobic surfaces. In contrast, the net reorientation of bonded OH groups occurs at a rate similar to that of bulk <span class="hlt">water</span>. This similarity in reorientation rate results from compensation of two effects: decreasing frequency of hydrogen-bond breaking/formation (i.e., hydrogen-bond <span class="hlt">exchange</span>) and faster rotation of intact hydrogen bonds. Both changes result from the decrease in density at the <span class="hlt">air/water</span> interface relative to the bulk. Interestingly, because of the presence of capillary waves, the slowdown of hydrogen-bond <span class="hlt">exchange</span> is significantly smaller than that reported for <span class="hlt">water</span> near extended hydrophobic surfaces, but it is almost identical to that reported for <span class="hlt">water</span> near small hydrophobic solutes. In this sense <span class="hlt">water</span> at the <span class="hlt">air/water</span> interface has characteristics of <span class="hlt">water</span> of hydration of both small and extended hydrophobic solutes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.epa.gov/watersense/cation-exchange-water-softeners','PESTICIDES'); return false;" href="https://www.epa.gov/watersense/cation-exchange-water-softeners"><span>Cation <span class="hlt">Exchange</span> <span class="hlt">Water</span> Softeners</span></a></p> <p><a target="_blank" href="http://www.epa.gov/pesticides/search.htm">EPA Pesticide Factsheets</a></p> <p></p> <p></p> <p><span class="hlt">Water</span>Sense released a notice of intent to develop a specification for cation <span class="hlt">exchange</span> <span class="hlt">water</span> softeners. The program has made the decision not to move forward with a spec at this time, but is making this information available.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26391334','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26391334"><span>A dynamic leaf <span class="hlt">gas-exchange</span> strategy is conserved in woody plants under changing ambient CO2 : evidence from carbon isotope discrimination in paleo and CO2 enrichment studies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Voelker, Steven L; Brooks, J Renée; Meinzer, Frederick C; Anderson, Rebecca; Bader, Martin K-F; Battipaglia, Giovanna; Becklin, Katie M; Beerling, David; Bert, Didier; Betancourt, Julio L; Dawson, Todd E; Domec, Jean-Christophe; Guyette, Richard P; Körner, Christian; Leavitt, Steven W; Linder, Sune; Marshall, John D; Mildner, Manuel; Ogée, Jérôme; Panyushkina, Irina; Plumpton, Heather J; Pregitzer, Kurt S; Saurer, Matthias; Smith, Andrew R; Siegwolf, Rolf T W; Stambaugh, Michael C; Talhelm, Alan F; Tardif, Jacques C; Van de Water, Peter K; Ward, Joy K; Wingate, Lisa</p> <p>2016-02-01</p> <p>Rising atmospheric [CO2 ], ca , is expected to affect stomatal regulation of leaf <span class="hlt">gas-exchange</span> of woody plants, thus influencing energy fluxes as well as carbon (C), <span class="hlt">water</span>, and nutrient cycling of forests. Researchers have proposed various strategies for stomatal regulation of leaf <span class="hlt">gas-exchange</span> that include maintaining a constant leaf internal [CO2 ], ci , a constant drawdown in CO2 (ca  - ci ), and a constant ci /ca . These strategies can result in drastically different consequences for leaf <span class="hlt">gas-exchange</span>. The accuracy of Earth systems models depends in part on assumptions about generalizable patterns in leaf <span class="hlt">gas-exchange</span> responses to varying ca . The concept of optimal stomatal behavior, exemplified by woody plants shifting along a continuum of these strategies, provides a unifying framework for understanding leaf <span class="hlt">gas-exchange</span> responses to ca . To assess leaf <span class="hlt">gas-exchange</span> regulation strategies, we analyzed patterns in ci inferred from studies reporting C stable isotope ratios (δ(13) C) or photosynthetic discrimination (∆) in woody angiosperms and gymnosperms that grew across a range of ca spanning at least 100 ppm. Our results suggest that much of the ca -induced changes in ci /ca occurred across ca spanning 200 to 400 ppm. These patterns imply that ca  - ci will eventually approach a constant level at high ca because assimilation rates will reach a maximum and stomatal conductance of each species should be constrained to some minimum level. These analyses are not consistent with canalization toward any single strategy, particularly maintaining a constant ci . Rather, the results are consistent with the existence of a broadly conserved pattern of stomatal optimization in woody angiosperms and gymnosperms. This results in trees being profligate <span class="hlt">water</span> users at low ca , when additional <span class="hlt">water</span> loss is small for each unit of C gain, and increasingly <span class="hlt">water</span>-conservative at high ca , when photosystems are saturated and <span class="hlt">water</span> loss is large for each unit C gain. </p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70178114','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70178114"><span>A dynamic leaf <span class="hlt">gas-exchange</span> strategy is conserved in woody plants under changing ambient CO2: evidence from carbon isotope discrimination in paleo and CO2 enrichment studies</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Voelker, Steven L.; Brooks, J. Renée; Meinzer, Frederick C.; Anderson, Rebecca D.; Bader, Martin K.-F.; Battipaglia, Giovanna; Becklin, Katie M.; Beerling, David; Bert, Didier; Betancourt, Julio L.; Dawson, Todd E.; Domec, Jean-Christophe; Guyette, Richard P.; Körner, Christian; Leavitt, Steven W.; Linder, Sune; Marshall, John D.; Mildner, Manuel; Ogée, Jérôme; Panyushkina, Irina P.; Plumpton, Heather J.; Pregitzer, Kurt S.; Saurer, Matthias; Smith, Andrew R.; Siegwolf, Rolf T.W.; Stambaugh, Michael C.; Talhelm, Alan F.; Tardif, Jacques C.; Van De Water, Peter K.; Ward, Joy K.; Wingate, Lisa</p> <p>2016-01-01</p> <p>Rising atmospheric [CO2], ca, is expected to affect stomatal regulation of leaf <span class="hlt">gas-exchange</span> of woody plants, thus influencing energy fluxes as well as carbon (C), <span class="hlt">water</span>, and nutrient cycling of forests. Researchers have proposed various strategies for stomatal regulation of leaf <span class="hlt">gas-exchange</span> that include maintaining a constant leaf internal [CO2], ci, a constant drawdown in CO2(ca − ci), and a constant ci/ca. These strategies can result in drastically different consequences for leaf <span class="hlt">gas-exchange</span>. The accuracy of Earth systems models depends in part on assumptions about generalizable patterns in leaf <span class="hlt">gas-exchange</span> responses to varying ca. The concept of optimal stomatal behavior, exemplified by woody plants shifting along a continuum of these strategies, provides a unifying framework for understanding leaf <span class="hlt">gas-exchange</span> responses to ca. To assess leaf <span class="hlt">gas-exchange</span> regulation strategies, we analyzed patterns in ci inferred from studies reporting C stable isotope ratios (δ13C) or photosynthetic discrimination (∆) in woody angiosperms and gymnosperms that grew across a range of ca spanning at least 100 ppm. Our results suggest that much of the ca-induced changes in ci/ca occurred across ca spanning 200 to 400 ppm. These patterns imply that ca − ci will eventually approach a constant level at high ca because assimilation rates will reach a maximum and stomatal conductance of each species should be constrained to some minimum level. These analyses are not consistent with canalization toward any single strategy, particularly maintaining a constant ci. Rather, the results are consistent with the existence of a broadly conserved pattern of stomatal optimization in woody angiosperms and gymnosperms. This results in trees being profligate <span class="hlt">water</span> users at low ca, when additional <span class="hlt">water</span> loss is small for each unit of C gain, and increasingly <span class="hlt">water</span>-conservative at high ca, when photosystems are saturated and <span class="hlt">water</span> loss is large for each unit C gain.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15266714','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15266714"><span>Variations in <span class="hlt">water</span> status, <span class="hlt">gas</span> <span class="hlt">exchange</span>, and growth in Rosmarinus officinalis plants infected with Glomus deserticola under drought conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sánchez-Blanco, Ma Jesús; Ferrández, Trinitario; Morales, Ma Angeles; Morte, Asunción; Alarcón, Juan José</p> <p>2004-06-01</p> <p>The influence of the arbuscular mycorrhizal fungus Glomus deserticola on the <span class="hlt">water</span> relations, <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters, and vegetative growth of Rosmarinus officinalis plants under <span class="hlt">water</span> stress was studied. Plants were grown with and without the mycorrhizal fungus under glasshouse conditions and subjected to <span class="hlt">water</span> stress by withholding irrigation <span class="hlt">water</span> for 14 days. Along the experimental period, a significant effect of the fungus on the plant growth was observed, and under <span class="hlt">water</span> stress, mycorrhizal plants showed an increase in aerial and root biomass compared to non-mycorrhizal plants. The decrease in the soil <span class="hlt">water</span> potential generated a decrease in leaf <span class="hlt">water</span> potential (psi(l)) and stem <span class="hlt">water</span> potential (psi(x)) of mycorrhizal and non-mycorrhizal plants, with this decrease being lower in mycorrhizal <span class="hlt">water</span>-stressed plants. Mycorrhization also had positive effects on the root hydraulic conductivity (Lp) of <span class="hlt">water</span> stressed plants. Furthermore, mycorrhizal-stressed plants showed a more important decrease in osmotic potential at full turgor (psi(os)) than did non-mycorrhizal-stressed plants, indicating the capacity of osmotic adjustment. Mycorrhizal infection also improved photosynthetic activity (Pn) and stomatal conductance (g(s)) in plants under <span class="hlt">water</span> stress compared to the non-mycorrhizal-stressed plants. A similar behaviour was observed in the photochemical efficiency of PSII (Fv/Fm) with this parameter being lower in non-mycorrhizal plants than in mycorrhizal plants under <span class="hlt">water</span> stress conditions. In the same way, under <span class="hlt">water</span> restriction, mycorrhizal plants showed higher values of chlorophyll content than did non-mycorrhizal plants. Thus, the results obtained indicated that the mycorrhizal symbiosis had a beneficial effect on the <span class="hlt">water</span> status and growth of Rosmarinus officinalis plants under <span class="hlt">water</span>-stress conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EPJWC..6702023D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EPJWC..6702023D"><span>Heat <span class="hlt">exchanger</span> design for hot <span class="hlt">air</span> ericsson-brayton piston engine</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ďurčanský, P.; Lenhard, R.; Jandačka, J.</p> <p>2014-03-01</p> <p>One of the solutions without negative consequences for the increasing energy consumption in the world may be use of alternative energy sources in micro-cogeneration. Currently it is looking for different solutions and there are many possible ways. Cogeneration is known for long time and is widely used. But the installations are often large and the installed output is more suitable for cities or industry companies. When we will speak about decentralization, the small machines have to be used. The article deals with the principle of hot-<span class="hlt">air</span> engines, their use in combined heat and electricity production from biomass and with heat <span class="hlt">exchangers</span> as primary energy transforming element. In the article is hot <span class="hlt">air</span> engine presented as a heat engine that allows the conversion of heat into mechanical energy while heat supply can be external. In the contribution are compared cycles of hot-<span class="hlt">air</span> engine. Then are compared suitable heat <span class="hlt">exchangers</span> for use with hot <span class="hlt">air</span> Ericsson-Brayton engine. In the final part is proposal of heat <span class="hlt">exchanger</span> for use in closed Ericsson-Brayton cycle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21389192','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21389192"><span>The effect of ambient humidity and metabolic rate on the <span class="hlt">gas-exchange</span> pattern of the semi-aquatic insect Aquarius remigis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Contreras, Heidy L; Bradley, Timothy J</p> <p>2011-04-01</p> <p>We have examined the effects of temperature on metabolic rate and respiratory pattern in the <span class="hlt">water</span> strider Aquarius remigis. As temperature was increased from 10 to 30°C, the metabolic rate of the insects increased and the respiratory pattern transitioned from discontinuous, to cyclic, to continuous. The discontinuous <span class="hlt">gas-exchange</span> cycle (DGC) was observed even in insects standing on <span class="hlt">water</span> when the respirometry chamber was being perfused with humid (>95% relative humidity) <span class="hlt">air</span>. Comparisons of insects at 20°C in humid and dry <span class="hlt">air</span> showed no statistically significant differences in metabolic rate or respiratory pattern (P>0.05). The proportion of time that the spiracles were closed was greater at 10°C than at 20°C (P<0.01), and greater at 20°C than at 30°C (P<0.05). These results are compatible with the hypothesis that the respiratory patterns of insects are determined by the relationship between oxygen supply and oxygen demand. There was no evidence in this insect that humidity had any effect on the respiratory pattern. The results are discussed in the context of the ongoing discussion in the literature of the origin, maintenance and adaptive significance of the DGC in insects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPS...375..442L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPS...375..442L"><span><span class="hlt">Water</span> permeation through anion <span class="hlt">exchange</span> membranes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Luo, Xiaoyan; Wright, Andrew; Weissbach, Thomas; Holdcroft, Steven</p> <p>2018-01-01</p> <p>An understanding of <span class="hlt">water</span> permeation through solid polymer electrolyte (SPE) membranes is crucial to offset the unbalanced <span class="hlt">water</span> activity within SPE fuel cells. We examine <span class="hlt">water</span> permeation through an emerging class of anion <span class="hlt">exchange</span> membranes, hexamethyl-p-terphenyl poly (dimethylbenzimidazolium) (HMT-PMBI), and compare it against series of membrane thickness for a commercial anion <span class="hlt">exchange</span> membrane (AEM), Fumapem® FAA-3, and a series of proton <span class="hlt">exchange</span> membranes, Nafion®. The HMT-PMBI membrane is found to possess higher <span class="hlt">water</span> permeabilities than Fumapem® FAA-3 and comparable permeability than Nafion (H+). By measuring <span class="hlt">water</span> permeation through membranes of different thicknesses, we are able to decouple, for the first time, internal and interfacial <span class="hlt">water</span> permeation resistances through anion <span class="hlt">exchange</span> membranes. Permeation resistances on liquid/membrane interface is found to be negligible compared to that for vapor/membrane for both series of AEMs. Correspondingly, the resistance of liquid <span class="hlt">water</span> permeation is found to be one order of magnitude smaller compared to that of vapor <span class="hlt">water</span> permeation. HMT-PMBI possesses larger effective internal <span class="hlt">water</span> permeation coefficient than both Fumapem® FAA-3 and Nafion® membranes (60 and 18% larger, respectively). In contrast, the effective interfacial permeation coefficient of HMT-PMBI is found to be similar to Fumapem® (±5%) but smaller than Nafion®(H+) (by 14%).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29251470','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29251470"><span>Effect of saline irrigation <span class="hlt">water</span> on <span class="hlt">gas</span> <span class="hlt">exchange</span> and proline metabolism in ber (Ziziphus).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bagdi, D L; Bagri, G K</p> <p>2016-09-01</p> <p>An experiment was conducted in pots of 25 kg capacity to study the effect of saline irrigation (EC 0,5,10,15 and 20 dSm-1) prepared by mixing NaCl, NaSO4, CaCl and MgCl2 in 3:1 ratio of chloride and sulphate on <span class="hlt">gas</span> <span class="hlt">exchange</span> traits, membrane stability, chlorophyll stability index and osmolytic defense mechanism in Ziziphus rotundifolia and Ziziphus nummularia species of Indian jujube (Z.mauritiana). Result showed that net photosynthetic rate (PN), transpiration (e) and stomatal conductance were comparatively lower in Ziziphus nummularia, which further declined with increasing level of saline irrigation <span class="hlt">water</span>. Chlorophyll stability and membrane stability also declined significantly in salt stress, with higher magnitude in Ziziphus nummularia. The activity of proline anabolic enzymes; Δ1-Pyrrolline-5-carboxylate reductase, Δ1-Pyrrolline-5-carboxylate synthetase and Ornithine-δ-aminotransferase were recorded higher in Ziziphus rotundifolia with decrease in proline dehydrogenase. The sodium content was observed higher in roots of Ziziphus rotundifolia and leaves of Ziziphus nummularia. Therefore, it is suggested that salt tolerance mechanism was more efficiently operative in Ziziphus rotundifolia owing to better management of physiological attributes, osmolytic defense mechanism and restricted translocation of sodium from root to leaves along with larger accumulation of potassium in its leaves.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1046750','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1046750"><span>Stirling <span class="hlt">Air</span> Conditioner for Compact Cooling</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>None</p> <p>2010-09-01</p> <p>BEETIT Project: Infinia is developing a compact <span class="hlt">air</span> conditioner that uses an unconventional high efficient Stirling cycle system (vs. conventional vapor compression systems) to produce cool <span class="hlt">air</span> that is energy efficient and does not rely on polluting refrigerants. The Stirling cycle system is a type of <span class="hlt">air</span> conditioning system that uses a motor with a piston to remove heat to the outside atmosphere using a <span class="hlt">gas</span> refrigerant. To date, Stirling systems have been expensive and have not had the right kind of heat <span class="hlt">exchanger</span> to help cool <span class="hlt">air</span> efficiently. Infinia is using chip cooling technology from the computer industry tomore » make improvements to the heat <span class="hlt">exchanger</span> and improve system performance. Infinia’s <span class="hlt">air</span> conditioner uses helium <span class="hlt">gas</span> as refrigerant, an environmentally benign <span class="hlt">gas</span> that does not react with other chemicals and does not burn. Infinia’s improvements to the Stirling cycle system will enable the cost-effective mass production of high-efficiency <span class="hlt">air</span> conditioners that use no polluting refrigerants.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3279239','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3279239"><span>Local Area <span class="hlt">Water</span> Removal Analysis of a Proton <span class="hlt">Exchange</span> Membrane Fuel Cell under <span class="hlt">Gas</span> Purge Conditions</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Lee, Chi-Yuan; Lee, Yu-Ming; Lee, Shuo-Jen</p> <p>2012-01-01</p> <p>In this study, local area <span class="hlt">water</span> content distribution under various <span class="hlt">gas</span> purging conditions are experimentally analyzed for the first time. The local high frequency resistance (HFR) is measured using novel micro sensors. The results reveal that the liquid <span class="hlt">water</span> removal rate in a membrane electrode assembly (MEA) is non-uniform. In the under-the-channel area, the removal of liquid <span class="hlt">water</span> is governed by both convective and diffusive flux of the through-plane drying. Thus, almost all of the liquid <span class="hlt">water</span> is removed within 30 s of purging with <span class="hlt">gas</span>. However, liquid <span class="hlt">water</span> that is stored in the under-the-rib area is not easy to remove during 1 min of <span class="hlt">gas</span> purging. Therefore, the re-hydration of the membrane by internal diffusive flux is faster than that in the under-the-channel area. Consequently, local fuel starvation and membrane degradation can degrade the performance of a fuel cell that is started from cold. PMID:22368495</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.ncbi.nlm.nih.gov/pubmed/22368495','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22368495"><span>Local area <span class="hlt">water</span> removal analysis of a proton <span class="hlt">exchange</span> membrane fuel cell under <span class="hlt">gas</span> purge conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lee, Chi-Yuan; Lee, Yu-Ming; Lee, Shuo-Jen</p> <p>2012-01-01</p> <p>In this study, local area <span class="hlt">water</span> content distribution under various <span class="hlt">gas</span> purging conditions are experimentally analyzed for the first time. The local high frequency resistance (HFR) is measured using novel micro sensors. The results reveal that the liquid <span class="hlt">water</span> removal rate in a membrane electrode assembly (MEA) is non-uniform. In the under-the-channel area, the removal of liquid <span class="hlt">water</span> is governed by both convective and diffusive flux of the through-plane drying. Thus, almost all of the liquid <span class="hlt">water</span> is removed within 30 s of purging with <span class="hlt">gas</span>. However, liquid <span class="hlt">water</span> that is stored in the under-the-rib area is not easy to remove during 1 min of <span class="hlt">gas</span> purging. Therefore, the re-hydration of the membrane by internal diffusive flux is faster than that in the under-the-channel area. Consequently, local fuel starvation and membrane degradation can degrade the performance of a fuel cell that is started from cold.