Sample records for observed nitrate concentrations

  1. Concentration of Nitrate near the Surface of Frozen Salt Solutions

    NASA Astrophysics Data System (ADS)

    Michelsen, R. R. H.; Marrocco, H. A.

    2017-12-01

    The photolysis of nitrate near the surface of snow and ice in Earth's environment results in the emission of nitrogen oxides (NO, NO2 and, in acidic snow, HONO) and OH radicals. As a result, nitrate photolysis affects the composition and oxidative capacity of the overlying atmosphere. Photolysis yields depend in part on how much nitrate is close enough to the surface to be photolyzed. These concentrations are assumed to be higher than the concentrations of nitrate that are measured in melted snow and ice samples. However, near-surface concentrations of nitrate have not been directly measured. In this work, laboratory studies of the concentration of nitrate in frozen aqueous solutions are described. Individual aqueous solutions of nitric acid, sodium nitrate, and magnesium nitrate were mixed. Attenuated total reflection infrared spectroscopy was utilized to measure the nitrate and liquid water signals within 200 - 400 nm of the lower surface of frozen samples. Temperature was varied from -18°C to -2°C. In addition to the amount of nitrate observed, changes to the frozen samples' morphology with annealing are discussed. Nitrate concentrations near the lower surface of these frozen solutions are high: close to 1 M at warmer temperatures and almost 4 M at the coldest temperature. Known freezing point depression data describe the observed concentrations better than ideal solution thermodynamics, which overestimate concentration significantly at colder temperatures. The implications for modeling the chemistry of snow are discussed. Extending and relating this work to the interaction of gas-phase nitric acid with the surfaces of vapor-deposited ice will also be explored.

  2. Effect of temperature & salt concentration on salt tolerant nitrate-perchlorate reducing bacteria: Nitrate degradation kinetics.

    PubMed

    Ebrahimi, Shelir; Nguyen, Thi Hau; Roberts, Deborah J

    2015-10-15

    The sustainability of nitrate-contaminated water treatment using ion-exchange processes can be achieved by regenerating the exhausted resin several times. Our previous study shows that the use of multi-cycle bioregeneration of resin enclosed in membrane is an effective and innovative regeneration method. In this research, the effects of two independent factors (temperature and salt concentration) on the biological denitrification rate were studied. The results of this research along with the experimental results of the previous study on the effect of the same factors on nitrate desorption rate from the resin allow the optimization of the bioregeneration process. The results of nitrate denitrification rate study show that the biodegradation rate at different temperature and salt concentration is independent of the initial nitrate concentration. At each specific salt concentration, the nitrate removal rate increased with increasing temperature with the average value of 0.001110 ± 0.0000647 mg-nitrate/mg-VSS.h.°C. However, the effect of different salt concentrations was dependent on the temperature; there is a significant interaction between salt concentration and temperature; within each group of temperatures, the nitrate degradation rate decreased with increasing the salt concentration. The temperature affected the tolerance to salinity and culture was less tolerant to high concentration of salt at low temperature. Evidenced by the difference between the minimum and maximum nitrate degradation rate being greater at lower temperature. At 35 °C, a 32% reduction in the nitrate degradation rate was observed while at 12 °C this reduction was 69%. This is the first published study to examine the interaction of salt concentration and temperature during biological denitrification. Copyright © 2015 Elsevier Ltd. All rights reserved.

  3. Observations of fine particulate nitrated phenols in four sites in northern China: concentrations, source apportionment, and secondary formation

    NASA Astrophysics Data System (ADS)

    Wang, Liwei; Wang, Xinfeng; Gu, Rongrong; Wang, Hao; Yao, Lan; Wen, Liang; Zhu, Fanping; Wang, Weihao; Xue, Likun; Yang, Lingxiao; Lu, Keding; Chen, Jianmin; Wang, Tao; Zhang, Yuanghang; Wang, Wenxing

    2018-03-01

    Filter samples of fine particulate matters were collected at four sites in northern China (urban, rural, and mountain) in summer and winter, and the contents of nine nitrated phenols were quantified in the laboratory with the use of ultra-high-performance liquid chromatography coupled with mass spectrometry. During the sampling periods, the concentrations of particulate nitrated phenols exhibited distinct temporal and spatial variation. On average, the total concentration of particulate nitrated phenols in urban Jinan in the wintertime reached 48.4 ng m-3, and those in the summertime were 9.8, 5.7, 5.9, and 2.5 ng m-3 in urban Jinan, rural Yucheng and Wangdu, and Mt. Tai, respectively. The elevated concentrations of nitrated phenols in wintertime and in urban areas demonstrate the apparent influences of anthropogenic sources. The positive matrix factorization receptor model was then applied to determine the origins of particulate nitrated phenols in northern China. The five major source factors were traffic, coal combustion, biomass burning, secondary formation, and aged coal combustion plume. Among them, coal combustion played a vital role, especially at the urban site in the wintertime, with a contribution of around 55 %. In the summertime, the observed nitrated phenols were highly influenced by aged coal combustion plumes at all of the sampling sites. Meanwhile, in remote areas, contributions from secondary formation were significant. Further correlation analysis indicates that nitrosalicylic acids were produced mostly from secondary formation that was dominated by NO2 nitration.

  4. High-nitrate vegetable diet increases plasma nitrate and nitrite concentrations and reduces blood pressure in healthy women.

    PubMed

    Ashworth, Ann; Mitchell, Klaus; Blackwell, Jamie R; Vanhatalo, Anni; Jones, Andrew M

    2015-10-01

    Epidemiological studies suggest that green leafy vegetables, which are high in dietary nitrate, are protective against CVD such as stroke. High blood pressure (BP) is a major risk factor for stroke and inorganic nitrate has been shown to reduce BP. The objective of the present study was to test the hypothesis that diets containing high-nitrate (HN) vegetables would increase plasma nitrate and nitrite concentrations and reduce BP in healthy women. A randomized, crossover trial, where participants received HN vegetables (HN diet) or avoided HN vegetables (Control diet) for 1 week. Before and after each intervention, resting BP and plasma nitrate and nitrite concentrations were measured. University of Exeter, UK. Nineteen healthy women (mean age 20 (sd 2) years; mean BMI 22·5 (sd 3·8) kg/m2). The HN diet significantly increased plasma nitrate concentration (before HN diet: mean 24·4 (sd 5·6) µmol/l; after HN diet: mean 61·0 (sd 44·1) µmol/l, P<0·05) and plasma nitrite concentration (before HN diet: mean 98 (sd 91) nmol/l; after HN diet: mean 185 (sd 34) nmol/l, P<0·05). No significant change in plasma nitrate or nitrite concentration was observed after the Control diet. The HN diet significantly reduced resting systolic BP (before HN diet: mean 107 (sd 9) mmHg; after HN diet: mean 103 (sd 6) mmHg, P<0·05). No significant change in systolic BP was observed after the Control diet (before Control diet: mean 106 (sd 8) mmHg; after Control diet: mean 106 (sd 8) mmHg). Consumption of HN vegetables significantly increased plasma nitrate and nitrite concentrations and reduced BP in normotensive women.

  5. Observational assessment of the role of nocturnal residual-layer chemistry in determining daytime surface particulate nitrate concentrations

    NASA Astrophysics Data System (ADS)

    Prabhakar, Gouri; Parworth, Caroline L.; Zhang, Xiaolu; Kim, Hwajin; Young, Dominique E.; Beyersdorf, Andreas J.; Ziemba, Luke D.; Nowak, John B.; Bertram, Timothy H.; Faloona, Ian C.; Zhang, Qi; Cappa, Christopher D.

    2017-12-01

    This study discusses an analysis of combined airborne and ground observations of particulate nitrate (NO3-(p)) concentrations made during the wintertime DISCOVER-AQ (Deriving Information on Surface Conditions from COlumn and VERtically resolved observations relevant to Air Quality) study at one of the most polluted cities in the United States - Fresno, CA - in the San Joaquin Valley (SJV) and focuses on developing an understanding of the various processes that impact surface nitrate concentrations during pollution events. The results provide an explicit case-study illustration of how nighttime chemistry can influence daytime surface-level NO3-(p) concentrations, complementing previous studies in the SJV. The observations exemplify the critical role that nocturnal chemical production of NO3-(p) aloft in the residual layer (RL) can play in determining daytime surface-level NO3-(p) concentrations. Further, they indicate that nocturnal production of NO3-(p) in the RL, along with daytime photochemical production, can contribute substantially to the buildup and sustaining of severe pollution episodes. The exceptionally shallow nocturnal boundary layer (NBL) heights characteristic of wintertime pollution events in the SJV intensify the importance of nocturnal production aloft in the residual layer to daytime surface concentrations. The observations also demonstrate that dynamics within the RL can influence the early-morning vertical distribution of NO3-(p), despite low wintertime wind speeds. This overnight reshaping of the vertical distribution above the city plays an important role in determining the net impact of nocturnal chemical production on local and regional surface-level NO3-(p) concentrations. Entrainment of clean free-tropospheric (FT) air into the boundary layer in the afternoon is identified as an important process that reduces surface-level NO3-(p) and limits buildup during pollution episodes. The influence of dry deposition of HNO3 gas to the surface on

  6. Antecedent flow conditions and nitrate concentrations in the Mississippi River basin

    USGS Publications Warehouse

    Murphy, Jennifer C.; Hirsch, Robert M.; Sprague, Lori A.

    2014-01-01

    The relationship between antecedent flow conditions and nitrate concentrations was explored at eight sites in the 2.9 million square kilometers (km2) Mississippi River basin, USA. Antecedent flow conditions were quantified as the ratio between the mean daily flow of the previous year and the mean daily flow from the period of record (Qratio), and the Qratio was statistically related to nitrate anomalies (the unexplained variability in nitrate concentration after filtering out season, long-term trend, and contemporaneous flow effects) at each site. Nitrate anomaly and Qratio were negatively related at three of the four major tributary sites and upstream in the Mississippi River, indicating that when mean daily streamflow during the previous year was lower than average, nitrate concentrations were higher than expected. The strength of these relationships increased when data were subdivided by contemporaneous flow conditions. Five of the eight sites had significant negative relationships (p ≤ 0.05) at high or moderately high contemporaneous flows, suggesting nitrate that accumulates in these basins during a drought is flushed during subsequent high flows. At half of the sites, when mean daily flow during the previous year was 50 percent lower than average, nitrate concentration can be from 9 to 27 percent higher than nitrate concentrations that follow a year with average mean daily flow. Conversely, nitrate concentration can be from 8 to 21 percent lower than expected when flow during the previous year was 50 percent higher than average. Previously documented for small, relatively homogenous basins, our results suggest that relationships between antecedent flows and nitrate concentrations are also observable at a regional scale. Relationships were not observed (using all contemporaneous flow data together) for basins larger than 1 million km2, suggesting that above this limit the overall size and diversity within these basins may necessitate the use of more

  7. Environmentally relevant concentrations of nitrate increase plasma testosterone concentrations in female American alligators (Alligator mississippiensis).

    PubMed

    Hamlin, Heather J; Edwards, Thea M; McCoy, Jessica; Cruze, Lori; Guillette, Louis J

    2016-11-01

    Anthropogenic nitrogen is a ubiquitous environmental contaminant that is contributing to the degradation of freshwater, estuarine, and coastal ecosystems worldwide. The effects of environmental nitrate, a principal form of nitrogen, on the health of aquatic life is of increasing concern. We exposed female American alligators to three concentrations of nitrate (0.7, 10 and 100mg/L NO 3 -N) for a duration of five weeks and five months from hatch. We assessed growth, plasma sex steroid and thyroid hormone concentrations, and transcription levels of key genes involved in steroidogenesis (StAR, 3β-HSD, and P450 scc ) and hepatic clearance (Cyp1a, Cyp3a). Exposure to 100mg/L NO 3 -N for both five weeks and five months resulted in significantly increased plasma testosterone (T) concentrations compared with alligators in the reference treatment. No differences in 17β-estradiol, progesterone, or thyroid hormones were observed, nor were there differences in alligator weight or the mRNA abundance of steroidogenic or hepatic genes. Plasma and urinary nitrate concentrations increased with increasing nitrate treatment levels, although relative plasma concentrations of nitrate were significantly lower in five month, versus five week old animals, possibly due to improved kidney function in older animals. These results indicate that environmentally relevant concentrations of nitrate can increase circulating concentrations of T in young female alligators. Copyright © 2016 Elsevier Inc. All rights reserved.

  8. Nitrate concentrations under irrigated agriculture

    USGS Publications Warehouse

    Zaporozec, A.

    1983-01-01

    In recent years, considerable interest has been expressed in the nitrate content of water supplies. The most notable toxic effect of nitrate is infant methemoglobinemia. The risk of this disease increases significantly at nitrate-nitrogen levels exceeding 10 mg/l. For this reason, this concentration has been established as a limit for drinking water in many countries. In natural waters, nitrate is a minor ionic constituent and seldom accounts for more than a few percent of the total anions. However, nitrate in a significant concentration may occur in the vicinity of some point sources such as septic tanks, manure pits, and waste-disposal sites. Non-point sources contributing to groundwater pollution are numerous and a majority of them are related to agricultural activities. The largest single anthropogenic input of nitrate into the groundwater is fertilizer. Even though it has not been proven that nitrogen fertilizers are responsible for much of nitrate pollution, they are generally recognized as the main threat to groundwater quality, especially when inefficiently applied to irrigated fields on sandy soils. The biggest challenge facing today's agriculture is to maintain the balance between the enhancement of crop productivity and the risk of groundwater pollution. ?? 1982 Springer-Verlag New York Inc.

  9. Modeling groundwater nitrate concentrations in private wells in Iowa

    USGS Publications Warehouse

    Wheeler, David C.; Nolan, Bernard T.; Flory, Abigail R.; DellaValle, Curt T.; Ward, Mary H.

    2015-01-01

    Contamination of drinking water by nitrate is a growing problem in many agricultural areas of the country. Ingested nitrate can lead to the endogenous formation of N-nitroso compounds, potent carcinogens. We developed a predictive model for nitrate concentrations in private wells in Iowa. Using 34,084 measurements of nitrate in private wells, we trained and tested random forest models to predict log nitrate levels by systematically assessing the predictive performance of 179 variables in 36 thematic groups (well depth, distance to sinkholes, location, land use, soil characteristics, nitrogen inputs, meteorology, and other factors). The final model contained 66 variables in 17 groups. Some of the most important variables were well depth, slope length within 1 km of the well, year of sample, and distance to nearest animal feeding operation. The correlation between observed and estimated nitrate concentrations was excellent in the training set (r-square = 0.77) and was acceptable in the testing set (r-square = 0.38). The random forest model had substantially better predictive performance than a traditional linear regression model or a regression tree. Our model will be used to investigate the association between nitrate levels in drinking water and cancer risk in the Iowa participants of the Agricultural Health Study cohort.

  10. Modeling groundwater nitrate concentrations in private wells in Iowa.

    PubMed

    Wheeler, David C; Nolan, Bernard T; Flory, Abigail R; DellaValle, Curt T; Ward, Mary H

    2015-12-01

    Contamination of drinking water by nitrate is a growing problem in many agricultural areas of the country. Ingested nitrate can lead to the endogenous formation of N-nitroso compounds, potent carcinogens. We developed a predictive model for nitrate concentrations in private wells in Iowa. Using 34,084 measurements of nitrate in private wells, we trained and tested random forest models to predict log nitrate levels by systematically assessing the predictive performance of 179 variables in 36 thematic groups (well depth, distance to sinkholes, location, land use, soil characteristics, nitrogen inputs, meteorology, and other factors). The final model contained 66 variables in 17 groups. Some of the most important variables were well depth, slope length within 1 km of the well, year of sample, and distance to nearest animal feeding operation. The correlation between observed and estimated nitrate concentrations was excellent in the training set (r-square=0.77) and was acceptable in the testing set (r-square=0.38). The random forest model had substantially better predictive performance than a traditional linear regression model or a regression tree. Our model will be used to investigate the association between nitrate levels in drinking water and cancer risk in the Iowa participants of the Agricultural Health Study cohort. Copyright © 2015 Elsevier B.V. All rights reserved.

  11. Digital spatial data for observed, predicted, and misclassification errors for observations in the training dataset for nitrate and arsenic concentrations in basin-fill aquifers in the Southwest Principal Aquifers study area

    USGS Publications Warehouse

    McKinney, Tim S.; Anning, David W.

    2012-01-01

    This product "Digital spatial data for observed, predicted, and misclassification errors for observations in the training dataset for nitrate and arsenic concentrations in basin-fill aquifers in the Southwest Principal Aquifers study area" is a 1:250,000-scale point spatial dataset developed as part of a regional Southwest Principal Aquifers (SWPA) study (Anning and others, 2012). The study examined the vulnerability of basin-fill aquifers in the southwestern United States to nitrate contamination and arsenic enrichment. Statistical models were developed by using the random forest classifier algorithm to predict concentrations of nitrate and arsenic across a model grid that represents local- and basin-scale measures of source, aquifer susceptibility, and geochemical conditions.

  12. High nitrate concentrations in some Midwest United States streams in 2013 after the 2012 drought

    USGS Publications Warehouse

    Van Metre, Peter C.; Frey, Jeffrey W.; Musgrove, MaryLynn; Nakagaki, Naomi; Qi, Sharon L.; Mahler, Barbara J.; Wieczorek, Michael; Button, Daniel T.

    2016-01-01

    Nitrogen sources in the Mississippi River basin have been linked to degradation of stream ecology and to Gulf of Mexico hypoxia. In 2013, the USGS and the USEPA characterized water quality stressors and ecological conditions in 100 wadeable streams across the midwestern United States. Wet conditions in 2013 followed a severe drought in 2012, a weather pattern associated with elevated nitrogen concentrations and loads in streams. Nitrate concentrations during the May to August 2013 sampling period ranged from <0.04 to 41.8 mg L−1 as N (mean, 5.31 mg L−1). Observed mean May to June nitrate concentrations at the 100 sites were compared with May to June concentrations predicted from a regression model developed using historical nitrate data. Observed concentrations for 17 sites, centered on Iowa and southern Minnesota, were outside the 95% confidence interval of the regression-predicted mean, indicating that they were anomalously high. The sites with a nitrate anomaly had significantly higher May to June nitrate concentrations than sites without an anomaly (means, 19.8 and 3.6 mg L−1, respectively) and had higher antecedent precipitation indices, a measure of the departure from normal precipitation, in 2012 and 2013. Correlations between nitrate concentrations and watershed characteristics and nitrogen and oxygen isotopes of nitrate indicated that fertilizer and manure used in crop production, principally corn, were the dominant sources of nitrate. The anomalously high nitrate levels in parts of the Midwest in 2013 coincide with reported higher-than-normal nitrate loads in the Mississippi River.

  13. [Nitrate concentrations in tap water in Spain].

    PubMed

    Vitoria, Isidro; Maraver, Francisco; Sánchez-Valverde, Félix; Armijo, Francisco

    2015-01-01

    To determine nitrate concentrations in drinking water in a sample of Spanish cities. We used ion chromatography to analyze the nitrate concentrations of public drinking water in 108 Spanish municipalities with more than 50,000 inhabitants (supplying 21,290,707 potential individuals). The samples were collected between January and April 2012. The total number of samples tested was 324. The median nitrate concentration was 3.47 mg/L (range: 0.38-66.76; interquartile range: 4.51). The water from 94% of the municipalities contained less than 15 mg/L. The concentration was higher than 25mg/L in only 3 municipalities and was greater than 50mg/L in one. Nitrate levels in most public drinking water supplies in municipalities inhabited by almost half of the Spanish population are below 15 mg/L. Copyright © 2014 SESPAS. Published by Elsevier Espana. All rights reserved.

  14. Serum nitrate/nitrite concentration correlates with gastric juice nitrate/nitrite: a possible marker for mutagenesis of the proximal stomach.

    PubMed

    Kishikawa, Hiroshi; Nishida, Jiro; Ichikawa, Hitoshi; Kaida, Shogo; Matsukubo, Takashi; Miura, Soichiro; Morishita, Tetsuo; Hibi, Toshifumi

    2011-01-01

    In the normal acid-secreting stomach, luminally generated nitric oxide, which contributes to carcinogenesis in the proximal stomach, is associated with the concentration of nitrate plus nitrite (nitrate/nitrite) in gastric juice. We investigated whether the serum nitrate/nitrite concentration is associated with that of gastric juice and whether it can be used as a serum marker. Serum and gastric juice nitrate/nitrite concentration, Helicobacter pylori antibody, and gastric pH were measured in 176 patients undergoing upper endoscopy. Multiple regression analysis revealed that serum nitrate/nitrite concentration was the best independent predictor of gastric juice nitrate/nitrite concentration. On single regression analysis, serum and gastric juice nitrate/nitrite concentration were significantly correlated, according to the following equation: gastric juice nitrate/nitrite concentration (μmol/l) = 3.93 - 0.54 × serum nitrate/nitrite concentration (μmol/l; correlation coefficient = 0.429, p < 0.001). In analyses confined to subjects with gastric pH less than 2.0, and in those with serum markers suggesting normal acid secretion (pepsinogen-I >30 ng/ml and negative H. pylori antibody), the serum nitrate/nitrite concentration was an independent predictor of the gastric juice nitrate/nitrite concentration (p < 0.001). Measuring the serum nitrate/nitrite concentration has potential in estimating the gastric juice nitrate/nitrite concentration. The serum nitrate/nitrite concentration could be useful as a marker for mutagenesis in the proximal stomach. Copyright © 2011 S. Karger AG, Basel.

  15. Relation of nitrate concentrations to baseflow in the Raccoon River, Iowa

    USGS Publications Warehouse

    Schilling, K.E.; Lutz, D.S.

    2004-01-01

    Excessive nitrate-nitrogen (nitrate) export from the Raccoon River in west central Iowa is an environmental concern to downstream receptors. The 1972 to 2000 record of daily streamflow and the results from 981 nitrate measurements were examined to describe the relation of nitrate to streamflow in the Raccoon River. No long term trends in streamflow and nitrate concentrations were noted in the 28-year record. Strong seasonal patterns were evident in nitrate concentrations, with higher concentrations occurring in spring and fall. Nitrate concentrations were linearly related to streamflow at daily, monthly, seasonal, and annual time scales. At all time scales evaluated, the relation was improved when baseflow was used as the discharge variable instead of total streamflow. Nitrate concentrations were found to be highly stratified according to flow, but there was little relation of nitrate to streamflow within each flow range. Simple linear regression models developed to predict monthly mean nitrate concentrations explained as much as 76 percent of the variability in the monthly nitrate concentration data for 2001. Extrapolation of current nitrate baseflow relations to historical conditions in the Raccoon River revealed that increasing baseflow over the 20th century could account for a measurable increase in nitrate concentrations.

  16. Nitrate-Rich Vegetables Increase Plasma Nitrate and Nitrite Concentrations and Lower Blood Pressure in Healthy Adults.

    PubMed

    Jonvik, Kristin L; Nyakayiru, Jean; Pinckaers, Philippe Jm; Senden, Joan Mg; van Loon, Luc Jc; Verdijk, Lex B

    2016-05-01

    Dietary nitrate is receiving increased attention due to its reported ergogenic and cardioprotective properties. The extent to which ingestion of various nitrate-rich vegetables increases postprandial plasma nitrate and nitrite concentrations and lowers blood pressure is currently unknown. We aimed to assess the impact of ingesting different nitrate-rich vegetables on subsequent plasma nitrate and nitrite concentrations and resting blood pressure in healthy normotensive individuals. With the use of a semirandomized crossover design, 11 men and 7 women [mean ± SEM age: 28 ± 1 y; mean ± SEM body mass index (BMI, in kg/m(2)): 23 ± 1; exercise: 1-10 h/wk] ingested 4 different beverages, each containing 800 mg (∼12.9 mmol) nitrate: sodium nitrate (NaNO3), concentrated beetroot juice, a rocket salad beverage, and a spinach beverage. Plasma nitrate and nitrite concentrations and blood pressure were determined before and up to 300 min after beverage ingestion. Data were analyzed using repeated-measures ANOVA. Plasma nitrate and nitrite concentrations increased after ingestion of all 4 beverages (P < 0.001). Peak plasma nitrate concentrations were similar for all treatments (all values presented as means ± SEMs: NaNO3: 583 ± 29 μmol/L; beetroot juice: 597 ± 23 μmol/L; rocket salad beverage: 584 ± 24 μmol/L; spinach beverage: 584 ± 23 μmol/L). Peak plasma nitrite concentrations were different between treatments (NaNO3: 580 ± 58 nmol/L; beetroot juice: 557 ± 57 nmol/L; rocket salad beverage: 643 ± 63 nmol/L; spinach beverage: 980 ± 160 nmol/L; P = 0.016). When compared with baseline, systolic blood pressure declined 150 min after ingestion of beetroot juice (from 118 ± 2 to 113 ± 2 mm Hg; P < 0.001) and rocket salad beverage (from 122 ± 3 to 116 ± 2 mm Hg; P = 0.007) and 300 min after ingestion of spinach beverage (from 118 ± 2 to 111 ± 3 mm Hg; P < 0.001), but did not change with NaNO3 Diastolic blood pressure declined 150 min after ingestion of all

  17. Biological denitrification of high concentration nitrate waste

    DOEpatents

    Francis, Chester W.; Brinkley, Frank S.

    1977-01-01

    Biological denitrification of nitrate solutions at concentrations of greater than one kilogram nitrate per cubic meter is accomplished anaerobically in an upflow column having as a packing material a support for denitrifying bacteria.

  18. Diminished Stream Nitrate Concentrations Linked to Dissolved Organic Carbon Dynamics After Leaf Fall

    NASA Astrophysics Data System (ADS)

    Sebestyen, S. D.; Shanley, J. B.; Boyer, E. W.; Doctor, D. H.; Kendall, C.

    2004-05-01

    Thermodynamic coupling of the nitrogen and carbon cycles has broad implications for controls on catchment nutrient fluxes. In the northeast US, leaf fall occurs in early October and the availability of organic carbon increases as the leaves decompose. At the Sleepers River Research Watershed in northeastern Vermont (USA), we sampled stream chemistry from seven nested catchments to determine how stream dissolved organic carbon (DOC) and nitrate vary as a function of flow conditions, land-use, and basin size in response to leaf fall. Following leaf fall, nitrate concentration patterns were quantitatively different from other times of the year. Under baseflow conditions, stream and soil water DOC concentrations were higher than normal, whereas nitrate concentrations declined sharply at the five smallest catchments and more modestly at the two largest catchments. Under high flow conditions, flushing of nitrate was observed, as is typical for stormflow response at Sleepers River. Our field data suggest that in-stream processing of nitrate is likely thermodynamically and kinetically favorable under baseflow but not at higher flow conditions when expanding variable source areas make hydrological connections between nitrate source areas and streams. We are working to evaluate this hypothesis with isotopic and other monitoring data, and to model the coupled interactions of water, DOC, and nitrate fluxes in these nested catchments.

  19. Effect of Nitrite/Nitrate concentrations on Corrosivity of Washed Precipitate

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

    Congdon, J.W.

    2001-03-28

    Cyclic polarization scans were performed using A-537 carbon steel in simulated washed precipitate solutions of various nitrite and nitrate concentrations. The results of this study indicate that nitrate is an aggressive anion in washed precipitate. Furthermore, a quantitative linear log-log relationship between the minimum effective nitrite concentration and the nitrate concentration was established for washed precipitate with other ions at their average compositions.

  20. Evaluating regional trends in ground-water nitrate concentrations of the Columbia Basin ground water management area, Washington

    USGS Publications Warehouse

    Frans, Lonna M.; Helsel, Dennis R.

    2005-01-01

    Trends in nitrate concentrations in water from 474 wells in 17 subregions in the Columbia Basin Ground Water Management Area (GWMA) in three counties in eastern Washington were evaluated using a variety of statistical techniques, including the Friedman test and the Kendall test. The Kendall test was modified from its typical 'seasonal' version into a 'regional' version by using well locations in place of seasons. No statistically significant trends in nitrate concentrations were identified in samples from wells in the GWMA, the three counties, or the 17 subregions from 1998 to 2002 when all data were included in the analysis. For wells in which nitrate concentrations were greater than 10 milligrams per liter (mg/L), however, a significant downward trend of -0.4 mg/L per year was observed between 1998 and 2002 for the GWMA as a whole, as well as for Adams County (-0.35 mg/L per year) and for Franklin County (-0.46 mg/L per year). Trend analysis for a smaller but longer-term 51-well dataset in Franklin County found a statistically significant upward trend in nitrate concentrations of 0.1 mg/L per year between 1986 and 2003. The largest increase of nitrate concentrations occurred between 1986 and 1991. No statistically significant differences were observed in this dataset between 1998 and 2003 indicating that the increase in nitrate concentrations has leveled off.

  1. Effects of over-winter green cover on soil solution nitrate concentrations beneath tillage land.

    PubMed

    Premrov, Alina; Coxon, Catherine E; Hackett, Richard; Kirwan, Laura; Richards, Karl G

    2014-02-01

    There is a growing need to reduce nitrogen losses from agricultural systems to increase food production while reducing negative environmental impacts. The efficacy of vegetation cover for reducing nitrate leaching in tillage systems during fallow periods has been widely investigated. Nitrate leaching reductions by natural regeneration (i.e. growth of weeds and crop volunteers) have been investigated to a lesser extent than reductions by planted cover crops. This study compares the efficacy of natural regeneration and a sown cover crop (mustard) relative to no vegetative cover under both a reduced tillage system and conventional plough-based system as potential mitigation measures for reducing over-winter soil solution nitrate concentrations. The study was conducted over three winter fallow seasons on well drained soil, highly susceptible to leaching, under temperate maritime climatic conditions. Mustard cover crop under both reduced tillage and conventional ploughing was observed to be an effective measure for significantly reducing nitrate concentrations. Natural regeneration under reduced tillage was found to significantly reduce the soil solution nitrate concentrations. This was not the case for the natural regeneration under conventional ploughing. The improved efficacy of natural regeneration under reduced tillage could be a consequence of potential stimulation of seedling germination by the autumn reduced tillage practices and improved over-winter plant growth. There was no significant effect of tillage practices on nitrate concentrations. This study shows that over winter covers of mustard and natural regeneration, under reduced tillage, are effective measures for reducing nitrate concentrations in free draining temperate soils. © 2013.