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PlST...20d4011Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PlST...20d4011Y"><span>Influence of <span class="hlt">water</span> content on the inactivation of P. digitatum spores using an <span class="hlt">air-water</span> plasma jet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Youyi, HU; Weidong, ZHU; Kun, LIU; Leng, HAN; Zhenfeng, ZHENG; Huimin, HU</p> <p>2018-04-01</p> <p>In order to investigate whether an <span class="hlt">air-water</span> plasma jet is beneficial to improve the efficiency of inactivation, a series of experiments were done using a ring-needle plasma jet. The <span class="hlt">water</span> content in the working <span class="hlt">gas</span> (<span class="hlt">air</span>) was accurately measured based on the Karl Fischer method. The effects of <span class="hlt">water</span> on the production of OH (A2Σ+-X2Πi) and O (3p5P-3s5S) were also studied by optical emission spectroscopy. The results show that the <span class="hlt">water</span> content is in the range of 2.53-9.58 mg l-1, depending on the <span class="hlt">gas/water</span> mixture ratio. The production of OH (A2Σ+-X2Πi) rises with the increase of <span class="hlt">water</span> content, whereas the O (3p5P-3s5S) shows a declining tendency with higher <span class="hlt">water</span> content. The sterilization experiments indicate that this <span class="hlt">air-water</span> plasma jet inactivates the P. digitatum spores very effectively and its efficiency rises with the increase of the <span class="hlt">water</span> content. It is possible that OH (A2Σ+-X2Πi) is a more effective species in inactivation than O (3p5P-3s5S) and the <span class="hlt">water</span> content benefit the spore germination inhibition through rising the OH (A2Σ+-X2Πi) production. The maximum of the inactivation efficacy is up to 93% when the applied voltage is -6.75 kV and the <span class="hlt">water</span> content is 9.58 mg l-1.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/6792674','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/6792674"><span>Functional morphology of the gills of the bowfin, Amia calva L., with special reference to their significance during <span class="hlt">air</span> exposure.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Daxboeck, C; Barnard, D K; Randall, D J</p> <p>1981-03-01</p> <p>The bowfin, Amia calva is a facultative <span class="hlt">air</span> breathing fish restricted to North America and is reported to estivate. The relative and functional gill surface areas of A. calva are not reduced, as in many amphibious fish, but have areas comparable to many completely aquatic species. The secondary lamellae are fused to form a lattice-work of rectangular pores, a gill arrangement unique among fresh-<span class="hlt">water</span> fishes. This highly modified gill structure imparts considerable rigidity such that these do not collapse upon <span class="hlt">air</span> exposure. In vivo blood <span class="hlt">gas</span> measurements from <span class="hlt">air</span> exposed Amia reveal that these gills must be free of <span class="hlt">water</span>, since there is both O2 uptake and CO2 excretion across them. The observed ventilatory motions therefore pass <span class="hlt">air</span> over the secondary lamellae for diffusive <span class="hlt">gas</span> <span class="hlt">exchange</span> during <span class="hlt">air</span> exposure. In the artificial conditions of our experiments, however, <span class="hlt">air</span> exposure was associated with a marked acidosis and the fish died within 2 hours of being returned to normoxic <span class="hlt">water</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27063719','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27063719"><span>Temperature Programmed Desorption of Quench-condensed Krypton and Acetone in <span class="hlt">Air</span>; Selective Concentration of Ultra-trace <span class="hlt">Gas</span> Components.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Suzuki, Taku T; Sakaguchi, Isao</p> <p>2016-01-01</p> <p>Selective concentration of ultra-trace components in <span class="hlt">air</span>-like gases has an important application in analyzing volatile organic compounds in the <span class="hlt">gas</span>. In the present study, we examined quench-condensation of the sample <span class="hlt">gas</span> on a ZnO substrate below 50 K followed by temperature programmed desorption (TPD) (low temperature TPD) as a selective <span class="hlt">gas</span> concentration technique. We studied two specific gases in the normal <span class="hlt">air</span>; krypton as an inert <span class="hlt">gas</span> and acetone as a reactive <span class="hlt">gas</span>. We evaluated the relationship between the operating condition of low temperature TPD and the lowest detection limit. In the case of krypton, we observed the selective concentration by exposing at 6 K followed by thermal desorption at about 60 K. On the other hand, no selectivity appeared for acetone although trace acetone was successfully concentrated. This is likely due to the solvent effect by a major component in the <span class="hlt">air</span>, which is suggested to be <span class="hlt">water</span>. We suggest that pre-condensation to remove the <span class="hlt">water</span> component may improve the selectivity in the trace acetone analysis by low temperature TPD.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020083268','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020083268"><span><span class="hlt">Air/Water</span> Purification</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1992-01-01</p> <p>After 18 years of research into <span class="hlt">air/water</span> pollution at Stennis Space Center, Dr. B. C. Wolverton formed his own company, Wolverton Environmental Services, Inc., to provide technology and consultation in <span class="hlt">air</span> and <span class="hlt">water</span> treatment. Common houseplants are used to absorb potentially harmful materials from bathrooms and kitchens. The plants are fertilized, <span class="hlt">air</span> is purified, and wastewater is converted to clean <span class="hlt">water</span>. More than 100 U.S. communities have adopted Wolverton's earlier <span class="hlt">water</span> hyacinth and artificial marsh applications. Catfish farmers are currently evaluating the artificial marsh technology as a purification system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JPhCS.655a2035D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JPhCS.655a2035D"><span>Modelling heat and mass transfer in a membrane-based <span class="hlt">air-to-air</span> enthalpy <span class="hlt">exchanger</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dugaria, S.; Moro, L.; Del, D., Col</p> <p>2015-11-01</p> <p>The diffusion of total energy recovery systems could lead to a significant reduction in the energy demand for building <span class="hlt">air</span>-conditioning. With these devices, sensible heat and humidity can be recovered in winter from the exhaust airstream, while, in summer, the incoming <span class="hlt">air</span> stream can be cooled and dehumidified by transferring the excess heat and moisture to the exhaust <span class="hlt">air</span> stream. Membrane based enthalpy <span class="hlt">exchangers</span> are composed by different channels separated by semi-permeable membranes. The membrane allows moisture transfer under vapour pressure difference, or <span class="hlt">water</span> concentration difference, between the two sides and, at the same time, it is ideally impermeable to <span class="hlt">air</span> and other contaminants present in exhaust <span class="hlt">air</span>. Heat transfer between the airstreams occurs through the membrane due to the temperature gradient. The aim of this work is to develop a detailed model of the coupled heat and mass transfer mechanisms through the membrane between the two airstreams. After a review of the most relevant models published in the scientific literature, the governing equations are presented and some simplifying assumptions are analysed and discussed. As a result, a steady-state, two-dimensional finite difference numerical model is setup. The developed model is able to predict temperature and humidity evolution inside the channels. Sensible and latent heat transfer rate, as well as moisture transfer rate, are determined. A sensitive analysis is conducted in order to determine the more influential parameters on the thermal and vapour transfer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70028198','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70028198"><span>Leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> characteristics of three neotropical mangrove species in response to varying hydroperiod</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Krauss, Ken W.; Twilley, Robert R.; Doyle, Thomas W.; Gardiner, Emile S.</p> <p>2006-01-01</p> <p>We determined how different hydroperiods affected leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> characteristics of greenhouse-grown seedlings (2002) and saplings (2003) of the mangrove species Avicennia germinans (L.) Stearn., Laguncularia racemosa (L.) Gaertn. f., and Rhizophora mangle L. Hydroperiod treatments included no flooding (unflooded), intermittent flooding (intermittent), and permanent flooding (flooded). Plants in the intermittent treatment were measured under both flooded and drained states and compared separately. In the greenhouse study, plants of all species maintained different leaf areas in the contrasting hydroperiods during both years. Assimilation–light response curves indicated that the different hydroperiods had little effect on leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> characteristics in either seedlings or saplings. However, short-term intermittent flooding for between 6 and 22 days caused a 20% reduction in maximum leaf-level carbon assimilation rate, a 51% lower light requirement to attain 50% of maximum assimilation, and a 38% higher demand from dark respiration. Although interspecific differences were evident for nearly all measured parameters in both years, there was little consistency in ranking of the interspecific responses. Species by hydroperiod interactions were significant only for sapling leaf area. In a field study, R. mangle saplings along the Shark River in the Everglades National Park either demonstrated no significant effect or slight enhancement of carbon assimilation and <span class="hlt">water</span>-use efficiency while flooded. We obtained little evidence that contrasting hydroperiods affect leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> characteristics of mangrove seedlings or saplings over long time intervals; however, intermittent flooding may cause short-term depressions in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>. The resilience of mangrove systems to flooding, as demonstrated in the permanently flooded treatments, will likely promote photosynthetic and morphological adjustment to slight hydroperiod shifts in many settings..</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27147352','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27147352"><span>Quantification of regional early stage <span class="hlt">gas</span> <span class="hlt">exchange</span> changes using hyperpolarized (129)Xe MRI in a rat model of radiation-induced lung injury.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Doganay, Ozkan; Stirrat, Elaine; McKenzie, Charles; Schulte, Rolf F; Santyr, Giles E</p> <p>2016-05-01</p> <p>To assess the feasibility of hyperpolarized (HP) (129)Xe MRI for detection of early stage radiation-induced lung injury (RILI) in a rat model involving unilateral irradiation by assessing differences in <span class="hlt">gas</span> <span class="hlt">exchange</span> dynamics between irradiated and unirradiated lungs. The dynamics of <span class="hlt">gas</span> <span class="hlt">exchange</span> between alveolar <span class="hlt">air</span> space and pulmonary tissue (PT), PT and red blood cells (RBCs) was measured using single-shot spiral iterative decomposition of <span class="hlt">water</span> and fat with echo asymmetry and least-squares estimation images of the right and left lungs of two age-matched cohorts of Sprague Dawley rats. The first cohort (n = 5) received 18 Gy irradiation to the right lung using a (60)Co source and the second cohort (n = 5) was not irradiated and served as the healthy control. Both groups were imaged two weeks following irradiation when radiation pneumonitis (RP) was expected to be present. The <span class="hlt">gas</span> <span class="hlt">exchange</span> data were fit to a theoretical <span class="hlt">gas</span> <span class="hlt">exchange</span> model to extract measurements of pulmonary tissue thickness (LPT) and relative blood volume (VRBC) from each of the right and left lungs of both cohorts. Following imaging, lung specimens were retrieved and percent tissue area (PTA) was assessed histologically to confirm RP and correlate with MRI measurements. Statistically significant differences in LPT and VRBC were observed between the irradiated and non-irradiated cohorts. In particular, LPT of the right and left lungs was increased approximately 8.2% and 5.0% respectively in the irradiated cohort. Additionally, VRBC of the right and left lungs was decreased approximately 36.1% and 11.7% respectively for the irradiated cohort compared to the non-irradiated cohort. PTA measurements in both right and left lungs were increased in the irradiated group compared to the non-irradiated cohort for both the left (P < 0.05) and right lungs (P < 0.01) confirming the presence of RP. PTA measurements also correlated with the MRI measurements for both the non-irradiated (r = 0.79, P < 0.01) and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20060021945','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20060021945"><span>Conceptual Design of a Condensing Heat <span class="hlt">Exchanger</span> for Space Systems Using Porous Media</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hasan, Mohammad M.; Khan, Lutful I.; Nayagam, Vedha; Balasubramaniam, Ramaswamy</p> <p>2006-01-01</p> <p>Condensing heat <span class="hlt">exchangers</span> are used in many space applications in the thermal and humidity control systems. In the International Space Station (ISS), humidity control is achieved by using a <span class="hlt">water</span> cooled fin surface over which the moist <span class="hlt">air</span> condenses, followed by "slurper bars" that take in both the condensate and <span class="hlt">air</span> into a rotary separator and separates the <span class="hlt">water</span> from <span class="hlt">air</span>. The use of a cooled porous substrate as the condensing surface provides and attractive alternative that combines both heat removal as well as liquid/<span class="hlt">gas</span> separation into a single unit. By selecting the pore sizes of the porous substrate a gravity independent operation may also be possible with this concept. Condensation of vapor into and on the porous surface from the flowing <span class="hlt">air</span> and the removal of condensate from the porous substrate are the critical processes involved in the proposed concept. This paper describes some preliminary results of the proposed condensate withdrawal process and discusses the on-going design and development work of a porous media based condensing heat <span class="hlt">exchanger</span> at the NASA Glenn Research Center in collaboration with NASA Johnson Space Center.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26163155','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26163155"><span>Ventilatory <span class="hlt">gas</span> <span class="hlt">exchange</span> and early response to cardiac resynchronization therapy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Chul-Ho; Olson, Lyle J; Shen, Win K; Cha, Yong-Mei; Johnson, Bruce D</p> <p>2015-11-01</p> <p>Cardiac resynchronization therapy (CRT) is an accepted intervention for chronic heart failure (HF), although approximately 30% of patients are non-responders. The purpose of this study was to determine whether exercise respiratory <span class="hlt">gas</span> <span class="hlt">exchange</span> obtained before CRT implantation predicts early response to CRT. Before CRT implantation, patients were assigned to either a mild-moderate group (Mod G, n = 33, age 67 ± 10 years) or a moderate-severe group (Sev G, n = 31, age 67 ± 10 years), based on abnormalities in exercise <span class="hlt">gas</span> <span class="hlt">exchange</span>. Severity of impaired <span class="hlt">gas</span> <span class="hlt">exchange</span> was based on a score from the measures of VE/VCO(2) slope, resting PETCO(2) and change of PETCO(2) from resting to peak. All measurements were performed before and 3 to 4 months after CRT implantation. Although Mod G did not have improved <span class="hlt">gas</span> <span class="hlt">exchange</span> (p > 0.05), Sev G improved significantly (p < 0.05) post-CRT. In addition, Mod G did not show improved right ventricular systolic pressure (RSVP; pre vs post: 37 ± 14 vs 36 ± 11 mm Hg, p > 0.05), yet Sev G showed significantly improved RVSP, by 23% (50 ± 14 vs 42 ± 12 mm Hg, p < 0.05). Both groups had improved left ventricular ejection fraction (p < 0.05), New York Heart Association class (p < 0.05) and quality of life (p < 0.05), but no significant differences were observed between groups (p > 0.05). No significant changes were observed in brain natriuretic peptide in either group post-CRT. Based on pre-CRT implantation ventilatory <span class="hlt">gas</span> <span class="hlt">exchange</span>, subjects with the most impaired values appeared to have more improvement post-CRT, possibly associated with a decrease in RVSP. Copyright © 2015 International Society for Heart and Lung Transplantation. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70014998','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70014998"><span><span class="hlt">Air</span> permeability and trapped-<span class="hlt">air</span> content in two soils</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Stonestrom, David A.; Rubin, Jacob</p> <p>1989-01-01</p> <p>To improve understanding of hysteretic <span class="hlt">air</span> permeability relations, a need exists for data on the <span class="hlt">water</span> content dependence of <span class="hlt">air</span> permeability, matric pressure, and <span class="hlt">air</span> trapping (especially for wetting-drying cycles). To obtain these data, a special instrument was designed. The instrument is a combination of a <span class="hlt">gas</span> permeameter (for <span class="hlt">air</span> permeability determination), a suction plate apparatus (for retentivity curve determination), and an <span class="hlt">air</span> pycnometer (for trapped-<span class="hlt">air</span>-volume determination). This design allowed values of <span class="hlt">air</span> permeability, matric pressure, and <span class="hlt">air</span> trapping to be codetermined, i.e., determined at the same values of <span class="hlt">water</span> content using the same sample and the same inflow-outflow boundaries. Such data were obtained for two nonswelling soils. The validity of the <span class="hlt">air</span> permeability determinations was repeatedly confirmed by rigorous tests of Darcy's law. During initial drying from complete <span class="hlt">water</span> saturation, supplementary measurements were made to assess the magnitude of <span class="hlt">gas</span> slip. The extended Darcy equation accurately described the measured flux gradient relations for each condition of absolute <span class="hlt">gas</span> pressure tested. <span class="hlt">Air</span> permeability functions exhibited zero-permeability regions at high <span class="hlt">water</span> contents as well as an abruptly appearing hysteresis at low <span class="hlt">water</span> contents. Measurements in the zero-permeability regions revealed that the total amount of <span class="hlt">air</span> in general exceeded the amount of trapped <span class="hlt">air</span>. This indicates that the medium' s <span class="hlt">air</span> space is partitioned into three measurable domains: through-flowing <span class="hlt">air</span>, locally accessible <span class="hlt">air</span> (i.e., <span class="hlt">air</span> accessible from only one flow boundary), and trapped <span class="hlt">air</span>. During repeated wetting and drying, the disappearance and reappearance of <span class="hlt">air</span> permeability coincided closely with the reappearance and disappearance, respectively, of trapped <span class="hlt">air</span>. The observed relation between critical features of the <span class="hlt">air</span> permeability functions and those of the <span class="hlt">air</span>-trapping functions suggest that <span class="hlt">water</span>-based blockages play a significant role in the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23755221','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23755221"><span>Oxygen and <span class="hlt">air</span> nanobubble <span class="hlt">water</span> solution promote the growth of plants, fishes, and mice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ebina, Kosuke; Shi, Kenrin; Hirao, Makoto; Hashimoto, Jun; Kawato, Yoshitaka; Kaneshiro, Shoichi; Morimoto, Tokimitsu; Koizumi, Kota; Yoshikawa, Hideki</p> <p>2013-01-01</p> <p>Nanobubbles (<200 nm in diameter) have several unique properties such as long lifetime in liquid owing to its negatively charged surface, and its high <span class="hlt">gas</span> solubility into the liquid owing to its high internal pressure. They are used in variety of fields including diagnostic aids and drug delivery, while there are no reports assessing their effects on the growth of lives. Nanobubbles of <span class="hlt">air</span> or oxygen <span class="hlt">gas</span> were generated using a nanobubble aerator (BUVITAS; Ligaric Company Limited, Osaka, Japan). Brassica campestris were cultured hydroponically for 4 weeks within <span class="hlt">air</span>-nanobubble <span class="hlt">water</span> or within normal <span class="hlt">water</span>. Sweetfish (for 3 weeks) and rainbow trout (for 6 weeks) were kept either within <span class="hlt">air</span>-nanobubble <span class="hlt">water</span> or within normal <span class="hlt">water</span>. Finally, 5 week-old male DBA1/J mice were bred with normal free-chaw and free-drinking either of oxygen-nanobubble <span class="hlt">water</span> or of normal <span class="hlt">water</span> for 12 weeks. Oxygen-nanobubble significantly increased the dissolved oxygen concentration of <span class="hlt">water</span> as well as concentration/size of nanobubbles which were relatively stable for 70 days. <span class="hlt">Air</span>-nanobubble <span class="hlt">water</span> significantly promoted the height (19.1 vs. 16.7 cm; P<0.05), length of leaves (24.4 vs. 22.4 cm; P<0.01), and aerial fresh weight (27.3 vs. 20.3 g; P<0.01) of Brassica campestris compared to normal <span class="hlt">water</span>. Total weight of sweetfish increased from 3.0 to 6.4 kg in normal <span class="hlt">water</span>, whereas it increased from 3.0 to 10.2 kg in <span class="hlt">air</span>-nanobubble <span class="hlt">water</span>. In addition, total weight of rainbow trout increased from 50.0 to 129.5 kg in normal <span class="hlt">water</span>, whereas it increased from 50.0 to 148.0 kg in <span class="hlt">air</span>-nanobubble <span class="hlt">water</span>. Free oral intake of oxygen-nanobubble <span class="hlt">water</span> significantly promoted the weight (23.5 vs. 21.8 g; P<0.01) and the length (17.0 vs. 16.1 cm; P<0.001) of mice compared to that of normal <span class="hlt">water</span>. We have demonstrated for the first time that oxygen and <span class="hlt">air</span>-nanobubble <span class="hlt">water</span> may be potentially effective tools for the growth of lives.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/14095','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/14095"><span>Effects of moisture and nitrogen stress on <span class="hlt">gas</span> <span class="hlt">exchange</span> and nutrient resorption in Quercus rubra seedlings</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>K. Francis Salifu; Douglass F. Jacobs</p> <p>2008-01-01</p> <p>The effects of simulated soil fertility at three levels (poor, medium, and rich soils) and moisture stress at two levels (well <span class="hlt">watered</span> versus moisture stressed) on <span class="hlt">gas</span> <span class="hlt">exchange</span> and foliar nutrient resorption in 1+0 bareroot northern red oak (Quercus rubra) seedlings were evaluated. Current nitrogen (N) uptake was labeled with the stable isotope</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/2386412','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/2386412"><span>A multiresidue method by high performance liquid chromatography-based fractionation and <span class="hlt">gas</span> chromatographic determination of trace levels of pesticides in <span class="hlt">air</span> and <span class="hlt">water</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Seiber, J N; Glotfelty, D E; Lucas, A D; McChesney, M M; Sagebiel, J C; Wehner, T A</p> <p>1990-01-01</p> <p>A multiresidue analytical method is described for pesticides, transformation products, and related toxicants based upon high performance liquid chromatographic (HPLC) fractionation of extracted residue on a Partisil silica gel normal phase column followed by selective-detector <span class="hlt">gas</span> chromatographic (GC) determination of components in each fraction. The HPLC mobile phase gradient (hexane to methyl t-butyl ether) gave good chromatographic efficiency, resolution, reproducibility and recovery for 61 test compounds, and allowed for collection in four fractions spanning polarities from low polarity organochlorine compounds (fraction 1) to polar N-methylcarbamates and organophosphorus oxons (fraction 4). The multiresidue method was developed for use with <span class="hlt">air</span> samples collected on XAD-4 and related trapping agents, and <span class="hlt">water</span> samples extracted with methylene chloride. Detection limits estimated from spiking experiments were generally 0.3-1 ng/m3 for high-volume <span class="hlt">air</span> samples, and 0.01-0.1 microgram/L for one-liter <span class="hlt">water</span> samples. Applications were made to determination of pesticides in fogwater and <span class="hlt">air</span> samples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1265948-energy-factor-analysis-gas-heat-pump-water-heaters','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1265948-energy-factor-analysis-gas-heat-pump-water-heaters"><span>Energy Factor Analysis for <span class="hlt">Gas</span> Heat Pump <span class="hlt">Water</span> Heaters</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>Gluesenkamp, Kyle R</p> <p>2016-01-01</p> <p><span class="hlt">Gas</span> heat pump <span class="hlt">water</span> heaters (HPWHs) can improve <span class="hlt">water</span> heating efficiency with zero GWP and zero ODP working fluids. The energy factor (EF) of a <span class="hlt">gas</span> HPWH is sensitive to several factors. In this work, expressions are derived for EF of <span class="hlt">gas</span> HPWHs, as a function of heat pump cycle COP, tank heat losses, burner efficiency, electrical draw, and effectiveness of supplemental heat <span class="hlt">exchangers</span>. The expressions are used to investigate the sensitivity of EF to each parameter. EF is evaluated on a site energy basis (as used by the US DOE for rating <span class="hlt">water</span> heater EF), and a primary energy-basismore » energy factor (PEF) is also defined and included. Typical ranges of values for the six parameters are given. For <span class="hlt">gas</span> HPWHs, using typical ranges for component performance, EF will be 59 80% of the heat pump cycle thermal COP (for example, a COP of 1.60 may result in an EF of 0.94 1.28). Most of the reduction in COP is due to burner efficiency and tank heat losses. <span class="hlt">Gas</span>-fired HPWHs are theoretically be capable of an EF of up to 1.7 (PEF of 1.6); while an EF of 1.1 1.3 (PEF of 1.0 1.1) is expected from an early market entry.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=248583','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=248583"><span>Carbon dioxide and <span class="hlt">water</span> vapour <span class="hlt">exchange</span> in a tropical dry forest as influenced by the North American Monsoon System (NAMS)</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>To better understand the effects and relationship between precipitation, net ecosystem carbon dioxide (NEE) and <span class="hlt">water</span> vapor <span class="hlt">exchange</span> (ET), we report a study conducted in the tropical dry forest (TDF) in the northwest of Mexico. Ecosystem <span class="hlt">gas</span> <span class="hlt">exchange</span> was measured using the eddy correlation technique...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20046684','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20046684"><span>The efficacy of fluid-<span class="hlt">gas</span> <span class="hlt">exchange</span> for the treatment of postvitrectomy retinal detachment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jang, Ji Hye; Kim, Yu Cheol; Kim, Kwang Soo</p> <p>2009-12-01</p> <p>This study was designed to evaluate the efficacy of fluid-<span class="hlt">gas</span> <span class="hlt">exchange</span> for the treatment of postvitrectomy retinal detachment. We retrospectively reviewed the records of 33 consecutive patients (35 eyes) who underwent fluid-<span class="hlt">gas</span> <span class="hlt">exchange</span> treatment for postvitrectomy retinal detachment using the two-needle pars plana approach technique. The retinal reattachment rate was 80.0% after complete intravitreal <span class="hlt">gas</span> disappearance following the fluid-<span class="hlt">gas</span> <span class="hlt">exchange</span>; the overall success rate was 65.7%. Visual acuity was improved or stable in 80.0% of cases; a two-line or greater vision improvement or a best-corrected visual acuity of 0.4 or better occurred in 62.9% of cases. The success rates for superior retinal detachments and posterior pole retinal detachments were 76.5% and 85.7%, respectively. Fluid-<span class="hlt">gas</span> <span class="hlt">exchange</span> represents a simple and cost-effective alternative outpatient procedure for retinal reattachment without reoperation for the treatment of superior and posterior pole retinal detachments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21292477','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21292477"><span>Experimental study of wood downdraft gasification for an improved producer <span class="hlt">gas</span> quality through an innovative two-stage <span class="hlt">air</span> and premixed <span class="hlt">air/gas</span> supply approach.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jaojaruek, Kitipong; Jarungthammachote, Sompop; Gratuito, Maria Kathrina B; Wongsuwan, Hataitep; Homhual, Suwan</p> <p>2011-04-01</p> <p>This study conducted experiments on three different downdraft gasification approaches: single stage, conventional two-stage, and an innovative two-stage <span class="hlt">air</span> and premixed <span class="hlt">air/gas</span> supply approach. The innovative two-stage approach has two nozzle locations, one for <span class="hlt">air</span> supply at combustion zone and the other located at the pyrolysis zone for supplying the premixed <span class="hlt">gas</span> (<span class="hlt">air</span> and producer <span class="hlt">gas</span>). The producer <span class="hlt">gas</span> is partially bypassed to mix with <span class="hlt">air</span> and supplied to burn at the pyrolysis zone. The result shows that producer <span class="hlt">gas</span> quality generated by the innovative two-stage approach improved as compared to conventional two-stage. The higher heating value (HHV) increased from 5.4 to 6.5 MJ/Nm(3). Tar content in producer <span class="hlt">gas</span> reduced to less than 45 mg/Nm(3). With this approach, <span class="hlt">gas</span> can be fed directly to an internal combustion engine. Furthermore, the gasification thermal efficiency also improved by approximately 14%. The approach gave double benefits on <span class="hlt">gas</span> qualities and energy savings. Copyright © 2010 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.5581L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.5581L"><span>Sum-Frequency Generation Spectroscopy for Studying Organic Layers at <span class="hlt">Water-Air</span> Interfaces: Microlayer Monitoring and Surface Reactivity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Laß, Kristian; Kleber, Joscha; Bange, Hermann; Friedrichs, Gernot</p> <p>2015-04-01</p> <p>The sea surface microlayer, according to commonly accepted terminology, comprises the topmost millimetre of the oceanic <span class="hlt">water</span> column. It is often enriched with organic matter and is directly influenced by sunlight exposure and <span class="hlt">gas</span> <span class="hlt">exchange</span> with the atmosphere, hence making it a place for active biochemistry and photochemistry as well as for heterogeneous reactions. In addition, surface active material either is formed or accumulates directly at the <span class="hlt">air-water</span> interface and gives rise to very thin layers, sometimes down to monomolecular thickness. This "sea surface nanolayer" determines the viscoelastic properties of the seawater surface and thus may impact the turbulent <span class="hlt">air</span>-sea <span class="hlt">gas</span> <span class="hlt">exchange</span> rates. To this effect, this small scale layer presumably plays an important role for large scale changes of atmospheric trace <span class="hlt">gas</span> concentrations (e.g., by modulating the ocean carbon sink characteristics) with possible implications for coupled climate models. To date, detailed knowledge about the composition, structure, and reactivity of the sea surface nanolayer is still scarce. Due to its small vertical dimension and the small amount of material, this surfactant layer is very difficult to separate and analyse. A way out is the application of second-order nonlinear optical methods, which make a direct surface-specific and background-free detection of this interfacial layer possible. In recent years, we have introduced the use of vibrational sum frequency generation (VSFG) spectroscopy to gain insight into natural and artificial organic monolayers at the <span class="hlt">air-water</span> interface. In this contribution, the application of VSFG spectroscopy for the analysis of the sea surface nanolayer will be illustrated. Resulting spectra are interpreted in terms of layer composition and surfactant classes, in particular with respect to carbohydrate-containing molecules such as glycolipids. The partitioning of the detected surfactants into soluble and non-soluble ("wet" and "dry") surfactants will be</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23481438','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23481438"><span>A novel, simplified ex vivo method for measuring <span class="hlt">water</span> <span class="hlt">exchange</span> performance of heat and moisture <span class="hlt">exchangers</span> for tracheostomy application.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>van den Boer, Cindy; Muller, Sara H; Vincent, Andrew D; Züchner, Klaus; van den Brekel, Michiel W M; Hilgers, Frans J M</p> <p>2013-09-01</p> <p>Breathing through a tracheostomy results in insufficient warming and humidification of inspired <span class="hlt">air</span>. This loss of <span class="hlt">air</span>-conditioning can be partially compensated for with the application of a heat and moisture <span class="hlt">exchanger</span> (HME) over the tracheostomy. In vitro (International Organization for Standardization [ISO] standard 9360-2:2001) and in vivo measurements of the effects of an HME are complex and technically challenging. The aim of this study was to develop a simple method to measure the ex vivo HME performance comparable with previous in vitro and in vivo results. HMEs were weighed at the end of inspiration and at the end of expiration at different breathing volumes. Four HMEs (Atos Medical, Hörby, Sweden) with known in vivo humidity and in vitro <span class="hlt">water</span> loss values were tested. The associations between weight change, volume, and absolute humidity were determined using both linear and non-linear mixed effects models. The rating between the 4 HMEs by weighing correlated with previous intra-tracheal measurements (R(2) = 0.98), and the ISO standard (R(2) = 0.77). Assessment of the weight change between end of inhalation and end of exhalation is a valid and simple method of measuring the <span class="hlt">water</span> <span class="hlt">exchange</span> performance of an HME.</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('https://ntrs.nasa.gov/search.jsp?R=19780027541&hterms=thought+experiments&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dthought%2Bexperiments','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19780027541&hterms=thought+experiments&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dthought%2Bexperiments"><span>The Viking <span class="hlt">gas</span> <span class="hlt">exchange</span> experiment results from Chryse and Utopia surface samples</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Oyama, V. I.; Berdahl, B. J.</p> <p>1977-01-01</p> <p>Immediate <span class="hlt">gas</span> changes occurred when untreated Martian surface samples were humidified and/or wet by an aqueous nutrient medium in the Viking lander <span class="hlt">gas</span> <span class="hlt">exchange</span> experiment. The evolutions of N2, CO2, and Ar are mainly associated with soil surface desorption caused by <span class="hlt">water</span> vapor, while O2 evolution is primarily associated with decomposition of superoxides inferred to be present on Mars. On recharges with fresh nutrient and test <span class="hlt">gas</span>, only CO2 was given off, and its rate of evolution decreased with each recharge. This CO2 evolution is thought to come from the oxidation of organics present in the nutrient by gamma Fe2O3 in the surface samples. Atmospheric analyses were also performed at both sites. The mean atmospheric composition from four analyses is N2, 2.3%; O2, not greater than 0.15%; Ar, 1.5% and CO2, 96.2%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24699994','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24699994"><span><span class="hlt">Air</span> quality concerns of unconventional oil and natural <span class="hlt">gas</span> production.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Field, R A; Soltis, J; Murphy, S</p> <p>2014-05-01</p> <p>Increased use of hydraulic fracturing ("fracking") in unconventional oil and natural <span class="hlt">gas</span> (O & NG) development from coal, sandstone, and shale deposits in the United States (US) has created environmental concerns over <span class="hlt">water</span> and <span class="hlt">air</span> quality impacts. In this perspective we focus on how the production of unconventional O & NG affects <span class="hlt">air</span> quality. We pay particular attention to shale <span class="hlt">gas</span> as this type of development has transformed natural <span class="hlt">gas</span> production in the US and is set to become important in the rest of the world. A variety of potential emission sources can be spread over tens of thousands of acres of a production area and this complicates assessment of local and regional <span class="hlt">air</span> quality impacts. We outline upstream activities including drilling, completion and production. After contrasting the context for development activities in the US and Europe we explore the use of inventories for determining <span class="hlt">air</span> emissions. Location and scale of analysis is important, as O & NG production emissions in some US basins account for nearly 100% of the pollution burden, whereas in other basins these activities make up less than 10% of total <span class="hlt">air</span> emissions. While emission inventories are beneficial to quantifying <span class="hlt">air</span> emissions from a particular source category, they do have limitations when determining <span class="hlt">air</span> quality impacts from a large area. <span class="hlt">Air</span> monitoring is essential, not only to validate inventories, but also to measure impacts. We describe the use of measurements, including ground-based mobile monitoring, network stations, airborne, and satellite platforms for measuring <span class="hlt">air</span> quality impacts. We identify nitrogen oxides, volatile organic compounds (VOC), ozone, hazardous <span class="hlt">air</span> pollutants (HAP), and methane as pollutants of concern related to O & NG activities. These pollutants can contribute to <span class="hlt">air</span> quality concerns and they may be regulated in ambient <span class="hlt">air</span>, due to human health or climate forcing concerns. Close to well pads, emissions are concentrated and exposure to a wide range of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930057514&hterms=reverse+osmosis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dreverse%2Bosmosis','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930057514&hterms=reverse+osmosis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dreverse%2Bosmosis"><span>Modeling of membrane processes for <span class="hlt">air</span> revitalization and <span class="hlt">water</span> recovery</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lange, Kevin E.; Foerg, Sandra L.; Dall-Bauman, Liese A.</p> <p>1992-01-01</p> <p><span class="hlt">Gas</span>-separation and reverse-osmosis membrane models are being developed in conjunction with membrane testing at NASA JSC. The completed <span class="hlt">gas</span>-separation membrane model extracts effective component permeabilities from multicomponent test data, and predicts the effects of flow configuration, operating conditions, and membrane dimensions on module performance. Variable feed- and permeate-side pressures are considered. The model has been applied to test data for hollow-fiber membrane modules with simulated cabin-<span class="hlt">air</span> feeds. Results are presented for a membrane designed for <span class="hlt">air</span> drying applications. Extracted permeabilities are used to predict the effect of operating conditions on <span class="hlt">water</span> enrichment in the permeate. A first-order reverse-osmosis model has been applied to test data for spiral wound membrane modules with a simulated hygiene <span class="hlt">water</span> feed. The model estimates an effective local component rejection coefficient under pseudosteady-state conditions. Results are used to define requirements for a detailed reverse-osmosis model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19920056264&hterms=gas+natural&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dgas%2Bnatural','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920056264&hterms=gas+natural&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dgas%2Bnatural"><span>Relationship between wind speed and <span class="hlt">gas</span> <span class="hlt">exchange</span> over the ocean</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wanninkhof, Rik</p> <p>1992-01-01</p> <p>A quadratic dependence of <span class="hlt">gas</span> <span class="hlt">exchange</span> on wind speed is employed to analyze the relationship between <span class="hlt">gas</span> transfer and wind speed with particular emphasizing variable and/or low wind speeds. The quadratic dependence is fit through <span class="hlt">gas</span>-transfer velocities over the ocean determined by methods based on the natural C-14 disequilibrium and the bomb C-14 inventory. The variation in the CO2 levels is related to these mechanisms, but the results show that other causes play significant roles. A weaker dependence of <span class="hlt">gas</span> transfer on wind is suggested for steady winds, and long-term averaged winds demonstrate a stronger dependence in the present model. The chemical enhancement of CO2 <span class="hlt">exchange</span> is also shown to play a role by increasing CO2 fluxes at low wind speeds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/53171','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/53171"><span>A dynamic leaf <span class="hlt">gas-exchange</span> strategy is conserved in woody plants under changing ambient CO2 : evidence from carbon isotope discrimination in paleo and CO2 enrichment studies</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Steven L. Voelker; J. Renee Brooks; Frederick C. Meinzer; Rebecca Anderson; Martin K.