  2. Estimating soil solution nitrate concentration from dielectric spectra using PLS analysis

    USDA-ARS?s Scientific Manuscript database

    Fast and reliable methods for in situ monitoring of soil nitrate-nitrogen concentration are vital for reducing nitrate-nitrogen losses to ground and surface waters from agricultural systems. While several studies have been done to indirectly estimate nitrate-nitrogen concentration from time domain s...

  3. Effects of mineral dust on global atmospheric nitrate concentrations

    NASA Astrophysics Data System (ADS)

    Karydis, V. A.; Tsimpidi, A. P.; Pozzer, A.; Astitha, M.; Lelieveld, J.

    2016-02-01

    This study assesses the chemical composition and global aerosol load of the major inorganic aerosol components, focusing on mineral dust and aerosol nitrate. The mineral dust aerosol components (i.e., Ca2+, Mg2+, K+, Na+) and their emissions are included in the ECHAM5/MESSy Atmospheric Chemistry model (EMAC). Gas/aerosol partitioning is simulated using the ISORROPIA-II thermodynamic equilibrium model that considers K+, Ca2+, Mg2+, NH4+, Na+, SO42-, NO3-, Cl-, and H2O aerosol components. Emissions of mineral dust are calculated online by taking into account the soil particle size distribution and chemical composition of different deserts worldwide. Presence of metallic ions can substantially affect the nitrate partitioning into the aerosol phase due to thermodynamic interactions. The model simulates highest fine aerosol nitrate concentration over urban and industrialized areas (1-3 µg m-3), while coarse aerosol nitrate is highest close to deserts (1-4 µg m-3). The influence of mineral dust on nitrate formation extends across southern Europe, western USA, and northeastern China. The tropospheric burden of aerosol nitrate increases by 44 % when considering interactions of nitrate with mineral dust. The calculated global average nitrate aerosol concentration near the surface increases by 36 %, while the coarse- and fine-mode concentrations of nitrate increase by 53 and 21 %, respectively. Other inorganic aerosol components are affected by reactive dust components as well (e.g., the tropospheric burden of chloride increases by 9 %, ammonium decreases by 41 %, and sulfate increases by 7 %). Sensitivity tests show that nitrate aerosol is most sensitive to the chemical composition of the emitted mineral dust, followed by the soil size distribution of dust particles, the magnitude of the mineral dust emissions, and the aerosol state assumption.

  4. Stability of nitrate-ion concentrations in simulated deposition samples used for quality-assurance activities by the U.S. Geological Survey

    USGS Publications Warehouse

    Willoughby, T.C.; See, R.B.; Schroder, L.J.

    1989-01-01

    Three experiments were conducted to determine the stability of nitrate-ion concentrations in simulated deposition samples. In the four experiment-A solutions, nitric acid provided nitrate-ion concentrations ranging from 0.6 to 10.0 mg/L and that had pH values ranging from 3.8 to 5.0. In the five experiment-B solutions, sodium nitrate provided nitrate-ion concentrations ranging from 0.5 to 3.0 mg/L. The pH was adjusted to about 4.5 for each of the solutions by addition of sulfuric acid. In the four experiment-C solutions, nitric acid provided nitrate-ion concentrations ranging from 0.5 to 3.0 mg/L. Major cation and anion concentrations were added to each solution to simulate natural deposition. Aliquots were removed from the 13 original solutions and analyzed by ion chromatography about once a week for 100 days to determine if any changes occurred in nitrate-ion concentrations throughout the study period. No substantial changes were observed in the nitrate-ion concentrations in solutions that had initial concentrations below 4.0 mg/L in experiments A and B, although most of the measured nitrate-ion concentrations for the 100-day study were below the initial concentrations. In experiment C, changes in nitrate-ion concentrations were much more pronounced; the measured nitrate-ion concentrations for the study period were less than the initial concentrations for 62 of the 67 analyses. (USGS)

  5. Dynamic regression modeling of daily nitrate-nitrogen concentrations in a large agricultural watershed.

    PubMed

    Feng, Zhujing; Schilling, Keith E; Chan, Kung-Sik

    2013-06-01

    Nitrate-nitrogen concentrations in rivers represent challenges for water supplies that use surface water sources. Nitrate concentrations are often modeled using time-series approaches, but previous efforts have typically relied on monthly time steps. In this study, we developed a dynamic regression model of daily nitrate concentrations in the Raccoon River, Iowa, that incorporated contemporaneous and lags of precipitation and discharge occurring at several locations around the basin. Results suggested that 95 % of the variation in daily nitrate concentrations measured at the outlet of a large agricultural watershed can be explained by time-series patterns of precipitation and discharge occurring in the basin. Discharge was found to be a more important regression variable than precipitation in our model but both regression parameters were strongly correlated with nitrate concentrations. The time-series model was consistent with known patterns of nitrate behavior in the watershed, successfully identifying contemporaneous dilution mechanisms from higher relief and urban areas of the basin while incorporating the delayed contribution of nitrate from tile-drained regions in a lagged response. The first difference of the model errors were modeled as an AR(16) process and suggest that daily nitrate concentration changes remain temporally correlated for more than 2 weeks although temporal correlation was stronger in the first few days before tapering off. Consequently, daily nitrate concentrations are non-stationary, i.e. of strong memory. Using time-series models to reliably forecast daily nitrate concentrations in a river based on patterns of precipitation and discharge occurring in its basin may be of great interest to water suppliers.

  6. Observations on particulate organic nitrates and unidentified components of NO y

    NASA Astrophysics Data System (ADS)

    Nielsen, Torben; Egeløv, Axel H.; Granby, Kit; Skov, Henrik

    A method to determine the total content of particulate organic nitrates (PON) has been developed and ambient air measurements of PON, NO, N02, HNO3, peroxyacetyl nitrate (PAN), peroxypropionyl nitrate (PPN), gas NOY and particulate inorganic nitrate have been performed in the spring and early summer at an agricultural site in Denmark and compared with measurements of ozone, H 2O 2, SO 2, formic acid, acetic acid and methane sulphonic acid. The gas NO y detector determines the sum NO + NO 2 + HNO 2 + HNO 3 + PAN + PPN + gas phase organic nitrates + 2 × N 2O 5 + NO 3. The content of residual gas NO y ( = gas NO y - NO - NO 2 - HNO 3 - PAN - PPN) was determined and a group of unidentified NO y compounds was found. The phenomenon was observed at a site with relatively high NO x/NO y ratios compared to previous observations in U.S.A. and Canada. The residual gas NO y made up 7 ± 6% of total NOY (total NO y = gas NO y + particulate inorganic nitrate). Residual gas NO y was much higher than the particulate fraction of organic nitrates (PON). PON was only 0.25 ± 0.11% of total NO y. Both residual gas NO y and particulate organic nitrates episodes occurred with elevated concentrations of photochemical oxidants in connection with high-pressure systems suggesting atmospheric processes being the major source. Clean marine air can be discarded as a source for PON and residual gas NO y.

  7. Nitrate concentrations, 1936-99, and pesticide concentrations, 1990-99, in the unconfined aquifer in the San Luis Valley, Colorado

    USGS Publications Warehouse

    Stogner, Sr., Robert W.

    2001-01-01

    The first documented analysis of nitrate concentrations for ground water in the unconfined aquifer was done in 1936. This valleywide investigation indicated that nitrate concentrations were 0.3 milligram per liter or less in water-quality samples from 38 wells completed in the unconfined aquifer. A valleywide study conducted in the late 1940's documented the first occurrences of nitrate concentrations greater than 3 mg/L. Up to this time, soil fertility was maintained primarily through the use of cattle and (or) sheep manure and crop rotation. Subsequent valleywide studies have documented several occurrences of elevated nitrate concentrations in the unconfined aquifer in a localized, intensively cultivated area north of the Rio Grande. The nitrate concentrations in water appear to have changed in response to increasing use of commercial inorganic fertilizers after the mid-1940's. A 1993 valleywide study evaluated the potential health risk associated with elevated nitrate concentrations in domestic water supplies. Water-quality samples from 14 percent of the wells sampled contained nitrate concentrations greater than 10 milligrams per liter. Most of the samples that contained concentrations greater than 10 milligrams per liter were collected from wells located in the intensively cultivated area north of the Rio Grande. During the 1990's, several local, small-scale, and field-scale investigations were conducted in the intensively cultivated area north of the Rio Grande. These studies identified spatial and temporal variations in nitrate concentration and evaluated the effectiveness of using shallow monitoring wells to determine nitrate leaching. Variations in nitrate concentration were attributed, in part, to seasonal recharge of the aquifer by surface water with low nitrate concentrations. Shallow monitoring wells were effective for determining the amount of nitrate leached, but because of the amount of residual nitrate in the soil from previous seasons, were

  8. Are groundwater nitrate concentrations reaching a turning point in some chalk aquifers?

    PubMed

    Smith, J T; Clarke, R T; Bowes, M J

    2010-09-15

    In past decades, there has been much scientific effort dedicated to the development of models for simulation and prediction of nitrate concentrations in groundwaters, but producing truly predictive models remains a major challenge. A time-series model, based on long-term variations in nitrate fertiliser applications and average rainfall, was calibrated against measured concentrations from five boreholes in the River Frome catchment of Southern England for the period spanning from the mid-1970s to 2003. The model was then used to "blind" predict nitrate concentrations for the period 2003-2008. To our knowledge, this represents the first "blind" test of a model for predicting nitrate concentrations in aquifers. It was found that relatively simple time-series models could explain and predict a significant proportion of the variation in nitrate concentrations in these groundwater abstraction points (R(2)=0.6-0.9 and mean absolute prediction errors 4.2-8.0%). The study highlighted some important limitations and uncertainties in this, and other modelling approaches, in particular regarding long-term nitrate fertiliser application data. In three of the five groundwater abstraction points (Hooke, Empool and Eagle Lodge), once seasonal variations were accounted for, there was a recent change in the generally upward historical trend in nitrate concentrations. This may be an early indication of a response to levelling-off (and declining) fertiliser application rates since the 1980s. There was no clear indication of trend change at the Forston and Winterbourne Abbas sites nor in the trend of nitrate concentration in the River Frome itself from 1965 to 2008. Copyright 2010 Elsevier B.V. All rights reserved.

  9. Knowledge discovery from high-frequency stream nitrate concentrations: hydrology and biology contributions.

    PubMed

    Aubert, Alice H; Thrun, Michael C; Breuer, Lutz; Ultsch, Alfred

    2016-08-30

    High-frequency, in-situ monitoring provides large environmental datasets. These datasets will likely bring new insights in landscape functioning and process scale understanding. However, tailoring data analysis methods is necessary. Here, we detach our analysis from the usual temporal analysis performed in hydrology to determine if it is possible to infer general rules regarding hydrochemistry from available large datasets. We combined a 2-year in-stream nitrate concentration time series (time resolution of 15 min) with concurrent hydrological, meteorological and soil moisture data. We removed the low-frequency variations through low-pass filtering, which suppressed seasonality. We then analyzed the high-frequency variability component using Pareto Density Estimation, which to our knowledge has not been applied to hydrology. The resulting distribution of nitrate concentrations revealed three normally distributed modes: low, medium and high. Studying the environmental conditions for each mode revealed the main control of nitrate concentration: the saturation state of the riparian zone. We found low nitrate concentrations under conditions of hydrological connectivity and dominant denitrifying biological processes, and we found high nitrate concentrations under hydrological recession conditions and dominant nitrifying biological processes. These results generalize our understanding of hydro-biogeochemical nitrate flux controls and bring useful information to the development of nitrogen process-based models at the landscape scale.

  10. Estimating the Probability of Elevated Nitrate Concentrations in Ground Water in Washington State

    USGS Publications Warehouse

    Frans, Lonna M.

    2008-01-01

    Logistic regression was used to relate anthropogenic (manmade) and natural variables to the occurrence of elevated nitrate concentrations in ground water in Washington State. Variables that were analyzed included well depth, ground-water recharge rate, precipitation, population density, fertilizer application amounts, soil characteristics, hydrogeomorphic regions, and land-use types. Two models were developed: one with and one without the hydrogeomorphic regions variable. The variables in both models that best explained the occurrence of elevated nitrate concentrations (defined as concentrations of nitrite plus nitrate as nitrogen greater than 2 milligrams per liter) were the percentage of agricultural land use in a 4-kilometer radius of a well, population density, precipitation, soil drainage class, and well depth. Based on the relations between these variables and measured nitrate concentrations, logistic regression models were developed to estimate the probability of nitrate concentrations in ground water exceeding 2 milligrams per liter. Maps of Washington State were produced that illustrate these estimated probabilities for wells drilled to 145 feet below land surface (median well depth) and the estimated depth to which wells would need to be drilled to have a 90-percent probability of drawing water with a nitrate concentration less than 2 milligrams per liter. Maps showing the estimated probability of elevated nitrate concentrations indicated that the agricultural regions are most at risk followed by urban areas. The estimated depths to which wells would need to be drilled to have a 90-percent probability of obtaining water with nitrate concentrations less than 2 milligrams per liter exceeded 1,000 feet in the agricultural regions; whereas, wells in urban areas generally would need to be drilled to depths in excess of 400 feet.

  11. Improved daily precipitation nitrate and ammonium concentration models for the Chesapeake Bay Watershed.

    PubMed

    Grimm, J W; Lynch, J A

    2005-06-01

    Daily precipitation nitrate and ammonium concentration models were developed for the Chesapeake Bay Watershed (USA) using a linear least-squares regression approach and precipitation chemistry data from 29 National Atmospheric Deposition Program/National Trends Network (NADP/NTN) sites. Only weekly samples that comprised a single precipitation event were used in model development. The most significant variables in both ammonium and nitrate models included: precipitation volume, the number of days since the last event, a measure of seasonality, latitude, and the proportion of land within 8km covered by forest or devoted to industry and transportation. Additional variables included in the nitrate model were the proportion of land within 0.8km covered by water and/or forest. Local and regional ammonia and nitrogen oxide emissions were not as well correlated as land cover. Modeled concentrations compared very well with event chemistry data collected at six NADP/AirMoN sites within the Chesapeake Bay Watershed. Wet deposition estimates were also consistent with observed deposition at selected sites. Accurately describing the spatial distribution of precipitation volume throughout the watershed is important in providing critical estimates of wet-fall deposition of ammonium and nitrate.

  12. The relationship between the nitrate concentration and hydrology of a small chalk spring; Israel

    NASA Astrophysics Data System (ADS)

    Burg, Avi; Heaton, Tim H. E.

    1998-01-01

    Discharge from a spring draining a small, perched, Cretaceous chalk aquifer in the Upper Galilee, Israel, was monitored over a period of two years. The water has elevated nitrate concentrations, with 15N/ 14N and chemical data suggesting that it is a mixture of low-nitrate and high-nitrate end-members; the latter derived from the sewage of a centuries-old village served by septic tanks. Hydrograph data allowed distinction between fissure flow during the period of winter rainfall, and matrix drainage during the dry summer months. These different flow types, however, did not have markedly different nitrate concentrations: a 50-fold increase in spring discharge due to fissure flow, compared with matrix drainage, was reflected in only a 35% decrease in nitrate concentrations. The relatively high nitrate concentrations in the fissure waters suggests that they have had close contact with, and are possibly displaced from the matrix. This should help to accelerate the decline in the spring's nitrate concentrations following the recent completion of the village's central sewage drainage system.

  13. Profound differences between humans and rodents in the ability to concentrate salivary nitrate: Implications for translational research.

    PubMed

    Montenegro, Marcelo F; Sundqvist, Michaela L; Nihlén, Carina; Hezel, Michael; Carlström, Mattias; Weitzberg, Eddie; Lundberg, Jon O

    2016-12-01

    In humans dietary circulating nitrate accumulates rapidly in saliva through active transport in the salivary glands. By this mechanism resulting salivary nitrate concentrations are 10-20 times higher than in plasma. In the oral cavity nitrate is reduced by commensal bacteria to nitrite, which is subsequently swallowed and further metabolized to nitric oxide (NO) and other bioactive nitrogen oxides in blood and tissues. This entero-salivary circulation of nitrate is central in the various NO-like effects observed after ingestion of inorganic nitrate. The very same system has also been the focus of toxicologists studying potential carcinogenic effects of nitrite-dependent nitrosamine formation. Whether active transport of nitrate and accumulation in saliva occurs also in rodents is not entirely clear. Here we measured salivary and plasma levels of nitrate and nitrite in humans, rats and mice after administration of a standardized dose of nitrate. After oral (humans) or intraperitoneal (rodents) sodium nitrate administration (0.1mmol/kg), plasma nitrate levels increased markedly reaching ~300µM in all three species. In humans ingestion of nitrate was followed by a rapid increase in salivary nitrate to >6000µM, ie 20 times higher than those found in plasma. In contrast, in rats and mice salivary nitrate concentrations never exceeded the levels in plasma. Nitrite levels in saliva and plasma followed a similar pattern, ie marked increases in humans but modest elevations in rodents. In mice there was also no accumulation of nitrate in the salivary glands as measured directly in whole glands obtained after acute administration of nitrate. This study suggests that in contrast to humans, rats and mice do not actively concentrate circulating nitrate in saliva. These apparent species differences should be taken into consideration when studying the nitrate-nitrite-nitric oxide pathway in rodents, when calculating doses, exploring physiological, therapeutic and toxicological

  14. Spatial and temporal study of nitrate concentration in groundwater by means of coregionalization

    USGS Publications Warehouse

    D'Agostino, V.; Greene, E.A.; Passarella, G.; Vurro, M.

    1998-01-01

    Spatial and temporal behavior of hydrochemical parameters in groundwater can be studied using tools provided by geostatistics. The cross-variogram can be used to measure the spatial increments between observations at two given times as a function of distance (spatial structure). Taking into account the existence of such a spatial structure, two different data sets (sampled at two different times), representing concentrations of the same hydrochemical parameter, can be analyzed by cokriging in order to reduce the uncertainty of the estimation. In particular, if one of the two data sets is a subset of the other (that is, an undersampled set), cokriging allows us to study the spatial distribution of the hydrochemical parameter at that time, while also considering the statistical characteristics of the full data set established at a different time. This paper presents an application of cokriging by using temporal subsets to study the spatial distribution of nitrate concentration in the aquifer of the Lucca Plain, central Italy. Three data sets of nitrate concentration in groundwater were collected during three different periods in 1991. The first set was from 47 wells, but the second and the third are undersampled and represent 28 and 27 wells, respectively. Comparing the result of cokriging with ordinary kriging showed an improvement of the uncertainty in terms of reducing the estimation variance. The application of cokriging to the undersampled data sets reduced the uncertainty in estimating nitrate concentration and at the same time decreased the cost of the field sampling and laboratory analysis.Spatial and temporal behavior of hydrochemical parameters in groundwater can be studied using tools provided by geostatistics. The cross-variogram can be used to measure the spatial increments between observations at two given times as a function of distance (spatial structure). Taking into account the existence of such a spatial structure, two different data sets (sampled

  15. Effects of Unsaturated Microtopography on Nitrate Concentrations in Tundra Ecosystems: Examples from Polygonal Terrain and Degraded Peat Plateaus

    NASA Astrophysics Data System (ADS)

    Heikoop, J. M.; Arendt, C. A.; Newman, B. D.; Charsley-Groffman, L.; Perkins, G.; Wilson, C. J.; Wullschleger, S.

    2017-12-01

    Under the auspices of the Next Generation Ecosystem Experiment - Arctic, we have been studying hydrogeochemical signals in Alaskan tundra ecosystems underlain by continuous permafrost (Barrow Environmental Observatory (BEO)) and discontinuous permafrost (Seward Peninsula). The Barrow site comprises largely saturated tundra associated with the low gradient Arctic Coastal Plain. Polygonal microtopography, however, can result in slightly raised areas that are unsaturated. In these areas we have previously demonstrated production and accumulation of nitrate, which, based on nitrate isotopic analysis, derives from microbial degradation. Our Seward Peninsula site is located in a much steeper and generally well-drained watershed. In lower-gradient areas at the top and bottom of the watershed, however, the tundra is generally saturated, likely because of the presence of underlying discontinuous permafrost inhibiting infiltration. These settings also contain microtopographic features, though in the form of degraded peat plateaus surrounded by wet graminoid sag ponds. Despite being very different microtopographic features in a very different setting with distinct vegetation, qualitatively similar nitrate accumulation patterns as seen in polygonal terrain were observed. The highest nitrate pore water concentration observed in an unsaturated peat plateau was approximately 5 mg/L, whereas subsurface pore water concentrations in surrounding sag ponds were generally below the limit of detection. Nitrate isotopes indicate this nitrate results from microbial mineralization and nitrification based on comparison to the nitrate isotopic composition of reduced nitrogen sources in the environment and the oxygen isotope composition of site pore water. Nitrate concentrations were most similar to those found in low-center polygon rims and flat-centered polygon centers at the BEO, but were significantly lower than the maximum concentrations seen in the highest and driest polygonal features

  16. Mining nitrate concentration patterns from high-frequency in situ monitoring: a step towards more detailed understanding of hydrological processes?

    NASA Astrophysics Data System (ADS)

    Aubert, Alice; Houska, Tobias; Plesca, Ina; Kraft, Philipp; Breuer, Lutz

    2015-04-01

    Recently developed sensing technics allow collecting a considerable amount of high-frequency data; not only for hydrologic parameters (water levels, rainfall, etc.) but also for water chemistry. With devices such as in situ spectrophotometer, nitrate concentration can be monitored down to sub-hourly intervals. Thus, opening the way to new questions: what about daily or sub-daily instream nitrate concentration variations? What do these newly observed variations tell us about hydrological processes? In the Vollnkirchener Bach catchment, a headwater creek flows through a human impacted landscape dominated by agricultural and forest use and including a small settlement. Since March 2013, a Pro-PS device has been installed at the gauging station (monitored since 2011). Nitrate concentration is measured every 15 minutes, discharge and water temperature every 5 minutes. Data mining, more precisely motif discovery, is performed on these time series to identify high-resolution patterns. Spectral analysis highlighted that, in data measured at sub-hourly sampling frequency, variations up to a few hours are more likely to be dominated by measurement noise rather than real-world fluctuations. Therefore, we focus on daily motifs and flood patterns (given the fact that hydrological conditions are changing during flood events, we assume that nitrate concentration changes are depicting real processes). Various flood motifs were extracted: (1) nitrate can either be diluted or (2) concentrated, or (3) both (dilution followed by a bumpy recession curve indicating nitrate enrichment at the end of the flood). In addition to these classical nutrient-discharge behaviors, a variety of other interesting motifs were highlighted. (4) A daily nitrate cycle is clearly observed, but only during a specific year period. (5) Lag to peak time between parameters differentiate flood patterns: sometimes nitrate peaks first, sometimes discharge peaks first. (6) Furthermore, we are able to pinpoint the

  17. Coniferous coverage as well as catchment steepness influences local stream nitrate concentrations within a nitrogen-saturated forest in central Japan.

    PubMed

    Watanabe, Mirai; Miura, Shingo; Hasegawa, Shun; Koshikawa, Masami K; Takamatsu, Takejiro; Kohzu, Ayato; Imai, Akio; Hayashi, Seiji

    2018-04-28

    High concentrations of nitrate have been detected in streams flowing from nitrogen-saturated forests; however, the spatial variations of nitrate leaching within those forests and its causes remain poorly explored. The aim of this study is to evaluate the influences of catchment topography and coniferous coverage on stream nitrate concentrations in a nitrogen-saturated forest. We measured nitrate concentrations in the baseflow of headwater streams at 40 montane forest catchments on Mount Tsukuba in central Japan, at three-month intervals for 1 year, and investigated their relationship with catchment topography and with coniferous coverage. Although stream nitrate concentrations varied from 0.5 to 3.0 mgN L -1 , those in 31 catchments consistently exceeded 1 mgN L -1 , indicating that this forest had experienced nitrogen saturation. A classification and regression tree analysis with multiple environmental factors showed that the mean slope gradient and coniferous coverage were the best and second best, respectively, at explaining inter-catchment variance of stream nitrate concentrations. This analysis suggested that the catchments with steep topography and high coniferous coverage tend to have high nitrate concentrations. Moreover, in the three-year observation period for five adjacent catchments, the two catchments with relatively higher coniferous coverage consistently had higher stream nitrate concentrations. Thus, the spatial variations in stream nitrate concentrations were primarily regulated by catchment steepness and, to a lesser extent, coniferous coverage in this nitrogen-saturated forest. Our results suggest that a decrease in coniferous coverage could potentially contribute to a reduction in nitrate leaching from this nitrogen-saturated forest, and consequently reduce the risk of nitrogen overload for the downstream ecosystems. This information will allow land managers and researchers to develop improved management plans for this and similar

  18. Assessment of regional change in nitrate concentrations in groundwater in the Central Valley, California, USA, 1950s-2000s

    USGS Publications Warehouse

    Burow, Karen R.; Jurgens, Bryant C.; Belitz, Kenneth; Dubrovsky, Neil M.

    2013-01-01

    A regional assessment of multi-decadal changes in nitrate concentrations was done using historical data and a spatially stratified non-biased approach. Data were stratified into physiographic subregions on the basis of geomorphology and soils data to represent zones of historical recharge and discharge patterns in the basin. Data were also stratified by depth to represent a shallow zone generally representing domestic drinking-water supplies and a deep zone generally representing public drinking-water supplies. These stratifications were designed to characterize the regional extent of groundwater with common redox and age characteristics, two factors expected to influence changes in nitrate concentrations over time. Overall, increasing trends in nitrate concentrations and the proportion of nitrate concentrations above 5 mg/L were observed in the east fans subregion of the Central Valley. Whereas the west fans subregion has elevated nitrate concentrations, temporal trends were not detected, likely due to the heterogeneous nature of the water quality in this area and geologic sources of nitrate, combined with sparse and uneven data coverage. Generally low nitrate concentrations in the basin subregion are consistent with reduced geochemical conditions resulting from low permeability soils and higher organic content, reflecting the distal portions of alluvial fans and historical groundwater discharge areas. Very small increases in the shallow aquifer in the basin subregion may reflect downgradient movement of high nitrate groundwater from adjacent areas or overlying intensive agricultural inputs. Because of the general lack of regionally extensive long-term monitoring networks, the results from this study highlight the importance of placing studies of trends in water quality into regional context. Earlier work concluded that nitrate concentrations were steadily increasing over time in the eastern San Joaquin Valley, but clearly those trends do not apply to other

  19. Distribution of Elevated Nitrate Concentrations in Ground Water in Washington State

    USGS Publications Warehouse

    Frans, Lonna

    2008-01-01

    More than 60 percent of the population of Washington State uses ground water for their drinking and cooking needs. Nitrate concentrations in ground water are elevated in parts of the State as a result of various land-use practices, including fertilizer application, dairy operations and ranching, and septic-system use. Shallow wells generally are more vulnerable to nitrate contamination than deeper wells (Williamson and others, 1998; Ebbert and others, 2000). In order to protect public health, the Washington State Department of Health requires that public water systems regularly measure nitrate in their wells. Public water systems serving more than 25 people collect water samples at least annually; systems serving from 2 to 14 people collect water samples at least every 3 years. Private well owners serving one residence may be required to sample when the well is first drilled, but are unregulated after that. As a result, limited information is available to citizens and public health officials about potential exposure to elevated nitrate concentrations for people whose primary drinking-water sources are private wells. The U.S. Geological Survey and Washington State Department of Health collaborated to examine water-quality data from public water systems and develop models that calculate the probability of detecting elevated nitrate concentrations in ground water. Maps were then developed to estimate ground water vulnerability to nitrate in areas where limited data are available.

  20. Effect of arbuscular mycorrhizal and bacterial inocula on nitrate concentration in mesocosms simulating a wastewater treatment system relying on phytodepuration.

    PubMed

    Lingua, Guido; Copetta, Andrea; Musso, Davide; Aimo, Stefania; Ranzenigo, Angelo; Buico, Alessandra; Gianotti, Valentina; Osella, Domenico; Berta, Graziella

    2015-12-01

    High nitrogen concentration in wastewaters requires treatments to prevent the risks of eutrophication in rivers, lakes and coastal waters. The use of constructed wetlands is one of the possible approaches to lower nitrate concentration in wastewaters. Beyond supporting the growth of the bacteria operating denitrification, plants can directly take up nitrogen. Since plant roots interact with a number of soil microorganisms, in the present work we report the monitoring of nitrate concentration in macrocosms with four different levels of added nitrate (0, 30, 60 and 90 mg l(-1)), using Phragmites australis, inoculated with bacteria or arbuscular mycorrhizal fungi, to assess whether the use of such inocula could improve wastewater denitrification. Higher potassium nitrate concentration increased plant growth and inoculation with arbuscular mycorrhizal fungi or bacteria resulted in larger plants with more developed root systems. In the case of plants inoculated with arbuscular mycorrhizal fungi, a faster decrease of nitrate concentration was observed, while the N%/C% ratio of the plants of the different treatments remained similar. At 90 mg l(-1) of added nitrate, only mycorrhizal plants were able to decrease nitrate concentration to the limits prescribed by the Italian law. These data suggest that mycorrhizal and microbial inoculation can be an additional tool to improve the efficiency of denitrification in the treatment of wastewaters via constructed wetlands.