-F. Bader; Giovanna Battipaglia; Katie M. Becklin; David Beerling; Didier Bert; Julio L. Betancourt; Todd E. Dawson; Jean-Christophe Domec; Richard P. Guyette; Christian K??rner; Steven W. Leavitt; Sune Linder; John D. Marshall; Manuel Mildner; Jerome Ogee; Irina Panyushkina; Heather J. Plumpton; Kurt S. Pregitzer; Matthias Saurer; Andrew R. Smith; Rolf T. W. Siegwolf; Michael C. Stambaugh; Alan F. Talhelm; Jacques C. Tardif; Peter K. Van de Water; Joy K. Ward; Lisa Wingate</p> <p>2016-01-01</p> <p>Rising atmospheric [CO2], ca, is expected to affect stomatal regulation of leaf <span class="hlt">gas-exchange</span> of woody plants, thus influencing energy fluxes as well as carbon (C), <span class="hlt">water</span>, and nutrient cycling of forests. Researchers have proposed various strategies for stomatal regulation of leaf <span class="hlt">gas-exchange</span> that include maintaining a constant leaf internal [CO...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19760022616','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19760022616"><span>Hydrogen rich <span class="hlt">gas</span> generator</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Houseman, J. (Inventor)</p> <p>1976-01-01</p> <p>A process and apparatus is described for producing a hydrogen rich <span class="hlt">gas</span> by introducing a liquid hydrocarbon fuel in the form of a spray into a partial oxidation region and mixing with a mixture of steam and <span class="hlt">air</span> that is preheated by indirect heat <span class="hlt">exchange</span> with the formed hydrogen rich <span class="hlt">gas</span>, igniting the hydrocarbon fuel spray mixed with the preheated mixture of steam and <span class="hlt">air</span> within the partial oxidation region to form a hydrogen rich <span class="hlt">gas</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JASMS..28..971H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JASMS..28..971H"><span>Regio-Selective Intramolecular Hydrogen/Deuterium <span class="hlt">Exchange</span> in <span class="hlt">Gas</span>-Phase Electron Transfer Dissociation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hamuro, Yoshitomo</p> <p>2017-05-01</p> <p>Protein backbone amide hydrogen/deuterium <span class="hlt">exchange</span> mass spectrometry (HDX-MS) typically utilizes enzymatic digestion after the <span class="hlt">exchange</span> reaction and before MS analysis to improve data resolution. <span class="hlt">Gas</span>-phase fragmentation of a peptic fragment prior to MS analysis is a promising technique to further increase the resolution. The biggest technical challenge for this method is elimination of intramolecular hydrogen/deuterium <span class="hlt">exchange</span> (scrambling) in the <span class="hlt">gas</span> phase. The scrambling obscures the location of deuterium. Jørgensen's group pioneered a method to minimize the scrambling in <span class="hlt">gas</span>-phase electron capture/transfer dissociation. Despite active investigation, the mechanism of hydrogen scrambling is not well-understood. The difficulty stems from the fact that the degree of hydrogen scrambling depends on instruments, various parameters of mass analysis, and peptide analyzed. In most hydrogen scrambling investigations, the hydrogen scrambling is measured by the percentage of scrambling in a whole molecule. This paper demonstrates that the degree of intramolecular hydrogen/deuterium <span class="hlt">exchange</span> depends on the nature of <span class="hlt">exchangeable</span> hydrogen sites. The deuterium on Tyr amide of neurotensin (9-13), Arg-Pro-Tyr-Ile-Leu, migrated significantly faster than that on Ile or Leu amides, indicating the loss of deuterium from the original sites is not mere randomization of hydrogen and deuterium but more site-specific phenomena. This more precise approach may help understand the mechanism of intramolecular hydrogen <span class="hlt">exchange</span> and provide higher confidence for the parameter optimization to eliminate intramolecular hydrogen/deuterium <span class="hlt">exchange</span> during <span class="hlt">gas</span>-phase fragmentation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28194737','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28194737"><span>Regio-Selective Intramolecular Hydrogen/Deuterium <span class="hlt">Exchange</span> in <span class="hlt">Gas</span>-Phase Electron Transfer Dissociation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hamuro, Yoshitomo</p> <p>2017-05-01</p> <p>Protein backbone amide hydrogen/deuterium <span class="hlt">exchange</span> mass spectrometry (HDX-MS) typically utilizes enzymatic digestion after the <span class="hlt">exchange</span> reaction and before MS analysis to improve data resolution. <span class="hlt">Gas</span>-phase fragmentation of a peptic fragment prior to MS analysis is a promising technique to further increase the resolution. The biggest technical challenge for this method is elimination of intramolecular hydrogen/deuterium <span class="hlt">exchange</span> (scrambling) in the <span class="hlt">gas</span> phase. The scrambling obscures the location of deuterium. Jørgensen's group pioneered a method to minimize the scrambling in <span class="hlt">gas</span>-phase electron capture/transfer dissociation. Despite active investigation, the mechanism of hydrogen scrambling is not well-understood. The difficulty stems from the fact that the degree of hydrogen scrambling depends on instruments, various parameters of mass analysis, and peptide analyzed. In most hydrogen scrambling investigations, the hydrogen scrambling is measured by the percentage of scrambling in a whole molecule. This paper demonstrates that the degree of intramolecular hydrogen/deuterium <span class="hlt">exchange</span> depends on the nature of <span class="hlt">exchangeable</span> hydrogen sites. The deuterium on Tyr amide of neurotensin (9-13), Arg-Pro-Tyr-Ile-Leu, migrated significantly faster than that on Ile or Leu amides, indicating the loss of deuterium from the original sites is not mere randomization of hydrogen and deuterium but more site-specific phenomena. This more precise approach may help understand the mechanism of intramolecular hydrogen <span class="hlt">exchange</span> and provide higher confidence for the parameter optimization to eliminate intramolecular hydrogen/deuterium <span class="hlt">exchange</span> during <span class="hlt">gas</span>-phase fragmentation. Graphical Abstract ᅟ.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5228218','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5228218"><span>Impacts of Changes of Indoor <span class="hlt">Air</span> Pressure and <span class="hlt">Air</span> <span class="hlt">Exchange</span> Rate in Vapor Intrusion Scenarios</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Shen, Rui; Suuberg, Eric M.</p> <p>2016-01-01</p> <p>There has, in recent years, been increasing interest in understanding the transport processes of relevance in vapor intrusion of volatile organic compounds (VOCs) into buildings on contaminated sites. These studies have included fate and transport modeling. Most such models have simplified the prediction of indoor <span class="hlt">air</span> contaminant vapor concentrations by employing a steady state assumption, which often results in difficulties in reconciling these results with field measurements. This paper focuses on two major factors that may be subject to significant transients in vapor intrusion situations, including the indoor <span class="hlt">air</span> pressure and the <span class="hlt">air</span> <span class="hlt">exchange</span> rate in the subject building. A three-dimensional finite element model was employed with consideration of daily and seasonal variations in these factors. From the results, the variations of indoor <span class="hlt">air</span> pressure and <span class="hlt">air</span> <span class="hlt">exchange</span> rate are seen to contribute to significant variations in indoor <span class="hlt">air</span> contaminant vapor concentrations. Depending upon the assumptions regarding the variations in these parameters, the results are only sometimes consistent with the reports of several orders of magnitude in indoor <span class="hlt">air</span> concentration variations from field studies. The results point to the need to examine more carefully the interplay of these factors in order to quantitatively understand the variations in potential indoor <span class="hlt">air</span> exposures. PMID:28090133</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28090133','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28090133"><span>Impacts of Changes of Indoor <span class="hlt">Air</span> Pressure and <span class="hlt">Air</span> <span class="hlt">Exchange</span> Rate in Vapor Intrusion Scenarios.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shen, Rui; Suuberg, Eric M</p> <p>2016-02-01</p> <p>There has, in recent years, been increasing interest in understanding the transport processes of relevance in vapor intrusion of volatile organic compounds (VOCs) into buildings on contaminated sites. These studies have included fate and transport modeling. Most such models have simplified the prediction of indoor <span class="hlt">air</span> contaminant vapor concentrations by employing a steady state assumption, which often results in difficulties in reconciling these results with field measurements. This paper focuses on two major factors that may be subject to significant transients in vapor intrusion situations, including the indoor <span class="hlt">air</span> pressure and the <span class="hlt">air</span> <span class="hlt">exchange</span> rate in the subject building. A three-dimensional finite element model was employed with consideration of daily and seasonal variations in these factors. From the results, the variations of indoor <span class="hlt">air</span> pressure and <span class="hlt">air</span> <span class="hlt">exchange</span> rate are seen to contribute to significant variations in indoor <span class="hlt">air</span> contaminant vapor concentrations. Depending upon the assumptions regarding the variations in these parameters, the results are only sometimes consistent with the reports of several orders of magnitude in indoor <span class="hlt">air</span> concentration variations from field studies. The results point to the need to examine more carefully the interplay of these factors in order to quantitatively understand the variations in potential indoor <span class="hlt">air</span> exposures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16665243','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16665243"><span>Photosynthesis Decrease and Stomatal Control of <span class="hlt">Gas</span> <span class="hlt">Exchange</span> in Abies alba Mill. in Response to Vapor Pressure Difference.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Guehl, J M; Aussenac, G</p> <p>1987-02-01</p> <p>The responses of steady state CO(2) assimilation rate (A), transpiration rate (E), and stomatal conductance (g(s)) to changes in leaf-to-<span class="hlt">air</span> vapor pressure difference (DeltaW) were examined on different dates in shoots from Abies alba trees growing outside. In Ecouves, a provenance representative of wet oceanic conditions in Northern France, both A and g(s) decreased when DeltaW was increased from 4.6 to 14.5 Pa KPa(-1). In Nebias, which represented the dry end of the natural range of A. alba in southern France, A and g(s) decreased only after reaching peak levels at 9.0 and 7.0 Pa KPa(-1), respectively. The representation of the data in assimilation rate (A) versus intercellular CO(2) partial pressure (C(i)) graphs allowed us to determine how stomata and mesophyll photosynthesis interacted when DeltaW was increased. Changes in A were primarily due to alterations in mesophyll photosynthesis. At high DeltaW, and especially in Ecouves when soil <span class="hlt">water</span> deficit prevailed, A declined, while C(i) remained approximately constant, which may be interpreted as an adjustment of g(s) to changes in mesophyll photosynthesis. Such a stomatal control of <span class="hlt">gas</span> <span class="hlt">exchange</span> appeared as an alternative to the classical feedforward interpretation of E versus DeltaW responses with a peak rate of E. The <span class="hlt">gas</span> <span class="hlt">exchange</span> response to DeltaW was also characterized by considerable deviations from the optimization theory of IR Cowan and GD Farquhar (1977 Symp Soc Exp Biol 31: 471-505).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.B51H1919C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.B51H1919C"><span>Tracking Seasonal and Diurnal Photosynthesis and Plant <span class="hlt">Water</span> Status in Maize Using SIF, Eddy Covariance Fluxes, PAM Fluorescence and <span class="hlt">Gas</span> <span class="hlt">Exchange</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chang, C.; Melkonian, J.; Riha, S. J.; Gu, L.; Sun, Y.</p> <p>2017-12-01</p> <p>Improving the sensitivity of methods for crop monitoring and yield forecasting is crucial as the frequency of extreme weather events increases. Conventional remote monitoring methods rely on greenness-based indices such as NDVI and EVI, which do not directly measure photosynthesis and are not sufficiently sensitive to rapid plant stress response. Solar-induced chlorophyll fluorescence (SIF) is a promising new technology that serves as a direct functional proxy of photosynthesis. We developed the first system utilizing dual QE Pro spectrometers to continuously measure the diurnal and seasonal cycle of SIF, and deployed the system in a corn field in upstate New York in 2017. To complement SIF, canopy-level measurements of carbon and <span class="hlt">water</span> fluxes were also measured, along with concurrent leaf-level measurements of <span class="hlt">gas</span> <span class="hlt">exchange</span> and PAM fluorescence, midday <span class="hlt">water</span> potential, leaf pigments, phenology, LAI, and soil moisture. We show that SIF is well correlated to GPP during the growing season and show that both are controlled by similar environmental conditions including PAR and <span class="hlt">water</span> availability. We also describe diurnal changes in photosynthesis and plant <span class="hlt">water</span> status and demonstrate the sensitivity of SIF to diurnal plant response.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26257361','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26257361"><span>Interactive response of photosynthetic characteristics in Haloxylon ammodendron and Hedysarum scoparium exposed to soil <span class="hlt">water</span> and <span class="hlt">air</span> vapor pressure deficits.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gong, Chunmei; Wang, Jiajia; Hu, Congxia; Wang, Junhui; Ning, Pengbo; Bai, Juan</p> <p>2015-08-01</p> <p>C4 plants possess better drought tolerance than C3 plants. However, Hedysarum scoparium, a C3 species, is dominant and widely distributed in the desert areas of northwestern China due to its strong drought tolerance. This study compared it with Haloxylon ammodendron, a C4 species, regarding the interactive effects of drought stress and different leaf-<span class="hlt">air</span> vapor pressure deficits. Variables of interest included <span class="hlt">gas</span> <span class="hlt">exchange</span>, the activity levels of key C4 photosynthetic enzymes, and cellular anatomy. In both species, <span class="hlt">gas</span> <span class="hlt">exchange</span> parameters were more sensitive to high vapor pressure deficit than to strong <span class="hlt">water</span> stress, and the net CO2 assimilation rate (An) was enhanced as vapor pressure deficits increased. A close relationship between An and stomatal conductance (gs) suggested that the species shared a similar response mechanism. In H. ammodendron, the activity levels of key C4 enzymes were higher, including those of phosphoenolpyruvate carboxylase (PEPC) and nicotinamide adenine dinucleotide phosphate-malate enzyme (NADP-ME), whereas in H. scoparium, the activity level of nicotinamide adenine dinucleotide-malate enzyme (NAD-ME) was higher. Meanwhile, H. scoparium utilized adaptive structural features, including a larger relative vessel area and a shorter distance from vein to stomata, which facilitated the movement of <span class="hlt">water</span>. These findings implied that some C4 biochemical pathways were present in H. scoparium to respond to environmental challenges. Copyright © 2015. Published by Elsevier B.V.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25355625','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25355625"><span><span class="hlt">Air</span> concentrations of volatile compounds near oil and <span class="hlt">gas</span> production: a community-based exploratory study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Macey, Gregg P; Breech, Ruth; Chernaik, Mark; Cox, Caroline; Larson, Denny; Thomas, Deb; Carpenter, David O</p> <p>2014-10-30</p> <p>Horizontal drilling, hydraulic fracturing, and other drilling and well stimulation technologies are now used widely in the United States and increasingly in other countries. They enable increases in oil and <span class="hlt">gas</span> production, but there has been inadequate attention to human health impacts. <span class="hlt">Air</span> quality near oil and <span class="hlt">gas</span> operations is an underexplored human health concern for five reasons: (1) prior focus on threats to <span class="hlt">water</span> quality; (2) an evolving understanding of contributions of certain oil and <span class="hlt">gas</span> production processes to <span class="hlt">air</span> quality; (3) limited state <span class="hlt">air</span> quality monitoring networks; (4) significant variability in <span class="hlt">air</span> emissions and concentrations; and (5) <span class="hlt">air</span> quality research that misses impacts important to residents. Preliminary research suggests that volatile compounds, including hazardous <span class="hlt">air</span> pollutants, are of potential concern. This study differs from prior research in its use of a community-based process to identify sampling locations. Through this approach, we determine concentrations of volatile compounds in <span class="hlt">air</span> near operations that reflect community concerns and point to the need for more fine-grained and frequent monitoring at points along the production life cycle. Grab and passive <span class="hlt">air</span> samples were collected by trained volunteers at locations identified through systematic observation of industrial operations and <span class="hlt">air</span> impacts over the course of resident daily routines. A total of 75 volatile organics were measured using EPA Method TO-15 or TO-3 by <span class="hlt">gas</span> chromatography/mass spectrometry. Formaldehyde levels were determined using UMEx 100 Passive Samplers. Levels of eight volatile chemicals exceeded federal guidelines under several operational circumstances. Benzene, formaldehyde, and hydrogen sulfide were the most common compounds to exceed acute and other health-based risk levels. <span class="hlt">Air</span> concentrations of potentially dangerous compounds and chemical mixtures are frequently present near oil and <span class="hlt">gas</span> production sites. Community-based research can provide an</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16508435','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16508435"><span>Retinal damage caused by <span class="hlt">air</span>-fluid <span class="hlt">exchange</span> during pars plana vitrectomy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Sam S; McDonald, H Richard; Everett, A I; Johnson, Robert N; Jumper, J Michael; Fu, Arthur D</p> <p>2006-03-01</p> <p>To report two cases of retinal damage associated with <span class="hlt">air</span> infusion during pars plana vitrectomy. Observational case report. The authors reviewed the course of two patients who had retinal damage during par plana vitrectomy and <span class="hlt">air</span>-fluid <span class="hlt">exchange</span> for the treatment of macular hole and optic pit-related macular detachment, respectively. The intraoperative observations, postoperative course, and outcomes were reported. As a result of high <span class="hlt">air</span> infusion flow during <span class="hlt">air</span>-fluid <span class="hlt">exchange</span>, retinal damage was created in the area contralateral to the infusion port. In Case 1, an oval area of whitening was noted on the first postoperative day. This area subsequently developed into a large retinal break associated with retinal detachment. In the second case, retinal whitening was noted intraoperatively. This region of pallor resolved quickly during the early postoperative period but resulted in a corresponding inferotemporal visual field defect. High infusion flow during <span class="hlt">air</span>-fluid <span class="hlt">exchange</span> in eyes undergoing vitrectomy surgery may result in significant retinal damage. This pressure-induced trauma initially causes retinal whitening that may be seen intraoperatively or during the early postoperative period. The region of damaged retina may develop a retinal break and detachment or a corresponding visual field defect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2012-07-30/pdf/2012-18434.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2012-07-30/pdf/2012-18434.pdf"><span>77 FR 44672 - Notice of Lodging of Consent Decree Under the Clean <span class="hlt">Water</span> and Clean <span class="hlt">Air</span> Acts</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2012-07-30</p> <p>... DEPARTMENT OF JUSTICE Notice of Lodging of Consent Decree Under the Clean <span class="hlt">Water</span> and Clean <span class="hlt">Air</span> Acts... a civil penalty of $1,750,000 to resolve its violations of the Clean <span class="hlt">Air</span> Act and the Clean <span class="hlt">Water</span> Act... of coke oven <span class="hlt">gas</span>. Under the Clean <span class="hlt">Water</span> Act, Plaintiffs allege that Shenango violated the effluent...