  1. Measurement of trace nitrate concentrations in seawater by ion chromatography with valve switching

    NASA Astrophysics Data System (ADS)

    Du, Juan; Fa, Yun; Zheng, Yue; Li, Xuebing; Du, Fanglin; Yang, Haiyan

    2014-05-01

    An ion chromatographic method with a valve switching facility was developed to determine trace nitrate concentrations in seawater using two pumps, two different suppressors, and two columns. A carbohydrate membrane desalter was used to reduce the high concentrations of sodium salts in samples. In this method, trace nitrate was eluted from the concentrator column to the analytical columns, while the matrix fl owed to waste. Neither chemical pre-treatment nor sample dilution was required. In the optimized separation conditions, the method showed good linearity ( R >0.99) in the 0.05 and 50 mg/L concentration range, and satisfactory repeatability (RSD<5%, n =6). The limit of detection for nitrate was 0.02 mg/L. Results showed that the valve switching system was suitable and practical for the determination of trace nitrate in seawater.

  2. The changing trend in nitrate concentrations in major aquifers due to historical nitrate loading from agricultural land across England and Wales from 1925 to 2150.

    PubMed

    Wang, L; Stuart, M E; Lewis, M A; Ward, R S; Skirvin, D; Naden, P S; Collins, A L; Ascott, M J

    2016-01-15

    Nitrate is necessary for agricultural productivity, but can cause considerable problems if released into aquatic systems. Agricultural land is the major source of nitrates in UK groundwater. Due to the long time-lag in the groundwater system, it could take decades for leached nitrate from the soil to discharge into freshwaters. However, this nitrate time-lag has rarely been considered in environmental water management. Against this background, this paper presents an approach to modelling groundwater nitrate at the national scale, to simulate the impacts of historical nitrate loading from agricultural land on the evolution of groundwater nitrate concentrations. An additional process-based component was constructed for the saturated zone of significant aquifers in England and Wales. This uses a simple flow model which requires modelled recharge values, together with published aquifer properties and thickness data. A spatially distributed and temporally variable nitrate input function was also introduced. The sensitivity of parameters was analysed using Monte Carlo simulations. The model was calibrated using national nitrate monitoring data. Time series of annual average nitrate concentrations along with annual spatially distributed nitrate concentration maps from 1925 to 2150 were generated for 28 selected aquifer zones. The results show that 16 aquifer zones have an increasing trend in nitrate concentration, while average nitrate concentrations in the remaining 12 are declining. The results are also indicative of the trend in the flux of groundwater nitrate entering rivers through baseflow. The model thus enables the magnitude and timescale of groundwater nitrate response to be factored into source apportionment tools and to be taken into account alongside current planning of land-management options for reducing nitrate losses. Copyright © 2015 The Authors. Published by Elsevier B.V. All rights reserved.

  3. Point source pollution and variability of nitrate concentrations in water from shallow aquifers

    NASA Astrophysics Data System (ADS)

    Nemčić-Jurec, Jasna; Jazbec, Anamarija

    2017-06-01

    Agriculture is one of the several major sources of nitrate pollution, and therefore the EU Nitrate Directive, designed to decrease pollution, has been implemented. Point sources like septic systems and broken sewage systems also contribute to water pollution. Pollution of groundwater by nitrate from 19 shallow wells was studied in a typical agricultural region, middle Podravina, in northwest Croatia. The concentration of nitrate ranged from <0.1 to 367 mg/l in water from wells, and 29.8 % of 253 total samples were above maximum acceptable value of 50 mg/l (MAV). Among regions R1-R6, there was no statistically significant difference in nitrate concentrations ( F = 1.98; p = 0.15) during the years 2002-2007. Average concentrations of nitrate in all 19 wells for all the analyzed years were between recommended limit value of 25 mg/l (RLV) and MAV except in 2002 (concentration was under RLV). The results of the repeated measures ANOVA showed statistically significant differences between the wells at the point source distance (proximity) of <10 m, compared to the wells at the point source distance of >20 m ( F = 10.6; p < 0.001). Average annual concentrations of nitrate during the years studied are not statistically different, but interaction between proximity and years is statistically significant ( F = 2.07; p = 0.04). Results of k-means clustering confirmed division into four clusters according to the pollution. Principal component analysis showed that there is only one significant factor, proximity, which explains 91.6 % of the total variability of nitrate. Differences in water quality were found as a result of different environmental factors. These results will contribute to the implementation of the Nitrate Directive in Croatia and the EU.

  4. Groundwater level and nitrate concentration trends on Mountain Home Air Force Base, southwestern Idaho

    USGS Publications Warehouse

    Williams, Marshall L.

    2014-01-01

    Mountain Home Air Force Base in southwestern Idaho draws most of its drinking water from the regional aquifer. The base is located within the State of Idaho's Mountain Home Groundwater Management Area and is adjacent to the State's Cinder Cone Butte Critical Groundwater Area. Both areas were established by the Idaho Department of Water Resources in the early 1980s because of declining water levels in the regional aquifer. The base also is listed by the Idaho Department of Environmental Quality as a nitrate priority area. The U.S. Geological Survey, in cooperation with the U.S. Air Force, began monitoring wells on the base in 1985, and currently monitors 25 wells for water levels and 17 wells for water quality, primarily nutrients. This report provides a summary of water-level and nitrate concentration data collected primarily between 2001 and 2013 and examines trends in those data. A Regional Kendall Test was run to combine results from all wells to determine an overall regional trend in water level. Groundwater levels declined at an average rate of about 1.08 feet per year. Nitrate concentration trends show that 3 wells (18 percent) are increasing in nitrate concentration trend, 3 wells (18 percent) show a decreasing nitrate concentration trend, and 11 wells (64 percent) show no nitrate concentration trend. Six wells (35 percent) currently exceed the U.S. Environmental Protection Agency's maximum contaminant limit of 10 milligrams per liter for nitrate (nitrite plus nitrate, measured as nitrogen).

  5. Fluoride, Nitrate, and Dissolved-Solids Concentrations in Ground Waters of Washington

    USGS Publications Warehouse

    Lum, W. E.; Turney, Gary L.

    1984-01-01

    This study provides basic data on ground-water quality throughout the State. It is intended for uses in planning and management by agencies and individuals who have responsibility for or interest in, public health and welfare. It also provides a basis for directing future studies of ground-water quality toward areas where ground-water quality problems may already exist. The information presented is a compilation of existing data from numerous sources including: the Washington Departments of Ecology and Social and Health Services, the Environmental Protection Agency, as well as many other local, county, state and federal agencies and private corporations. Only data on fluoride, nitrate, and dissolved-solids concentrations in ground water are presented, as these constituents are among those commonly used to determine the suitability of water for drinking or other purposes. They also reflect both natural and man-imposed effects on water quality and are the most readily available water-quality data for the State of Washington. The percentage of wells with fluoride, nitrate, or dissolved-solids concentrations exceeding U.S. Environmental Protection Agency Primary and Secondary Drinking Water Regulations were about 1, about 3, and about 3, respectively. Most high concentrations occurred in widely separated wells. Two exceptions were: high concentrations of nitrate and dissolved solids in wells on the Hanford Department of Energy Facility and high concentrations of nitrate in the lower Yakima River basin. (USGS)

  6. Effects of watershed-scale land use change on stream nitrate concentrations

    USGS Publications Warehouse

    Schilling, K.E.; Spooner, J.

    2006-01-01

    The Walnut Creek Watershed Monitoring Project was conducted from 1995 through 2005 to evaluate the response of stream nitrate concentrations to changing land use patterns in paired 5000-ha Iowa watersheds. A large portion of the Walnut Creek watershed is being converted from row crop agriculture to native prairie and savanna by the U.S. Fish and Wildlife Service at the Neal Smith National Wildlife Refuge (NSNWR). Before restoration, land use in both Walnut Creek (treatment) and Squaw Creek (control) watersheds consisted of 70% row crops. Between 1990 and 2005, row crop area decreased 25.4% in Walnut Creek due to prairie restoration but increased 9.2% in Squaw Creek due to Conservation Reserve Program (CRP) grassland conversion back to row crop. Nitrate concentrations ranged between <0.5 to 14 mg L-1 at the Walnut Creek outlet and 2.1 to 15 mg L-1 at the downstream Squaw Creek outlet. Nitrate concentrations decreased 1.2 mg L-1 over 10 yr in the Walnut Creek watershed but increased 1.9 mg L-1 over 10 yr in Squaw Creek. Changes in nitrate were easier to detect and more pronounced in monitored subbasins, decreasing 1.2 to 3.4 mg L-1 in three Walnut Creek subbasins, but increasing up to 8.0 and 11.6 mg L-1 in 10 yr in two Squaw Creek subbasins. Converting row crop lands to grass reduced stream nitrate levels over time in Walnut Creek, but stream nitrate rapidly increased in Squaw Creek when CRP grasslands were converted back to row crop. Study results highlight the close association of stream nitrate to land use change and emphasize that grasslands or other perennial vegetation placed in agricultural settings should be part of a long-term solution to water quality problems. ?? ASA, CSSA, SSSA.

  7. SPATIO-TEMPORAL ANALYSIS OF TOTAL NITRATE CONCENTRATIONS USING DYNAMIC STATISTICAL MODELS

    EPA Science Inventory

    Atmospheric concentrations of total nitrate (TNO3), defined here as gas-phase nitric acid plus particle-phase nitrate, are difficult to simulate in numerical air quality models due to the presence of a variety of formation pathways and loss mechanisms, some of which ar...

  8. Effect of nitrate concentration on filamentous bulking under low level of dissolved oxygen in an airlift inner circular anoxic-aerobic incorporate reactor.

    PubMed

    Su, Yiming; Zhang, Yalei; Zhou, Xuefei; Jiang, Ming

    2013-09-01

    This laboratory research investigated a possible cause of filamentous bulking under low level of dissolved oxygen conditions (dissolved oxygen value in aerobic zone maintained between 0.6-0.8 mg O2/L) in an airlift inner-circular anoxic-aerobic reactor. During the operating period, it was observed that low nitrate concentrations affected sludge volume index significantly. Unlike the existing hypothesis, the batch tests indicated that filamentous bacteria (mainly Thiothrix sp.) could store nitrate temporarily under carbon restricted conditions. When nitrate concentration was below 4 mg/L, low levels of carbon substrates and dissolved oxygen in the aerobic zone stimulated the nitrate-storing capacity of filaments. When filamentous bacteria riched in nitrate reached the anoxic zone, where they were exposed to high levels of carbon but limited nitrate, they underwent denitrification. However, when nonfilamentous bacteria were exposed to similar conditions, denitrification was restrained due to their intrinsic nitrate limitation. Hence, in order to avoid filamentous bulking, the nitrate concentration in the return sludge (from aerobic zone to the anoxic zone) should be above 4 mg/L, or alternatively, the nitrate load in the anoxic zone should be kept at levels above 2.7 mg NO(3-)-N/g SS.

  9. Evaluation of the Source and Transport of High Nitrate Concentrations in Ground Water, Warren Subbasin, California

    USGS Publications Warehouse

    Nishikawa, Tracy; Densmore, Jill N.; Martin, Peter; Matti, Jonathan

    2003-01-01

    , the midwest, the mideast, the east and the northeast hydrogeologic units. Water-quality analyses indicate that septage from septic tanks is the primary source of the high-nitrate concentrations measured in the Warren ground-water basin. Water-quality and stable-isotope data, collected after the start of the artificial recharge program, indicate that mixing occurs between imported water and native ground water, with the highest recorded nitrate concentrations in the midwest and the mideast hydrogeologic units. In general, the timing of the increase in measured nitrate concentrations in the midwest hydrogeologic unit is directly related to the distance of the monitoring well from a recharge site, indicating that the increase in nitrate concentrations is related to the artificial recharge program. Nitrate-to-chloride and nitrogen-isotope data indicate that septage is the source of the measured increase in nitrate concentrations in the midwest and the mideast hydrogeologic units. Samples from four wells in the Warren ground-water basin were analyzed for caffeine and selected human pharmaceutical products; these analyses suggest that septage is reaching the water table. There are two possible conceptual models that explain how high-nitrate septage reaches the water table: (1) the continued downward migration of septage through the unsaturated zone to the water table and (2) rising water levels, a result of the artificial recharge program, entraining septage in the unsaturated zone. The observations that nitrate concentrations increase in ground-water samples from wells soon after the start of the artificial recharge program in 1995 and that the largest increase in nitrate concentrations occur in the midwest and mideast hydrogeologic units where the largest increase in water levels occur indicate the validity of the second conceptual model (rising water levels). The potential nitrate concentration resulting from a water-level rise in the midwest and

  10. Effects of groundwater-flow paths on nitrate concentrations across two riparian forest corridors

    USGS Publications Warehouse

    Speiran, Gary K.

    2010-01-01

    Groundwater levels, apparent age, and chemistry from field sites and groundwater-flow modeling of hypothetical aquifers collectively indicate that groundwater-flow paths contribute to differences in nitrate concentrations across riparian corridors. At sites in Virginia (one coastal and one Piedmont), lowland forested wetlands separate upland fields from nearby surface waters (an estuary and a stream). At the coastal site, nitrate concentrations near the water table decreased from more than 10 mg/L beneath fields to 2 mg/L beneath a riparian forest buffer because recharge through the buffer forced water with concentrations greater than 5 mg/L to flow deeper beneath the buffer. Diurnal changes in groundwater levels up to 0.25 meters at the coastal site reflect flow from the water table into unsaturated soil where roots remove water and nitrate dissolved in it. Decreases in aquifer thickness caused by declines in the water table and decreases in horizontal hydraulic gradients from the uplands to the wetlands indicate that more than 95% of the groundwater discharged to the wetlands. Such discharge through organic soil can reduce nitrate concentrations by denitrification. Model simulations are consistent with field results, showing downward flow approaching toe slopes and surface waters to which groundwater discharges. These effects show the importance of buffer placement over use of fixed-width, streamside buffers to control nitrate concentrations.

  11. The relationship of nitrate concentrations in streams to row crop land use in Iowa

    USGS Publications Warehouse

    Schilling, K.E.; Libra, R.D.

    2000-01-01

    The relationship between row crop land use and nitrate N concentrations in surface water was evaluated for 15 Iowa watersheds ranging from 1002 to 2774 km2 and 10 smaller watersheds ranging from 47 to 775 km2 for the period 1996 to 1998. The percentage of land in row crop varied from 24 to >87% in the 15 large watersheds, and mean annual NO3-N concentrations ranged from 0.5 to 10.8 mg/L. In the small watersheds, row crop percentage varied from 28 to 87% and mean annual NO3-N concentrations ranged from 3.0 to 10.5 mg/L. In both cases, nitrate N concentrations were directly related to the percentage of row crop in the watershed (p 87% in the 15 large watersheds, and mean annual NO3-N concentrations ranged from 0.5 to 10.8 mg/L. In the small watersheds, row crop percentage varied from 28 to 87% and mean annual NO3-N concentrations ranged from 3.0 to 10.5 mg/L. In both cases, nitrate N concentrations were directly related to the percentage of row crop in the watershed (p<0.0003). Linear regression showed similar slope for both sets of watersheds (0.11) suggesting that average annual surface water nitrate concentrations in Iowa, and possibly similar agricultural areas in the midwestern USA, can be approximated by multiplying a watershed's row crop percentage by 0.1. Comparing the Iowa watershed data with similar data collected at a subwatershed scale in Iowa (0.1 to 8.1 km2) and a larger midcontinent scale (7300 to 237 100 km2) suggests that watershed scale affects the relationship of nitrate concentration and land use. The slope of nitrate concentration versus row crop percentage decreases with increasing watershed size.Mean nitrate concentrations and row crop land use were summarized for 15 larger and ten smaller watersheds in Iowa, and the relationship between NO3 concentration and land use was examined. Linear regression of mean NO3 concentration and percent row crop was highly significant for both sets of watershed data, but a stronger correlation was noted in the

  12. Secondary formation of nitrated phenols: insights from observations during the Uintah Basin Winter Ozone Study (UBWOS) 2014

    NASA Astrophysics Data System (ADS)

    Yuan, B.; Liggio, J.; Wentzell, J.; Li, S.-M.; Stark, H.; Roberts, J. M.; Gilman, J.; Lerner, B.; Warneke, C.; Li, R.; Leithead, A.; Osthoff, H. D.; Wild, R.; Brown, S. S.; de Gouw, J. A.

    2015-10-01

    We describe the results from online measurements of nitrated phenols using a time of flight chemical ionization mass spectrometer (ToF-CIMS) with acetate as reagent ion in an oil and gas production region in January and February of 2014. Strong diurnal profiles were observed for nitrated phenols, with concentration maxima at night. Based on known markers (CH4, NOx, CO2), primary emissions of nitrated phenols were not important in this study. A box model was used to simulate secondary formation of phenol, nitrophenol (NP) and dinitrophenols (DNP). The box model results indicate that oxidation of aromatics in the gas phase can explain the observed concentrations of NP and DNP in this study. Photolysis was the most efficient loss pathway for NP in the gas phase. We show that aqueous-phase reactions and heterogeneous reactions were minor sources of nitrated phenols in our study. This study demonstrates that the emergence of new ToF-CIMS (including PTR-TOF) techniques allows for the measurement of intermediate oxygenates at low levels and these measurements improve our understanding of the evolution of primary VOCs in the atmosphere.

  13. Secondary formation of nitrated phenols: insights from observations during the Uintah Basin Winter Ozone Study (UBWOS) 2014

    NASA Astrophysics Data System (ADS)

    Yuan, Bin; Liggio, John; Wentzell, Jeremy; Li, Shao-Meng; Stark, Harald; Roberts, James M.; Gilman, Jessica; Lerner, Brian; Warneke, Carsten; Li, Rui; Leithead, Amy; Osthoff, Hans D.; Wild, Robert; Brown, Steven S.; de Gouw, Joost A.

    2016-02-01

    We describe the results from online measurements of nitrated phenols using a time-of-flight chemical ionization mass spectrometer (ToF-CIMS) with acetate as reagent ion in an oil and gas production region in January and February of 2014. Strong diurnal profiles were observed for nitrated phenols, with concentration maxima at night. Based on known markers (CH4, NOx, CO2), primary emissions of nitrated phenols were not important in this study. A box model was used to simulate secondary formation of phenol, nitrophenol (NP), and dinitrophenols (DNP). The box model results indicate that oxidation of aromatics in the gas phase can explain the observed concentrations of NP and DNP in this study. Photolysis was the most efficient loss pathway for NP in the gas phase. We show that aqueous-phase reactions and heterogeneous reactions were minor sources of nitrated phenols in our study. This study demonstrates that the emergence of new ToF-CIMS (including PTR-TOF) techniques allows for the measurement of intermediate oxygenates at low levels and these measurements improve our understanding on the evolution of primary VOCs in the atmosphere.

  14. Sulfate and nitrate in Asian dust particles observed in desert, coastal and marine air

    NASA Astrophysics Data System (ADS)

    Zhang, D.; Wu, F.; Junji, C.

    2016-12-01

    Sulfate and nitrate in dust particles are believed to be two key species which can largely alter the physical and chemical properties of the particles in the atmosphere, in particular under humid conditions. Their occurrence in the particles has usually been considered to be the consequence of particles' aging during their long-distance travel in the air although they are present in some crustal minerals. Our observations at two deserts in China during dust episodes revealed that there were soil-derived sulfate and background-like nitrate in atmospheric dust samples. Sulfate in dust samples was proportional to samples' mass and comprised at steady mass percentages in differently sized samples. In contrast, nitrate concentration was approximately stable and independent from dust loading. Our observations at inland and coastal areas of China during dust episodes revealed that sulfate and nitrate were hardly produced on the surface of dust particles that were originated from the deserts areas in northwestern China. This is because the dust particles were in the postfrontal air, where the temperature was low and the relative humidity was small due to the adiabatic properties of the air mass. There are a number studies reporting that sulfate and nitrate had been efficiently produced on mineral particles in inland areas of China. However, those mineral particles were more likely from the local areas rather than from the desert areas. Our observations in the coastal areas of Japan, which is located in the downstream areas of the Asian continent and surrounded by sea areas revealed that dust particles appearing there frequently contained sulfate and nitrate, indicating sulfate and nitrate had been efficiently produced on the surface of the particles when the particles traveled in the marine air between China and Japan.

  15. Laboratory study of nitrate photolysis in Antarctic snow. I. Observed quantum yield, domain of photolysis, and secondary chemistry

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

    Meusinger, Carl; Johnson, Matthew S.; Berhanu, Tesfaye A.

    2014-06-28

    Post-depositional processes alter nitrate concentration and nitrate isotopic composition in the top layers of snow at sites with low snow accumulation rates, such as Dome C, Antarctica. Available nitrate ice core records can provide input for studying past atmospheres and climate if such processes are understood. It has been shown that photolysis of nitrate in the snowpack plays a major role in nitrate loss and that the photolysis products have a significant influence on the local troposphere as well as on other species in the snow. Reported quantum yields for the main reaction spans orders of magnitude – apparently amore » result of whether nitrate is located at the air-ice interface or in the ice matrix – constituting the largest uncertainty in models of snowpack NO{sub x} emissions. Here, a laboratory study is presented that uses snow from Dome C and minimizes effects of desorption and recombination by flushing the snow during irradiation with UV light. A selection of UV filters allowed examination of the effects of the 200 and 305 nm absorption bands of nitrate. Nitrate concentration and photon flux were measured in the snow. The quantum yield for loss of nitrate was observed to decrease from 0.44 to 0.003 within what corresponds to days of UV exposure in Antarctica. The superposition of photolysis in two photochemical domains of nitrate in snow is proposed: one of photolabile nitrate, and one of buried nitrate. The difference lies in the ability of reaction products to escape the snow crystal, versus undergoing secondary (recombination) chemistry. Modeled NO{sub x} emissions may increase significantly above measured values due to the observed quantum yield in this study. The apparent quantum yield in the 200 nm band was found to be ∼1%, much lower than reported for aqueous chemistry. A companion paper presents an analysis of the change in isotopic composition of snowpack nitrate based on the same samples as in this study.« less

  16. Groundwater nitrate concentration evolution under climate change and agricultural adaptation scenarios: Prince Edward Island, Canada

    NASA Astrophysics Data System (ADS)

    Paradis, Daniel; Vigneault, Harold; Lefebvre, René; Savard, Martine M.; Ballard, Jean-Marc; Qian, Budong

    2016-03-01

    Nitrate (N-NO3) concentration in groundwater, the sole source of potable water in Prince Edward Island (PEI, Canada), currently exceeds the 10 mg L-1 (N-NO3) health threshold for drinking water in 6 % of domestic wells. Increasing climatic and socio-economic pressures on PEI agriculture may further deteriorate groundwater quality. This study assesses how groundwater nitrate concentration could evolve due to the forecasted climate change and its related potential changes in agricultural practices. For this purpose, a tridimensional numerical groundwater flow and mass transport model was developed for the aquifer system of the entire Island (5660 km2). A number of different groundwater flow and mass transport simulations were made to evaluate the potential impact of the projected climate change and agricultural adaptation. According to the simulations for year 2050, N-NO3 concentration would increase due to two main causes: (1) the progressive attainment of steady-state conditions related to present-day nitrogen loadings, and (2) the increase in nitrogen loadings due to changes in agricultural practices provoked by future climatic conditions. The combined effects of equilibration with loadings, climate and agricultural adaptation would lead to a 25 to 32 % increase in N-NO3 concentration over the Island aquifer system. The change in groundwater recharge regime induced by climate change (with current agricultural practices) would only contribute 0 to 6 % of that increase for the various climate scenarios. Moreover, simulated trends in groundwater N-NO3 concentration suggest that an increased number of domestic wells (more than doubling) would exceed the nitrate drinking water criteria. This study underlines the need to develop and apply better agricultural management practices to ensure sustainability of long-term groundwater resources. The simulations also show that observable benefits from positive changes in agricultural practices would be delayed in time due to

  17. Groundwater nitrate concentration evolution under climate change and agricultural adaptation scenarios: Prince Edward Island, Canada

    NASA Astrophysics Data System (ADS)

    Paradis, D.; Vigneault, H.; Lefebvre, R.; Savard, M. M.; Ballard, J.-M.; Qian, B.

    2015-08-01

    Nitrate (N-NO3) concentration in groundwater, the sole source of potable water in Prince Edward Island (PEI, Canada), currently exceeds the 10 mg L-1 (N-NO3) health threshold for drinking water in 6 % of domestic wells. Increasing climatic and socio-economic pressures on PEI agriculture may further deteriorate groundwater quality. This study assesses how groundwater nitrate concentrations could evolve due to the forecasted climate change and its related potential changes in agricultural practices. For this purpose, a tridimensional numerical groundwater flow and mass transport model was developed for the aquifer system of the entire Island (5660 km2). A number of different groundwater flow and mass transport simulations were made to evaluate the potential impact of the projected climate change and agricultural adaptation. According to the simulations for year 2050, N-NO3 concentration would increase due to two main causes: (1) the progressive attainment of steady-state conditions related to present-day nitrogen loadings, and (2) the increase in nitrogen loadings due to changes in agricultural practices provoked by future climatic conditions. The combined effects of equilibration with loadings, climate and agricultural adaptation would lead to a 25 to 32 % increase in N-NO3 concentration over the Island aquifer system. Climate change alone (practices maintained at their current level) would contribute only 0 to 6 % to that increase according to the various climate scenarios. Moreover, simulated trends in groundwater N-NO3 concentration suggest that an increased number of domestic wells (more than doubling) would exceed the nitrate drinking water criteria. This study underlines the need to develop and apply better agricultural management practices to ensure sustainability of long-term groundwater resources. The simulations also show that observable benefits from positive changes in agricultural practices would be delayed in time due to the slow dynamics of nitrate

  18. Deep-sea coral evidence for lower Southern Ocean surface nitrate concentrations during the last ice age

    NASA Astrophysics Data System (ADS)

    Wang, Xingchen Tony; Sigman, Daniel M.; Prokopenko, Maria G.; Adkins, Jess F.; Robinson, Laura F.; Hines, Sophia K.; Chai, Junyi; Studer, Anja S.; Martínez-García, Alfredo; Chen, Tianyu; Haug, Gerald H.

    2017-03-01

    The Southern Ocean regulates the ocean’s biological sequestration of CO2 and is widely suspected to underpin much of the ice age decline in atmospheric CO2 concentration, but the specific changes in the region are debated. Although more complete drawdown of surface nutrients by phytoplankton during the ice ages is supported by some sediment core-based measurements, the use of different proxies in different regions has precluded a unified view of Southern Ocean biogeochemical change. Here, we report measurements of the 15N/14N of fossil-bound organic matter in the stony deep-sea coral Desmophyllum dianthus, a tool for reconstructing surface ocean nutrient conditions. The central robust observation is of higher 15N/14N across the Southern Ocean during the Last Glacial Maximum (LGM), 18-25 thousand years ago. These data suggest a reduced summer surface nitrate concentration in both the Antarctic and Subantarctic Zones during the LGM, with little surface nitrate transport between them. After the ice age, the increase in Antarctic surface nitrate occurred through the deglaciation and continued in the Holocene. The rise in Subantarctic surface nitrate appears to have had both early deglacial and late deglacial/Holocene components, preliminarily attributed to the end of Subantarctic iron fertilization and increasing nitrate input from the surface Antarctic Zone, respectively.

  19. Deep-sea coral evidence for lower Southern Ocean surface nitrate concentrations during the last ice age

    PubMed Central

    Sigman, Daniel M.; Prokopenko, Maria G.; Adkins, Jess F.; Robinson, Laura F.; Hines, Sophia K.; Chai, Junyi; Studer, Anja S.; Martínez-García, Alfredo; Chen, Tianyu; Haug, Gerald H.

    2017-01-01

    The Southern Ocean regulates the ocean’s biological sequestration of CO2 and is widely suspected to underpin much of the ice age decline in atmospheric CO2 concentration, but the specific changes in the region are debated. Although more complete drawdown of surface nutrients by phytoplankton during the ice ages is supported by some sediment core-based measurements, the use of different proxies in different regions has precluded a unified view of Southern Ocean biogeochemical change. Here, we report measurements of the 15N/14N of fossil-bound organic matter in the stony deep-sea coral Desmophyllum dianthus, a tool for reconstructing surface ocean nutrient conditions. The central robust observation is of higher 15N/14N across the Southern Ocean during the Last Glacial Maximum (LGM), 18–25 thousand years ago. These data suggest a reduced summer surface nitrate concentration in both the Antarctic and Subantarctic Zones during the LGM, with little surface nitrate transport between them. After the ice age, the increase in Antarctic surface nitrate occurred through the deglaciation and continued in the Holocene. The rise in Subantarctic surface nitrate appears to have had both early deglacial and late deglacial/Holocene components, preliminarily attributed to the end of Subantarctic iron fertilization and increasing nitrate input from the surface Antarctic Zone, respectively. PMID:28298529

  20. The role of climate on inter-annual variation in stream nitrate fluxes and concentrations.

    PubMed

    Gascuel-Odoux, Chantal; Aurousseau, Pierre; Durand, Patrick; Ruiz, Laurent; Molenat, Jérôme

    2010-11-01

    In recent decades, temporal variations in nitrate fluxes and concentrations in temperate rivers have resulted from the interaction of anthropogenic and climatic factors. The effect of climatic drivers remains unclear, while the relative importance of the drivers seems to be highly site dependent. This paper focuses on 2-6 year variations called meso-scale variations, and analyses the climatic drivers of these variations in a study site characterized by high N inputs from intensive animal farming systems and shallow aquifers with impervious bedrock in a temperate climate. Three approaches are developed: 1) an analysis of long-term records of nitrate fluxes and nitrate concentrations in 30 coastal rivers of Western France, which were well-marked by meso-scale cycles in the fluxes and concentration with a slight hysteresis; 2) a test of the climatic control using a lumped two-box model, which demonstrates that hydrological assumptions are sufficient to explain these meso-scale cycles; and 3) a model of nitrate fluxes and concentrations in two contrasted catchments subjected to recent mitigation measures, which analyses nitrate fluxes and concentrations in relation to N stored in groundwater. In coastal rivers, hydrological drivers (i.e., effective rainfall), and particularly the dynamics of the water table and rather stable nitrate concentration, explain the meso-scale cyclic patterns. In the headwater catchment, agricultural and hydrological drivers can interact according to their settings. The requirements to better distinguish the effect of climate and human changes in integrated water management are addressed: long-term monitoring, coupling the analysis and the modelling of large sets of catchments incorporating different sizes, land uses and environmental factors. Copyright © 2009 Elsevier B.V. All rights reserved.

  1. Predicted nitrate and arsenic concentrations in basin-fill aquifers of the Southwestern United States

    USGS Publications Warehouse

    Anning, David W.; Paul, Angela P.; McKinney, Tim S.; Huntington, Jena M.; Bexfield, Laura M.; Thiros, Susan A.

    2012-01-01

    The National Water-Quality Assessment (NAWQA) Program of the U.S. Geological Survey (USGS) is conducting a regional analysis of water quality in the principal aquifer systems across the United States. The Southwest Principal Aquifers (SWPA) study is building a better understanding of the susceptibility and vulnerability of basin-fill aquifers in the region to groundwater contamination by synthesizing baseline knowledge of groundwater-quality conditions in 16 basins previously studied by the NAWQA Program. The improved understanding of aquifer susceptibility and vulnerability to contamination is assisting in the development of tools that water managers can use to assess and protect the quality of groundwater resources.Human-health concerns and economic considerations associated with meeting drinking-water standards motivated a study of the vulnerability of basin-fill aquifers to nitrate con­tamination and arsenic enrichment in the southwestern United States. Statistical models were developed by using the random forest classifier algorithm to predict concentrations of nitrate and arsenic across a model grid that represents about 190,600 square miles of basin-fill aquifers in parts of Arizona, California, Colorado, Nevada, New Mexico, and Utah. The statistical models, referred to as classifiers, reflect natural and human-related factors that affect aquifer vulnerability to contamina­tion and relate nitrate and arsenic concentrations to explana­tory variables representing local- and basin-scale measures of source, aquifer susceptibility, and geochemical conditions. The classifiers were unbiased and fit the observed data well, and misclassifications were primarily due to statistical sampling error in the training datasets.The classifiers were designed to predict concentrations to be in one of six classes for nitrate, and one of seven classes for arsenic. Each classification scheme allowed for identification of areas with concentrations that were equal to or exceeding

  2. Nitrous oxide and nitrate concentration in under-drainage from arable fields subject to diffuse pollution mitigation measures

    NASA Astrophysics Data System (ADS)

    Hama-Aziz, Zanist; Hiscock, Kevin; Adams, Christopher; Reid, Brian

    2016-04-01

    Atmospheric nitrous oxide concentrations are increasing by 0.3% annually and a major source of this greenhouse gas is agriculture. Indirect emissions of nitrous oxide (e.g. from groundwater and surface water) account for about quarter of total nitrous oxide emissions. However, these indirect emissions are subject to uncertainty, mainly due to the range in reported emission factors. It's hypothesised in this study that cover cropping and implementing reduced (direct drill) cultivation in intensive arable systems will reduce dissolved nitrate concentration and subsequently indirect nitrous oxide emissions. To test the hypothesis, seven fields with a total area of 102 ha in the Wensum catchment in eastern England have been chosen for experimentation together with two fields (41 ha) under conventional cultivation (deep inversion ploughing) for comparison. Water samples from field under-drainage have been collected for nitrate and nitrous oxide measurement on a weekly basis from April 2013 for two years from both cultivation areas. A purge and trap preparation line connected to a Shimadzu GC-8A gas chromatograph fitted with an electron capture detector was used for dissolved nitrous oxide analysis. Results revealed that with an oilseed radish cover crop present, the mean concentration of nitrate, which is the predominant form of N, was significantly depleted from 13.9 mg N L-1 to 2.5 mg N L-1. However, slightly higher mean nitrous oxide concentrations under the cover crop of 2.61 μg N L-1 compared to bare fields of 2.23 μg N L-1 were observed. Different inversion intensity of soil tended to have no effect on nitrous oxide and nitrate concentrations. The predominant production mechanism for nitrous oxide was nitrification process and the significant reduction of nitrate was due to plant uptake rather than denitrification. It is concluded that although cover cropping might cause a slight increase of indirect nitrous oxide emission, it can be a highly effective

  3. Concentrated Animal Feeding Operations, Row Crops and their Relationship to Nitrate in Eastern Iowa Rivers

    PubMed Central

    Weldon, Mark B.; Hornbuckle, Keri C.

    2009-01-01

    Concentrated animal feeding operations (CAFO) and fertilizer application to row crops may contribute to poor water quality in surface waters. To test this hypothesis, we evaluated nutrient concentrations and fluxes in four Eastern Iowa watersheds sampled between 1996-2004. We found that these watersheds contribute nearly 10% of annual nitrate flux entering the Gulf of Mexico, while representing only 1.5% of the contributing drainage basin. Mass budget analysis shows stream flow to be a major loss of nitrogen (18% of total N output), second only to crop harvest (63%). The major watershed inputs of nitrogen include applied fertilizer for corn (54% of total N input) and nitrogen fixation by soybeans (26%). Despite the relatively small input from animal manure (~5%), the results of spatial analysis indicate that row crop and CAFO densities are significantly and independently correlated to higher nitrate concentration in streams. Pearson correlation coefficients of 0.59 and 0.89 were found between nitrate concentration and row crop and CAFO density, respectively. Multiple linear regression analysis produced a correlation for nitrate concentration with an R2 value of 85%. High spatial density of row crops and CAFOs are linked to the highest river nitrate concentrations (up to 15 mg/l normalized over five years). PMID:16749677

  4. Evaluation of Nitrate Concentrations and Sources in the Elk Creek Watershed, Southwestern Ohio, 2003-2004

    USGS Publications Warehouse

    Schumann, Thomas L.; Pletsch, Bruce A.

    2006-01-01

    Nitrate concentrations exceeding the U.S. Environmental Protection Agency maximum contaminant level of 10 milligrams per liter have been reported in ground water near the City of Trenton, Ohio, in the southern part of the Elk Creek watershed. A study of nitrate concentrations and sources in surface and ground water within the Elk Creek watershed was conducted during 2003 and 2004. Nitrate concentrations in the Elk Creek watershed range from less than 0.06 to 11 milligrams per liter. The likely sources of elevated nitrate in the ground water near the City of Trenton appear to be soil organic matter and ammonia fertilizer. Land use is predominantly (93 percent) agricultural, with no identified point sources of nitrate. Likely sources of nitrate in the surface water appear to be manure and septic system effluent, soil organic matter, and ammonia fertilizer. Water-quality constituents, including nitrate, were sampled in water from 38 wells and at 6 surface-water sites. The wells were all shallow (less than 105 feet deep), with open intervals in aquifers of glacial origin, that include tills, outwash, and alluvium. Nitrate concentrations (median of 0.06 milligrams per liter) in the ground water of the upper section of the watershed were lower than those in the lower section of the watershed (median of 4.2 milligrams per liter). Nitrate was analyzed for nitrogen and oxygen isotope values. The d15N and d18O range from -22.36 to +32.29 per mil, and -6.27 to +17.72 per mil, respectively. A positive correlation of d15N and d18O enrichment indicates that denitrification is a prevalent process within the watershed.

  5. Seasonal dynamics of nitrate and ammonium ion concentrations in soil solutions collected using MacroRhizon suction cups.

    PubMed

    Kabala, Cezary; Karczewska, Anna; Gałka, Bernard; Cuske, Mateusz; Sowiński, Józef

    2017-07-01

    The aims of the study were to analyse the concentration of nitrate and ammonium ions in soil solutions obtained using MacroRhizon miniaturized composite suction cups under field conditions and to determine potential nitrogen leaching from soil fertilized with three types of fertilizers (standard urea, slow-release urea, and ammonium nitrate) at the doses of 90 and 180 kg ha -1 , applied once or divided into two rates. During a 3-year growing experiment with sugar sorghum, the concentration of nitrate and ammonium ions in soil solutions was the highest with standard urea fertilization and the lowest in variants fertilized with slow-release urea for most of the months of the growing season. Higher concentrations of both nitrogen forms were noted at the fertilizer dose of 180 kg ha -1 . One-time fertilization, at both doses, resulted in higher nitrate concentrations in June and July, while dividing the dose into two rates resulted in higher nitrate concentrations between August and November. The highest potential for nitrate leaching during the growing season was in July. The tests confirmed that the miniaturized suction cups MacroRhizon are highly useful for routine monitoring the concentration of nitrate and ammonium ions in soil solutions under field conditions.

  6. Relation of pathways and transit times of recharge water to nitrate concentrations using stable isotopes

    USGS Publications Warehouse

    Landon, M.K.; Delin, G.N.; Komor, S.C.; Regan, C.P.

    2000-01-01

    Oxygen and hydrogen stable isotope values of precipitation, irrigation water, soil water, and ground water were used with soil-moisture contents and water levels to estimate transit times and pathways of recharge water in the unsaturated zone of a sand and gravel aquifer. Nitrate-nitrogen (nitrate) concentrations in ground water were also measured to assess their relation to seasonal recharge. Stable isotope values indicated that recharge water usually had a transit time through the unsaturated zone of several weeks to months. However, wetting fronts usually moved through the unsaturated zone in hours to weeks. The much slower transit of isotopic signals than that of wetting fronts indicates that recharge was predominantly composed of older soil water that was displaced downward by more recent infiltrating water. Comparison of observed and simulated isotopic values from pure-piston flow and mixing-cell water and isotope mass balance models indicates that soil water isotopic values were usually highly mixed. Thus, movement of recharge water did not occur following a pure piston-flow displacement model but rather follows a hydrid model involving displacement of mixed older soil water with new infiltration water. An exception to this model occurred in a topographic depression, where movement of water along preferential flowpaths to the water table occurred within hours to days following spring thaw as result of depression-focused infiltration of snow melt. In an adjacent upland area, recharge of snow melt occurred one to two months later. Increases in nitrate concentrations at the water table during April-May 1993 and 1994 in a topographic lowland within a corn field were related to recharge of water that had infiltrated the previous summer and was displaced from the unsaturated zone by spring infiltration. Increases in nitrate concentrations also occurred during July-August 1994 in response to recharge of water that infiltrated during May-August 1994. These results

  7. Probability of detecting atrazine/desethyl-atrazine and elevated concentrations of nitrate in ground water in Colorado

    USGS Publications Warehouse

    Rupert, Michael G.

    2003-01-01

    Draft Federal regulations may require that each State develop a State Pesticide Management Plan for the herbicides atrazine, alachlor, metolachlor, and simazine. Maps were developed that the State of Colorado could use to predict the probability of detecting atrazine and desethyl-atrazine (a breakdown product of atrazine) in ground water in Colorado. These maps can be incorporated into the State Pesticide Management Plan and can help provide a sound hydrogeologic basis for atrazine management in Colorado. Maps showing the probability of detecting elevated nitrite plus nitrate as nitrogen (nitrate) concentrations in ground water in Colorado also were developed because nitrate is a contaminant of concern in many areas of Colorado. Maps showing the probability of detecting atrazine and(or) desethyl-atrazine (atrazine/DEA) at or greater than concentrations of 0.1 microgram per liter and nitrate concentrations in ground water greater than 5 milligrams per liter were developed as follows: (1) Ground-water quality data were overlaid with anthropogenic and hydrogeologic data using a geographic information system to produce a data set in which each well had corresponding data on atrazine use, fertilizer use, geology, hydrogeomorphic regions, land cover, precipitation, soils, and well construction. These data then were downloaded to a statistical software package for analysis by logistic regression. (2) Relations were observed between ground-water quality and the percentage of land-cover categories within circular regions (buffers) around wells. Several buffer sizes were evaluated; the buffer size that provided the strongest relation was selected for use in the logistic regression models. (3) Relations between concentrations of atrazine/DEA and nitrate in ground water and atrazine use, fertilizer use, geology, hydrogeomorphic regions, land cover, precipitation, soils, and well-construction data were evaluated, and several preliminary multivariate models with various

  8. Geochemical controls on microbial nitrate-dependent U(IV) oxidation

    USGS Publications Warehouse

    Senko, John M.; Suflita, Joseph M.; Krumholz, Lee R.

    2005-01-01

    After reductive immobilization of uranium, the element may be oxidized and remobilized in the presence of nitrate by the activity of dissimilatory nitrate-reducing bacteria. We examined controls on microbially mediated nitrate-dependent U(IV) oxidation in landfill leachate-impacted subsurface sediments. Nitrate-dependent U(IV)-oxidizing bacteria were at least two orders of magnitude less numerous in these sediments than glucose- or Fe(II)-oxidizing nitrate-reducing bacteria and grew more slowly than the latter organisms, suggesting that U(IV) is ultimately oxidized by Fe(III) produced by nitrate-dependent Fe(II)-oxidizing bacteria or by oxidation of Fe(II) by nitrite that accumulates during organotrophic dissimilatory nitrate reduction. We examined the effect of nitrate and reductant concentration on nitrate-dependent U(IV) oxidation in sediment incubations and used the initial reductive capacity (RDC = [reducing equivalents] - [oxidizing equivalents]) of the incubations as a unified measurement of the nitrate or reductant concentration. When we lowered the RDC with progressively higher nitrate concentrations, we observed a corresponding increase in the extent of U(IV) oxidation, but did not observe this relationship between RDC and U(IV) oxidation rate, especially when RDC > 0, suggesting that nitrate concentration strongly controls the extent, but not the rate of nitrate-dependent U(IV) oxidation. On the other hand, when we raised the RDC in sediment incubations with progressively higher reductant (acetate, sulfide, soluble Fe(II), or FeS) concentrations, we observed progressively lower extents and rates of nitrate-dependent U(IV) oxidation. Acetate was a relatively poor inhibitor of nitrate-dependent U(IV) oxidation, while Fe(II) was the most effective inhibitor. Based on these results, we propose that it may be possible to predict the stability of U(IV) in a bioremediated aquifer based on the geochemical characteristics of that aquifer.

  9. Nitrate sources and sinks in Elkhorn Slough, California: Results from long-term continuous in situ nitrate analyzers

    USGS Publications Warehouse

    Chapin, T.P.; Caffrey, J.M.; Jannasch, H.W.; Coletti, L.J.; Haskins, J.C.; Johnson, K.S.

    2004-01-01

    Nitrate and water quality parameters (temperature, salinity, dissolved oxygen, turbidity, and depth) were measured continuously with in situ NO 3 analyzers and water quality sondes at two sites in Elkhorn Slough in Central California. The Main Channel site near the mouth of Elkhorn Slough was sampled from February to September 2001. Azevedo Pond, a shallow tidal pond bordering agricultural fields further inland, was sampled from December 1999 to July 2001. Nitrate concentrations were recorded hourly while salinity, temperature, depth, oxygen, and turbidity were recorded every 30 min. Nitrate concentrations at the Main Channel site ranged from 5 to 65 ??M. The propagation of an internal wave carrying water from ???100 m depth up the Monterey Submarine Canyon and into the lower section of Elkhorn Slough on every rising tide was a major source of nitrate, accounting for 80-90% of the nitrogen load during the dry summer period. Nitrate concentrations in Azevedo Pond ranged from 0-20 ??M during the dry summer months. Nitrate in Azevedo Pond increased to over 450 ??M during a heavy winter precipitation event, and interannual variability driven by differences in precipitation was observed. At both sites, tidal cycling was the dominant forcing, often changing nitrate concentrations by 5-fold or more within a few hours. Water volume flux estimates were combined with observed nitrate concentrations to obtain nitrate fluxes. Nitrate flux calculations indicated a loss of 4 mmol NO3 m -2 d-1 for the entire Elkhorn Slough and 1 mmol NO 3 m-2 d-1 at Azevedo Pond. These results suggested that the waters of Elkhorn Slough were not a major source of nitrate to Monterey Bay but actually a nitrate sink during the dry season. The limited winter data at the Main Channel site suggest that nitrate was exported from Elkhorn Slough during the wet season. Export of ammonium or dissolved organic nitrogen, which we did not monitor, may balance some or all of the NO 3 flux.

  10. Nitrite and Nitrate Concentrations and Metabolism in Breast Milk, Infant Formula, and Parenteral Nutrition

    PubMed Central

    Jones, Jesica A.; Ninnis, Janet R.; Hopper, Andrew O.; Ibrahim, Yomna; Merritt, T. Allen; Wan, Kim-Wah; Power, Gordon G.; Blood, Arlin B.

    2015-01-01

    Dietary nitrate and nitrite are sources of gastric NO, which modulates blood flow, mucus production, and microbial flora. However, the intake and importance of these anions in infants is largely unknown. Nitrate and nitrite levels were measured in breast milk of mothers of preterm and term infants, infant formulas, and parenteral nutrition. Nitrite metabolism in breast milk was measured after freeze-thawing, at different temperatures, varying oxygen tensions, and after inhibition of potential nitrite-metabolizing enzymes. Nitrite concentrations averaged 0.07 ± 0.01 μM in milk of mothers of preterm infants, less than that of term infants (0.13 ± 0.02 μM) (P < .01). Nitrate concentrations averaged 13.6 ± 3.7 μM and 12.7 ± 4.9 μM, respectively. Nitrite and nitrate concentrations in infant formulas varied from undetectable to many-fold more than breast milk. Concentrations in parenteral nutrition were equivalent to or lower than those of breast milk. Freeze-thawing decreased nitrite concentration ∼64%, falling with a half-life of 32 minutes at 37°C. The disappearance of nitrite was oxygen-dependent and prevented by ferricyanide and 3 inhibitors of lactoperoxidase. Nitrite concentrations in breast milk decrease with storage and freeze-thawing, a decline likely mediated by lactoperoxidase. Compared to adults, infants ingest relatively little nitrite and nitrate, which may be of importance in the modulation of blood flow and the bacterial flora of the infant GI tract, especially given the protective effects of swallowed nitrite. PMID:23894175

  11. Nitrite and nitrate concentrations and metabolism in breast milk, infant formula, and parenteral nutrition.

    PubMed

    Jones, Jesica A; Ninnis, Janet R; Hopper, Andrew O; Ibrahim, Yomna; Merritt, T Allen; Wan, Kim-Wah; Power, Gordon G; Blood, Arlin B

    2014-09-01

    Dietary nitrate and nitrite are sources of gastric NO, which modulates blood flow, mucus production, and microbial flora. However, the intake and importance of these anions in infants is largely unknown. Nitrate and nitrite levels were measured in breast milk of mothers of preterm and term infants, infant formulas, and parenteral nutrition. Nitrite metabolism in breast milk was measured after freeze-thawing, at different temperatures, varying oxygen tensions, and after inhibition of potential nitrite-metabolizing enzymes. Nitrite concentrations averaged 0.07 ± 0.01 μM in milk of mothers of preterm infants, less than that of term infants (0.13 ± 0.02 μM) (P < .01). Nitrate concentrations averaged 13.6 ± 3.7 μM and 12.7 ± 4.9 μM, respectively. Nitrite and nitrate concentrations in infant formulas varied from undetectable to many-fold more than breast milk. Concentrations in parenteral nutrition were equivalent to or lower than those of breast milk. Freeze-thawing decreased nitrite concentration ~64%, falling with a half-life of 32 minutes at 37°C. The disappearance of nitrite was oxygen-dependent and prevented by ferricyanide and 3 inhibitors of lactoperoxidase. Nitrite concentrations in breast milk decrease with storage and freeze-thawing, a decline likely mediated by lactoperoxidase. Compared to adults, infants ingest relatively little nitrite and nitrate, which may be of importance in the modulation of blood flow and the bacterial flora of the infant GI tract, especially given the protective effects of swallowed nitrite. © 2013 American Society for Parenteral and Enteral Nutrition.

  12. Blood plasma response and urinary excretion of nitrite and nitrate in milk-fed calves after oral nitrite and nitrate administration.

    PubMed

    Hüsler, B R.; Blum, J W.

    2001-05-01

    There is marked endogenous production of nitrate in young calves. Here we have studied the contribution of exogenous nitrate and nitrite to plasma concentrations and urinary excretion of nitrite and nitrate in milk-fed calves. In experiment 1, calves were fed 0 or 200 &mgr;mol nitrate or nitrite/kg(0.75) or 100 &mgr;mol nitrite plus 100 &mgr;mol nitrate/kg(0.75) with milk for 3 d. In experiment 2, calves were fed 400 &mgr;mol nitrate or nitrite/kg(0.75) with milk for 1 d. Plasma nitrate rapidly and comparably increased after feeding nitrite, nitrate or nitrite plus nitrate. The rise of plasma nitrate was greater if 400 than 200 &mgr;mol nitrate or nitrite/kg(0.75) were fed. Plasma nitrate decreased slowly after the 3-d administration of 200 &mgr;mol nitrate or nitrite/kg(0.75) and reached pre-experimental concentrations 4 d later. Urinary nitrate excretions nearly identically increased if nitrate, nitrite or nitrite plus nitrate were administered and excreted amounts were greater if 400 than 200 &mgr;mol nitrate or nitrite/kg(0.75) were fed. After nitrite ingestion plasma nitrite only transiently increased after 2 and 4 h and urinary excretion rates remained unchanged. Plasma nitrate concentration remained unchanged if milk was not supplemented with nitrite or nitrate. Nitrate concentrations were stable for 24 h after addition of nitrite to full blood in vitro, whereas nitrite concentrations decreased within 2 h. In conclusion, plasma nitrate concentrations and urinary nitrate excretions are enhanced dose-dependently by feeding low amounts of nitrate and nitrite, whereas after ingested nitrite only a transient and small rise of plasma nitrite is observed because of rapid conversion to nitrate.

  13. Streamwater nitrate concentrations in six agricultural catchments in Scotland.

    PubMed

    Hooda, P S; Moynagh, M; Svoboda, I F; Thurlow, M; Stewart, M; Thomson, M; Anderson, H A

    1997-08-01

    The concentrations of nitrate-N (NO3-N) in catchment inputs and outputs have been compared and contrasted between 6 farm catchments in Scotland, 3 in the West and 3 in the North-East. Forms of intensive animal farming ranging between beef and dairy cattle, sheep and poultry give different sources for potential NO3-N leakage from the systems. While stream reaches bordered by intensive cereal production give rise to the largest inputs to surface waters, climatic influences result in the more-efficient use of fertilizer- and farm waste-N in the West, and an enhanced potential for N-loss to waters in the cooler North-East, regardless of the N-inputs being considerably lower in the latter region. Although the EC Nitrate Directive limit of 11.3 mg NO3-N 1(-1) was not exceeded, peak values occurring during summer baseflows and autumn soil rewetting were commonly larger than the 'target' maximum concentration of 5.65 mg NO3-N 1-1.

  14. A GIS-based groundwater travel time model to evaluate stream nitrate concentration reductions from land use change

    USGS Publications Warehouse

    Schilling, K.E.; Wolter, C.F.

    2007-01-01

    Excessive nitrate-nitrogen (nitrate) loss from agricultural watersheds is an environmental concern. A common conservation practice to improve stream water quality is to retire vulnerable row croplands to grass. In this paper, a groundwater travel time model based on a geographic information system (GIS) analysis of readily available soil and topographic variables was used to evaluate the time needed to observe stream nitrate concentration reductions from conversion of row crop land to native prairie in Walnut Creek watershed, Iowa. Average linear groundwater velocity in 5-m cells was estimated by overlaying GIS layers of soil permeability, land slope (surrogates for hydraulic conductivity and gradient, respectively) and porosity. Cells were summed backwards from the stream network to watershed divide to develop a travel time distribution map. Results suggested that groundwater from half of the land planted in prairie has reached the stream network during the 10 years of ongoing water quality monitoring. The mean travel time for the watershed was estimated to be 10.1 years, consistent with results from a simple analytical model. The proportion of land in the watershed and subbasins with prairie groundwater reaching the stream (10-22%) was similar to the measured reduction of stream nitrate (11-36%). Results provide encouragement that additional nitrate reductions in Walnut Creek are probable in the future as reduced nitrate groundwater from distal locations discharges to the stream network in the coming years. The high spatial resolution of the model (5-m cells) and its simplicity may make it potentially applicable for land managers interested in communicating lag time issues to the public, particularly related to nitrate concentration reductions over time. ?? 2007 Springer-Verlag.

  15. Assessing bottled water nitrate concentrations to evaluate total drinking water nitrate exposure and risk of birth defects.

    PubMed

    Weyer, Peter J; Brender, Jean D; Romitti, Paul A; Kantamneni, Jiji R; Crawford, David; Sharkey, Joseph R; Shinde, Mayura; Horel, Scott A; Vuong, Ann M; Langlois, Peter H

    2014-12-01

    Previous epidemiologic studies of maternal exposure to drinking water nitrate did not account for bottled water consumption. The objective of this National Birth Defects Prevention Study (NBDPS) (USA) analysis was to assess the impact of bottled water use on the relation between maternal exposure to drinking water nitrate and selected birth defects in infants born during 1997-2005. Prenatal residences of 1,410 mothers reporting exclusive bottled water use were geocoded and mapped; 326 bottled water samples were collected and analyzed using Environmental Protection Agency Method 300.0. Median bottled water nitrate concentrations were assigned by community; mothers' overall intake of nitrate in mg/day from drinking water was calculated. Odds ratios for neural tube defects, limb deficiencies, oral cleft defects, and heart defects were estimated using mixed-effects models for logistic regression. Odds ratios (95% CIs) for the highest exposure group in offspring of mothers reporting exclusive use of bottled water were: neural tube defects [1.42 (0.51, 3.99)], limb deficiencies [1.86 (0.51, 6.80)], oral clefts [1.43 (0.61, 3.31)], and heart defects [2.13, (0.87, 5.17)]. Bottled water nitrate had no appreciable impact on risk for birth defects in the NBDPS.

  16. Assessing bottled water nitrate concentrations to evaluate total drinking water nitrate exposure and risk of birth defects

    PubMed Central

    Weyer, Peter J.; Brender, Jean D.; Romitti, Paul A.; Kantamneni, Jiji R.; Crawford, David; Sharkey, Joseph R.; Shinde, Mayura; Horel, Scott A.; Vuong, Ann M.; Langlois, Peter H.

    2016-01-01

    Previous epidemiologic studies of maternal exposure to drinking water nitrate did not account for bottled water consumption. The objective of this National Birth Defects Prevention Study (NBDPS) (USA) analysis was to assess the impact of bottled water use on the relation between maternal exposure to drinking water nitrate and selected birth defects in infants born during 1997–2005. Prenatal residences of 1,410 mothers reporting exclusive bottled water use were geocoded and mapped; 326 bottled water samples were collected and analyzed using Environmental Protection Agency Method 300.0. Median bottled water nitrate concentrations were assigned by community; mothers’ overall intake of nitrate in mg/day from drinking water was calculated. Odds ratios for neural tube defects, limb deficiencies, oral cleft defects, and heart defects were estimated using mixed-effects models for logistic regression. Odds ratios (95% CIs) for the highest exposure group in offspring of mothers reporting exclusive use of bottled water were: neural tube defects [1.42 (0.51, 3.99)], limb deficiencies [1.86 (0.51, 6.80)], oral clefts [1.43 (0.61, 3.31)], and heart defects [2.13, (0.87, 5.17)]. Bottled water nitrate had no appreciable impact on risk for birth defects in the NBDPS. PMID:25473985

  17. Effect of Exogenous and Endogenous Nitrate Concentration on Nitrate Utilization by Dwarf Bean 1

    PubMed Central

    Breteler, Hans; Nissen, Per

    1982-01-01

    The effect of the exogenous and endogenous NO3− concentration on net uptake, influx, and efflux of NO3− and on nitrate reductase activity (NRA) in roots was studied in Phaseolus vulgaris L. cv. Witte Krombek. After exposure to NO3−, an apparent induction period of about 6 hours occurred regardless of the exogenous NO3− level. A double reciprocal plot of the net uptake rate of induced plants versus exogenous NO3− concentration yielded four distinct phases, each with simple Michaelis-Menten kinetics, and separated by sharp breaks at about 45, 80, and 480 micromoles per cubic decimeter. Influx was estimated as the accumulation of 15N after 1 hour exposure to 15NO3−. The isotherms for influx and net uptake were similar and corresponded to those for alkali cations and Cl−. Efflux of NO3− was a constant proportion of net uptake during initial NO3− supply and increased with exogenous NO3− concentration. No efflux occurred to a NO3−-free medium. The net uptake rate was negatively correlated with the NO3− content of roots. Nitrate efflux, but not influx, was influenced by endogenous NO3−. Variations between experiments, e.g. in NO3− status, affected the values of Km and Vmax in the various concentration phases. The concentrations at which phase transitions occurred, however, were constant both for influx and net uptake. The findings corroborate the contention that separate sites are responsible for uptake and transitions between phases. Beyond 100 micromoles per cubic decimeter, root NRA was not affected by exogenous NO3− indicating that NO3− uptake was not coupled to root NRA, at least not at high concentrations. PMID:16662570

  18. Mortality of nitrate fertiliser workers.

    PubMed Central

    Al-Dabbagh, S; Forman, D; Bryson, D; Stratton, I; Doll, R

    1986-01-01

    An epidemiological cohort study was conducted to investigate the mortality patterns among a group of workers engaged in the production of nitrate based fertilisers. This study was designed to test the hypothesis that individuals exposed to high concentrations of nitrates might be at increased risk of developing cancers, particularly gastric cancer. A total of 1327 male workers who had been employed in the production of fertilisers between 1946 and 1981 and who had been occupationally exposed to nitrates for at least one year were followed up until 1 March 1981. In total, 304 deaths were observed in this group and these were compared with expected numbers calculated from mortality rates in the northern region of England, where the factory was located. Analysis was also carried out separately for a subgroup of the cohort who had been heavily exposed to nitrates--that is, working in an environment likely to contain more than 10 mg nitrate/m3 for a year or longer. In neither the entire cohort nor the subgroup was any significant excess observed for all causes of mortality or for mortality from any of five broad categories of cause or from four specific types of cancer. A small excess of lung cancer was noted more than 20 years after first exposure in men heavily exposed for more than 10 years. That men were exposed to high concentrations of nitrate was confirmed by comparing concentrations of nitrates in the saliva of a sample of currently employed men with control men, employed at the same factory but not in fertiliser production. The men exposed to nitrate had substantially raised concentrations of nitrate in their saliva compared with both controls within the industry and with men in the general population and resident nearby. The results of this study therefore weight against the idea that exposure to nitrates in the environment leads to the formation in vivo of material amounts of carcinogens. PMID:3015194

  19. Long-term trends in dissolved iron and DOC concentration linked to nitrate depletion in riparian soils

    NASA Astrophysics Data System (ADS)

    Musolff, Andreas; Selle, Benny; Fleckenstein, Jan H.; Oosterwoud, Marieke R.; Tittel, Jörg

    2016-04-01

    The instream concentrations of dissolved organic carbon (DOC) are rising in many catchments of the northern hemisphere. Elevated concentrations of DOC, mainly in the form of colored humic components, increase efforts and costs of drinking water purification. In this study, we evaluated a long-term dataset of 110 catchments draining into German drinking water reservoirs in order to assess sources of DOC and drivers of a potential long-term change. The average DOC concentrations across the wide range of different catchments were found to be well explained by the catchment's topographic wetness index. Higher wetness indices were connected to higher average DOC concentrations, which implies that catchments with shallow topography and pronounced riparian wetlands mobilize more DOC. Overall, 37% of the investigated catchments showed a significant long-term increase in DOC concentrations, while 22% exhibited significant negative trends. Moreover, we found that increasing trends in DOC were positively correlated to trends in dissolved iron concentrations at pH≤6 due to remobilization of DOC previously sorbed to iron minerals. Both, increasing trends in DOC and dissolve iron were found to be connected to decreasing trends and low concentrations of nitrate (below ~6 mg/L). This was especially observed in forested catchments where atmospheric N-depositions were the major source for nitrate availability. In these catchments, we also found long-term increases of phosphate concentrations. Therefore, we argue that dissolved iron, DOC and phosphate were jointly released under iron-reducing conditions when nitrate as a competing electron acceptor was too low in concentrations to prevent the microbial iron reduction. In contrast, we could not explain the observed increasing trends in DOC, iron and phosphate concentrations by the long-term trends of pH, sulfate or precipitation. Altogether this study gives strong evidence that both, source and long-term increases in DOC are

  20. Numerical model simulations of nitrate concentrations in groundwater using various nitrogen input scenarios, mid-Snake region, south-central Idaho

    USGS Publications Warehouse

    Skinner, Kenneth D.; Rupert, Michael G.