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27835767','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27835767"><span>Improving respiration measurements with <span class="hlt">gas</span> <span class="hlt">exchange</span> analyzers.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Montero, R; Ribas-Carbó, M; Del Saz, N F; El Aou-Ouad, H; Berry, J A; Flexas, J; Bota, J</p> <p>2016-12-01</p> <p>Dark respiration measurements with open-flow <span class="hlt">gas</span> <span class="hlt">exchange</span> analyzers are often questioned for their low accuracy as their low values often reach the precision limit of the instrument. Respiration was measured in five species, two hypostomatous (Vitis Vinifera L. and Acanthus mollis) and three amphistomatous, one with similar amount of stomata in both sides (Eucalyptus citriodora) and two with different stomata density (Brassica oleracea and Vicia faba). CO 2 differential (ΔCO 2 ) increased two-fold with no change in apparent R d , when the two leaves with higher stomatal density faced outside. These results showed a clear effect of the position of stomata on ΔCO 2 . Therefore, it can be concluded that leaf position is important to guarantee the improvement of respiration measurements increasing ΔCO 2 without affecting the respiration results by leaf or mass units. This method will help to increase the accuracy of leaf respiration measurements using <span class="hlt">gas</span> <span class="hlt">exchange</span> analyzers. Copyright © 2016 Elsevier GmbH. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25783787','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25783787"><span>Development of <span class="hlt">gas</span> <span class="hlt">exchange</span> and ion regulation in two species of <span class="hlt">air</span>-breathing fish, Betta splendens and Macropodus opercularis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huang, Chun-Yen; Lin, Cheng-Huang; Lin, Hui-Chen</p> <p>2015-07-01</p> <p>Aquatic <span class="hlt">air</span>-breathing anabantoids, a group of fish species characterized by the presence of a labyrinth organ and some gills, exhibit morphological variations. This study aimed to examine whether unequal gill growth begins during the early stages and described the sequence of the early gill developmental events in Betta splendens and Macropodus opercularis. To determine when the ion regulatory and <span class="hlt">gas</span> <span class="hlt">exchange</span> abilities first appear in the gills, mitochondria-rich cells (MRCs) and neuroepithelial cells (NECs) were examined in young B. splendens. To evaluate the relative importance of the gills and the labyrinth organ under different levels of oxygen uptake stress, the levels of carbonic anhydrase II (CAII) and Na(+)/K(+)-ATPase (NKA) protein expressions in 2 gills and the labyrinth organ were examined in M. opercularis. We found that the first 3 gills developed earlier than the 4th gill in both species, an indication that the morphological variation begins early in life. In B. splendens, the MRCs and NECs clearly appeared in the first 3 gills at 4 dph and were first found in the 4th gill until 11 dph. The oxygen-sensing ability of the gills was concordant with the ionoregulatory function. In M. opercularis, the hypoxic group had a significantly higher <span class="hlt">air</span>-breathing frequency. CAII protein expression was higher in the labyrinth organ in the hypoxic group. The gills exhibited increased NKA protein expression in the hypoxic and restricted groups, respectively. Functional plasticity in CAII and NKA protein expressions was found between the gills and the labyrinth organ in adult M. opercularis. Copyright © 2015 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26188268','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26188268"><span>Hypoxia and hypercarbia in endophagous insects: Larval position in the plant <span class="hlt">gas</span> <span class="hlt">exchange</span> network is key.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pincebourde, Sylvain; Casas, Jérôme</p> <p>2016-01-01</p> <p><span class="hlt">Gas</span> composition is an important component of any micro-environment. Insects, as the vast majority of living organisms, depend on O2 and CO2 concentrations in the <span class="hlt">air</span> they breathe. Low O2 (hypoxia), and high CO2 (hypercarbia) levels can have a dramatic effect. For phytophagous insects that live within plant tissues (endophagous lifestyle), <span class="hlt">gas</span> is <span class="hlt">exchanged</span> between ambient <span class="hlt">air</span> and the atmosphere within the insect habitat. The insect larva contributes to the modification of this environment by expiring CO2. Yet, knowledge on the <span class="hlt">gas</span> <span class="hlt">exchange</span> network in endophagous insects remains sparse. Our study identified mechanisms that modulate <span class="hlt">gas</span> composition in the habitat of endophagous insects. Our aim was to show that the mere position of the insect larva within plant tissues could be used as a proxy for estimating risk of occurrence of hypoxia and hypercarbia, despite the widely diverse life history traits of these organisms. We developed a conceptual framework for a <span class="hlt">gas</span> diffusion network determining <span class="hlt">gas</span> composition in endophagous insect habitats. We applied this framework to mines, galls and insect tunnels (borers) by integrating the numerous obstacles along O2 and CO2 pathways. The nature and the direction of <span class="hlt">gas</span> transfers depended on the physical structure of the insect habitat, the photosynthesis activity as well as stomatal behavior in plant tissues. We identified the insect larva position within the <span class="hlt">gas</span> diffusion network as a predictor of risk exposure to hypoxia and hypercarbia. We ranked endophagous insect habitats in terms of risk of exposure to hypoxia and/or hypercarbia, from the more to the less risky as cambium mines>borer tunnels≫galls>bark mines>mines in aquatic plants>upper and lower surface mines. Furthermore, we showed that the photosynthetically active tissues likely assimilate larval CO2 produced. In addition, temperature of the microhabitat and atmospheric CO2 alter <span class="hlt">gas</span> composition in the insect habitat. We predict that (i) hypoxia indirectly favors</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1084048','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1084048"><span>Coaxial fuel and <span class="hlt">air</span> premixer for a <span class="hlt">gas</span> turbine combustor</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>York, William D; Ziminsky, Willy S; Lacy, Benjamin P</p> <p>2013-05-21</p> <p>An <span class="hlt">air</span>/fuel premixer comprising a peripheral wall defining a mixing chamber, a nozzle disposed at least partially within the peripheral wall comprising an outer annular wall spaced from the peripheral wall so as to define an outer <span class="hlt">air</span> passage between the peripheral wall and the outer annular wall, an inner annular wall disposed at least partially within and spaced from the outer annular wall, so as to define an inner <span class="hlt">air</span> passage, and at least one fuel <span class="hlt">gas</span> annulus between the outer annular wall and the inner annular wall, the at least one fuel <span class="hlt">gas</span> annulus defining at least one fuel <span class="hlt">gas</span> passage, at least one <span class="hlt">air</span> inlet for introducing <span class="hlt">air</span> through the inner <span class="hlt">air</span> passage and the outer <span class="hlt">air</span> passage to the mixing chamber, and at least one fuel inlet for injecting fuel through the fuel <span class="hlt">gas</span> passage to the mixing chamber to form an <span class="hlt">air</span>/fuel mixture.</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('https://www.ncbi.nlm.nih.gov/pubmed/21941230','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21941230"><span>Sildenafil citrate, bronchopulmonary dysplasia and disordered pulmonary <span class="hlt">gas</span> <span class="hlt">exchange</span>: any benefits?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nyp, M; Sandritter, T; Poppinga, N; Simon, C; Truog, W E</p> <p>2012-01-01</p> <p>The objective of this study is to determine the effects that sildenafil citrate has on <span class="hlt">gas</span> <span class="hlt">exchange</span> in infants with bronchopulmonary dysplasia (BPD)-associated pulmonary hypertension (PH). A retrospective review was performed from 2005 to 2009. Infants treated with sildenafil citrate for greater than 48  h were included. Standard patient data was collected, including echocardiogram, inspired oxygen and systemic blood pressure, before and during administration of sildenafil citrate. Sildenafil citrate was used in 21 preterm infants with BPD-associated PH. A significant reduction in estimated right ventricular peak systolic pressure was seen after initiation of sildenafil citrate, with the majority of infants showing no improvement in <span class="hlt">gas</span> <span class="hlt">exchange</span> at 48  h of treatment. Four infants died during treatment. Sildenafil citrate reduced estimated pulmonary artery pressures, but this reduction was not reflected in improved <span class="hlt">gas</span> <span class="hlt">exchange</span> within the first 48  h.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25597683','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25597683"><span><span class="hlt">Air</span>-soil <span class="hlt">exchange</span> of organochlorine pesticides in a sealed chamber.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Bing; Han, Baolu; Xue, Nandong; Zhou, Lingli; Li, Fasheng</p> <p>2015-01-01</p> <p>So far little is known about <span class="hlt">air</span>-soil <span class="hlt">exchange</span> under any sealed circumstances (e.g., in plastic and glass sheds), which however has huge implications for the soil-<span class="hlt">air</span>-plant pathways of persistent organic pollutants including organochlorine pesticides (OCPs). A newly designed passive <span class="hlt">air</span> sampler was tested in a sealed chamber for measuring the vertical concentration profiles of gaseous phase OCPs (hexachlorocyclohexanes (HCHs) and dichlorodiphenyltrichloroethanes (DDTs)). <span class="hlt">Air</span> was sampled at 5, 15, and 30 cm above ground level every 10th day during a 60-day period by deploying polyurethane foam cylinders housed in acrylonitrile butadiene styrene-covered cartridges. Concentrations and compositions of OCPs along the vertical sections indicated a clear relationship with proximity to the mixture of HCHs and DDTs which escapes from the soils. In addition, significant positive correlations were found between <span class="hlt">air</span> temperatures and concentrations of HCHs and DDTs. These results indicated revolatilization and re-deposition being at or close to dynamic pseudo-equilibrium with the overlying <span class="hlt">air</span>. The sampler used for addressing <span class="hlt">air</span>-soil <span class="hlt">exchange</span> of persistent organic pollutants in any sealed conditions is discussed. Copyright © 2014. Published by Elsevier B.V.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H11A1162P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H11A1162P"><span>High-resolution (noble) <span class="hlt">gas</span> time series for aquatic research</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Popp, A. L.; Brennwald, M. S.; Weber, U.; Kipfer, R.</p> <p>2017-12-01</p> <p>We developed a portable mass spectrometer (miniRUEDI) for on-site quantification of <span class="hlt">gas</span> concentrations (He, Ar, Kr, N2, O2, CO2, CH4, etc.) in terrestrial gases [1,2]. Using the <span class="hlt">gas</span>-equilibrium membrane-inlet technique (GE-MIMS), the miniRUEDI for the first time also allows accurate on-site and long-term dissolved-<span class="hlt">gas</span> analysis in <span class="hlt">water</span> bodies. The miniRUEDI is designed for operation in the field and at remote locations, using battery power and ambient <span class="hlt">air</span> as a calibration <span class="hlt">gas</span>. In contrast to conventional sampling and subsequent lab analysis, the miniRUEDI provides real-time and continuous time series of <span class="hlt">gas</span> concentrations with a time resolution of a few seconds.Such high-resolution time series and immediate data availability open up new opportunities for research in highly dynamic and heterogeneous environmental systems. In addition the combined analysis of inert and reactive <span class="hlt">gas</span> species provides direct information on the linkages of physical and biogoechemical processes, such as the <span class="hlt">air/water</span> <span class="hlt">gas</span> <span class="hlt">exchange</span>, excess <span class="hlt">air</span> formation, O2 turnover, or N2 production by denitrification [1,3,4].We present the miniRUEDI instrument and discuss its use for environmental research based on recent applications of tracking <span class="hlt">gas</span> dynamics related to rapid and short-term processes in aquatic systems. [1] Brennwald, M.S., Schmidt, M., Oser, J., and Kipfer, R. (2016). Environmental Science and Technology, 50(24):13455-13463, doi: 10.1021/acs.est.6b03669[2] Gasometrix GmbH, gasometrix.com[3] Mächler, L., Peter, S., Brennwald, M.S., and Kipfer, R. (2013). Excess <span class="hlt">air</span> formation as a mechanism for delivering oxygen to groundwater. <span class="hlt">Water</span> Resources Research, doi:10.1002/wrcr.20547[4] Mächler, L., Brennwald, M.S., and Kipfer, R. (2013). Argon Concentration Time-Series As a Tool to Study <span class="hlt">Gas</span> Dynamics in the Hyporheic Zone. Environmental Science and Technology, doi: 10.1021/es305309b</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.afdc.energy.gov/case/1','SCIGOVWS'); return false;" href="https://www.afdc.energy.gov/case/1"><span>Alternative Fuels Data Center: Natural <span class="hlt">Gas</span> Street Sweepers Improve <span class="hlt">Air</span></span></a></p> <p><a target="_blank" href="http://www.science.gov/aboutsearch.html">Science.gov Websites</a></p> <p></p> <p></p> <p>Quality in New York</A> Natural <em><span class="hlt">Gas</span></em> Street Sweepers Improve <span class="hlt">Air</span> Quality in New York to someone by E -mail Share Alternative Fuels Data Center: Natural <em><span class="hlt">Gas</span></em> Street Sweepers Improve <span class="hlt">Air</span> Quality in New York on Facebook Tweet about Alternative Fuels Data Center: Natural <em><span class="hlt">Gas</span></em> Street Sweepers Improve <span class="hlt">Air</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22429767-highly-tritiated-water-processing-isotopic-exchange','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22429767-highly-tritiated-water-processing-isotopic-exchange"><span>Highly tritiated <span class="hlt">water</span> processing by isotopic <span class="hlt">exchange</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>Shu, W.M.; Willms, R.S.; Glugla, M.</p> <p>2015-03-15</p> <p>Highly tritiated <span class="hlt">water</span> (HTW) is produced in fusion machines and one of the promising technologies to process it is isotopic <span class="hlt">exchange</span>. 3 kinds of Pt-catalyzed zeolite (13X-APG, CBV-100-CY and HiSiv-1000) were tested as candidates for isotopic <span class="hlt">exchange</span> of highly tritiated <span class="hlt">water</span> (HTW), and CBV-100-CY (Na-Y type with a SiO{sub 2}/Al{sub 2}O{sub 3} ratio of ∼ 5.0) shows the best performance. Small-scale tritium testing indicates that this method is efficient for reaching an <span class="hlt">exchange</span> factor (EF) of 100. Full-scale non-tritium testing implies that an EF of 300 can be achieved in 24 hours of operation if a temperature gradient is appliedmore » along the column. For the isotopic <span class="hlt">exchange</span>, deuterium recycled from the Isotope Separation System (deuterium with 1% T and/or 200 ppm T) should be employed, and the tritiated <span class="hlt">water</span> regenerated from the Pt-catalyzed zeolite bed after isotopic <span class="hlt">exchange</span> should be transferred to <span class="hlt">Water</span> Detritiation System (WDS) for further processing.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1851b0077D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1851b0077D"><span>Performance evaluation of cross-flow single-phase liquid-to-<span class="hlt">gas</span> polymer tube heat <span class="hlt">exchanger</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dewanjee, Sujan; Hossain, Md. Rakibul; Rahman, Md. Ashiqur</p> <p>2017-06-01</p> <p>Reduced core weight and material cost, higher corrosion resistance are some of the major eye catching properties to study polymers over metal in heat <span class="hlt">exchanger</span> applications in spite of the former's relatively low thermal conductivity and low strength. In the present study, performance of polymer parallel thin tube heat <span class="hlt">exchanger</span> is numerically evaluated for cross flow liquid to <span class="hlt">air</span> applications for a wide range of design and operating parameters such as tube diameter, thickness, fluid velocity and temperature, etc. using Computational Fluid Dynamics (CFD). Among a range of available polymeric materials, those with a moderate to high thermal conductivity and strength are selected for this study. A 90 cm × 1 cm single unit of polymer tubes, with appropriate number of tubes such that at least a gap of 5 mm is maintained in between the tubes, is used as a basic unit and multiple combination in the transverse direction of this single unit is simulated to measure the effect. The tube inner diameter is varied from 2 mm to 4 mm and the pressure drop is measured to have a relative idea of pumping cost. For each inner diameter the thickness is varied from .5 mm to 2.5 mm. The <span class="hlt">water</span> velocity and the <span class="hlt">air</span> velocity are varied from 0.4 m/s to 2 m/s and 1 m/s to 5 m/s, respectively. The performance of the polymer heat <span class="hlt">exchanger</span> is compared with that of metal heat <span class="hlt">exchanger</span> through and an optimum design for polymer heat <span class="hlt">exchanger</span> is sought out.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18751738','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18751738"><span>Pulvinus activity, leaf movement and leaf <span class="hlt">water</span>-use efficiency of bush bean (Phaseplus vulgaris L.) in a hot environment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Raeini-Sarjaz, Mahmoud; Chalavi, Vida</p> <p>2008-11-01</p> <p>Pulvinus activity of Phaseolus species in response to environmental stimuli plays an essential role in heliotropic leaf movement. The aims of this study were to monitor the continuous daily pulvinus movement and pulvinus temperature, and to evaluate the effects of leaf movements, on a hot day, on instantaneous leaf <span class="hlt">water</span>-use efficiency (WUEi), leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>, and leaf temperature. Potted plants of Phaseolus vulgaris L. var. Provider were grown in Chicot sandy loam soil under well-<span class="hlt">watered</span> conditions in a greenhouse. When the second trifoliate leaf was completely extended, one plant was selected to measure pulvinus movement using a beta-ray gauging (BRG) meter with a point source of thallium-204 (204Tl). Leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements took place on similar leaflets of three plants at an <span class="hlt">air</span> temperature interval of 33-42 degrees C by a steady-state LI-6200 photosynthesis system. A copper-constantan thermocouple was used to monitor pulvinus temperature. Pulvinus bending followed the daily diurnal rhythm. Significant correlations were found between the leaf-incident angle and the stomatal conductance (R2 = 0.54; P < 0.01), and photosynthesis rate (R2 = 0.84; P < 0.01). With a reduction in leaf-incidence angle and increase in <span class="hlt">air</span> temperature, WUEi was reduced. During the measurements, leaf temperature remained below <span class="hlt">air</span> temperature and was a significant function of <span class="hlt">air</span> temperature (r = 0.92; P < 0.01). In conclusion, pulvinus bending followed both light intensity and <span class="hlt">air</span> temperature and influenced leaf <span class="hlt">gas</span> <span class="hlt">exchange</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70157933','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70157933"><span>Natural radium and radon tracers to quantify <span class="hlt">water</span> <span class="hlt">exchange</span> and movement in reservoirs</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Smith, Christopher G.; Baskaran, Mark</p> <p>2011-01-01</p> <p>Radon and radium isotopes are routinely used to quantify <span class="hlt">exchange</span> rates between different hydrologic reservoirs. Since their recognition as oceanic tracers in the 1960s, both radon and radium have been used to examine processes such as <span class="hlt">air</span>-sea <span class="hlt">exchange</span>, deep oceanic mixing, benthic inputs, and many others. Recently, the application of radon-222 and the radium-quartet (223,224,226,228Ra) as coastal tracers has seen a revelation with the growing interest in coastal groundwater dynamics. The enrichment of these isotopes in benthic fluids including groundwater makes both radium and radon ideal tracers of coastal benthic processes (e.g. submarine groundwater discharge). In this chapter we review traditional and recent advances in the application of radon and radium isotopes to understand mixing and <span class="hlt">exchange</span> between various hydrologic reservoirs, specifically: (1) atmosphere and ocean, (2) deep and shallow oceanic <span class="hlt">water</span> masses, (3) coastal groundwater/benthic pore <span class="hlt">waters</span> and surface ocean, and (4) aquifer-lakes. While the isotopes themselves and their distribution in the environment provide qualitative information about the <span class="hlt">exchange</span> processes, it is mixing/<span class="hlt">exchange</span> and transport models for these isotopes that provide specific quantitative information about these processes. Brief introductions of these models and mixing parameters are provided for both historical and more recent studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3673973','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3673973"><span>Oxygen and <span class="hlt">Air</span> Nanobubble <span class="hlt">Water</span> Solution Promote the Growth of Plants, Fishes, and Mice</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ebina, Kosuke; Shi, Kenrin; Hirao, Makoto; Hashimoto, Jun; Kawato, Yoshitaka; Kaneshiro, Shoichi; Morimoto, Tokimitsu; Koizumi, Kota; Yoshikawa, Hideki</p> <p>2013-01-01</p> <p>Nanobubbles (<200 nm in diameter) have several unique properties such as long lifetime in liquid owing to its negatively charged surface, and its high <span class="hlt">gas</span> solubility into the liquid owing to its high internal pressure. They are used in variety of fields including diagnostic aids and drug delivery, while there are no reports assessing their effects on the growth of lives. Nanobubbles of <span class="hlt">air</span> or oxygen <span class="hlt">gas</span> were generated using a nanobubble aerator (BUVITAS; Ligaric Company Limited, Osaka, Japan). Brassica campestris were cultured hydroponically for 4 weeks within <span class="hlt">air</span>-nanobubble <span class="hlt">water</span> or within normal <span class="hlt">water</span>. Sweetfish (for 3 weeks) and rainbow trout (for 6 weeks) were kept either within <span class="hlt">air</span>-nanobubble <span class="hlt">water</span> or within normal <span class="hlt">water</span>. Finally, 5 week-old male DBA1/J mice were bred with normal free-chaw and free-drinking either of oxygen-nanobubble <span class="hlt">water</span> or of normal <span class="hlt">water</span> for 12 weeks. Oxygen-nanobubble significantly increased the dissolved oxygen concentration of <span class="hlt">water</span> as well as concentration/size of nanobubbles which were relatively stable for 70 days. <span class="hlt">Air</span>-nanobubble <span class="hlt">water</span> significantly promoted the height (19.1 vs. 16.7 cm; P<0.05), length of leaves (24.4 vs. 22.4 cm; P<0.01), and aerial fresh weight (27.3 vs. 20.3 g; P<0.01) of Brassica campestris compared to normal <span class="hlt">water</span>. Total weight of sweetfish increased from 3.0 to 6.4 kg in normal <span class="hlt">water</span>, whereas it increased from 3.0 to 10.2 kg in <span class="hlt">air</span>-nanobubble <span class="hlt">water</span>. In addition, total weight of rainbow trout increased from 50.0 to 129.5 kg in normal <span class="hlt">water</span>, whereas it increased from 50.0 to 148.0 kg in <span class="hlt">air</span>-nanobubble <span class="hlt">water</span>. Free oral intake of oxygen-nanobubble <span class="hlt">water</span> significantly promoted the weight (23.5 vs. 21.8 g; P<0.01) and the length (17.0 vs. 16.1 cm; P<0.001) of mice compared to that of normal <span class="hlt">water</span>. We have demonstrated for the first time that oxygen and <span class="hlt">air</span>-nanobubble <span class="hlt">water</span> may be potentially effective tools for the growth of lives. PMID:23755221</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=307512&Lab=NRMRL&keyword=jet+OR+turbine+OR+turbo+OR+turbofan+OR+turbojet+AND+aerofoil+OR+aerofoils+OR+airfoil+OR+airfoils+OR+blade+OR+blades+OR+vane+OR+vanes+AND+aeroengine+OR+aeronautical+OR+aeroturbine+OR+aircraft+OR+aviation+AND+engine&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=307512&Lab=NRMRL&keyword=jet+OR+turbine+OR+turbo+OR+turbofan+OR+turbojet+AND+aerofoil+OR+aerofoils+OR+airfoil+OR+airfoils+OR+blade+OR+blades+OR+vane+OR+vanes+AND+aeroengine+OR+aeronautical+OR+aeroturbine+OR+aircraft+OR+aviation+AND+engine&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>Intensive measurements of <span class="hlt">gas</span>, <span class="hlt">water</span>, and energy <span class="hlt">exchange</span> between vegetation and troposphere during the MONTES Campaign in a vegetation gradient from short semi-desertic shrublands to tall wet temperate forests in the NW Mediterranean basin</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>MONTES (“Woodlands”) was a multidisciplinary international field campaign aimed at measuring energy, <span class="hlt">water</span> and especially <span class="hlt">gas</span> <span class="hlt">exchange</span> between vegetation and atmosphere in a gradient from short semi-desertic shrublands to tall wet temperate forests in NE Spain in the North Wester...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10926641','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10926641"><span>Response-time enhancement of a clinical <span class="hlt">gas</span> analyzer facilitates measurement of breath-by-breath <span class="hlt">gas</span> <span class="hlt">exchange</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Farmery, A D; Hahn, C E</p> <p>2000-08-01</p> <p>Tidal ventilation <span class="hlt">gas-exchange</span> models in respiratory physiology and medicine not only require solution of mass balance equations breath-by-breath but also may require within-breath measurements, which are instantaneous functions of time. This demands a degree of temporal resolution and fidelity of integration of <span class="hlt">gas</span> flow and concentration signals that cannot be provided by most clinical <span class="hlt">gas</span> analyzers because of their slow response times. We have characterized the step responses of the Datex Ultima (Datex Instrumentation, Helsinki, Finland) <span class="hlt">gas</span> analyzer to oxygen, carbon dioxide, and nitrous oxide in terms of a Gompertz four-parameter sigmoidal function. By inversion of this function, we were able to reduce the rise times for all these gases almost fivefold, and, by its application to real on-line respiratory <span class="hlt">gas</span> signals, it is possible to achieve a performance comparable to the fastest mass spectrometers. With the use of this technique, measurements required for non-steady-state and tidal <span class="hlt">gas-exchange</span> models can be made easily and reliably in the clinical setting.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhDT........99M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhDT........99M"><span>Proton Transfers at the <span class="hlt">Air-Water</span> Interface</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mishra, Himanshu</p> <p></p> <p>Proton transfer reactions at the interface of <span class="hlt">water</span> with hydrophobic media, such as <span class="hlt">air</span> or lipids, are ubiquitous on our planet. These reactions orchestrate a host of vital phenomena in the environment including, for example, acidification of clouds, enzymatic catalysis, chemistries of aerosol and atmospheric gases, and bioenergetic transduction. Despite their importance, however, quantitative details underlying these interactions have remained unclear. Deeper insight into these interfacial reactions is also required in addressing challenges in green chemistry, improved <span class="hlt">water</span> quality, self-assembly of materials, the next generation of micro-nanofluidics, adhesives, coatings, catalysts, and electrodes. This thesis describes experimental and theoretical investigation of proton transfer reactions at the <span class="hlt">air-water</span> interface as a function of hydration gradients, electrochemical potential, and electrostatics. Since emerging insights hold at the lipid-<span class="hlt">water</span> interface as well, this work is also expected to aid understanding of complex biological phenomena associated with proton migration across membranes. Based on our current understanding, it is known that the physicochemical properties of the <span class="hlt">gas</span>-phase <span class="hlt">water</span> are drastically different from those of bulk <span class="hlt">water</span>. For example, the <span class="hlt">gas</span>-phase hydronium ion, H3O +(g), can protonate most (non-alkane) organic species, whereas H 3O+(aq) can neutralize only relatively strong bases. Thus, to be able to understand and engineer <span class="hlt">water</span>-hydrophobe interfaces, it is imperative to investigate this fluctuating region of molecular thickness wherein the 'function' of chemical species transitions from one phase to another via steep gradients in hydration, dielectric constant, and density. Aqueous interfaces are difficult to approach by current experimental techniques because designing experiments to specifically sample interfacial layers (< 1 nm thick) is an arduous task. While recent advances in surface-specific spectroscopies have provided</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20010082932&hterms=water+purification&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dwater%2Bpurification','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20010082932&hterms=water+purification&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dwater%2Bpurification"><span>Regenerable <span class="hlt">Air</span> Purification System for <span class="hlt">Gas</span>-Phase Contaminant Control</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Constantinescu, Ileana C.; Qi, Nan; LeVan, M. Douglas; Finn, Cory K.; Finn, John E.; Luna, Bernadette (Technical Monitor)</p> <p>2000-01-01</p> <p>A regenerable <span class="hlt">air</span> purification system (RAPS) that uses <span class="hlt">water</span> vapor to displace adsorbed contaminants from an. adsorbent column into a closed oxidation loop is under development through cooperative R&D between Vanderbilt University and NASA Ames Research Center. A unit based on this design can be used for removing trace <span class="hlt">gas</span>-phase contaminants from spacecraft cabin <span class="hlt">air</span> or from polluted process streams including incinerator exhaust. Recent work has focused on fabrication and operation of a RAPS breadboard at NASA Ames, and on measurement of adsorption isotherm data for several important organic compounds at Vanderbilt. These activities support the use and validation of RAPS modeling software also under development at Vanderbilt, which will in turn be used to construct a prototype system later in the project.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1915057L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1915057L"><span>Cavity Enhanced Spectrometer performance assessment for greenhouse <span class="hlt">gas</span> dry mole fraction measurement in humid <span class="hlt">air</span>.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Laurent, Olivier; Yver Kwok, Camille; Guemri, Ali; Philippon, Carole; Rivier, Leonard; Ramonet, Michel</p> <p>2017-04-01</p> <p>Due to the high variability of the <span class="hlt">water</span> vapor content in the atmosphere, the mole fraction of trace <span class="hlt">gas</span> such as greenhouse <span class="hlt">gas</span> (GHG) in the atmosphere is usually presented as mole fraction in dry <span class="hlt">air</span>. In consequence, the first technology used for GHG measurement, <span class="hlt">gas</span> chromatography or non-dispersive infra-red spectroscopy, required to dry the <span class="hlt">air</span> sample prior to analysis at a dew point lower than -50°C. The emergence of new GHG analyzers using infrared Enhanced Cavity Spectroscopy which measure the <span class="hlt">water</span> vapor content in the <span class="hlt">air</span> sample, allows providing the dry mole fraction of GHG without any drying system upstream by applying appropriate correction of the <span class="hlt">water</span> vapor effects (dilution, pressure broadening…). In the framework of ICOS, a European research infrastructure aiming to provide harmonized high precision data for advanced research on carbon cycle and GHG budgets over Europe, the Metrology Lab of the Atmosphere Thematic Centre (ATC), located at LSCE in France, is mainly dedicated to elaborating measurement protocols and evaluating performance of GHG analyzers. Among the different tests conducted to characterize the metrological performance, the Metrology Lab focuses on the <span class="hlt">water</span> vapor correction to apply on the GHG measurement. Most of the analyzers tested at the Metrology Lab are based on Cavity Enhanced Spectroscopy measuring the ICOS mandatory species, CO2, CH4 and CO. This presentation presents the results of the performance assessment of the manufacturer built-in <span class="hlt">water</span> vapor correction and the possible improvement. Thanks to the large number of instrument tested, the presentation provides a performance overview of the GHG analyzers deployed in the ICOS atmospheric station network. Finally the performance of the <span class="hlt">water</span> vapor correction will be discussed in regard of the performance obtained by using a drying system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27558796','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27558796"><span>Genotypic variation in biomass allocation in response to field drought has a greater affect on yield than <span class="hlt">gas</span> <span class="hlt">exchange</span> or phenology.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Edwards, Christine E; Ewers, Brent E; Weinig, Cynthia</p> <p>2016-08-24</p> <p>Plant performance in agricultural and natural settings varies with moisture availability, and understanding the range of potential drought responses and the underlying genetic architecture is important for understanding how plants will respond to both natural and artificial selection in various <span class="hlt">water</span> regimes. Here, we raised genotypes of Brassica rapa under well-<span class="hlt">watered</span> and drought treatments in the field. Our primary goal was to understand the genetic architecture and yield effects of different drought-escape and dehydration-avoidance strategies. Drought treatments reduced soil moisture by 62 % of field capacity. Drought decreased biomass accumulation and fruit production by as much as 48 %, whereas instantaneous <span class="hlt">water</span>-use efficiency and root:shoot ratio increased. Genotypes differed in the mean value of all traits and in the sensitivity of biomass accumulation, root:shoot ratio, and fruit production to drought. Bivariate correlations involving <span class="hlt">gas-exchange</span> and phenology were largely constant across environments, whereas those involving root:shoot varied across treatments. Although root:shoot was typically unrelated to <span class="hlt">gas-exchange</span> or yield under well-<span class="hlt">watered</span> conditions, genotypes with low to moderate increases in root:shoot allocation in response to drought survived the growing season, maintained maximum photosynthesis levels, and produced more fruit than genotypes with the greatest root allocation under drought. QTL for <span class="hlt">gas-exchange</span> and yield components (total biomass or fruit production) had common effects across environments while those for root:shoot were often environment-specific. Increases in root allocation beyond those needed to survive and maintain favorable <span class="hlt">water</span> relations came at the cost of fruit production. The environment-specific effects of root:shoot ratio on yield and the differential expression of QTL for this trait across <span class="hlt">water</span> regimes have important implications for efforts to improve crops for drought resistance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29574258','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29574258"><span><span class="hlt">Water</span> relation, leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and chlorophyll a fluorescence imaging of soybean leaves infected with Colletotrichum truncatum.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dias, Carla Silva; Araujo, Leonardo; Alves Chaves, Joicy Aparecida; DaMatta, Fábio M; Rodrigues, Fabrício A</p> <p>2018-06-01</p> <p>Considering the potential of anthracnose to decrease soybean yield and the need to gain more information regarding its effect on soybean physiology, the present study performed an in-depth analysis of the photosynthetic performance of soybean leaflets challenged with Colletotrichum truncatum by combining chlorophyll a fluorescence images with <span class="hlt">gas-exchange</span> measurements and photosynthetic pigment pools. There were no significant differences between non-inoculated and inoculated plants in leaf <span class="hlt">water</span> potential, apparent hydraulic conductance, net CO 2 assimilation rate, stomatal conductance to <span class="hlt">water</span> vapor and transpiration rate. For internal CO 2 concentration, significant difference between non-inoculated and inoculated plants occurred only at 36 h after inoculation. Reductions in the values of the chlorophyll a fluorescence parameters [initial fluorescence (F 0 ), maximal fluorescence (F m ), maximal photosystem II quantum yield (F v /F m ), quantum yield of regulated energy dissipation (Y(NPQ))] and increases in effective PS II quantum yield (Y(II)), quantum yield of non-regulated energy dissipation Y(NO) and photochemical quenching coefficient (q P ) were noticed on the necrotic vein tissue in contrast to the surrounding leaf tissue. It appears that the impact of the infection by C. truncatum on the photosynthetic performance of the leaflets was minimal considering the preference of the fungus to colonize the veins. Copyright © 2018 Elsevier Masson SAS. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23320654','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23320654"><span>Reduction of molecular <span class="hlt">gas</span> diffusion through gaskets in leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> cuvettes by leaf-mediated pores.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Boesgaard, Kristine S; Mikkelsen, Teis N; Ro-Poulsen, Helge; Ibrom, Andreas</p> <p>2013-07-01</p> <p>There is an ongoing debate on how to correct leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> measurements for the unavoidable diffusion leakage that occurs when measurements are done in non-ambient CO2 concentrations. In this study, we present a theory on how the CO2 diffusion gradient over the gasket is affected by leaf-mediated pores (LMP) and how LMP reduce diffusive <span class="hlt">exchange</span> across the gaskets. Recent discussions have so far neglected the processes in the quasi-laminar boundary layer around the gasket. Counter intuitively, LMP reduce the leakage through gaskets, which can be explained by assuming that the boundary layer at the exterior of the cuvette is enriched with <span class="hlt">air</span> from the inside of the cuvette. The effect can thus be reduced by reducing the boundary layer thickness. The theory clarifies conflicting results from earlier studies. We developed leaf adaptor frames that eliminate LMP during measurements on delicate plant material such as grass leaves with circular cross section, and the effectiveness is shown with respiration measurements on a harp of Deschampsia flexuosa leaves. We conclude that the best solution for measurements with portable photosynthesis systems is to avoid LMP rather than trying to correct for the effects. © 2013 John Wiley & Sons Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16452079','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16452079"><span>Seasonal patterns of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> relations in dry rain forest trees of contrasting leaf phenology.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Choat, Brendan; Ball, Marilyn C; Luly, Jon G; Donnelly, Christine F; Holtum, Joseph A M</p> <p>2006-05-01</p> <p>Diurnal and seasonal patterns of leaf <span class="hlt">gas</span> <span class="hlt">exchange</span> and <span class="hlt">water</span> relations were examined in tree species of contrasting leaf phenology growing in a seasonally dry tropical rain forest in north-eastern Australia. Two drought-deciduous species, Brachychiton australis (Schott and Endl.) A. Terracc. and Cochlospermum gillivraei Benth., and two evergreen species, Alphitonia excelsa (Fenzal) Benth. and Austromyrtus bidwillii (Benth.) Burret. were studied. The deciduous species had higher specific leaf areas and maximum photosynthetic rates per leaf dry mass in the wet season than the evergreens. During the transition from wet season to dry season, total canopy area was reduced by 70-90% in the deciduous species and stomatal conductance (g(s)) and assimilation rate (A) were markedly lower in the remaining leaves. Deciduous species maintained daytime leaf <span class="hlt">water</span> potentials (Psi(L)) at close to or above wet season values by a combination of stomatal regulation and reduction in leaf area. Thus, the timing of leaf drop in deciduous species was not associated with large negative values of daytime Psi(L) (greater than -1.6 MPa) or predawn Psi(L) (greater than -1.0 MPa). The deciduous species appeared sensitive to small perturbations in soil and leaf <span class="hlt">water</span> status that signalled the onset of drought. The evergreen species were less sensitive to the onset of drought and g(s) values were not significantly lower during the transitional period. In the dry season, the evergreen species maintained their canopies despite increasing <span class="hlt">water</span>-stress; however, unlike Eucalyptus species from northern Australian savannas, A and g(s) were significantly lower than wet season values.