    2012-01-01

    As part of the U.S. Geological Survey’s National Water Quality Assessment (NAWQA) program nitrate transport in groundwater was modeled in the mid-Snake River region in south-central Idaho to project future concentrations of nitrate. Model simulation results indicated that nitrate concentrations would continue to increase over time, eventually exceeding the U.S. Environmental Protection Agency maximum contaminant level for drinking water of 10 milligrams per liter in some areas. A subregional groundwater model simulated the change of nitrate concentrations in groundwater over time in response to three nitrogen input scenarios: (1) nitrogen input fixed at 2008 levels; (2) nitrogen input increased from 2008 to 2028 using the same rate of increase as the average rate of increase during the previous 10 years (1998 through 2008); after 2028, nitrogen input is fixed at 2028 levels; and (3) nitrogen input related to agriculture completely halted, with only nitrogen input from precipitation remaining. Scenarios 1 and 2 project that nitrate concentrations in groundwater continue to increase from 10 to 50 years beyond the year nitrogen input is fixed, depending on the location in the model area. Projected nitrate concentrations in groundwater increase by as much as 2–4 milligrams per liter in many areas, with nitrate concentrations in some areas reaching 10 milligrams per liter. Scenario 3, although unrealistic, estimates how long (20–50 years) it would take nitrate in groundwater to return to background concentrations—the “flushing time” of the system. The amount of nitrate concentration increase cannot be explained solely by differences in nitrogen input; in fact, some areas with the highest amount of nitrogen input have the lowest increase in nitrate concentration. The geometry of the aquifer and the pattern of regional groundwater flow through the aquifer greatly influence nitrate concentrations. The aquifer thins toward discharge areas along the Snake River

  1. Survey of Nitrate Ion Concentrations in Vegetables Cultivated in Plant Factories: Comparison with Open-Culture Vegetables.

    PubMed

    Oka, Yuka; Hirayama, Izumi; Yoshikawa, Mitsuhide; Yokoyama, Tomoko; Iida, Kenji; Iwakoshi, Katsushi; Suzuki, Ayana; Yanagihara, Midori; Segawa, Yukino; Kukimoto, Sonomi; Hamada, Humika; Matsuzawa, Satomi; Tabata, Setsuko; Sasamoto, Takeo

    2017-01-01

    A survey of nitrate-ion concentrations in plant-factory-cultured leafy vegetables was conducted. 344 samples of twenty-one varieties of raw leafy vegetables were examined using HPLC. The nitrate-ion concentrations in plant-factory-cultured leafy vegetables were found to be LOD-6,800 mg/kg. Furthermore, the average concentration values varied among different leafy vegetables. The average values for plant-factory-cultured leafy vegetables were higher than those of open-cultured leafy vegetables reported in previous studies, such as the values listed in the Standard Tables of Food Composition in Japan- 2015 - (Seventh revised edition). For some plant-factory-cultured leafy vegetables, such as salad spinach, the average values were above the maximum permissible levels of nitrate concentration in EC No 1258/2011; however, even when these plant-factory-cultured vegetables were routinely eaten, the intake of nitrate ions in humans did not exceed the ADI.

  2. Impact of Vitamin B12 and Nitrate Availability on the Concentration of Particulate Dimethylsulfoniopropionate in Phytoplankton

    NASA Astrophysics Data System (ADS)

    Zavala, J.; Lee, P. A.; Schanke, N. L.; Pound, H.; Penta, W. B.; Shore, S. K.

    2016-02-01

    The production of particulate dimethylsulfoniopropionate (DMSPp) was examined in natural phytoplankton communities from the South Atlantic Bight near Savannah, Georgia, during an expedition in June 2015. Vitamin B12 and nitrate were added to seawater samples from a coastal and an oceanic site, both of which contained low-biomass, cyanobacteria-dominated communities. Under nitrate-limited conditions, irrespective of changes in B12 levels, DMSPp concentrations increased. DMSPp concentrations of these mixed phytoplankton communities did not appear to be limited by the availability of B12. In a laboratory experiment, DMSPp concentrations in the diatom Phaeodactylum tricornutum were measured after the removal of vitamin B12 and nitrate from a synthetic seawater culture media. DMSPp concentrations increased under nitrate-limited conditions, irrespective of changes in B12 levels, and are argued to be the result of increased biosynthesis. DMSPp concentrations in P. tricornutum were unaffected by B12 limitation. It is hypothesized that P. tricornutum is using the B12-independent methionine synthase MetE to synthesize DMSPp rather than the B12-dependent methionine synthase MetH.

  3. Annual nitrate drawdown observed by SOCCOM profiling floats and the relationship to annual net community production

    NASA Astrophysics Data System (ADS)

    Johnson, Kenneth S.; Plant, Joshua N.; Dunne, John P.; Talley, Lynne D.; Sarmiento, Jorge L.