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20100040574&hterms=common+good&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcommon%2Bgood','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20100040574&hterms=common+good&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcommon%2Bgood"><span>International Space Station Common Cabin <span class="hlt">Air</span> Assembly Condensing Heat <span class="hlt">Exchanger</span> Hydrophilic Coating Failures and Lessons Learned</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Balistreri, Steven F.; Shaw, Laura A.; Laliberte, Yvon</p> <p>2010-01-01</p> <p>The ability to control the temperature and humidity of an environment or habitat is critical for human survival. These factors are important to maintaining human health and comfort, as well as maintaining mechanical and electrical equipment in good working order to support the human and to accomplish mission objectives. The temperature and humidity of the International Space Station (ISS) United States On-orbit Segment (USOS) cabin <span class="hlt">air</span> is controlled by the Common Cabin <span class="hlt">Air</span> Assembly (CCAA). The CCAA consists of a fan, a condensing heat <span class="hlt">exchanger</span> (CHX), an <span class="hlt">air/water</span> separator, temperature and liquid sensors, and electrical controlling hardware and software. The CHX is the primary component responsible for control of temperature and humidity. The CCAA CHX contains a chemical coating that was developed to be hydrophilic and thus attract <span class="hlt">water</span> from the humid influent <span class="hlt">air</span>. This attraction forms the basis for <span class="hlt">water</span> removal and therefore cabin humidity control. However, there have been several instances of CHX coatings becoming hydrophobic and repelling <span class="hlt">water</span>. When this behavior is observed in an operational CHX, the unit s ability to remove moisture from the <span class="hlt">air</span> is compromised and the result is liquid <span class="hlt">water</span> carryover into downstream ducting and systems. This <span class="hlt">water</span> carryover can have detrimental effects on the cabin atmosphere quality and on the health of downstream hardware. If the <span class="hlt">water</span> carryover is severe and widespread, this behavior can result in an inability to maintain humidity levels in the USOS. This paper will describe the operation of the five CCAAs within in the USOS, the potential causes of the hydrophobic condition, and the impacts of the resulting <span class="hlt">water</span> carryover to downstream systems. It will describe the history of this behavior and the actual observed impacts to the ISS USOS. Information on mitigation steps to protect the health of future CHX hydrophilic coatings and potential remediation techniques will also be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25985421','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25985421"><span>Impact of <span class="hlt">air</span> and <span class="hlt">water</span> vapor environments on the hydrophobicity of surfaces.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Weisensee, Patricia B; Neelakantan, Nitin K; Suslick, Kenneth S; Jacobi, Anthony M; King, William P</p> <p>2015-09-01</p> <p>Droplet wettability and mobility play an important role in dropwise condensation heat transfer. Heat <span class="hlt">exchangers</span> and heat pipes operate at liquid-vapor saturation. We hypothesize that the wetting behavior of liquid <span class="hlt">water</span> on microstructures surrounded by pure <span class="hlt">water</span> vapor differs from that for <span class="hlt">water</span> droplets in <span class="hlt">air</span>. The static and dynamic contact angles and contact angle hysteresis of <span class="hlt">water</span> droplets were measured in <span class="hlt">air</span> and pure <span class="hlt">water</span> vapor environments inside a pressure vessel. Pressures ranged from 60 to 1000 mbar, with corresponding saturation temperatures between 36 and 100°C. The wetting behavior was studied on four hydrophobic surfaces: flat Teflon-coated, micropillars, micro-scale meshes, and nanoparticle-coated with hierarchical micro- and nanoscale roughness. Static advancing contact angles are 9° lower in the <span class="hlt">water</span> vapor environment than in <span class="hlt">air</span> on a flat surface. One explanation for this reduction in contact angles is <span class="hlt">water</span> vapor adsorption to the Teflon. On microstructured surfaces, the vapor environment has little effect on the static contact angles. In all cases, variations in pressure and temperature do not influence the wettability and mobility of the <span class="hlt">water</span> droplets. In most cases, advancing contact angles increase and contact angle hysteresis decreases when the droplets are sliding or rolling down an inclined surface. Copyright © 2015 Elsevier Inc. 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_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('http://hdl.handle.net/2060/19850012858','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850012858"><span>Combustion <span class="hlt">gas</span> properties. 2: Natural <span class="hlt">gas</span> fuel and dry <span class="hlt">air</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wear, J. D.; Jones, R. E.; Trout, A. M.; Mcbride, B. J.</p> <p>1985-01-01</p> <p>A series of computations has been made to produce the equilibrium temperature and <span class="hlt">gas</span> composition for natural <span class="hlt">gas</span> fuel and dry <span class="hlt">air</span>. The computed tables and figures provide combustion <span class="hlt">gas</span> property data for pressures from 0.5 to 50 atmospheres and equivalence ratios from 0 to 2.0. Only samples tables and figures are provided in this report. The complete set of tables and figures is provided on four microfiche films supplied with this report.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26406492','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26406492"><span>Cryogenic separation of an oxygen-argon mixture in natural <span class="hlt">air</span> samples for the determination of isotope and molecular ratios.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Keedakkadan, Habeeb Rahman; Abe, Osamu</p> <p>2015-04-30</p> <p>The separation and purification of oxygen-argon mixtures are critical in the high-precision analysis of Δ(17) O and δ(O2 /Ar) for geochemical applications. At present, chromatographic methods are used for the separation and purification of oxygen-argon mixtures or pure oxygen, but these methods require the use of high-purity helium as a carrier <span class="hlt">gas</span>. Considerable interest has been expressed in the development of a helium-free cryogenic separation of oxygen-argon mixtures in natural <span class="hlt">air</span> samples. The precise and simplified cryogenic separation of oxygen-argon mixtures from natural <span class="hlt">air</span> samples presented here was made possible using a single 5A (30/60 mesh) molecular sieve column. The method involves the trapping of eluted gases using molecular sieves at liquid nitrogen temperature, which is associated with isotopic fractionation. We tested the proposed method for the determination of isotopic fractionations during the <span class="hlt">gas</span> <span class="hlt">exchange</span> between <span class="hlt">water</span> and atmospheric <span class="hlt">air</span> at equilibrium. The dependency of fractionation was studied at different <span class="hlt">water</span> temperatures and for different methods of equilibration (bubbling and stirring). Isotopic and molecular fractionations during <span class="hlt">gas</span> desorption from molecular sieves were studied for different amounts and types of molecular sieves. Repeated measurements of atmospheric <span class="hlt">air</span> yielded a reproducibility (±SD) of 0.021 ‰, 0.044 ‰, 15 per meg and 1.9 ‰ for δ(17) O, δ(18) O, Δ(17) O and δ(O2 /Ar) values, respectively. We applied the method to determine equilibrium isotope fractionation during <span class="hlt">gas</span> <span class="hlt">exchange</span> between <span class="hlt">air</span> and <span class="hlt">water</span>. Consistent δ(18) O and Δ(17) O results were obtained with the latest two studies, whereas there was a significant difference in δ(18) O values between seawater and deionized <span class="hlt">water</span>. We have revised a helium-free, cryogenic separation of oxygen-argon mixtures in natural <span class="hlt">air</span> samples for isotopic and molecular ratio analysis. The use of a single 13X (1/8" pellet) molecular sieve yielded the smallest isotopic</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26286697','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26286697"><span>Deriving C4 photosynthetic parameters from combined <span class="hlt">gas</span> <span class="hlt">exchange</span> and chlorophyll fluorescence using an Excel tool: theory and practice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bellasio, Chandra; Beerling, David J; Griffiths, Howard</p> <p>2016-06-01</p> <p>The higher photosynthetic potential of C4 plants has led to extensive research over the past 50 years, including C4 -dominated natural biomes, crops such as maize, or for evaluating the transfer of C4 traits into C3 lineages. Photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> can be measured in <span class="hlt">air</span> or in a 2% Oxygen mixture using readily available commercial <span class="hlt">gas</span> <span class="hlt">exchange</span> and modulated PSII fluorescence systems. Interpretation of these data, however, requires an understanding (or the development) of various modelling approaches, which limit the use by non-specialists. In this paper we present an accessible summary of the theory behind the analysis and derivation of C4 photosynthetic parameters, and provide a freely available Excel Fitting Tool (EFT), making rigorous C4 data analysis accessible to a broader audience. Outputs include those defining C4 photochemical and biochemical efficiency, the rate of photorespiration, bundle sheath conductance to CO2 diffusion and the in vivo biochemical constants for PEP carboxylase. The EFT compares several methodological variants proposed by different investigators, allowing users to choose the level of complexity required to interpret data. We provide a complete analysis of <span class="hlt">gas</span> <span class="hlt">exchange</span> data on maize (as a model C4 organism and key global crop) to illustrate the approaches, their analysis and interpretation. © 2015 John Wiley & Sons Ltd. © 2016 John Wiley & Sons Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMPP31A2205C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMPP31A2205C"><span>Helium Isotopes and Noble <span class="hlt">Gas</span> Abundances of Cave Dripping <span class="hlt">Water</span> in Three Caves 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>Chen, A. T.; Shen, C. C.; Tan, M.; Li, T.; Uemura, R.; Asami, R.</p> <p>2015-12-01</p> <p>Paleo-temperature recorded in nature archives is a critical parameter to understand climate change in the past. With advantages of unique inert chemical characteristics and sensitive solubilities with temperature, dissolved noble gases in speleothem inclusion <span class="hlt">water</span> were recently proposed to retrieve terrestrial temperature history. In order to accurately apply this newly-developed speleothem noble <span class="hlt">gas</span> temperature (NGT) as a reliable proxy, a fundamental issue about behaviors of noble gases in the karst should be first clarified. In this study, we measured noble <span class="hlt">gas</span> contents in <span class="hlt">air</span> and dripping <span class="hlt">water</span> to evaluate any ratio deviation between noble gases. Cave dripping <span class="hlt">water</span> samples was collected from three selected caves, Shihua Cave in northern China, Furong Cave in southwestern, and Gyukusen Cave in an island located in the western Pacific. For these caves are characterized by a thorough mixing and long-term storage of <span class="hlt">waters</span> in a karst aquifer by the absence of seasonal oxygen isotope shifts. Ratios of dripping <span class="hlt">water</span> noble gases are statistically insignificant from <span class="hlt">air</span> data. Helium isotopic ratios in the dripping <span class="hlt">water</span> samples match <span class="hlt">air</span> value. The results indicate that elemental and isotopic signatures of noble gases from <span class="hlt">air</span> can be frankly preserved in the epikarst and support the fidelity of NGT techniques.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008APS..MARJ17010M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008APS..MARJ17010M"><span>Dynamics of <span class="hlt">Gas</span> <span class="hlt">Exchange</span> through the Fractal Architecture of the Human Lung, Modeled as an Exactly Solvable Hierarchical Tree</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mayo, Michael; Pfeifer, Peter; Gheorghiu, Stefan</p> <p>2008-03-01</p> <p>The acinar airways lie at the periphery of the human lung and are responsible for the transfer of oxygen from <span class="hlt">air</span> to the blood during respiration. This transfer occurs by the diffusion-reaction of oxygen over the irregular surface of the alveolar membranes lining the acinar airways. We present an exactly solvable diffusion-reaction model on a hierarchically branched tree, allowing a quantitative prediction of the oxygen current over the entire system of acinar airways responsible for the <span class="hlt">gas</span> <span class="hlt">exchange</span>. We discuss the effect of diffusional screening, which is strongly coupled to oxygen transport in the human lung. We show that the oxygen current is insensitive to a loss of permeability of the alveolar membranes over a wide range of permeabilities, similar to a ``constant-current source'' in an electric network. Such fault tolerance has been observed in other treatments of the <span class="hlt">gas</span> <span class="hlt">exchange</span> in the lung and is obtained here as a fully analytical result.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1993ONERA..75.....A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993ONERA..75.....A"><span>High temperature heat <span class="hlt">exchangers</span> for <span class="hlt">gas</span> turbines and future hypersonic <span class="hlt">air</span> breathing propulsion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Avran, Patrick; Bernard, Pierre</p> <p></p> <p>After surveying the results of ONERA's investigations to date of metallic and ceramic heat <span class="hlt">exchangers</span> applicable to automotive and aircraft powerplants, which are primarily of finned-tube counterflow configuration, attention is given to the influence of heat-<span class="hlt">exchanger</span> effectiveness on fuel consumption and <span class="hlt">exchanger</span> dimensions and weight. Emphasis is placed on the results of studies of cryogenic heat <span class="hlt">exchangers</span> used by airbreathing hypersonic propulsion systems. The numerical codes developed by ONERA for the modeling of heat <span class="hlt">exchanger</span> thermodynamics are evaluated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFM.P21A0220H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFM.P21A0220H"><span>Laboratory Measurements of Oxygen <span class="hlt">Gas</span> Release from Basaltic Minerals Exposed to UV- Radiation: Implications for the Viking <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Experiments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hurowitz, J. A.; Yen, A. S.</p> <p>2007-12-01</p> <p>The biology experiments onboard the Viking Landers determined that the Martian soils at Chryse and Utopia Planitia contain an unknown chemical compound of a highly oxidizing nature. The <span class="hlt">Gas</span> <span class="hlt">Exchange</span> Experiments (GEx) demonstrated that the humidification of a 1-cc Martian soil sample resulted in the production of as much as 790 nanomoles of oxygen <span class="hlt">gas</span>. Yen et al. (2000) have provided experimental evidence that superoxide radicals can be generated on plagioclase feldspar (labradorite) grain surfaces by exposure to ultraviolet (UV) light in the presence of oxygen <span class="hlt">gas</span>. Adsorbed superoxide radicals are thought to react readily with <span class="hlt">water</span> vapor, and produce oxygen <span class="hlt">gas</span> in quantities sufficient to explain the Viking GEx results. Direct evidence for the formation of oxygen <span class="hlt">gas</span>, however, was not provided in the experiments of Yen et al (2000). Accordingly, the motivation of this study is to determine whether superoxide radicals adsorbed on labradorite surfaces are capable of producing oxygen <span class="hlt">gas</span> upon exposure to <span class="hlt">water</span> vapor. We have constructed an experimental apparatus that is capable of monitoring oxygen <span class="hlt">gas</span> release from basaltic mineral powders that have been exposed to UV-radiation under Martian atmospheric pressure conditions. The apparatus consists of a stainless-steel vacuum chamber with a UV- transparent window where sample radiation exposures are performed. The vacuum chamber has multiple valved ports for injection of gases and <span class="hlt">water</span> vapor. The vacuum chamber is connected via a precision leak valve to a quadrupole mass spectrometer, which measures changes in the composition of the headspace gases over our mineral samples. We will report on the results of our experiments, which are aimed at detecting and quantifying oxygen <span class="hlt">gas</span> release from UV-exposed basaltic mineral samples using this new experimental facility. These results will further constrain whether superoxide ions adsorbed on mineral surfaces provide a viable explanation for the Viking GEx results, which have</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGRC..121.7369F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGRC..121.7369F"><span>Surface shear stress dependence of <span class="hlt">gas</span> transfer velocity parameterizations using DNS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fredriksson, S. T.; Arneborg, L.; Nilsson, H.; Handler, R. A.</p> <p>2016-10-01</p> <p><span class="hlt">Air-water</span> <span class="hlt">gas-exchange</span> is studied in direct numerical simulations (DNS) of free-surface flows driven by natural convection and weak winds. The wind is modeled as a constant surface-shear-stress and the <span class="hlt">gas</span>-transfer is modeled via a passive scalar. The simulations are characterized via a Richardson number Ri=Bν/u*4 where B, ν, and u* are the buoyancy flux, kinematic viscosity, and friction velocity respectively. The simulations comprise 0<Ri<∞ ranging from convection-dominated to shear-dominated cases. The results are used to: (i) evaluate parameterizations of the <span class="hlt">air-water</span> <span class="hlt">gas-exchange</span>, (ii) determine, for a given buoyancy flux, the wind speed at which <span class="hlt">gas</span> transfer becomes primarily shear driven, and (iii) find an expression for the <span class="hlt">gas</span>-transfer velocity for flows driven by both convection and shear. The evaluated <span class="hlt">gas</span> transfer-velocity parametrizations are based on either the rate of turbulent kinetic energy dissipation, the surface flow-divergence, the surface heat-flux, or the wind-speed. The parametrizations based on dissipation or divergence show an unfavorable Ri dependence for flows with combined forcing whereas the parametrization based on heat-flux only shows a limited Ri dependence. The two parametrizations based on wind speed give reasonable estimates for the transfer-velocity, depending however on the surface heat-flux. The transition from convection- to shear-dominated <span class="hlt">gas</span>-transfer-velocity is shown to be at Ri≈0.004. Furthermore, the <span class="hlt">gas</span>-transfer is shown to be well represented by two different approaches: (i) additive forcing expressed as kg,sum =AShearu*|Ri/Ric+1| 1/4Sc-n where Ric=|AShear/ABuoy|4, and (ii) either buoyancy or shear dominated expressed as, kg=ABuoy|Bν| 1/4Sc-n, Ri>Ric or kg=AShearu*Sc-n, Ri<Ric. Here ABuoy=0.4 and AShear=0.1 are constants, and n is an exponent that depends on the <span class="hlt">water</span> surface-characteristics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010cosp...38.3367R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010cosp...38.3367R"><span><span class="hlt">Air</span> Circulation and Heat <span class="hlt">Exchange</span> under Reduced Pressures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rygalov, Vadim; Wheeler, Raymond; Dixon, Mike; Hillhouse, Len; Fowler, Philip</p> <p></p> <p>Low pressure atmospheres were suggested for Space Greenhouses (SG) design to minimize sys-tem construction and re-supply materials, as well as system manufacturing and deployment costs. But rarified atmospheres modify heat <span class="hlt">exchange</span> mechanisms what finally leads to alter-ations in thermal control for low pressure closed environments. Under low atmospheric pressures (e.g., lower than 25 kPa compare to 101.3 kPa for normal Earth atmosphere), convection is becoming replaced by diffusion and rate of heat <span class="hlt">exchange</span> reduces significantly. During a period from 2001 to 2009, a series of hypobaric experiments were conducted at Space Life Sciences Lab (SLSLab) NASA's Kennedy Space Center and the Department of Space Studies, University of North Dakota. Findings from these experiments showed: -<span class="hlt">air</span> circulation rate decreases non-linearly with lowering of total atmospheric pressure; -heat <span class="hlt">exchange</span> slows down with pressure decrease creating risk of thermal stress (elevated leaf tem-peratures) for plants in closed environments; -low pressure-induced thermal stress could be reduced by either lowering system temperature set point or increasing forced convection rates (circulation fan power) within certain limits; <span class="hlt">Air</span> circulation is an important constituent of controlled environments and plays crucial role in material and heat <span class="hlt">exchange</span>. Theoretical schematics and mathematical models are developed from a series of observations. These models can be used to establish optimal control algorithms for low pressure environments, such as a space greenhouse, as well as assist in fundamental design concept developments for these or similar habitable structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/14567951','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14567951"><span>A simple bubbling system for measuring radon (222Rn) <span class="hlt">gas</span> concentrations in <span class="hlt">water</span> samples based on the high solubility of radon in olive oil.