    2017-08-01

    Annual nitrate cycles have been measured throughout the pelagic waters of the Southern Ocean, including regions with seasonal ice cover and southern hemisphere subtropical zones. Vertically resolved nitrate measurements were made using in situ ultraviolet spectrophotometer (ISUS) and submersible ultraviolet nitrate analyzer (SUNA) optical nitrate sensors deployed on profiling floats. Thirty-one floats returned 40 complete annual cycles. The mean nitrate profile from the month with the highest winter nitrate minus the mean profile from the month with the lowest nitrate yields the annual nitrate drawdown. This quantity was integrated to 200 m depth and converted to carbon using the Redfield ratio to estimate annual net community production (ANCP) throughout the Southern Ocean south of 30°S. A well-defined, zonal mean distribution is found with highest values (3-4 mol C m-2 yr-1) from 40 to 50°S. Lowest values are found in the subtropics and in the seasonal ice zone. The area weighted mean was 2.9 mol C m-2 yr-1 for all regions south of 40°S. Cumulative ANCP south of 50°S is 1.3 Pg C yr-1. This represents about 13% of global ANCP in about 14% of the global ocean area.Plain Language SummaryThis manuscript reports on 40 annual cycles of <span class="hlt">nitrate</span> <span class="hlt">observed</span> by chemical sensors on SOCCOM profiling floats. The annual drawdown in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> by phytoplankton is used to assess the spatial variability of annual net community production in the Southern Ocean. This ANCP is a key component of the global carbon cycle and it exerts an important control on atmospheric carbon dioxide. We show that the results are consistent with our prior understanding of Southern Ocean ANCP, which has required decades of <span class="hlt">observations</span> to accumulate. The profiling floats now enable annual resolution of this key process. The results also highlight spatial variability in ANCP in the Southern Ocean.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.water.usgs.gov/wri034152/','USGSPUBS'); return false;" href="http://pubs.water.usgs.gov/wri034152/"><span>Trends in <span class="hlt">nitrate</span> and dissolved-solids <span class="hlt">concentrations</span> in ground water, Carson Valley, Douglas County, Nevada, 1985-2001</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Rosen, Michael R.</p> <p>2003-01-01</p> <p>Analysis of trends in <span class="hlt">nitrate</span> and total dissolved-solids <span class="hlt">concentrations</span> over time in Carson Valley, Nevada, indicates that 56 percent of 27 monitoring wells that have long-term records of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> show increasing trends, 11 percent show decreasing trends, and 33 percent have not changed. Total dissolved-solids <span class="hlt">concentrations</span> have increased in 52 percent of these wells and are stable in 48 percent. None of these wells show decreasing trends in total dissolved-solids <span class="hlt">concentrations</span>. The wells showing increasing trends in <span class="hlt">nitrate</span> and total dissolved-solids <span class="hlt">concentrations</span> were always in areas that use septic waste-disposal systems. Therefore, the primary cause of these increases is likely the increase in septic-tank usage over the past 40 years.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFMOS31A1232J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFMOS31A1232J"><span>Long-Term <span class="hlt">Observations</span> of Ocean Biogeochemistry with <span class="hlt">Nitrate</span> and Oxygen Sensors in Apex Profiling Floats</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Johnson, K. S.; Coletti, L.; Jannasch, H.; Martz, T.; Swift, D.; Riser, S.</p> <p>2008-12-01</p> <p>Long-term, autonomous <span class="hlt">observations</span> of ocean biogeochemical cycles are now feasible with chemical sensors in profiling floats. These sensors will enable decadal-scale <span class="hlt">observations</span> of trends in global ocean biogeochemical cycles. Here, we focus on measurements on <span class="hlt">nitrate</span> and dissolved oxygen. The ISUS (In Situ Ultraviolet Spectrophotometer) optical <span class="hlt">nitrate</span> sensor has been adapted to operate in a Webb Research, Apex profiling float. The Apex float is of the type used in the Argo array and is designed for multi-year, expendable deployments in the ocean. Floats park at 1000 m depth and make 60 <span class="hlt">nitrate</span> and oxygen measurements at depth intervals ranging from 50 m below 400 m to 5 m in the upper 100 m as they profile to the surface. All data are transmitted to shore using the Iridium telemetry system and they are available on the Internet in near-real time. Floats equipped with ISUS and an Aanderaa oxygen sensor are capable of making 280 vertical profiles from 1000 m. At a 5 day cycle time, the floats should have nearly a four year endurance. Three floats have now been deployed at the Hawaii Ocean Time series station (HOT), Ocean Station Papa (OSP) in the Gulf of Alaska and at 50 South, 30 East in the Southern Ocean. Two additional floats are designated for deployment at the Bermuda Atlantic Time Series station (BATS) and in the Drake Passage. The HOT float has made 56 profiles over 260 days and should continue operating for 3 more years. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> are in excellent agreement with the long-term mean <span class="hlt">observed</span> at HOT. No significant long-term drift in sensor response has occurred. A variety of features have been <span class="hlt">observed</span> in the HOT <span class="hlt">nitrate</span> data that are linked to contemporaneous changes in oxygen production and mesoscale dynamics. The impacts of these features will be briefly described. The Southern Ocean float has operated for 200 days and is now <span class="hlt">observing</span> reinjection of <span class="hlt">nitrate</span> into surface waters as winter mixing occurs(surface <span class="hlt">nitrate</span> > 24 micromolar). We</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28409413','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28409413"><span>Impact of ammonium <span class="hlt">nitrate</span> and sodium <span class="hlt">nitrate</span> on tadpoles of Alytes obstetricans.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Garriga, Núria; Montori, A; Llorente, G A</p> <p>2017-07-01</p> <p>The presence of pesticides, herbicides and fertilisers negatively affect aquatic communities in general, and particularly amphibians in their larval phase, even though sensitivity to pollutants is highly variable among species. The Llobregat Delta (Barcelona, Spain) has experienced a decline of amphibian populations, possibly related to the reduction in water quality due to the high levels of farming activity, but also to habitat loss and alteration. We studied the effects of increasing ammonium <span class="hlt">nitrate</span> and sodium <span class="hlt">nitrate</span> levels on the survival and growth rate of Alytes obstetricans tadpoles under experimental conditions. We exposed larvae to increasing <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and ammonium for 14 days and then exposed them to water without pollutants for a further 14 days. Only the higher <span class="hlt">concentrations</span> of ammonium (>33.75 mg/L) caused larval mortality. The growth rate of larvae was reduced at ≥22.5 mg/L NH 4 + , although individuals recovered and even increased their growth rate once exposure to the pollutant ended. The effect of <span class="hlt">nitrate</span> on growth rate was detected at ≥80 mg/L <span class="hlt">concentrations</span>, and the growth rate reduction in tadpoles was even <span class="hlt">observed</span> during the post-exposure phase. The <span class="hlt">concentrations</span> of ammonium with adverse effects on larvae are within the range levels found in the study area, while the <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> with some adverse effect are close to the upper range limit of current <span class="hlt">concentrations</span> in the study area. Therefore, only the presence of ammonium in the study area is likely to be considered of concern for the population of this species, even though the presence of <span class="hlt">nitrate</span> could cause some sublethal effects. These negative effects could have an impact on population dynamics, which in this species is highly sensitive to larval mortality due to its small clutch size and prolonged larval period compared to other anuran amphibians.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.H53K1555E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.H53K1555E"><span>Trends in Surface-Water <span class="hlt">Nitrate</span>-N <span class="hlt">Concentrations</span> and Loads from Predominantly-Forested Watersheds of the Chesapeake Bay Basin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eshleman, K. N.</p> <p>2011-12-01</p> <p> absolute changes in <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span> corresponded to the least forested watersheds, the largest percentage changes in <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span> were actually <span class="hlt">observed</span> for those watersheds with the greatest percentages of forestland. This result suggests that the natural dynamics of forests may be playing a very important (and under-appreciated) role in improving water quality throughout the Bay watershed. A second interesting finding was that the statistically significant reductions in annual <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span> at the Potomac River RIM station could be entirely explained by commensurate improvements at the upstream (Hancock) station; in fact, no trend in <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span> associated with the eastern portion of the basin was found (after subtracting out the influence of the upstream portion). Additional research is needed to understand why nitrogen retention by forested lands may be increasing and thus helping restore water quality throughout the Chesapeake Bay watershed. The results also have obvious implications for meeting local water quality goals as well as the basin-wide goal of the Chesapeake Bay TMDL for nitrogen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70019018','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70019018"><span>Effects of agricultural practices and vadose zone stratigraphy on <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in ground water in Kansas, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Townsend, M.A.; Sleezer, R.O.; Macko, S.A.; ,</p> <p>1996-01-01</p> <p>Differences in <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> in,around water in Kansas can be explained by variations in agricultural practices and vadose-zone stratigraphy. In northwestern Kansas, past use of a local stream for tailwater runoff from irrigation and high fertilizer applications for sugar-beet farming resulted in high <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> (12-60 mg L-1; in both soil and ground water. Nitrogen isotope values from the soil and ground water range from +4 to +8? which is typical for a fertilizer source. In parts of south-central Kansas, the use of crop rotation and the presence of both continuous fine-textured layers and a reducing ground-water chemistry resulted in ground-water <span class="hlt">nitrate</span>-N values of 10 mg L-1; in both soil and grounwater. Nitrogen isotope values of +3 to +7? indicate a fertilizer source. Crop rotation decreased <span class="hlt">nitrate</span>-N values in the shallow ground water (9 m). However, deeper ground water showed increasing <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> as a result of past farming practices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA157688','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA157688"><span>Ipso <span class="hlt">Nitration</span>. Regiospecific <span class="hlt">Nitration</span> via Ipso <span class="hlt">Nitration</span> Products.</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1985-05-20</p> <p>products of <span class="hlt">nitration</span> of alkylbenzenes and alkylphenol derivatives. The general pattern envisioned is shown in Scheme 1. In order to realize this...we have also explored solid state <span class="hlt">nitration</span> of various alkylphenols . This procedure involves adsorbing <span class="hlt">concentrated</span> nitric acid on alumina, followed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23974533','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23974533"><span>Modeling of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater using artificial intelligence approach--a case study of Gaza coastal aquifer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Alagha, Jawad S; Said, Md Azlin Md; Mogheir, Yunes</p> <p>2014-01-01</p> <p><span class="hlt">Nitrate</span> <span class="hlt">concentration</span> in groundwater is influenced by complex and interrelated variables, leading to great difficulty during the modeling process. The objectives of this study are (1) to evaluate the performance of two artificial intelligence (AI) techniques, namely artificial neural networks and support vector machine, in modeling groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span> using scant input data, as well as (2) to assess the effect of data clustering as a pre-modeling technique on the developed models' performance. The AI models were developed using data from 22 municipal wells of the Gaza coastal aquifer in Palestine from 2000 to 2010. Results indicated high simulation performance, with the correlation coefficient and the mean average percentage error of the best model reaching 0.996 and 7 %, respectively. The variables that strongly influenced groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span> were previous <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, groundwater recharge, and on-ground nitrogen load of each land use land cover category in the well's vicinity. The results also demonstrated the merit of performing clustering of input data prior to the application of AI models. With their high performance and simplicity, the developed AI models can be effectively utilized to assess the effects of future management scenarios on groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, leading to more reasonable groundwater resources management and decision-making.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70187880','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70187880"><span>A hybrid machine learning model to predict and visualize <span class="hlt">nitrate</span> <span class="hlt">concentration</span> throughout the Central Valley aquifer, California, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ransom, Katherine M.; Nolan, Bernard T.; Traum, Jonathan A.; Faunt, Claudia; Bell, Andrew M.; Gronberg, Jo Ann M.; Wheeler, David C.; Zamora, Celia; Jurgens, Bryant; Schwarz, Gregory E.; Belitz, Kenneth; Eberts, Sandra; Kourakos, George; Harter, Thomas</p> <p>2017-01-01</p> <p>Intense demand for water in the Central Valley of California and related increases in groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span> threaten the sustainability of the groundwater resource. To assess contamination risk in the region, we developed a hybrid, non-linear, machine learning model within a statistical learning framework to predict <span class="hlt">nitrate</span> contamination of groundwater to depths of approximately 500 m below ground surface. A database of 145 predictor variables representing well characteristics, historical and current field and landscape-scale nitrogen mass balances, historical and current land use, oxidation/reduction conditions, groundwater flow, climate, soil characteristics, depth to groundwater, and groundwater age were assigned to over 6000 private supply and public supply wells measured previously for <span class="hlt">nitrate</span> and located throughout the study area. The boosted regression tree (BRT) method was used to screen and rank variables to predict <span class="hlt">nitrate</span> <span class="hlt">concentration</span> at the depths of domestic and public well supplies. The novel approach included as predictor variables outputs from existing physically based models of the Central Valley. The top five most important predictor variables included two oxidation/reduction variables (probability of manganese <span class="hlt">concentration</span> to exceed 50 ppb and probability of dissolved oxygen <span class="hlt">concentration</span> to be below 0.5 ppm), field-scale adjusted unsaturated zone nitrogen input for the 1975 time period, average difference between precipitation and evapotranspiration during the years 1971–2000, and 1992 total landscape nitrogen input. Twenty-five variables were selected for the final model for log-transformed <span class="hlt">nitrate</span>. In general, increasing probability of anoxic conditions and increasing precipitation relative to potential evapotranspiration had a corresponding decrease in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> predictions. Conversely, increasing 1975 unsaturated zone nitrogen leaching flux and 1992 total landscape nitrogen input had an increasing relative</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1999/4106/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1999/4106/"><span><span class="hlt">Concentrations</span> and possible sources of <span class="hlt">nitrate</span> in water from the Silurian-Devonian aquifer, Cedar Falls, Iowa</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Schaap, Bryan D.</p> <p>1999-01-01</p> <p>Nitrogen fertilizer sales in Iowa have been higher in recent years than during the mid- 1970’s. This suggests that <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in water from well 9 may persist at present levels or could increase in future years if fertilizer use increases and if higher <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are directly related to higher nitrogen fertilizer use.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2010/5100/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2010/5100/"><span>Relations that affect the probability and prediction of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in private wells in the glacial aquifer system in the United States</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Warner, Kelly L.; Arnold, Terri L.</p> <p>2010-01-01</p> <p><span class="hlt">Nitrate</span> in private wells in the glacial aquifer system is a concern for an estimated 17 million people using private wells because of the proximity of many private wells to nitrogen sources. Yet, less than 5 percent of private wells sampled in this study contained <span class="hlt">nitrate</span> in <span class="hlt">concentrations</span> that exceeded the U.S. Environmental Protection Agency (USEPA) Maximum Contaminant Level (MCL) of 10 mg/L (milligrams per liter) as N (nitrogen). However, this small group with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above the USEPA MCL includes some of the highest <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> detected in groundwater from private wells (77 mg/L). Median <span class="hlt">nitrate</span> <span class="hlt">concentration</span> measured in groundwater from private wells in the glacial aquifer system (0.11 mg/L as N) is lower than that in water from other unconsolidated aquifers and is not strongly related to surface sources of <span class="hlt">nitrate</span>. Background <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> is less than 1 mg/L as N. Although overall <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in private wells was low relative to the MCL, <span class="hlt">concentrations</span> were highly variable over short distances and at various depths below land surface. Groundwater from wells in the glacial aquifer system at all depths was a mixture of old and young water. Oxidation and reduction potential changes with depth and groundwater age were important influences on <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in private wells. A series of 10 logistic regression models was developed to estimate the probability of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> above various thresholds. The threshold <span class="hlt">concentration</span> (1 to 10 mg/L) affected the number of variables in the model. Fewer explanatory variables are needed to predict <span class="hlt">nitrate</span> at higher threshold <span class="hlt">concentrations</span>. The variables that were identified as significant predictors for <span class="hlt">nitrate</span> <span class="hlt">concentration</span> above 4 mg/L as N included well characteristics such as open-interval diameter, open-interval length, and depth to top of open interval. Environmental variables in the models were mean percent silt in soil, soil type, and mean depth to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008JHyd..357...42S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008JHyd..357...42S"><span>Building factorial regression models to explain and predict <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater under agricultural land</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stigter, T. Y.; Ribeiro, L.; Dill, A. M. M. Carvalho</p> <p>2008-07-01</p> <p>SummaryFactorial regression models, based on correspondence analysis, are built to explain the high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater beneath an agricultural area in the south of Portugal, exceeding 300 mg/l, as a function of chemical variables, electrical conductivity (EC), land use and hydrogeological setting. Two important advantages of the proposed methodology are that qualitative parameters can be involved in the regression analysis and that multicollinearity is avoided. Regression is performed on eigenvectors extracted from the data similarity matrix, the first of which clearly reveals the impact of agricultural practices and hydrogeological setting on the groundwater chemistry of the study area. Significant correlation exists between response variable NO3- and explanatory variables Ca 2+, Cl -, SO42-, depth to water, aquifer media and land use. Substituting Cl - by the EC results in the most accurate regression model for <span class="hlt">nitrate</span>, when disregarding the four largest outliers (model A). When built solely on land use and hydrogeological setting, the regression model (model B) is less accurate but more interesting from a practical viewpoint, as it is based on easily obtainable data and can be used to predict <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater in other areas with similar conditions. This is particularly useful for conservative contaminants, where risk and vulnerability assessment methods, based on assumed rather than established correlations, generally produce erroneous results. Another purpose of the models can be to predict the future evolution of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> under influence of changes in land use or fertilization practices, which occur in compliance with policies such as the <span class="hlt">Nitrates</span> Directive. Model B predicts a 40% decrease in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater of the study area, when horticulture is replaced by other land use with much lower fertilization and irrigation rates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19945739','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19945739"><span>Spatial and temporal trends in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the River Derwent, North Yorkshire, and its need for NVZ status.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mian, Ishaq A; Begum, Shaheen; Riaz, Muhammad; Ridealgh, Mike; McClean, Colin J; Cresser, Malcolm S</p> <p>2010-01-15</p> <p>Long-term spatial and temporal variations in <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> along the River Derwent have been examined using Environment Agency data to investigate the relative importance of impacts of atmospheric N deposition, land use, and changes in management. Where moorland and rough grazing dominate upstream of Forge Valley and Malton, over the 20 years since 1988 mean <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> were initially increasing significantly, but are now levelling off, with peaks at ca. 4.5 mg Nl(-1). As expected in a catchment in a <span class="hlt">nitrate</span> vulnerable zone (NVZ), more agricultural land use increases mean <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and the occurrence of distinct winter maxima, though the latter have become markedly less pronounced since 2001. It is suggested that this improvement is a combined effect of imposition of NVZ designation in the lower reaches in 2002, animal number declines associated with the Foot & Mouth outbreak in the region in 2001, and the impact of farmers' responses to increasing fertilizer prices and to beneficial pollutant mineral N inputs from the atmosphere. Minima in <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> in summer have become much less pronounced over the past decade and are typically ca. 60% higher in <span class="hlt">concentration</span> than a decade earlier. This probably is attributable to the effects of pollutant-N leaching to depths in soil below the rooting zone when near surface biotic uptake is low in winter. The resultant N mineralization in summer enhances summer <span class="hlt">nitrate</span> leaching. The Derwent is a relatively clean river; however, its entire catchment was designated justifiably as a NVZ in January 2009, apparently based upon a projected 95 percentile <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span> >11.29 mg l(-1) for 2010 based upon forward projection of data from 1990 to 2004 for Derwent Bridge. A survey of water quality in March 2009 showed that some agricultural areas are still making a significant contribution to the total <span class="hlt">nitrate</span> level well downstream, at the point responsible for implementation of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28282914','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28282914"><span><span class="hlt">Nitrate</span>, Nitrite, and Ammonium Variability in Drinking Water Distribution Systems.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schullehner, Jörg; Stayner, Leslie; Hansen, Birgitte</p> <p>2017-03-09</p> <p>Accurate assessments of exposure to <span class="hlt">nitrate</span> in drinking water is a crucial part of epidemiological studies investigating long-term adverse human health effects. However, since drinking water <span class="hlt">nitrate</span> measurements are usually collected for regulatory purposes, assumptions on (1) the intra-distribution system variability and (2) short-term (seasonal) <span class="hlt">concentration</span> variability have to be made. We assess <span class="hlt">concentration</span> variability in the distribution system of <span class="hlt">nitrate</span>, nitrite, and ammonium, and seasonal variability in all Danish public waterworks from 2007 to 2016. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> at the exit of the waterworks are highly correlated with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> within the distribution net or at the consumers' taps, while nitrite and ammonium <span class="hlt">concentrations</span> are generally lower within the net compared with the exit of the waterworks due to nitrification. However, nitrification of nitrite and ammonium in the distribution systems only results in a relatively small increase in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. No seasonal variation for <span class="hlt">nitrate</span>, nitrite, or ammonium was <span class="hlt">observed</span>. We conclude that <span class="hlt">nitrate</span> measurements taken at the exit of the waterworks are suitable to calculate exposures for all consumers connected to that waterworks and that sampling frequencies in the national monitoring programme are sufficient to describe temporal variations in longitudinal studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5369112','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5369112"><span><span class="hlt">Nitrate</span>, Nitrite, and Ammonium Variability in Drinking Water Distribution Systems</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Schullehner, Jörg; Stayner, Leslie; Hansen, Birgitte</p> <p>2017-01-01</p> <p>Accurate assessments of exposure to <span class="hlt">nitrate</span> in drinking water is a crucial part of epidemiological studies investigating long-term adverse human health effects. However, since drinking water <span class="hlt">nitrate</span> measurements are usually collected for regulatory purposes, assumptions on (1) the intra-distribution system variability and (2) short-term (seasonal) <span class="hlt">concentration</span> variability have to be made. We assess <span class="hlt">concentration</span> variability in the distribution system of <span class="hlt">nitrate</span>, nitrite, and ammonium, and seasonal variability in all Danish public waterworks from 2007 to 2016. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> at the exit of the waterworks are highly correlated with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> within the distribution net or at the consumers’ taps, while nitrite and ammonium <span class="hlt">concentrations</span> are generally lower within the net compared with the exit of the waterworks due to nitrification. However, nitrification of nitrite and ammonium in the distribution systems only results in a relatively small increase in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. No seasonal variation for <span class="hlt">nitrate</span>, nitrite, or ammonium was <span class="hlt">observed</span>. We conclude that <span class="hlt">nitrate</span> measurements taken at the exit of the waterworks are suitable to calculate exposures for all consumers connected to that waterworks and that sampling frequencies in the national monitoring programme are sufficient to describe temporal variations in longitudinal studies. PMID:28282914</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.H13C0937B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.H13C0937B"><span>Stream <span class="hlt">Nitrate</span> <span class="hlt">Concentrations</span> in a Small Catchment in South West England over a Period of 35 Years (1970-2005)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burt, T.; Worrall, F.</p> <p>2008-12-01</p> <p>A 35-year record of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> for the Slapton Wood stream, a small agricultural catchment in south west England, is presented. The study reconsiders earlier work in order to assess whether upward trends have been maintained and how controls on catchment <span class="hlt">nitrate</span> processes have altered. The study has shown that: (i) the catchment has reached a new position of equilibrium and increases in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> have levelled off; (ii) the occurrence of severe droughts means that records of less than a decade are misleading and only longer records can illustrate changes of system state; (iii) the change of state <span class="hlt">observed</span> in the catchment is illustrated in the switching of long-term memory effects from a negative to a positive annual memory; (iv) a significant long-term impulsivity relationship with rainfall becomes insignificant over the course of the study period. The study shows the importance of long records in exposing changes in state in catchment systems and understanding the time constants of a range of driving processes. The study by its very nature also demonstrates the importance of maintaining long-term monitoring programmes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2557725','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2557725"><span>Adaptation of cytochrome-b5 reductase activity and methaemoglobinaemia in areas with a high <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in drinking-water.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Gupta, S. K.; Gupta, R. C.; Seth, A. K.; Gupta, A. B.; Bassin, J. K.; Gupta, A.</p> <p>1999-01-01</p> <p>An epidemiological investigation was undertaken in India to assess the prevalence of methaemoglobinaemia in areas with high <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in drinking-water and the possible association with an adaptation of cytochrome-b5 reductase. Five areas were selected, with average <span class="hlt">nitrate</span> ion <span class="hlt">concentrations</span> in drinking-water of 26, 45, 95, 222 and 459 mg/l. These areas were visited and house schedules were prepared in accordance with a statistically designed protocol. A sample of 10% of the total population was selected in each of the areas, matched for age and weight, giving a total of 178 persons in five age groups. For each subject, a detailed history was documented, a medical examination was conducted and blood samples were taken to determine methaemoglobin level and cytochrome-b5 reductase activity. Collected data were subjected to statistical analysis to test for a possible relationship between <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, cytochrome-b5 reductase activity and methaemoglobinaemia. High <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> caused methaemoglobinaemia in infants and adults. The reserve of cytochrome-b5 reductase activity (i.e. the enzyme activity not currently being used, but which is available when needed; for example, under conditions of increased <span class="hlt">nitrate</span> ingestion) and its adaptation with increasing water <span class="hlt">nitrate</span> <span class="hlt">concentration</span> to reduce methaemoglobin were more pronounced in children and adolescents. PMID:10534899</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22606534-comparative-study-glycine-single-crystals-additive-potassium-nitrate-different-concentration-ratios','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22606534-comparative-study-glycine-single-crystals-additive-potassium-nitrate-different-concentration-ratios"><span>Comparative study of glycine single crystals with additive of potassium <span class="hlt">nitrate</span> in different <span class="hlt">concentration</span> ratios</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>Gujarati, Vivek P., E-mail: vivekgujarati@gmail.com; Deshpande, M. P., E-mail: vishwadeshpande@yahoo.co.in; Patel, Kamakshi R.</p> <p>2016-05-06</p> <p>Semi-organic crystals of Glycine Potassium <span class="hlt">Nitrate</span> (GPN) with potential applications in Non linear optics (NLO) were grown using slow evaporation technique. Glycine and Potassium <span class="hlt">Nitrate</span> were taken in three different <span class="hlt">concentration</span> ratios of 3:1, 2:1 and 1:1 respectively. We checked the solubility of the material in distilled water at different temperatures and could <span class="hlt">observe</span> the growth of crystals in 7 weeks time. Purity of the grown crystals was confirmed by Energy Dispersive X-ray Analysis (EDAX) and CHN analysis. GSN Powder X-ray diffraction pattern was recorded to confirm the crystalline nature. To confirm the applications of grown crystals in opto-electronics field,more » UV-Vis-NIR study was carried out. Dielectric properties of the samples were studied in between the frequency range 1Hz to 100 KHz.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_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('https://pubs.er.usgs.gov/publication/70019427','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70019427"><span>Predicting the probability of elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the Puget Sound Basin: Implications for aquifer susceptibility and vulnerability</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Tesoriero, A.J.; Voss, F.D.</p> <p>1997-01-01</p> <p>The occurrence and distribution of elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> (≥ 3 mg/l) in ground water in the Puget Sound Basin, Washington, were determined by examining existing data from more than 3000 wells. Models that estimate the probability that a well has an elevated <span class="hlt">nitrate</span> <span class="hlt">concentration</span> were constructed by relating the occurrence of elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> to both natural and anthropogenic variables using logistic regression. The variables that best explain the occurrence of elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were well depth, surficial geology, and the percentage of urban and agricultural land within a radius of 3.2 kilometers of the well. From these relations, logistic regression models were developed to assess aquifer susceptibility (relative ease with which contaminants will reach aquifer) and ground-water vulnerability (relative ease with which contaminants will reach aquifer for a given set of land-use practices). Both models performed well at predicting the probability of elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in an independent data set. This approach to assessing aquifer susceptibility and ground-water vulnerability has the advantages of having both model variables and coefficient values determined on the basis of existing water quality information and does not depend on the assignment of variables and weighting factors based on qualitative criteria.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sim/3234/pdf/sim3234booklet.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sim/3234/pdf/sim3234booklet.pdf"><span>Maps of estimated <span class="hlt">nitrate</span> and arsenic <span class="hlt">concentrations</span> in basin-fill aquifers of the southwestern United States</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Beisner, Kimberly R.; Anning, David W.; Paul, Angela P.; McKinney, Tim S.; Huntington, Jena M.; Bexfield, Laura M.; Thiros, Susan A.</p> <p>2012-01-01</p> <p>Human-health concerns and economic considerations associated with meeting drinking-water standards motivated a study of the vulnerability of basin-fill aquifers to <span class="hlt">nitrate</span> contamination and arsenic enrichment in the southwestern United States. Statistical models were developed by using the random forest classifier algorithm to predict <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and arsenic across a model grid representing about 190,600 square miles of basin-fill aquifers in parts of Arizona, California, Colorado, Nevada, New Mexico, and Utah. The statistical models, referred to as classifiers, reflect natural and human-related factors that affect aquifer vulnerability to contamination and relate <span class="hlt">nitrate</span> and arsenic <span class="hlt">concentrations</span> to explanatory variables representing local- and basin-scale measures of source and aquifer susceptibility conditions. Geochemical variables were not used in <span class="hlt">concentration</span> predictions because they were not available for the entire study area. The models were calibrated to assess model accuracy on the basis of measured values.Only 2 percent of the area underlain by basin-fill aquifers in the study area was predicted to equal or exceed the U.S. Environmental Protection Agency drinking-water standard for <span class="hlt">nitrate</span> as N (10 milligrams per liter), whereas 43 percent of the area was predicted to equal or exceed the standard for arsenic (10 micrograms per liter). Areas predicted to equal or exceed the drinking-water standard for <span class="hlt">nitrate</span> include basins in central Arizona near Phoenix; the San Joaquin Valley, the Santa Ana Inland, and San Jacinto Basins of California; and the San Luis Valley of Colorado. Much of the area predicted to equal or exceed the drinking-water standard for arsenic is within a belt of basins along the western portion of the Basin and Range Physiographic Province that includes almost all of Nevada and parts of California and Arizona. Predicted <span class="hlt">nitrate</span> and arsenic <span class="hlt">concentrations</span> are substantially lower than the drinking-water standards in much of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMOS42A..05J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMOS42A..05J"><span>The annual cycle of <span class="hlt">nitrate</span> and net community production in surface waters of the Southern Ocean <span class="hlt">observed</span> with SOCCOM profiling floats</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Johnson, K. S.; Plant, J. N.; Sakamoto, C.; Coletti, L. J.; Sarmiento, J. L.; Riser, S.; Talley, L. D.</p> <p>2016-12-01</p> <p>Sixty profiling floats with ISUS and SUNA <span class="hlt">nitrate</span> sensors have been deployed in the Southern Ocean (south of 30 degrees S) as part of the SOCCOM (Southern Ocean Carbon and Climate <span class="hlt">Observations</span> and Modeling) program and earlier efforts. These floats have produced detailed records of the annual cycle of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> throughout the region from the surface to depths near 2000 m. In surface waters, there are clear cycles in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> that result from uptake of <span class="hlt">nitrate</span> during austral spring and summer. These changes in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> were used to compute the annual net community production over this region. NCP was computed using a simplified version of the approach detailed by Plant et al. (2016, Global Biogeochemical Cycles, 30, 859-879, DOI: 10.1002/2015GB005349). At the time the abstract was written 41 complete annual cycles were available from floats deployed before the austral summer of 2015/2016. After filtering the data to remove floats that crossed distinct frontal boundaries, floats with other anomalies, and floats in sub-tropical waters, 23 cycles were available. A preliminary assessment of the data yields an NCP of 2.8 +/- 0.95 (1 SD) mol C/m2/y after integrating to 100 m depth and converting <span class="hlt">nitrate</span> uptake to carbon using the Redfield ratio. This preliminary assessment ignores vertical transport across the nitracline and is, therefore, a minimum estimate. The number of cycles available for analysis will increase rapidly, as 32 of the floats were deployed in the austral summer of 2015/2016 and have not yet been analyzed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22854088','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22854088"><span><span class="hlt">Nitrate</span> dynamics in agricultural catchments deduced from groundwater dating and long-term <span class="hlt">nitrate</span> monitoring in surface- and groundwaters.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Aquilina, L; Vergnaud-Ayraud, V; Labasque, T; Bour, O; Molénat, J; Ruiz, L; de Montety, V; De Ridder, J; Roques, C; Longuevergne, L</p> <p>2012-10-01</p> <p>Although <span class="hlt">nitrate</span> export in agricultural catchments has been simulated using various types of models, the role of groundwater in <span class="hlt">nitrate</span> dynamics has rarely been fully taken into account. We used groundwater dating methods (CFC analyses) to reconstruct the original <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the groundwater recharge in Brittany (Western France) from 1950 to 2009. This revealed a sharp increase in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> from 1977 to 1990 followed by a slight decrease. The recharge <span class="hlt">concentration</span> curve was then compared with past chronicles of groundwater <span class="hlt">concentration</span>. Groundwater can be interpreted as resulting from the annual dilution of recharge water in an uncontaminated aquifer. Two aquifers were considered: the weathered aquifer and the deeper fractured aquifer. The <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> <span class="hlt">observed</span> in the upper part of the weathered aquifer implied an annual renewal rate of 27 to 33% of the reservoir volume while those in the lower part indicated an annual renewal rate of 2-3%. The <span class="hlt">concentrations</span> in the deep fractured aquifer showed an annual renewal rate of 0.1%. The river <span class="hlt">concentration</span> can be simulated by combining these various groundwater reservoirs with the recharge. Winter and summer waters contain i) recharge water, or water from the variably saturated zone with rapid transfer and high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, and ii) a large contribution (from 35 to 80% in winter and summer, respectively) from the lower part of the aquifer (lower weathered aquifer and deep fractured aquifer). This induces not only a relatively rapid response of the catchment to variations in agricultural pressure, but also a potential inertia which has to be taken into account. Copyright © 2012 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=103998&keyword=Atlantic+AND+forest&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=103998&keyword=Atlantic+AND+forest&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><span class="hlt">NITRATE</span> AND NITROUS OXIDE <span class="hlt">CONCENTRATIONS</span> IN SMALL STREAMS OF THE GEORGIA PIEDMONT</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>We are measuring dissolved <span class="hlt">nitrate</span> and nitrous oxide <span class="hlt">concentrations</span> and related parameters in 17 headwater streams in the South Fork Broad River, Georgia watershed on a monthly basis. The selected small streams drain watersheds dominated by forest, pasture, residential, or mixed...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2010/5098/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2010/5098/"><span><span class="hlt">Nitrate</span> Loads and <span class="hlt">Concentrations</span> in Surface-Water Base Flow and Shallow Groundwater for Selected Basins in the United States, Water Years 1990-2006</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Spahr, Norman E.; Dubrovsky, Neil M.; Gronberg, JoAnn M.; Franke, O. Lehn; Wolock, David M.