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Al-Azmi, D; Snopek, B; Sayed, A M; Domanski, T</p> <p>2004-01-01</p> <p>Based on the different levels of solubility of radon <span class="hlt">gas</span> in organic solvents and <span class="hlt">water</span>, a bubbling system has been developed to transfer radon <span class="hlt">gas</span>, dissolving naturally in <span class="hlt">water</span> samples, to an organic solvent, i.e. olive oil, which is known to be a good solvent of radon <span class="hlt">gas</span>. The system features the application of a fixed volume of bubbling <span class="hlt">air</span> by introducing a fixed volume of <span class="hlt">water</span> into a flask mounted above the system, to displace an identical volume of <span class="hlt">air</span> from an <span class="hlt">air</span> cylinder. Thus a gravitational flow of <span class="hlt">water</span> is provided without the need for pumping. Then, the flushing <span class="hlt">air</span> (radon-enriched <span class="hlt">air</span>) is directed through a vial containing olive oil, to achieve deposition of the radon <span class="hlt">gas</span> by another bubbling process. Following this, the vial (containing olive oil) is measured by direct use of gamma ray spectrometry, without the need of any chemical or physical processing of the samples. Using a standard solution of 226Ra/222Rn, a lowest measurable concentration (LMC) of radon in <span class="hlt">water</span> samples of 9.4 Bq L(-1) has been achieved (below the maximum contaminant level of 11 Bq L(-1)).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.8968S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.8968S"><span>Methane <span class="hlt">gas</span> seepage - Disregard of significant <span class="hlt">water</span> column filter processes?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schneider von Deimling, Jens; Schmale, Oliver</p> <p>2016-04-01</p> <p>Marine methane seepage represents a potential contributor for greenhouse <span class="hlt">gas</span> in the atmosphere and is discussed as a driver for climate change. The ultimate question is how much methane is released from the seafloor on a global scale and what fraction may reach the atmosphere? Dissolved fluxes from methane seepage sites on the seabed were found to be very efficiently reduced by benthic microbial oxidation, whereas transport of free <span class="hlt">gas</span> bubbles from the seabed is considered to bypass the effective benthic methane filter. Numerical models are available today to predict the fate of such methane <span class="hlt">gas</span> bubble release to the <span class="hlt">water</span> column in regard to <span class="hlt">gas</span> <span class="hlt">exchange</span> with the ambient <span class="hlt">water</span> column, respective bubble lifetime and rise height. However, the fate of rising <span class="hlt">gas</span> bubbles and dissolved methane in the <span class="hlt">water</span> column is not only governed by dissolution, but is also affected by lateral oceanographic currents and vertical bubble-induced upwelling, microbial oxidation, and physico-chemical processes that remain poorly understood so far. According to this gap of knowledge we present data from two study sites - the anthropogenic North Sea 22/4b Blowout and the natural Coal Oil point seeps - to shed light into two new processes gathered with hydro-acoustic multibeam <span class="hlt">water</span> column imaging and microbial investigations. The newly discovered processes are hereafter termed Spiral Vortex and Bubble Transport Mechanism. Spiral Vortex describes the evolution of a complex vortical fluid motion of a bubble plume in the wake of an intense <span class="hlt">gas</span> release site (Blowout, North Sea). It appears very likely that it dramatically changes the dissolution kinetics of the seep <span class="hlt">gas</span> bubbles. Bubble Transport Mechanism prescribes the transport of sediment-hosted bacteria into the <span class="hlt">water</span> column via rising <span class="hlt">gas</span> bubbles. Both processes act as filter mechanisms in regard to vertical transport of seep related methane, but have not been considered before. Spiral Vortex and Bubble Transport Mechanism represent the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28177581','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28177581"><span>Harvesting Hydrogen <span class="hlt">Gas</span> from <span class="hlt">Air</span> Pollutants with an Unbiased <span class="hlt">Gas</span> Phase Photoelectrochemical Cell.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Verbruggen, Sammy W; Van Hal, Myrthe; Bosserez, Tom; Rongé, Jan; Hauchecorne, Birger; Martens, Johan A; Lenaerts, Silvia</p> <p>2017-04-10</p> <p>The concept of an all-<span class="hlt">gas</span>-phase photoelectrochemical (PEC) cell producing hydrogen <span class="hlt">gas</span> from volatile organic contaminated <span class="hlt">gas</span> and light is presented. Without applying any external bias, organic contaminants are degraded and hydrogen <span class="hlt">gas</span> is produced in separate electrode compartments. The system works most efficiently with organic pollutants in inert carrier <span class="hlt">gas</span>. In the presence of oxygen, the cell performs less efficiently but still significant photocurrents are generated, showing the cell can be run on organic contaminated <span class="hlt">air</span>. The purpose of this study is to demonstrate new application opportunities of PEC technology and to encourage further advancement toward PEC remediation of <span class="hlt">air</span> pollution with the attractive feature of simultaneous energy recovery and pollution abatement. © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=47600&Lab=NHEERL&keyword=physiology+AND+stress&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=47600&Lab=NHEERL&keyword=physiology+AND+stress&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>JOINT ACTION OF O3 AND SO2 IN MODIFYING PLANT <span class="hlt">GAS</span> <span class="hlt">EXCHANGE</span></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>The joint action of O3 and SO2 stress on plants was investigated. <span class="hlt">Gas</span> <span class="hlt">exchange</span> measurements of O3, SO2, and H2O vapor were made for garden pea. Plants were grown under controlled environments; O3, SO2, H2O vapor fluxes were evaluated with a whole-plant <span class="hlt">gas</span> <span class="hlt">exchange</span> chamber using ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25562933','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25562933"><span>Impact of emissions from natural <span class="hlt">gas</span> production facilities on ambient <span class="hlt">air</span> quality in the Barnett Shale area: a pilot study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zielinska, Barbara; Campbell, Dave; Samburova, Vera</p> <p>2014-12-01</p> <p>Rapid and extensive development of shale <span class="hlt">gas</span> resources in the Barnett Shale region of Texas in recent years has created concerns about potential environmental impacts on <span class="hlt">water</span> and <span class="hlt">air</span> quality. The purpose of this study was to provide a better understanding of the potential contributions of emissions from <span class="hlt">gas</span> production operations to population exposure to <span class="hlt">air</span> toxics in the Barnett Shale region. This goal was approached using a combination of chemical characterization of the volatile organic compound (VOC) emissions from active wells, saturation monitoring for gaseous and particulate pollutants in a residential community located near active <span class="hlt">gas</span>/oil extraction and processing facilities, source apportionment of VOCs measured in the community using the Chemical Mass Balance (CMB) receptor model, and direct measurements of the pollutant gradient downwind of a <span class="hlt">gas</span> well with high VOC emissions. Overall, the study results indicate that <span class="hlt">air</span> quality impacts due to individual <span class="hlt">gas</span> wells and compressor stations are not likely to be discernible beyond a distance of approximately 100 m in the downwind direction. However, source apportionment results indicate a significant contribution to regional VOCs from <span class="hlt">gas</span> production sources, particularly for lower-molecular-weight alkanes (< C6). Although measured ambient VOC concentrations were well below health-based safe exposure levels, the existence of urban-level mean concentrations of benzene and other mobile source <span class="hlt">air</span> toxics combined with soot to total carbon ratios that were high for an area with little residential or commercial development may be indicative of the impact of increased heavy-duty vehicle traffic related to <span class="hlt">gas</span> production. Implications: Rapid and extensive development of shale <span class="hlt">gas</span> resources in recent years has created concerns about potential environmental impacts on <span class="hlt">water</span> and <span class="hlt">air</span> quality. This study focused on directly measuring the ambient <span class="hlt">air</span> pollutant levels occurring at residential properties located near</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5874683-method-dehydrating-natural-gas','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5874683-method-dehydrating-natural-gas"><span>Method of dehydrating natural <span class="hlt">gas</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>Wells, R. E.</p> <p>1985-01-01</p> <p>A method for dehydration of natural <span class="hlt">gas</span> is provided wherein well head <span class="hlt">gas</span> is supplied to a three-phase inlet separator, the vapor mixture of natural <span class="hlt">gas</span> and <span class="hlt">water</span> removed from that inlet separator means is supplied to a turboexpander, and the resulting refrigerated mixture of natural <span class="hlt">gas</span> and condensed <span class="hlt">water</span> vapor is supplied to a multi-phase outlet separator. The turboexpander may have integral means for subsequent compression of the refrigerated mixture and may be coupled through reduction gears to a means for generating electricity. A portion of the refrigerated mixture may be connected to a heat <span class="hlt">exchanger</span> for cooling themore » well head natural <span class="hlt">gas</span> prior to entry into the inlet separator. The flow of refrigerated mixture to this heat <span class="hlt">exchanger</span> may be controlled by a temperature sensitive valve downstream of the heat <span class="hlt">exchanger</span>. Methanol may be injected into the vapor mixture prior to entry into the turboexpander. The flow of methanol into the vapor mixture may be controlled by a valve sensitive to the flow rate of the vapor mixture and the <span class="hlt">water</span> vapor content of the refrigerated mixture. Natural <span class="hlt">gas</span> vapor from the outlet separator may be recirculated through the turboexpander if the output <span class="hlt">water</span> vapor content of the natural <span class="hlt">gas</span> vapor stream is too high.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.H31G0704O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.H31G0704O"><span>Sensitivity of Photosynthetic <span class="hlt">Gas</span> <span class="hlt">Exchange</span> and Growth of Lodgepole Pine to Climate Variability Depends on the Age of Pleistocene Glacial Surfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Osborn, B.; Chapple, W.; Ewers, B. E.; Williams, D. G.</p> <p>2014-12-01</p> <p>The interaction between soil conditions and climate variability plays a central role in the ecohydrological functions of montane conifer forests. Although soil moisture availability to trees is largely dependent on climate, the depth and texture of soil exerts a key secondary influence. Multiple Pleistocene glacial events have shaped the landscape of the central Rocky Mountains creating a patchwork of soils differing in age and textural classification. This mosaic of soil conditions impacts hydrological properties, and montane conifer forests potentially respond to climate variability quite differently depending on the age of glacial till and soil development. We hypothesized that the age of glacial till and associated soil textural changes exert strong control on growth and photosynthetic <span class="hlt">gas</span> <span class="hlt">exchange</span> of lodgepole pine. We examined physiological and growth responses of lodgepole pine to interannual variation in maximum annual snow <span class="hlt">water</span> equivalence (SWEmax) of montane snowpack and growing season <span class="hlt">air</span> temperature (Tair) and vapor pressure deficit (VPD) across a chronosequence of Pleistocene glacial tills ranging in age from 700k to 12k years. Soil textural differences across the glacial tills illustrate the varying degrees of weathering with the most well developed soils with highest clay content on the oldest till surfaces. We show that sensitivity of growth and carbon isotope discrimination, an integrated measure of canopy <span class="hlt">gas</span> <span class="hlt">exchange</span> properties, to interannual variation SWEmax , Tair and VPD is greatest on young till surfaces, whereas trees on old glacial tills with well-developed soils are mostly insensitive to these interannual climate fluctuations. Tree-ring widths were most sensitive to changes in SWEmax on young glacial tills (p < 0.01), and less sensitive on the oldest till (p < 0.05). Tair correlates strongly with δ13C values on the oldest and youngest tills sites, but shows no significant relationship on the middle aged glacial till. It is clear that</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29938338','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29938338"><span>Calculation algorithms for breath-by-breath alveolar <span class="hlt">gas</span> <span class="hlt">exchange</span>: the unknowns!</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Golja, Petra; Cettolo, Valentina; Francescato, Maria Pia</p> <p>2018-06-25</p> <p>Several papers (algorithm papers) describe computational algorithms that assess alveolar breath-by-breath <span class="hlt">gas</span> <span class="hlt">exchange</span> by accounting for changes in lung <span class="hlt">gas</span> stores. It is unclear, however, if the effects of the latter are actually considered in literature. We evaluated dissemination of algorithm papers and the relevant provided information. The list of documents investigating exercise transients (in 1998-2017) was extracted from Scopus database. Documents citing the algorithm papers in the same period were analyzed in full text to check consistency of the relevant information provided. Less than 8% (121/1522) of documents dealing with exercise transients cited at least one algorithm paper; the paper of Beaver et al. (J Appl Physiol 51:1662-1675, 1981) was cited most often, with others being cited tenfold less. Among the documents citing the algorithm paper of Beaver et al. (J Appl Physiol 51:1662-1675, 1981) (N = 251), only 176 cited it for the application of their algorithm/s; in turn, 61% (107/176) of them stated the alveolar breath-by-breath <span class="hlt">gas</span> <span class="hlt">exchange</span> measurement, but only 1% (1/107) of the latter also reported the assessment of volunteers' functional residual capacity, a crucial parameter for the application of the algorithm. Information related to <span class="hlt">gas</span> <span class="hlt">exchange</span> was provided consistently in the methods and in the results in 1 of the 107 documents. Dissemination of algorithm papers in literature investigating exercise transients is by far narrower than expected. The information provided about the actual application of <span class="hlt">gas</span> <span class="hlt">exchange</span> algorithms is often inadequate and/or ambiguous. Some guidelines are provided that can help to improve the quality of future publications in the field.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25063854','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25063854"><span>A hierarchy of factors influence discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> in the grasshopper Paracinema tricolor (Orthoptera: Acrididae).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Groenewald, Berlizé; Chown, Steven L; Terblanche, John S</p> <p>2014-10-01</p> <p>The evolutionary origin and maintenance of discontinuous <span class="hlt">gas</span> <span class="hlt">exchange</span> (DGE) in tracheate arthropods are poorly understood and highly controversial. We investigated prioritization of abiotic factors in the <span class="hlt">gas</span> <span class="hlt">exchange</span> control cascade by examining oxygen, <span class="hlt">water</span> and haemolymph pH regulation in the grasshopper Paracinema tricolor. Using a full-factorial design, grasshoppers were acclimated to hypoxic or hyperoxic (5% O2, 40% O2) <span class="hlt">gas</span> conditions, or dehydrated or hydrated, whereafter their CO2 release was measured under a range of O2 and relative humidity (RH) conditions (5%, 21%, 40% O2 and 5%, 60%, 90% RH). DGE was significantly less common in grasshoppers acclimated to dehydrating conditions compared with the other acclimations (hypoxia, 98%; hyperoxia, 100%; hydrated, 100%; dehydrated, 67%). Acclimation to dehydrating conditions resulted in a significant decrease in haemolymph pH from 7.0±0.3 to 6.6±0.1 (mean ± s.d., P=0.018) and also significantly increased the open (O)-phase duration under 5% O2 treatment conditions (5% O2, 44.1±29.3 min; 40% O2, 15.8±8.0 min; 5% RH, 17.8±1.3 min; 60% RH, 24.0±9.7 min; 90% RH, 20.6±8.9 min). The observed acidosis could potentially explain the extension of the O-phase under low RH conditions, when it would perhaps seem more useful to reduce the O-phase to lower respiratory <span class="hlt">water</span> loss. The results confirm that DGE occurrence and modulation are affected by multiple abiotic factors. A hierarchical framework for abiotic factors influencing DGE is proposed in which the following stressors are prioritized in decreasing order of importance: oxygen supply, CO2 excretion and pH modulation, oxidative damage protection and <span class="hlt">water</span> savings. © 2014. Published by The Company of Biologists Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=261773&keyword=Human+AND+interaction&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=261773&keyword=Human+AND+interaction&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>Spatiotemporally‐Resolved <span class="hlt">Air</span> <span class="hlt">Exchange</span> Rate as a Modifier of Acute <span class="hlt">Air</span> Pollution‐Related Morbidity in AtlantaMorbidity in Atlanta</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>Epidemiological studies frequently use central site concentrations as surrogates of exposure to <span class="hlt">air</span> pollutants. Variability in <span class="hlt">air</span> pollutant infiltration due to differential <span class="hlt">air</span> <span class="hlt">exchange</span> rates (AERs) is potentially a major factor affecting the relationship between central site c...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5292819','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5292819"><span>Effect of Leaf <span class="hlt">Water</span> Potential on Internal Humidity and CO2 Dissolution: Reverse Transpiration and Improved <span class="hlt">Water</span> Use Efficiency under Negative Pressure</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Vesala, Timo; Sevanto, Sanna; Grönholm, Tiia; Salmon, Yann; Nikinmaa, Eero; Hari, Pertti; Hölttä, Teemu</p> <p>2017-01-01</p> <p>The pull of <span class="hlt">water</span> from the soil to the leaves causes <span class="hlt">water</span> in the transpiration stream to be under negative pressure decreasing the <span class="hlt">water</span> potential below zero. The osmotic concentration also contributes to the decrease in leaf <span class="hlt">water</span> potential but with much lesser extent. Thus, the surface tension force is approximately balanced by a force induced by negative <span class="hlt">water</span> potential resulting in concavely curved <span class="hlt">water-air</span> interfaces in leaves. The lowered <span class="hlt">water</span> potential causes a reduction in the equilibrium <span class="hlt">water</span> vapor pressure in internal (sub-stomatal/intercellular) cavities in relation to that over <span class="hlt">water</span> with the potential of zero, i.e., over the flat surface. The curved surface causes a reduction also in the equilibrium vapor pressure of dissolved CO2, thus enhancing its physical solubility to <span class="hlt">water</span>. Although the <span class="hlt">water</span> vapor reduction is acknowledged by plant physiologists its consequences for <span class="hlt">water</span> vapor <span class="hlt">exchange</span> at low <span class="hlt">water</span> potential values have received very little attention. Consequences of the enhanced CO2 solubility to a leaf <span class="hlt">water</span>-carbon budget have not been considered at all before this study. We use theoretical calculations and modeling to show how the reduction in the vapor pressures affects transpiration and carbon assimilation rates. Our results indicate that the reduction in vapor pressures of <span class="hlt">water</span> and CO2 could enhance plant <span class="hlt">water</span> use efficiency up to about 10% at a leaf <span class="hlt">water</span> potential of −2 MPa, and much more when <span class="hlt">water</span> potential decreases further. The low <span class="hlt">water</span> potential allows for a direct stomatal <span class="hlt">water</span> vapor uptake from the ambient <span class="hlt">air</span> even at sub-100% relative humidity values. This alone could explain the observed rates of foliar <span class="hlt">water</span> uptake by e.g., the coastal redwood in the fog belt region of coastal California provided the stomata are sufficiently open. The omission of the reduction in the <span class="hlt">water</span> vapor pressure causes a bias in the estimates of the stomatal conductance and leaf internal CO2 concentration based on leaf <span class="hlt">gas</span> <span class="hlt">exchange</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_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>