</p> <p>2010-01-01</p> <p> nutrient management practices designed to reduce nutrient transport to streams by runoff. Conversely, sites with potential for shallow or deep groundwater contribution (some combination of permeable soils or permeable bedrock) had significantly greater contributions of <span class="hlt">nitrate</span> from base flow. Effective nutrient management strategies would consider groundwater <span class="hlt">nitrate</span> contributions in these areas. Mean annual base-flow <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were compared to shallow-groundwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> for 27 sites. <span class="hlt">Concentrations</span> in groundwater tended to be greater than base-flow <span class="hlt">concentrations</span> for this group of sites. Sites where groundwater <span class="hlt">concentrations</span> were much greater than base-flow <span class="hlt">concentrations</span> were found in areas of high infiltration and oxic groundwater conditions. The lack of correspondingly high <span class="hlt">concentrations</span> in the base flow of the paired surface-water sites may have multiple causes. In some settings, there has not been sufficient time for enough high-<span class="hlt">nitrate</span> shallow groundwater to migrate to the nearby stream. In these cases, the stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> lag behind those in the shallow groundwater, and <span class="hlt">concentrations</span> may increase in the future as more high-<span class="hlt">nitrate</span> groundwater reaches the stream. Alternatively, some of these sites may have processes that rapidly remove <span class="hlt">nitrate</span> as water moves from the aquifer into the stream channel. Partitioning streamflow and <span class="hlt">nitrate</span> load between the quick-flow and base-flow portions of the hydrograph coupled with relative scales of soil permeability can infer the importance of surface water compared to groundwater <span class="hlt">nitrate</span> sources. Study of the relation of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> to base-flow index and the comparison of groundwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> to stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> during times when base-flow index is high can provide evidence of potential <span class="hlt">nitrate</span> transport mechanisms. Accounting for the surface-water and groundwater contributions of <span class="hlt">nitrate</span> is crucial to effective management and remediat</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/146817-global-perspective-nitrate-flux-ice-cores','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/146817-global-perspective-nitrate-flux-ice-cores"><span>Global perspective of <span class="hlt">nitrate</span> flux in ice cores</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>Yang, O.; Mayewski, P.A.; Whitlow, S.</p> <p>1995-03-20</p> <p>The relationships between the <span class="hlt">concentration</span> and the flux of chemical species (Cl{sup {minus}}, NO{sub 3}{sup {minus}}, SO{sub 4}{sup 2{minus}}, Na{sup +}, K{sup +}, NH{sub 4}{sup +}, Mg{sub 2+}, Ca{sup 2+}) versus snow accumulation rate were examined at GISP2 and 20D in Greenland, Mount Logan from the St. Elias Range, Yukon Territory, Canada, and Sentik Glacier from the northwest end of the Zanskar Range in the Indian Himalayas. At all sites, only <span class="hlt">nitrate</span> flux is significantly ({alpha}=0.05) related to snow accumulation rate. Of all the chemical series, only <span class="hlt">nitrate</span> <span class="hlt">concentration</span> data are normally distributed. Therefore the authors suggest that <span class="hlt">nitrate</span> concentrationmore » in snow is affected by postdepositional exchange with the atmosphere over a broad range of environmental conditions. The persistant summer maxima in <span class="hlt">nitrate</span> <span class="hlt">observed</span> in Greenland snow over the entire range of record studied (the last 800 years) may be mainly due to NO{sub x} released from peroxyacetyl <span class="hlt">nitrate</span> by thermal decomposition in the presence of higher OH <span class="hlt">concentrations</span> in summer. The late winter/early spring <span class="hlt">nitrate</span> peak <span class="hlt">observed</span> in modern Greenland snow may be related to the buildup of anthropogenically derived NO{sub y} in the Arctic troposphere during the long polar winter. 58 refs., 3 figs., 4 tabs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70027815','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70027815"><span>Algal productivity and <span class="hlt">nitrate</span> assimilation in an effluent dominated concrete lined stream</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kent, Robert; Belitz, Kenneth; Burton, Carmen</p> <p>2005-01-01</p> <p>This study examined algal productivity and <span class="hlt">nitrate</span> assimilation in a 2.85 km reach of Cucamonga Creek, California, a concrete lined channel receiving treated municipal wastewater. Stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> <span class="hlt">observed</span> at two stations indicated nearly continuous loss throughout the diel study. <span class="hlt">Nitrate</span> loss in the reach was approximately 11 mg/L/d or 1.0 g/m2/d as N, most of which occurred during daylight. The peak rate of <span class="hlt">nitrate</span> loss (1.13 mg/l/hr) occurred just prior to an afternoon total CO2 depletion. Gross primary productivity, as estimated by a model using the <span class="hlt">observed</span> differences in dissolved oxygen between the two stations, was 228 mg/L/d, or 21 g/m2/d as O2. The <span class="hlt">observed</span> diel variations in productivity, <span class="hlt">nitrate</span> loss, pH, dissolved oxygen, and CO2indicate that <span class="hlt">nitrate</span> loss was primarily due to algal assimilation. The <span class="hlt">observed</span> levels of productivity and <span class="hlt">nitrate</span> assimilation were exceptionally high on a mass per volume basis compared to studies on other streams; these rates occurred because of the shallow stream depth. This study suggests that concrete‐lined channels can provide an important environmental service: lowering of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> similar to rates <span class="hlt">observed</span> in biological treatment systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27483266','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27483266"><span>Practical Application of Electrochemical <span class="hlt">Nitrate</span> Sensor under Laboratory and Forest Nursery Conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Caron, William-Olivier; Lamhamedi, Mohammed S; Viens, Jeff; Messaddeq, Younès</p> <p>2016-07-28</p> <p>The reduction of <span class="hlt">nitrate</span> leaching to ensure greater protection of groundwater quality has become a global issue. The development of new technologies for more accurate dosing of <span class="hlt">nitrates</span> helps optimize fertilization programs. This paper presents the practical application of a newly developed electrochemical sensor designed for in situ quantification of <span class="hlt">nitrate</span>. To our knowledge, this paper is the first to report the use of electrochemical impedance to determine <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in growing media under forest nursery conditions. Using impedance measurements, the sensor has been tested in laboratory and compared to colorimetric measurements of the <span class="hlt">nitrate</span>. The developed sensor has been used in water-saturated growing medium and showed good correlation to certified methods, even in samples obtained over a multi-ion fertilisation season. A linear and significant relationship was <span class="hlt">observed</span> between the resistance and the <span class="hlt">concentration</span> of <span class="hlt">nitrates</span> (R² = 0.972), for a range of <span class="hlt">concentrations</span> of <span class="hlt">nitrates</span>. We also <span class="hlt">observed</span> stability of the sensor after exposure of one month to the real environmental conditions of the forest nursery.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5017356','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5017356"><span>Practical Application of Electrochemical <span class="hlt">Nitrate</span> Sensor under Laboratory and Forest Nursery 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>Caron, William-Olivier; Lamhamedi, Mohammed S.; Viens, Jeff; Messaddeq, Younès</p> <p>2016-01-01</p> <p>The reduction of <span class="hlt">nitrate</span> leaching to ensure greater protection of groundwater quality has become a global issue. The development of new technologies for more accurate dosing of <span class="hlt">nitrates</span> helps optimize fertilization programs. This paper presents the practical application of a newly developed electrochemical sensor designed for in situ quantification of <span class="hlt">nitrate</span>. To our knowledge, this paper is the first to report the use of electrochemical impedance to determine <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in growing media under forest nursery conditions. Using impedance measurements, the sensor has been tested in laboratory and compared to colorimetric measurements of the <span class="hlt">nitrate</span>. The developed sensor has been used in water-saturated growing medium and showed good correlation to certified methods, even in samples obtained over a multi-ion fertilisation season. A linear and significant relationship was <span class="hlt">observed</span> between the resistance and the <span class="hlt">concentration</span> of <span class="hlt">nitrates</span> (R2 = 0.972), for a range of <span class="hlt">concentrations</span> of <span class="hlt">nitrates</span>. We also <span class="hlt">observed</span> stability of the sensor after exposure of one month to the real environmental conditions of the forest nursery. PMID:27483266</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.B31A0404W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.B31A0404W"><span>Changes in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and loads in United Kingdom watersheds since 1868: Evidence for land use change as a dominant driver and ineffective mitigation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Whelan, M. J.; Burt, T. P.; Howden, N. K.; Worrall, F.</p> <p>2012-12-01</p> <p><span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> and fluxes in many of the world's rivers have increased over the latter part of the 20th Century leading to freshwater and, more frequently, marine eutrophication. This has largely been linked to agricultural intensification which has increased food production via a combination of improved methods, crop varieties, pest control technologies, mechanisation and fertiliser use. The area of land under intensive production has also increased in many watersheds. Here we analyse long term water quality records from a number of UK rivers to assess temporal patterns of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and fluxes and to identify driving factors. The data, which include the world's longest record of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> from the River Thames, London (1868 to 2008), show that dramatic and sustained increases in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and fluxes occurred after periods of substantial land use change including the conversion of significant areas of permanent pasture to arable production in the 1940s and 1960s and increases in fertiliser use in the 1960s. Grassland to arable conversion can release large quantities of <span class="hlt">nitrate</span> due to the enhanced mineralisation of the extra soil organic nitrogen which accumulates in pasture systems. Recent attempts to reverse these increases, such as widespread restrictions of nitrogen inputs have, thus far, been relatively ineffective, although there have been recent indications that levels have stabilised in some watersheds. Our analysis clearly shows the dominance of land use change over other potential drivers such as waste water emission, climatic and hydrological changes and suggests that current measures for the mitigation of large-scale diffuse pollution based on limiting fertiliser use are not drastic enough to reverse significantly the increases in riverine <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> which have been <span class="hlt">observed</span>. Analysis of the long term data also shows that trend detection is problematic for periods less than about 15 years</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70047546','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70047546"><span>Trends in <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and total dissolved solids in public supply wells of the Bunker Hill, Lytle, Rialto, and Colton groundwater subbasins, San Bernardino County, California: Influence of legacy land use</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kent, Robert; Landon, Matthew K.</p> <p>2013-01-01</p> <p><span class="hlt">Concentrations</span> and temporal changes in <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and total dissolved solids (TDS) in groundwater of the Bunker Hill, Lytle, Rialto, and Colton groundwater subbasins of the Upper Santa Ana Valley Groundwater Basin were evaluated to identify trends and factors that may be affecting trends. One hundred, thirty-one public-supply wells were selected for analysis based on the availability of data spanning at least 11 years between the late 1980s and the 2000s. Forty-one of the 131 wells (31%) had a significant (p < 0.10) increase in <span class="hlt">nitrate</span> and 14 wells (11%) had a significant decrease in <span class="hlt">nitrate</span>. For TDS, 46 wells (35%) had a significant increase and 8 wells (6%) had a significant decrease. Slopes for the <span class="hlt">observed</span> significant trends ranged from − 0.44 to 0.91 mg/L/yr for <span class="hlt">nitrate</span> (as N) and − 8 to 13 mg/L/yr for TDS. Increasing <span class="hlt">nitrate</span> trends were associated with greater well depth, higher percentage of agricultural land use, and being closer to the distal end of the flow system. Decreasing <span class="hlt">nitrate</span> trends were associated with the occurrence of volatile organic compounds (VOCs); VOC occurrence decreases with increasing depth. The relations of <span class="hlt">nitrate</span> trends to depth, lateral position, and VOCs imply that increasing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are associated with <span class="hlt">nitrate</span> loading from historical agricultural land use and that more recent urban land use is generally associated with lower <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and greater VOC occurrence. Increasing TDS trends were associated with relatively greater current <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and relatively greater amounts of urban land. Decreasing TDS trends were associated with relatively greater amounts of natural land use. Trends in TDS <span class="hlt">concentrations</span> were not related to depth, lateral position, or VOC occurrence, reflecting more complex factors affecting TDS than <span class="hlt">nitrate</span> in the study area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26988767','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26988767"><span>Is beetroot juice more effective than sodium <span class="hlt">nitrate</span>? The effects of equimolar <span class="hlt">nitrate</span> dosages of <span class="hlt">nitrate</span>-rich beetroot juice and sodium <span class="hlt">nitrate</span> on oxygen consumption during exercise.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Flueck, Joelle Leonie; Bogdanova, Anna; Mettler, Samuel; Perret, Claudio</p> <p>2016-04-01</p> <p>Dietary <span class="hlt">nitrate</span> has been reported to lower oxygen consumption in moderate- and severe-intensity exercise. To date, it is unproven that sodium <span class="hlt">nitrate</span> (NaNO3(-); NIT) and <span class="hlt">nitrate</span>-rich beetroot juice (BR) have the same effects on oxygen consumption, blood pressure, and plasma <span class="hlt">nitrate</span> and nitrite <span class="hlt">concentrations</span> or not. The aim of this study was to compare the effects of different dosages of NIT and BR on oxygen consumption in male athletes. Twelve healthy, well-trained men (median [minimum; maximum]; peak oxygen consumption: 59.4 mL·min(-1)·kg(-1) [40.5; 67.0]) performed 7 trials on different days, ingesting different <span class="hlt">nitrate</span> dosages and placebo (PLC). Dosages were 3, 6, and 12 mmol <span class="hlt">nitrate</span> as <span class="hlt">concentrated</span> BR or NIT dissolved in plain water. Plasma <span class="hlt">nitrate</span> and nitrite <span class="hlt">concentrations</span> were measured before, 3 h after ingestion, and postexercise. Participants cycled for 5 min at moderate intensity and further 8 min at severe intensity. End-exercise oxygen consumption at moderate intensity was not significantly different between the 7 trials (p = 0.08). At severe-intensity exercise, end-exercise oxygen consumption was ~4% lower in the 6-mmol BR trial compared with the 6-mmol NIT (p = 0.003) trial as well as compared with PLC (p = 0.010). Plasma nitrite and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were significantly increased after the ingestion of BR and NIT with the highest <span class="hlt">concentrations</span> in the 12-mmol trials. Plasma nitrite <span class="hlt">concentration</span> between NIT and BR did not significantly differ in the 6-mmol (p = 0.27) and in the 12-mmol (p = 0.75) trials. In conclusion, BR might reduce oxygen consumption to a greater extent compared with NIT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4011814','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4011814"><span>Developmental stage- and <span class="hlt">concentration</span>-specific sodium nitroprusside application results in <span class="hlt">nitrate</span> reductase regulation and the modification of <span class="hlt">nitrate</span> metabolism in leaves of Medicago truncatula 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>Antoniou, Chrystalla; Filippou, Panagiota; Mylona, Photini; Fasoula, Dionysia; Ioannides, Ioannis; Polidoros, Alexios; Fotopoulos, Vasileios</p> <p>2013-01-01</p> <p>Nitric oxide (NO) is a bioactive molecule involved in numerous biological events that has been reported to display both pro-oxidant and antioxidant properties in plants. Several reports exist which demonstrate the protective action of sodium nitroprusside (SNP), a widely used NO donor, which acts as a signal molecule in plants responsible for the expression regulation of many antioxidant enzymes. This study attempts to provide a novel insight into the effect of application of low (100 μΜ) and high (2.5 mM) <span class="hlt">concentrations</span> of SNP on the nitrosative status and <span class="hlt">nitrate</span> metabolism of mature (40 d) and senescing (65 d) Medicago truncatula plants. Higher <span class="hlt">concentrations</span> of SNP resulted in increased NO content, cellular damage levels and reactive oxygen species (ROS) <span class="hlt">concentration</span>, further induced in older tissues. Senescing M. truncatula plants demonstrated greater sensitivity to SNP-induced oxidative and nitrosative damage, suggesting a developmental stage-dependent suppression in the plant’s capacity to cope with free oxygen and nitrogen radicals. In addition, measurements of the activity of <span class="hlt">nitrate</span> reductase (NR), a key enzyme involved in the generation of NO in plants, indicated a differential regulation in a dose and time-dependent manner. Furthermore, expression levels of NO-responsive genes (NR, <span class="hlt">nitrate</span>/nitrite transporters) involved in nitrogen assimilation and NO production revealed significant induction of NR and <span class="hlt">nitrate</span> transporter during long-term 2.5 mM SNP application in mature plants and overall gene suppression in senescing plants, supporting the differential nitrosative response of M. truncatula plants treated with different <span class="hlt">concentrations</span> of SNP. PMID:23838961</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1367981','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1367981"><span><span class="hlt">Nitrate</span> vulnerability projections from Bayesian inference of multiple groundwater age tracers</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>Alikhani, Jamal; Deinhart, Amanda L.; Visser, Ate</p> <p></p> <p><span class="hlt">Nitrate</span> is a major source of contamination of groundwater in the United States and around the world. We tested the applicability of multiple groundwater age tracers ( 3H, 3He, 4He, 14C, 13C, and 85Kr) in projecting future trends of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in 9 long-screened, public drinking water wells in Turlock, California, where <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are increasing toward the regulatory limit. Very low 85Kr <span class="hlt">concentrations</span> and apparent 3H/ 3He ages point to a relatively old modern fraction (40–50 years), diluted with pre-modern groundwater, corroborated by the onset and slope of increasing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. An inverse Gaussian–Dirac model was chosen to representmore » the age distribution of the sampled groundwater at each well. Model parameters were estimated using a Bayesian inference, resulting in the posterior probability distribution – including the associated uncertainty – of the parameters and projected <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Three scenarios were considered, including combined historic <span class="hlt">nitrate</span> and age tracer data, the sole use of <span class="hlt">nitrate</span> and the sole use of age tracer data. Each scenario was evaluated based on the ability of the model to reproduce the data and the level of reliability of the <span class="hlt">nitrate</span> projections. The tracer-only scenario closely reproduced tracer <span class="hlt">concentrations</span>, but not <span class="hlt">observed</span> trends in the <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. Both cases that included <span class="hlt">nitrate</span> data resulted in good agreement with historical <span class="hlt">nitrate</span> trends. Use of combined tracers and <span class="hlt">nitrate</span> data resulted in a narrower range of projections of future <span class="hlt">nitrate</span> levels. However, use of combined tracer and <span class="hlt">nitrate</span> resulted in a larger discrepancy between modeled and measured tracers for some of the tracers. In conclusion, despite <span class="hlt">nitrate</span> trend slopes between 0.56 and 1.73 mg/L/year in 7 of the 9 wells, the probability that <span class="hlt">concentrations</span> will increase to levels above the MCL by 2040 are over 95% for only two of the wells, and below 15% in the other wells, due to a leveling off</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1367981-nitrate-vulnerability-projections-from-bayesian-inference-multiple-groundwater-age-tracers','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1367981-nitrate-vulnerability-projections-from-bayesian-inference-multiple-groundwater-age-tracers"><span><span class="hlt">Nitrate</span> vulnerability projections from Bayesian inference of multiple groundwater age tracers</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Alikhani, Jamal; Deinhart, Amanda L.; Visser, Ate; ...</p> <p>2016-04-20</p> <p><span class="hlt">Nitrate</span> is a major source of contamination of groundwater in the United States and around the world. We tested the applicability of multiple groundwater age tracers ( 3H, 3He, 4He, 14C, 13C, and 85Kr) in projecting future trends of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in 9 long-screened, public drinking water wells in Turlock, California, where <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are increasing toward the regulatory limit. Very low 85Kr <span class="hlt">concentrations</span> and apparent 3H/ 3He ages point to a relatively old modern fraction (40–50 years), diluted with pre-modern groundwater, corroborated by the onset and slope of increasing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. An inverse Gaussian–Dirac model was chosen to representmore » the age distribution of the sampled groundwater at each well. Model parameters were estimated using a Bayesian inference, resulting in the posterior probability distribution – including the associated uncertainty – of the parameters and projected <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Three scenarios were considered, including combined historic <span class="hlt">nitrate</span> and age tracer data, the sole use of <span class="hlt">nitrate</span> and the sole use of age tracer data. Each scenario was evaluated based on the ability of the model to reproduce the data and the level of reliability of the <span class="hlt">nitrate</span> projections. The tracer-only scenario closely reproduced tracer <span class="hlt">concentrations</span>, but not <span class="hlt">observed</span> trends in the <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. Both cases that included <span class="hlt">nitrate</span> data resulted in good agreement with historical <span class="hlt">nitrate</span> trends. Use of combined tracers and <span class="hlt">nitrate</span> data resulted in a narrower range of projections of future <span class="hlt">nitrate</span> levels. However, use of combined tracer and <span class="hlt">nitrate</span> resulted in a larger discrepancy between modeled and measured tracers for some of the tracers. In conclusion, despite <span class="hlt">nitrate</span> trend slopes between 0.56 and 1.73 mg/L/year in 7 of the 9 wells, the probability that <span class="hlt">concentrations</span> will increase to levels above the MCL by 2040 are over 95% for only two of the wells, and below 15% in the other wells, due to a leveling off</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2015/5075/support/sir20155075.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2015/5075/support/sir20155075.pdf"><span>Median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater in the New Jersey Highlands Region estimated using regression models and land-surface characteristics</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Baker, Ronald J.; Chepiga, Mary M.; Cauller, Stephen J.</p> <p>2015-01-01</p> <p>The Kaplan-Meier method of estimating summary statistics from left-censored data was applied in order to include nondetects (left-censored data) in median <span class="hlt">nitrate-concentration</span> calculations. Median <span class="hlt">concentrations</span> also were determined using three alternative methods of handling nondetects. Treatment of the 23 percent of samples that were nondetects had little effect on estimated median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> because method detection limits were mostly less than median values.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JSR...127...26N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JSR...127...26N"><span><span class="hlt">Nitrate</span> consumption in sediments of the German Bight (North Sea)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Neumann, Andreas; van Beusekom, Justus E. E.; Holtappels, Moritz; Emeis, Kay-Christian</p> <p>2017-09-01</p> <p>Denitrification on continental margins and in coastal sediments is a major sink of reactive N in the present nitrogen cycle and a major ecosystem service of eutrophied coastal waters. We analyzed the <span class="hlt">nitrate</span> removal in surface sediments of the Elbe estuary, Wadden Sea, and adjacent German Bight (SE North Sea) during two seasons (spring and summer) along a eutrophication gradient ranging from a high riverine <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> at the Elbe Estuary to offshore areas with low <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. The gradient encompassed the full range of sediment types and organic carbon <span class="hlt">concentrations</span> of the southern North Sea. Based on <span class="hlt">nitrate</span> penetration depth and <span class="hlt">concentration</span> gradient in the porewater we estimated benthic <span class="hlt">nitrate</span> consumption rates assuming either diffusive transport in cohesive sediments or advective transport in permeable sediments. For the latter we derived a mechanistic model of porewater flow. During the peak <span class="hlt">nitrate</span> discharge of the river Elbe in March, the highest rates of diffusive <span class="hlt">nitrate</span> uptake were <span class="hlt">observed</span> in muddy sediments (up to 2.8 mmol m- 2 d- 1). The highest advective uptake rate in that period was <span class="hlt">observed</span> in permeable sediment and was tenfold higher (up to 32 mmol m- 2 d- 1). The intensity of both diffusive and advective <span class="hlt">nitrate</span> consumption dropped with the <span class="hlt">nitrate</span> availability and thus decreased from the Elbe estuary towards offshore stations, and were further decreased during late summer (minimum <span class="hlt">nitrate</span> discharge) compared to late winter (maximum <span class="hlt">nitrate</span> discharge). In summary, our rate measurements indicate that the permeable sediment accounts for up to 90% of the total benthic reactive nitrogen consumption in the study area due to the high efficiency of advective <span class="hlt">nitrate</span> transport into permeable sediment. Extrapolating the averaged <span class="hlt">nitrate</span> consumption of different sediment classes to the areas of Elbe Estuary, Wadden Sea and eastern German Bight amounts to an N-loss of 3.1 ∗ 106 mol N d- 1 from impermeable, diffusion</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5992502','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5992502"><span>Knock-Down of CsNRT2.1, a Cucumber <span class="hlt">Nitrate</span> Transporter, Reduces <span class="hlt">Nitrate</span> Uptake, Root length, and Lateral Root Number at Low External <span class="hlt">Nitrate</span> <span class="hlt">Concentration</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>Li, Yang; Li, Juanqi; Yan, Yan; Liu, Wenqian; Zhang, Wenna; Gao, Lihong; Tian, Yongqiang</p> <p>2018-01-01</p> <p>Nitrogen (N) is a macronutrient that plays a crucial role in plant growth and development. <span class="hlt">Nitrate</span> (NO3-) is the most abundant N source in aerobic soils. Plants have evolved two adaptive mechanisms such as up-regulation of the high-affinity transport system (HATS) and alteration of the root system architecture (RSA), allowing them to cope with the temporal and spatial variation of NO3-. However, little information is available regarding the <span class="hlt">nitrate</span> transporter in cucumber, one of the most important fruit vegetables in the world. In this study we isolated a <span class="hlt">nitrate</span> transporter named CsNRT2.1 from cucumber. Analysis of the expression profile of the CsNRT2.1 showed that CsNRT2.1 is a high affinity <span class="hlt">nitrate</span> transporter which mainly located in mature roots. Subcellular localization analysis revealed that CsNRT2.1 is a plasma membrane transporter. In N-starved CsNRT2.1 knock-down plants, both of the constitutive HATS (cHATS) and inducible HATS (iHATS) were impaired under low external NO3- <span class="hlt">concentration</span>. Furthermore, the CsNRT2.1 knock-down plants showed reduced root length and lateral root numbers. Together, our results demonstrated that CsNRT2.1 played a dual role in regulating the HATS and RSA to acquire NO3- effectively under N limitation. PMID:29911677</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70029316','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70029316"><span>Temporal variations and scaling of streamflow and baseflow and their <span class="hlt">nitrate</span>-nitrogen <span class="hlt">concentrations</span> and loads</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Zhang, Y.-K.; Schilling, K.</p> <p>2005-01-01</p> <p>The patterns of temporal variations of precipitation (P), streamflow (SF) and baseflow (BF) as well as their <span class="hlt">nitrate</span>-nitrogen (<span class="hlt">nitrate</span>) <span class="hlt">concentrations</span> (C) and loads (L) from a long-term record (28 years) in the Raccoon River, Iowa, were analyzed using variogram and spectral analyses. The daily P is random but scaling may exist in the daily SF and BF with a possible break point in the scaling at about 18 days and 45 days, respectively. The <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and loads are shown to have a half-year cycle while daily P, SF, and BF have a one-year cycle. Furthermore, there may be a low-frequency cycle of 6-8 years in C. The power spectra of C and L in both SF and BF exhibit fractal 1/f scaling with two characteristic frequencies of half-year and one-year, and are fitted well with the spectrum of the gamma distribution. The <span class="hlt">nitrate</span> input to SF and BF at the Raccoon watershed seems likely to be a white noise process superimposed on another process with a half-year and one-year cycle. ?? 2005 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_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/2018JEnM...36..191L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JEnM...36..191L"><span>Mechanochemical <span class="hlt">Nitration</span> of Aromatic Compounds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lagoviyer, Oleg S.; Krishtopa, Larisa; Schoenitz, Mirko; Trivedi, Nirupam J.; Dreizin, Edward L.</p> <p>2018-04-01</p> <p><span class="hlt">Nitration</span> of organic compounds is necessary to produce many energetic materials, such as TNT and nitrocellulose. The conventional <span class="hlt">nitration</span> process uses a mixture of <span class="hlt">concentrated</span> sulfuric and nitric acids as <span class="hlt">nitrating</span> agents and multiple solvents. The chemicals are corrosive and require special handling and disposal procedures. In this study, aromatic <span class="hlt">nitration</span> has been achieved using solvent-free mechanochemical processing of environmentally benign precursors. Mononitrotoluene was synthesized by milling toluene with sodium <span class="hlt">nitrate</span> and molybdenum trioxide as a Lewis acid catalyst. Several parameters affecting the desired product yield were identified and varied. A number of byproducts, i.e., dimers of toluene were also produced, but the selectivity was <span class="hlt">observed</span> to increase with increasing mononitrotoluene yield. Both absolute mononitrotoluene yields and selectivity of its production increased with the increase in the energy transferred to the material from the milling tools.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.water.usgs.gov/wri034084/','USGSPUBS'); return false;" href="http://pubs.water.usgs.gov/wri034084/"><span>Percentile Distributions of Median Nitrite Plus <span class="hlt">Nitrate</span> as Nitrogen, Total Nitrogen, and Total Phosphorus <span class="hlt">Concentrations</span> in Oklahoma Streams, 1973-2001</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Haggard, Brian E.; Masoner, Jason R.; Becker, Carol J.</p> <p>2003-01-01</p> <p>Nutrients are one of the primary causes of water-quality impairments in streams, lakes, reservoirs, and estuaries in the United States. The U.S. Environmental Protection Agency has developed regional-based nutrient criteria using ecoregions to protect streams in the United States from impairment. However, nutrient criteria were based on nutrient <span class="hlt">concentrations</span> measured in large aggregated nutrient ecoregions with little relevance to local environmental conditions in states. The Oklahoma Water Resources Board is using a dichotomous process known as Use Support Assessment Protocols to define nutrient criteria in Oklahoma streams. The Oklahoma Water Resources Board is modifying the Use Support Assessment Protocols to reflect nutrient informa-tion and environmental characteristics relevant to Oklahoma streams, while considering nutrient information grouped by geographic regions based on level III ecoregions and state boundaries. Percentile distributions of median nitrite plus <span class="hlt">nitrate</span> as nitrogen, total nitrogen, and total phosphorous <span class="hlt">concentrations</span> were calculated from 563 sites in Oklahoma and 4 sites in Arkansas near the Oklahoma and Arkansas border to facilitate development of nutrient criteria for Oklahoma streams. Sites were grouped into four geographic regions and were categorized into eight stream categories by stream slope and stream order. The 50th percentiles of median nitrite plus <span class="hlt">nitrate</span> as nitrogen, total nitrogen, and total phosphorus <span class="hlt">concentrations</span> were greater in the Ozark Highland ecoregion and were less in the Ouachita Mountains ecoregion when compared to other geographic areas used to group sites. The 50th percentiles of median <span class="hlt">concentrations</span> of nitrite plus <span class="hlt">nitrate</span> as nitrogen, total nitrogen, and total phosphorus were least in first, second, and third order streams. The 50th percentiles of median nitrite plus <span class="hlt">nitrate</span> as nitrogen, total nitrogen and total phosphorus <span class="hlt">concentrations</span> in the Ozark Highland and Ouachita Mountains ecoregions were least in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25358487','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25358487"><span>Inhibition of <span class="hlt">nitrate</span> reduction by NaCl adsorption on a nano-zero-valent iron surface during a <span class="hlt">concentrate</span> treatment for water reuse.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hwang, Yuhoon; Kim, Dogun; Shin, Hang-Sik</p> <p>2015-01-01</p> <p>Nanoscale zero-valent iron (NZVI) has been considered as a possible material to treat water and wastewater. However, it is necessary to verify the effect of the matrix components in different types of target water. In this study, different effects depending on the sodium chloride (NaCl) <span class="hlt">concentration</span> on reductions of <span class="hlt">nitrates</span> and on the characteristics of NZVI were investigated. Although NaCl is known as a promoter of iron corrosion, a high <span class="hlt">concentration</span> of NaCl (>3 g/L) has a significant inhibition effect on the degree of NZVI reactivity towards <span class="hlt">nitrate</span>. The experimental results were interpreted by a Langmuir-Hinshelwood-Hougen-Watson reaction in terms of inhibition, and the decreased NZVI reactivity could be explained by the increase in the inhibition constant. As a result of a chloride <span class="hlt">concentration</span> analysis, it was verified that 7.7-26.5% of chloride was adsorbed onto the surface of NZVI. Moreover, the change of the iron corrosion product under different NaCl <span class="hlt">concentrations</span> was investigated by a surface analysis of spent NZVI. Magnetite was the main product, with a low NaCl <span class="hlt">concentration</span> (0.5 g/L), whereas amorphous iron hydroxide was <span class="hlt">observed</span> at a high <span class="hlt">concentration</span> (12 g/L). Though the surface was changed to permeable iron hydroxide, the Fe(0) in the core was not completely oxidized. Therefore, the inhibition effect of NaCl could be explained as the competitive adsorption of chloride and <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/ds/1084/ds1084.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/ds/1084/ds1084.pdf"><span><span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> in drinking water in the lower Yakima River Basin, Groundwater Management Area, Yakima County, Washington, 2017</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Huffman, Raegan L.</p> <p>2018-05-29</p> <p>The U.S. Geological Survey, in cooperation with the lower Yakima River Basin Groundwater Management Area (GWMA) group, conducted an intensive groundwater sampling collection effort of collecting <span class="hlt">nitrate</span> <span class="hlt">concentration</span> data in drinking water to provide a baseline for future <span class="hlt">nitrate</span> assessments within the GWMA. About every 6 weeks from April through December 2017, a total of 1,059 samples were collected from 156 wells and 24 surface-water drains. The domestic wells were selected based on known location, completion depth, ability to collect a sample prior to treatment on filtration, and distribution across the GWMA. The drains were pre-selected by the GWMA group, and further assessed based on ability to access sites and obtain a representative sample. More than 20 percent of samples from the domestic wells and 12.8 percent of drain samples had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> that exceeded the maximum contaminant level (MCL) of 10 milligrams per liter established by the U.S. Environmental Protection Agency. At least one <span class="hlt">nitrate</span> <span class="hlt">concentration</span> above the MCL was detected in 26 percent of wells and 33 percent of drains sampled. <span class="hlt">Nitrate</span> was not detected in 13 percent of all samples collected.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ACP....1211213W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ACP....1211213W"><span>Simulation of <span class="hlt">nitrate</span>, sulfate, and ammonium aerosols over the United States</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Walker, J. M.; Philip, S.; Martin, R. V.; Seinfeld, J. H.</p> <p>2012-11-01</p> <p>Atmospheric <span class="hlt">concentrations</span> of inorganic gases and aerosols (<span class="hlt">nitrate</span>, sulfate, and ammonium) are simulated for 2009 over the United States using the chemical transport model GEOS-Chem. Predicted aerosol <span class="hlt">concentrations</span> are compared with surface-level measurement data from the Interagency Monitoring of Protected Visual Environments (IMPROVE), the Clean Air Status and Trends Network (CASTNET), and the California Air Resources Board (CARB). Sulfate predictions nationwide are in reasonably good agreement with <span class="hlt">observations</span>, while <span class="hlt">nitrate</span> and ammonium are over-predicted in the East and Midwest, but under-predicted in California, where <span class="hlt">observed</span> <span class="hlt">concentrations</span> are the highest in the country. Over-prediction of <span class="hlt">nitrate</span> in the East and Midwest is consistent with results of recent studies, which suggest that nighttime nitric acid formation by heterogeneous hydrolysis of N2O5 is over-predicted based on current values of the N2O5 uptake coefficient, γ, onto aerosols. After reducing the value of γ by a factor of 10, predicted <span class="hlt">nitrate</span> levels in the US Midwest and East still remain higher than those measured, and over-prediction of <span class="hlt">nitrate</span> in this region remains unexplained. Comparison of model predictions with satellite measurements of ammonia from the Tropospheric Emissions Spectrometer (TES) indicates that ammonia emissions in GEOS-Chem are underestimated in California and that the nationwide seasonality applied to ammonia emissions in GEOS-Chem does not represent California very well, particularly underestimating winter emissions. An ammonia sensitivity study indicates that GEOS-Chem simulation of <span class="hlt">nitrate</span> is ammonia-limited in southern California and much of the state, suggesting that an underestimate of ammonia emissions is likely the main cause for the under-prediction of <span class="hlt">nitrate</span> aerosol in many areas of California. An approximate doubling of ammonia emissions is needed to reproduce <span class="hlt">observed</span> <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in southern California and in other ammonia sensitive areas</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27639783','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27639783"><span><span class="hlt">Nitration</span> of pollen aeroallergens by <span class="hlt">nitrate</span> ion in conditions simulating the liquid water phase of atmospheric particles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ghiani, Alessandra; Bruschi, Maurizio; Citterio, Sandra; Bolzacchini, Ezio; Ferrero, Luca; Sangiorgi, Giorgia; Asero, Riccardo; Perrone, Maria Grazia</p> <p>2016-12-15</p> <p>Pollen aeroallergens are present in atmospheric particulate matter (PM) where they can be found in coarse biological particles such as pollen grains (aerodynamic diameter d ae >10μm), as well as fragments in the finest respirable particles (PM2.5; d ae <2.5μm). <span class="hlt">Nitration</span> of tyrosine residues in pollen allergenic proteins can occur in polluted air, and inhalation and deposition of these <span class="hlt">nitrated</span> proteins in the human respiratory tract may lead to adverse health effects by enhancing the allergic response in population. Previous studies investigated protein <span class="hlt">nitration</span> by atmospheric gaseous pollutants such as nitrogen dioxide and ozone. In this work we report, for the first time, a study on protein <span class="hlt">nitration</span> by <span class="hlt">nitrate</span> ion in aqueous solution, at <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and pH conditions simulating those occurring in the atmospheric aerosol liquid water phase. Experiments have been carried out on the Bovine serum albumin (BSA) protein and the recombinant Phleum pratense allergen (Phl p 2) both in the dark and under UV-A irradiation (range 4-90Wm -2 ) to take into account thermal and/or photochemical <span class="hlt">nitration</span> processes. For the latter protein, modifications in the allergic response after treatment with <span class="hlt">nitrate</span> solutions have been evaluated by immunoblot analyses using sera from grass-allergic patients. Experimental results in bulk solutions showed that protein <span class="hlt">nitration</span> in the dark occurs only in dilute <span class="hlt">nitrate</span> solutions and under very acidic conditions (pH<3 for BSA; pH<2.2 for Phl p 2), while <span class="hlt">nitration</span> is always <span class="hlt">observed</span> (at pH0.5-5) under UV-A irradiation, both in dilute and <span class="hlt">concentrated</span> <span class="hlt">nitrate</span> solutions, being significantly enhanced at the lowest pH values. In some cases, protein <span class="hlt">nitration</span> resulted in an increase of the allergic response. Copyright © 2016. Published by Elsevier B.V.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29324081','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29324081"><span>Determination of endogenous <span class="hlt">concentrations</span> of nitrites and <span class="hlt">nitrates</span> in different types of cheese in the United States: method development and validation using ion chromatography.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Genualdi, Susan; Jeong, Nahyun; DeJager, Lowri</p> <p>2018-04-01</p> <p>Nitrites and <span class="hlt">nitrates</span> can be present in dairy products from both endogenous and exogenous sources. In the European Union (EU), 150 mg kg - 1 of <span class="hlt">nitrates</span> are allowed to be added to the cheese milk during the manufacturing process. The CODEX General Standard for Food Additives has a maximum permitted level of 50 mg kg - 1 residue in cheese, while in the United States (U.S.) <span class="hlt">nitrates</span> are unapproved for use as food additives in cheese. In order to be able to investigate imported cheeses for <span class="hlt">nitrates</span> intentionally added as preservatives and the endogenous <span class="hlt">concentrations</span> of <span class="hlt">nitrates</span> and nitrites present in cheeses in the U.S. marketplace, a method was developed and validated using ion chromatography with conductivity detection. A market sampling of cheese samples purchased in the Washington DC metro area was performed. In 64 samples of cheese, <span class="hlt">concentrations</span> ranged from below the method detection limit (MDL) to 26 mg kg - 1 for <span class="hlt">nitrates</span> and no <span class="hlt">concentrations</span> of nitrites were found in any of the cheese samples above the MDL of 0.1 mg kg - 1 . A majority of the samples (93%) had <span class="hlt">concentrations</span> below 10 mg kg - 1 , which indicate the presence of endogenous <span class="hlt">nitrates</span>. The samples with <span class="hlt">concentrations</span> above 10 mg kg - 1 were mainly processed cheese spread, which can contain additional ingredients often of plant-based origin. These ingredients are likely the cause of the elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. The analysis of 12 additional cheese samples that are liable to the intentional addition of <span class="hlt">nitrates</span>, 9 of which were imported, indicated that in this limited study, <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> in the U.S.-produced cheeses did not differ from those in imported samples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28986080','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28986080"><span>Chronic <span class="hlt">nitrate</span> exposure alters reproductive physiology in fathead minnows.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kellock, Kristen A; Moore, Adrian P; Bringolf, Robert B</p> <p>2018-01-01</p> <p><span class="hlt">Nitrate</span> is a ubiquitous aquatic pollutant that is commonly associated with eutrophication and dead zones in estuaries around the world. At high <span class="hlt">concentrations</span> <span class="hlt">nitrate</span> is toxic to aquatic life but at environmental <span class="hlt">concentrations</span> it has also been purported as an endocrine disruptor in fish. To investigate the potential for <span class="hlt">nitrate</span> to cause endocrine disruption in fish, we conducted a lifecycle study with fathead minnows (Pimephales promelas) exposed to <span class="hlt">nitrate</span> (0, 11.3, and 56.5 mg/L (total <span class="hlt">nitrate</span>-nitrogen (NO 3 -N)) from <24 h post hatch to sexual maturity (209 days). Body mass, condition factor, gonadal somatic index (GSI), incidence of intersex, and vitellogenin induction were determined in mature male and female fish and plasma 11-keto testosterone (11-KT) was measured in males only. In <span class="hlt">nitrate</span>-exposed males both 11-KT and vitellogenin were significantly induced when compared with controls. No significant differences occurred for body mass, condition factor, or GSI among males and intersex was not <span class="hlt">observed</span> in any of the <span class="hlt">nitrate</span> treatments. <span class="hlt">Nitrate</span>-exposed females also had significant increases in vitellogenin compared to controls but no significant differences for mass, condition factor, or GSI were <span class="hlt">observed</span> in <span class="hlt">nitrate</span> exposed groups. Estradiol was used as a positive control for vitellogenin induction. Our findings suggest that environmentally relevant <span class="hlt">nitrate</span> levels may disrupt steroid hormone synthesis and/or metabolism in male and female fish and may have implications for fish reproduction, watershed management, and regulation of nutrient pollution. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009ACP.....9.1863L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009ACP.....9.1863L"><span><span class="hlt">Observation</span> of <span class="hlt">nitrate</span> coatings on atmospheric mineral dust particles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, W. J.; Shao, L. Y.</p> <p>2009-03-01</p> <p><span class="hlt">Nitrate</span> compounds have received much attention because of their ability to alter the hygroscopic properties and cloud condensation nuclei (CCN) activity of mineral dust particles in the atmosphere. However, very little is known about specific characteristics of ambient <span class="hlt">nitrate</span>-coated mineral particles on an individual particle scale. In this study, sample collection was conducted during brown haze and dust episodes between 24 May and 21 June 2007 in Beijing, northern China. Sizes, morphologies, and compositions of 332 mineral dust particles together with their coatings were analyzed using transmission electron microscopy (TEM) coupled with energy-dispersive X-ray (EDX) microanalyses. Structures of some mineral particles were verified using selected-area electron diffraction (SAED). TEM <span class="hlt">observation</span> indicates that approximately 90% of the collected mineral particles are covered by visible coatings in haze samples whereas only 5% are coated in the dust sample. 92% of the analyzed mineral particles are covered with Ca-, Mg-, and Na-rich coatings, and 8% are associated with K- and S-rich coatings. The majority of coatings contain Ca, Mg, O, and N with minor amounts of S and Cl, suggesting that they are possibly <span class="hlt">nitrates</span> mixed with small amounts of sulfates and chlorides. These <span class="hlt">nitrate</span> coatings are strongly correlated with the presence of alkaline mineral components (e.g., calcite and dolomite). CaSO4 particles with diameters from 10 to 500 nm were also detected in the coatings including Ca(NO3)2 and Mg(NO3)2. Our results indicate that mineral particles in brown haze episodes were involved in atmospheric heterogeneous reactions with two or more acidic gases (e.g., SO2, NO2, HCl, and HNO3). Mineral particles that acquire hygroscopic <span class="hlt">nitrate</span> coatings tend to be more spherical and larger, enhancing their light scattering and CCN activity, both of which have cooling effects on the climate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26913813','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26913813"><span>Concurrent microbial reduction of high <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and perchlorate in an ion exchange membrane bioreactor.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fox, Shalom; Bruner, Tali; Oren, Yoram; Gilron, Jack; Ronen, Zeev</p> <p>2016-09-01</p> <p>We investigated effective simultaneous removal of high loads of <span class="hlt">nitrate</span> and perchlorate from synthetic groundwater using an ion exchange membrane bioreactor (IEMB). The aim of this research was to characterize both transport aspects and biodegradation mechanisms involved in the treatment process of high loads of the two anions. Biodegradation process was proven to be efficient with over 99% efficiency of both perchlorate and <span class="hlt">nitrate</span>, regardless of their load. The maximum biodegradation rates were 18.3 (mmol m(-2)  h(-1) ) and 5.5 (mmol m(-2)  h(-1) ) for <span class="hlt">nitrate</span> and perchlorate, respectively. The presence of a biofilm on the bio-side of the membrane only slightly increased the <span class="hlt">nitrate</span> and perchlorate transmembrane flux as compared to the measured flux during a Donnan dialysis experiment where there is no biodegradation of perchlorate and <span class="hlt">nitrate</span> in the bio-compartment. The <span class="hlt">nitrate</span> flux in presence of a biofilm was 18.3 (±1.9) (mmole m(-2)  h(-1) ), while without the biofilm, the flux was 16.9 (±1.5) (mmole m(-2)  h(-1) ) for the same feed inlet <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of 4 mM. The perchlorate transmembrane flux increased similarly by an average of 5%. Samples of membrane biofilm and suspended bacteria from the bio-reactor were analyzed for diversity and abundance of the perchlorate and <span class="hlt">nitrate</span> reducing bacteria. Klebsiella oxytoca, known as a glycerol fermenter, accounted for 70% of the suspended bacteria. In contrast, perchlorate and <span class="hlt">nitrate</span> reducing bacteria predominated in the biofilm present on the membrane. These results are consistent with our proposed two stage biodegradation mechanism where glycerol is first fermented in the suspended phase of the bio-reactor and the fermentation products drive perchlorate and <span class="hlt">nitrate</span> bio-reduction in the biofilm attached to the membrane. These results suggest that the niche exclusion of microbial populations in between the reactor and membrane is controlled by the fluxes of the electron donors and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70010032','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70010032"><span>Near-decadal changes in <span class="hlt">nitrate</span> and pesticide <span class="hlt">concentrations</span> in the South Platte River alluvial aquifer, 1993-2004</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Paschke, S.S.; Schaffrath, K.R.; Mashbum, S.L.</p> <p>2008-01-01</p> <p>The lower South Platte River basin of Colorado and Nebraska is an area of intense agriculture supported by surface-water diversions from the river and ground-water pumping from a valley-fill alluvial aquifer. Two well networks consisting of 45 wells installed in the South Platte alluvial aquifer were sampled in the early 1990s and again in the early 2000s to examine near-decadal ground-water quality changes in irrigated agricultural areas. Ground-water age generally increases and dissolved-oxygen content decreases with distance along flow paths and with depdi below the water table, and denitrification is an important natural mitigation mechanism for <span class="hlt">nitrate</span> in downgradient areas. Ground-water travel time from upland areas to the river ranges from 12 to 31 yr on the basis of apparent ground-water ages. Ground-water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> for agricultural land-use wells increased significantly for oxidized samples over the decade, and nitrogen isotope ratios for oxidized samples indicate synthetic fertilizer as the predominant <span class="hlt">nitrate</span> source. Ground-water <span class="hlt">concentrations</span> of atrazine, DEA, and prometon decreased significandy. The decrease in pesticide <span class="hlt">concentrations</span> and a significant increase in the ratio of DEA to atrazine suggest decreases in pesticide <span class="hlt">concentrations</span> are likely caused by local decreases in application rates and/or degradation processes and that atrazine degradation is promoted by oxidizing conditions. The difference between results for oxidizing and <span class="hlt">nitrate</span>-reducing conditions indicates redox state is an important variable to consider when evaluating ground-water quality trends for redox-sensitive constituents such as <span class="hlt">nitrate</span> and pesticides in the South Platte alluvial aquifer. Copyright ?? 2008 by the American Society of Agronomy, Crop Science Society of America, and Soil Science Society of America. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994JGR....9925369M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994JGR....9925369M"><span>Relationships between organic <span class="hlt">nitrates</span> and surface ozone destruction during Polar Sunrise Experiment 1992</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Muthuramu, K.; Shepson, P. B.; Bottenheim, J. W.; Jobson, B. T.; Niki, H.; Anlauf, K. G.</p> <p>1994-12-01</p> <p>Concurrent measurements of total reactive odd nitrogen species (i.e., NOy) and its major components, including organic <span class="hlt">nitrates</span>, were carried out during 1992 Polar Sunrise Experiment (PSE92) at Alert, Northwest Territories, Canada, to investigate the episodic depletion of surface level ozone following polar sunrise. A series of C3-C7 alkyl <span class="hlt">nitrates</span> formed from the atmospheric oxidation of hydrocarbons was measured daily during the 13-week study period (January 22 to April 22). In addition, a large number of gas chromatography/electron capture detector (GC/ECD) peaks with retention times greater than those of the hexyl <span class="hlt">nitrates</span> were also identified as species containing -ONO2 group(s), using a nitrogen specific detector. The total <span class="hlt">concentrations</span> of these organic <span class="hlt">nitrates</span> ranged from 34 to 128 parts per trillion by volume and the distribution in the dark period was found to be similar to that found for rural lower-latitude air masses. In contrast to <span class="hlt">observations</span> made at lower latitudes where alkyl <span class="hlt">nitrates</span> make a relatively small contribution to NOy, the organic <span class="hlt">nitrates</span> at Alert were found to contribute between 7 and 20% of the total odd nitrogen species. After polar sunrise the total <span class="hlt">concentrations</span> of these organic <span class="hlt">nitrates</span> decreased steadily, due primarily to the consumption of larger (>C4) alkyl <span class="hlt">nitrates</span>. The C3 alkyl <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> showed little variation during this study. During ozone depletion episodes in April there was a positive correlation between the <span class="hlt">concentration</span> of the larger organic <span class="hlt">nitrates</span> and ozone. Most surprisingly, the ratio of <span class="hlt">concentrations</span> of isomeric alkyl <span class="hlt">nitrates</span> with carbon numbers ≥5, and in particular those involving the C5 isomers, was found to show substantial variations coinciding with the O3 depletion events. This change in the isomeric alkyl <span class="hlt">nitrate</span> ratios implies a substantial chemical processing of the air masses exhibiting ozone depletion. The possible mechanisms, which must involve consumption of the organic <span class="hlt">nitrates</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19350901','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19350901"><span>Transformation of benzalkonium chloride under <span class="hlt">nitrate</span> reducing conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tezel, Ulas; Pavlostathis, Spyros G</p> <p>2009-03-01</p> <p>The effect and transformation potential of benzalkonium chlorides (BAC) under <span class="hlt">nitrate</span> reducing conditions were investigated at <span class="hlt">concentrations</span> up to 100 mg/L in batch assays using a mixed, mesophilic (35 degrees C) methanogenic culture. Glucose was used as the carbon and energy source and the initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was 70 mg N/L Dissimilatory <span class="hlt">nitrate</span> reduction to ammonia (DNRA) and to dinitrogen (DNRN) were <span class="hlt">observed</span> at BAC <span class="hlt">concentrations</span> up to 25 mg/L At and above 50 mg BAC/L, DNRA was inhibited and DNRN was incomplete resulting in accumulation of nitrous oxide. Long-term inhibition of methanogenesis and accumulation of volatile fatty acids were <span class="hlt">observed</span> at and above 50 mg BAC/L Over 99% of the added BAC was recovered from all cultures except the one amended with 100 mg BAC/L where 37% of the initially added BAC was transformed during the 100 day incubation period. Abiotic and biotic assays performed with 100 mg/L of BAC and 5 mM (in the liquid phase) of either <span class="hlt">nitrate</span>, nitrite, or nitric oxide demonstrated that BAC transformation was abiotic and followed the modified Hofmann degradation pathway, i.e., bimolecular nucleophilic substitution with nitrite. Alkyl dimethyl amines (tertiary amines) were produced at equamolar levels to BAC transformed, but were not further degraded. This is the first report demonstrating the transformation of BAC under <span class="hlt">nitrate</span> reducing conditions and elucidating the BAC transformation pathway.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/24284','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/24284"><span>Temporal and spatial trends in streamwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the San Bernardino mountains, southern California</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Mark E. Fenn; Mark A. Poth</p> <p>1999-01-01</p> <p>We report streamwater <span class="hlt">nitrate</span> (NO,) <span class="hlt">concentrations</span> for December 1995 to September 1998 from 19 sampling sites across a N deposition gradient in the San Bernardino Mountains. Streamwater NO3- <span class="hlt">concentrations</span> in Devil Canyon (DC), a high-pollution area, and in previously reported data from the San Gabriel Mountains 40 km...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005PCE....30..712S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005PCE....30..712S"><span>Investigation of processes leading to <span class="hlt">nitrate</span> enrichment in soils in the Kalahari Region, Botswana</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schwiede, M.; Duijnisveld, W. H. M.; Böttcher, J.</p> <p></p> <p>In Southern Africa elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are <span class="hlt">observed</span> in mostly uninhabited semi-arid areas. In the Kalahari of Botswana groundwater locally exhibits <span class="hlt">concentrations</span> up to 600 mg/l. It is assumed, that <span class="hlt">nitrate</span> found in the groundwater originates mainly from nitrogen input and transformations in the soils. Our investigations in the Kalahari between Serowe and Orapa show that cattle raising is an important source for enhanced <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the soils (Arenosols). But also in termite mounds very high <span class="hlt">nitrate</span> stocks were found, and under natural vegetation (acacia trees and shrubs) <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were mostly unexpectedly high. This <span class="hlt">nitrate</span> enrichment in the soils poses a serious threat to the groundwater quality. However, calculated soil water age distributions in the unsaturated zone clearly show that today’s <span class="hlt">nitrate</span> pollution of the groundwater below the investigation area could originate from natural sources, but cannot be caused by the current land use for cattle raising.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70021177','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70021177"><span>Simultaneous reduction of <span class="hlt">nitrate</span> and selenate by cell suspensions of selenium-respiring bacteria</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Oremland, R.S.; Blum, J.S.; Bindi, A.B.; Dowdle, P.R.; Herbel, M.; Stolz, J.F.</p> <p>1999-01-01</p> <p>Washed-cell suspensions of Sulfurospirillum barnesii reduced selenate [Se(VI)] when cells were cultured with <span class="hlt">nitrate</span>, thiosulfate, arsenate, or fumarate as the electron acceptor. When the <span class="hlt">concentration</span> of the electron donor was limiting, Se(VI) reduction in whole cells was approximately fourfold greater in Se(VI)-grown cells than was <span class="hlt">observed</span> in <span class="hlt">nitrate</span>-grown cells; correspondingly, <span class="hlt">nitrate</span> reduction was ~11-fold higher in <span class="hlt">nitrate</span>-grown cells than in Se(VI)-grown cells. However, a simultaneous reduction of <span class="hlt">nitrate</span> and Se(VI) was <span class="hlt">observed</span> in both cases. At nonlimiting electron donor <span class="hlt">concentrations</span>, <span class="hlt">nitrate</span>- grown cells suspended with equimolar <span class="hlt">nitrate</span> and selenate achieved a complete reductive removal of nitrogen and selenium oxyanions, with the bulk of <span class="hlt">nitrate</span> reduction preceding that of selenate reduction. Chloramphenicol did not inhibit these reductions. The Se(VI)-respiring haloalkaliphile Bacillus arsenicoselenatis gave similar results, but its Se(VI) reductase was not constitutive in <span class="hlt">nitrate</span>-grown cells. No reduction of Se(VI) was noted for Bacillus selenitireducens, which respires selenite. The results of kinetic experiments with cell membrane preparations of S. barnesii suggest the presence of constitutive selenate and <span class="hlt">nitrate</span> reduction, as well as an inducible, high- affinity <span class="hlt">nitrate</span> reductase in <span class="hlt">nitrate</span>-grown cells which also has a low affinity for selenate. The simultaneous reduction of micromolar Se(VI) in the presence of millimolar <span class="hlt">nitrate</span> indicates that these organisms may have a functional use in bioremediating <span class="hlt">nitrate</span>-rich, seleniferous agricultural wastewaters. Results with 75Se-selenate tracer show that these organisms can lower ambient Se(VI) <span class="hlt">concentrations</span> to levels in compliance with new regulations proposed for release of selenium oxyanions into the environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2006/5062/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2006/5062/"><span>Evaluation of hydrologic conditions and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the Rio Nigua de Salinas alluvial fan aquifer, Salinas, Puerto Rico, 2002-03</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Rodriguez, Jose M.</p> <p>2006-01-01</p> <p>A ground-water quality study to define the potential sources and <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> in the Rio Nigua de Salinas alluvial fan aquifer was conducted between January 2002 and March 2003. The study area covers about 3,600 hectares of the coastal plain within the municipality of Salinas in southern Puerto Rico, extending from the foothills to the Caribbean Sea. Agriculture is the principal land use and includes cultivation of diverse crops, turf grass, bioengineered crops for seed production, and commercial poultry farms. Ground-water withdrawal in the alluvial fan was estimated to be about 43,500 cubic meters per day, of which 49 percent was withdrawn for agriculture, 42 percent for public supply, and 9 percent for industrial use. Ground-water flow in the study area was primarily to the south and toward a cone of depression within the south-central part of the alluvial fan. The presence of that cone of depression and a smaller one located in the northeastern quadrant of the study area may contribute to the increase in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> within a total area of about 545 hectares by 'recycling' ground water used for irrigation of cultivated lands. In an area that covers about 405 hectares near the center of the Salinas alluvial fan, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increased from 0.9 to 6.7 milligrams per liter as nitrogen in 1986 to 8 to 12 milligrams per liter as nitrogen in 2002. Principal sources of <span class="hlt">nitrate</span> in the study area are fertilizers (used in the cultivated farmlands) and poultry farm wastes. The highest nitrogen <span class="hlt">concentrations</span> were found at poultry farms in the foothills area. In the area of disposed poultry farm wastes, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in ground water ranged from 25 to 77 milligrams per liter as nitrogen. Analyses for the stable isotope ratios of nitrogen-15/nitrogen-14 in <span class="hlt">nitrate</span> were used to distinguish the source of <span class="hlt">nitrate</span> in the coastal plain alluvial fan aquifer. Potential <span class="hlt">nitrate</span> loads from areas under cultivation were estimated for the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.A31A0062R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.A31A0062R"><span><span class="hlt">Observed</span> secondary organic aerosol (SOA) and organic <span class="hlt">nitrate</span> yields from NO3 oxidation of isoprene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rollins, A. W.; Fry, J. L.; Kiendler-Scharr, A.; Wooldridge, P. J.; Brown, S. S.; Fuchs, H.; Dube, W.; Mensah, A.; Tillmann, R.; Dorn, H.; Brauers, T.; Cohen, R. C.</p> <p>2008-12-01</p> <p>Formation of organic <span class="hlt">nitrates</span> and secondary organic aerosol (SOA) from the NO3 oxidation of isoprene has been studied at atmospheric <span class="hlt">concentrations</span> of VOC (10 ppb) and oxidant (<100 ppt NO3) in the presence of ammonium sulfate seed aerosol in the atmosphere simulation chamber SAPHIR at Forschungszentrum Jülich. Cavity Ringdown (CaRDS) and thermal dissociation - CaRDS measurements of NO3 and N2O5 as well as Thermal Dissociation - Laser Induced Fluorescence (TD-LIF) detection of alkyl <span class="hlt">nitrates</span> (RONO2) and Aerodyne Aerosol Mass Spectrometer (AMS) measurements of aerosol composition were all used in comparison to a Master Chemical Mechanism (MCM) based chemical kinetics box model to quantify the product yields from two stages in isoprene oxidation. We find significant yields of organic <span class="hlt">nitrate</span> formation from both the initial isoprene + NO3 reaction (71%) as well as from the reaction of NO3 with the initial oxidation products (30% - 60%). Under these low <span class="hlt">concentration</span> conditions (~1 μg / m3), measured SOA production was greater than instrument noise only for the second oxidation step. Based on the modeled chemistry, we estimate an SOA mass yield of 10% (relative to isoprene mass reacted) for the reaction of the initial oxidation products with NO3. This yield is found to be consistent with the estimated saturation <span class="hlt">concentration</span> (C*) of the presumed gas products of the doubly oxidized isoprene, where both oxidations lead to the addition of <span class="hlt">nitrate</span>, carbonyl, and hydroxyl groups.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AGUFM.H11C0250T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AGUFM.H11C0250T"><span><span class="hlt">Nitrate</span> and Phosphate <span class="hlt">Concentration</span> Trends in Selected Puerto Rico Rivers over the Past Four Decades--The Impact of Human Activity on Tropical Island Landscapes and Water Quality</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Troester, J. W.</p> <p>2001-12-01</p> <p>For more than four decades, the U.S. Geological Survey (USGS) has collected riverine nutrient <span class="hlt">concentration</span> data in Puerto Rico, a mountainous Caribbean tropical island. During the last forty years the population of this 9043 square km island has increased from about 2.4 to 3.8 million people. Much of the island has been developed for agriculture, and later for industry and urbanization. Data from gaging stations located within four of the larger, mixed land-use drainage basins of Puerto Rico were compiled and analyzed. The stations selected were the Rio Grande de Manati at Highway 2 (Station 50038100), Rio de la Plata at Highway 2 (Station 50046000), Rio Grande de Patillas near Patillas (Station 50092000), and Rio Grande de Anasco near San Sebastian (Station 50144000). Analytical results were compared with a shorter-term data set from smaller forested watersheds (that are part of the USGS Water, Energy, and Biogeochemical Budgets (WEBB) Program) to evaluate the impact of human activity on the water quality. During the 1960's, discharge weighted average <span class="hlt">concentrations</span> (DWAC) of dissolved <span class="hlt">nitrate</span>-nitrogen (<span class="hlt">nitrate</span>-N) ranged from 0.10 to 0.51 mg/L in the four rivers. DWAC of <span class="hlt">nitrate</span>-N increased and peaked in the 1970's and 1980's (range of 0.35 to 1.00 mg/L), and have subsequently decreased (range of 0.30 to 0.95 mg/L). DWAC of <span class="hlt">nitrate</span>-N declined, even though the average <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span> continued to increase in three of these rivers. The decrease in DWAC of <span class="hlt">nitrate</span>-N may reflect the changes in land use from the 1960's to present, which includes an increase in forest and a decrease in cropland throughout much of Puerto Rico. However, the largest decrease (from 0.77 to 0.34 mg/L) occurred in the Rio de la Plata after it was dammed in 1974. DWAC of <span class="hlt">nitrate</span>-N in the four rivers were several times higher than the total <span class="hlt">nitrate</span>-N <span class="hlt">observed</span> at gaging stations in undisturbed forested watersheds, such as at the Rio Mameyes near Sabana (Station 50065500) and the Rio</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/ds/698/pdf/ds698.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/ds/698/pdf/ds698.pdf"><span>Digital spatial data for predicted <span class="hlt">nitrate</span> and arsenic <span class="hlt">concentrations</span> in basin-fill aquifers of the Southwest Principal Aquifers study area</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>McKinney, Tim S.; Anning, David W.</p> <p>2012-01-01</p> <p>This product "Digital spatial data for predicted <span class="hlt">nitrate</span> and arsenic <span class="hlt">concentrations</span> in basin-fill aquifers of the Southwest Principal Aquifers study area" is a 1:250,000-scale vector spatial dataset developed as part of a regional Southwest Principal Aquifers (SWPA) study (Anning and others, 2012). The study examined the vulnerability of basin-fill aquifers in the southwestern United States to <span class="hlt">nitrate</span> contamination and arsenic enrichment. Statistical models were developed by using the random forest classifier algorithm to predict <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and arsenic across a model grid that represents local- and basin-scale measures of source, aquifer susceptibility, and geochemical conditions.</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('https://www.fs.usda.gov/treesearch/pubs/15744','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/15744"><span>Soil water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in giant cane and forest riparian buffer zones</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Jon E. Schoonover; Karl W. J. Williard; James J. Zaczek; Jean C. Mangun; Andrew D. Carver</p> <p>2003-01-01</p> <p>Soil water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in giant cane and forest riparian buffer zones along Cypress Creek in southern Illinois were compared to determine if the riparian zones were sources or sinks for nitrogen in the rooting zone. Suction lysimeters were used to collect soil water samples from the lower rooting zone in each of the two vegetation types. The cane riparian...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70191671','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70191671"><span>New insights into <span class="hlt">nitrate</span> dynamics in a karst groundwater system gained from in situ high-frequency optical sensor measurements</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Opsahl, Stephen P.; Musgrove, MaryLynn; Slattery, Richard N.</p> <p>2017-01-01</p> <p> within the urbanized area of the aquifer. Changes in specific conductance were <span class="hlt">observed</span> at both sites during groundwater recharge, and a significant correlation between specific conductance and <span class="hlt">nitrate</span> (correlation coefficient [R] = 0.455) was evident at the urban site where large (3-fold) changes in <span class="hlt">nitrate</span> occurred. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> and specific conductance measured during a depth profile indicated that the water column was generally homogeneous as expected for this karst environment, but changes were <span class="hlt">observed</span> in the most productive zone of the aquifer that might indicate some heterogeneity within the complex network of flow paths. Resolving the timing and magnitude of changes and characterizing fine-scale vertical differences would not be possible using conventional sampling techniques. The patterns <span class="hlt">observed</span> in situ provided new insight into the dynamic nature of <span class="hlt">nitrate</span> in a karst groundwater system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870026463&hterms=js&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Djs','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870026463&hterms=js&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Djs"><span><span class="hlt">Observation</span> of several chlorine <span class="hlt">nitrate</span> (ClONO2) bands in stratospheric infrared spectra</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zander, R.; Rinsland, C. P.; Farmer, C. B.; Brown, L. R.; Norton, R. H.</p> <p>1986-01-01</p> <p>Four of the most prominent and sharpest infrared absorption features of chlorine <span class="hlt">nitrate</span> at 780.2, 807.7, 809.4, and 1292.6/cm have been <span class="hlt">observed</span> in a series of infrared solar spectra obtained at an unapodized spectral resolution of 0.01/cm, using the Atmospheric Trace Molecule Spectroscopy instrument from on-board Sapcelab 3. A quantitative analysis of the nu4 Q branch at 780.2/cm has provided insight into the <span class="hlt">concentration</span> of ClONO2 between 19 and 40 km altitude. While the mean profile deduced from three sunset occultations near 30 deg N latitude exhibits a shape close to that predicted by model calculations, its <span class="hlt">concentrations</span> in the 20 to 32 km altitude range are, however, about 30 percent larger, reaching a peak <span class="hlt">concentration</span> of 9 x 10 to the 8th molecules/cu cm at 25 km. The <span class="hlt">concentrations</span> above 32 km, deduced from one sunrise occultation at 47 deg JS, are even larger than the corresponding sunset values at 30 deg N latitude. Some of these discrepancies may be caused by the rather large uncertainty in the assumed Q branch strength.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1994/4174/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1994/4174/report.pdf"><span>Relation of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in ground water to well depth, well use, and land use in Franklin Township, Gloucester County, New Jersey, 1970-85</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>MacLeod, Cecilia Louise; Barringer, T.H.; Vowinkel, E.F.; Price, C.V.</p> <p>1995-01-01</p> <p>A water-quality data base was developed to permit the investigation of the relation of <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> (as nitrogen) in ground water to well depth, well use, and land use (agricultural, residential, urban nonresidential, and undeveloped) in Franklin Township. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in water from 868 wells tended to decrease with depth. A rank-order regression model of <span class="hlt">nitrate</span> concen- trations and land-use percentages was fitted to data from 98 shallow domestic wells. The model, which explains about 25 percent of the variance in the data, indicated that <span class="hlt">nitrate</span> <span class="hlt">concentration</span> increased with the percentage of developed land in a well's buffer zone. Further stratification of the data based on well use (commercial, domestic, or agricultural/irrigation) indicated that elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were more common in water from agricultural/irrigation wells than in water from domestic or commercial wells. <span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> were indicative of human activities in water from about one-third of the wells sampled but exceeded the U.S. Environmental Protection Agency's maximum contaminant level of 10 milligrams per liter in water from only 1 percent of the wells. A sampling strategy in which water from wells of different depths located within areas in each of the four land-use categories is sampled yearly and analyzed for <span class="hlt">nitrate</span> and other constituents would facilitate determination of the effects of human activities on ground-water quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/15732','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/15732"><span>Assessing the extent of nitrogen saturation in northern West Virginia forested watersheds: a survey of stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span></span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Karl W. J. Williard; David R. DeWalle; Pamela J. Edwards</p> <p>2003-01-01</p> <p>Twenty-seven forested watersheds in northern West Virginia were sampled for stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> during summer 1997 and fall 1998 baseflow periods to determine if Fernow watershed 4, an often-cited and studied nitrogen saturated basin, was anomalous or regionally representative in terms of stream <span class="hlt">nitrate</span> levels. Baseflow stream NO3-N...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2757690','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2757690"><span>Aminoethyl <span class="hlt">nitrate</span> – the novel super <span class="hlt">nitrate</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>Bauersachs, Johann</p> <p>2009-01-01</p> <p>Long-term use of most organic <span class="hlt">nitrates</span> is limited by development of tolerance, induction of oxidative stress and endothelial dysfunction. In this issue of the BJP, Schuhmacher et al. characterized a novel class of organic <span class="hlt">nitrates</span> with amino moieties (aminoalkyl <span class="hlt">nitrates</span>). Aminoethyl <span class="hlt">nitrate</span> was identified as a novel organic mononitrate with high potency but devoid of induction of mitochondrial oxidative stress. Cross-tolerance to nitroglycerin or the endothelium-dependent agonist acetylcholine after in vivo treatment was not <span class="hlt">observed</span>. Like all <span class="hlt">nitrates</span>, aminoethyl <span class="hlt">nitrate</span> induced vasorelaxation by activation of soluble guanylate cyclase. Thus, in contrast to the prevailing view, high potency in an organic <span class="hlt">nitrate</span> is not necessarily accompanied by induction of oxidative stress or endothelial dysfunction. This work from Daiber's group is an important step forward in the understanding of <span class="hlt">nitrate</span> bioactivation, tolerance phenomena and towards the development of better organic <span class="hlt">nitrates</span> for clinical use. PMID:19732062</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22279888','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22279888"><span>[Assessment of shallow groundwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in typical terrestrial ecosystems of Chinese Ecosystem Research Network (CERN) during 2004-2009].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Zhi-Wei; Zhang, Xin-Yu; Sun, Xiao-Min; Yuan, Guo-Fu; Wang, Sheng-Zhong; Liu, Wen-Hua</p> <p>2011-10-01</p> <p>The <span class="hlt">nitrate</span>-N (NO3(-) -N) <span class="hlt">concentrations</span> of 38 shallow groundwater wells from 31 of the typical terrestrial ecosystems on Chinese Ecosystem Research Network (CERN) were assessed using the monitoring data from 2004 to 2009. The results showed that the average values of NO3(-) -N <span class="hlt">concentrations</span> were significantly higher in the agricultural (4.85 mg x L(-1) +/- 0.42 mg x L(-1)), desert (oasis) (3.72 mg x L(-1) +/- 0.42 mg x L(-1)) and urban ecosystems (3.77 mg x L(-1) 0.51 mg x L(-1)) than in the grass (1.59 mg x L(-1) +/- 0.35 mg L(-1)) and forest ecosystems (0.39 mg x L(-1) +/- 0.03 mg x L(-1)). <span class="hlt">Nitrate</span> was the major form of nitrogen, with between 56% to 88% of nitrogen in the <span class="hlt">nitrate</span>-N form in the shallow groundwater of desert (oasis), urban and agricultural ecosystems. <span class="hlt">Nitrate</span>-N <span class="hlt">concentrations</span> for some agricultural ecosystems (Ansai, Yanting, Yucheng) and desert (oasis) ecosystems (Cele, Linze, Akesu) analysis exceeded the 10 mg x L(-1) World Health Organization drinking water standards between 14.3% and 84.6%. Significant seasonality was found in Ansai, Fengqiu, Yanting agricultural ecosystems and the Beijing urban ecosystem using the relatively high frequency monitoring data, with the higher <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> usually found during summer and winter months. The monitoring results indicated that the shallow groundwater of agricultural ecosystems was contaminated by agricultural management practices, i.e. fertilization, while the shallow groundwater of forest ecosystems was under natural condition with no contamination from human activities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70185475','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70185475"><span>Modeling biotic uptake by periphyton and transient hyporrheic storage of <span class="hlt">nitrate</span> in a natural stream</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kim, Brian K.A.; Jackman, Alan P.; Triska, Frank J.</p> <p>1992-01-01</p> <p>To a convection-dispersion hydrologic transport model we coupled a transient storage submodel (Bencala, 1984) and a biotic uptake submodel based on Michaelis-Menten kinetics (Kim et al., 1990). Our purpose was threefold: (1) to simulate <span class="hlt">nitrate</span> retention in response to change in load in a third-order stream, (2) to differentiate biotic versus hydrologie factors in <span class="hlt">nitrate</span> retention, and (3) to produce a research tool whose properties are consistent with laboratory and field <span class="hlt">observations</span>. Hydrodynamic parameters were fitted from chloride <span class="hlt">concentration</span> during a 20-day chloride-<span class="hlt">nitrate</span> coinjection (Bencala, 1984), and biotic uptake kinetics were based on flume studies by Kim et al. (1990) and Triska et al. (1983). <span class="hlt">Nitrate</span> <span class="hlt">concentration</span> from the 20-day coinjection experiment served as a base for model validation. The complete transport retention model reasonably predicted the <span class="hlt">observed</span> <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. However, simulations which lacked either the transient storage submodel or the biotic uptake submodel poorly predicted the <span class="hlt">observed</span> <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. Model simulations indicated that transient storage in channel and hyporrheic interstices dominated <span class="hlt">nitrate</span> retention within the first 24 hours, whereas biotic uptake dominated thereafter. A sawtooth function for Vmax ranging from 0.10 to 0.17 μg NO3-N s−1 gAFDM−1 (grams ash free dry mass) slightly underpredicted <span class="hlt">nitrate</span> retention in simulations of 2–7 days. This result was reasonable since uptake by other <span class="hlt">nitrate</span>-demanding processes were not included. The model demonstrated how ecosystem retention is an interaction between physical and biotic processes and supports the validity of coupling separate hydrodynamic and reactive submodels to established solute transport models in biological studies of fluvial ecosystems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016HESS...20.2353O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016HESS...20.2353O"><span>A meta-analysis and statistical modelling of <span class="hlt">nitrates</span> in groundwater at the African scale</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ouedraogo, Issoufou; Vanclooster, Marnik</p> <p>2016-06-01</p> <p>Contamination of groundwater with <span class="hlt">nitrate</span> poses a major health risk to millions of people around Africa. Assessing the space-time distribution of this contamination, as well as understanding the factors that explain this contamination, is important for managing sustainable drinking water at the regional scale. This study aims to assess the variables that contribute to <span class="hlt">nitrate</span> pollution in groundwater at the African scale by statistical modelling. We compiled a literature database of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater (around 250 studies) and combined it with digital maps of physical attributes such as soil, geology, climate, hydrogeology, and anthropogenic data for statistical model development. The maximum, medium, and minimum <span class="hlt">observed</span> <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were analysed. In total, 13 explanatory variables were screened to explain <span class="hlt">observed</span> <span class="hlt">nitrate</span> pollution in groundwater. For the mean <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, four variables are retained in the statistical explanatory model: (1) depth to groundwater (shallow groundwater, typically < 50 m); (2) recharge rate; (3) aquifer type; and (4) population density. The first three variables represent intrinsic vulnerability of groundwater systems to pollution, while the latter variable is a proxy for anthropogenic pollution pressure. The model explains 65 % of the variation of mean <span class="hlt">nitrate</span> contamination in groundwater at the African scale. Using the same proxy information, we could develop a statistical model for the maximum <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> that explains 42 % of the <span class="hlt">nitrate</span> variation. For the maximum <span class="hlt">concentrations</span>, other environmental attributes such as soil type, slope, rainfall, climate class, and region type improve the prediction of maximum <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> at the African scale. As to minimal <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, in the absence of normal distribution assumptions of the data set, we do not develop a statistical model for these data. The data-based statistical model presented here represents an important</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApWS....7...71E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApWS....7...71E"><span>Occurrence of <span class="hlt">nitrate</span> in Tanzanian groundwater aquifers: 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>Elisante, Eliapenda; Muzuka, Alfred N. N.</p> <p>2017-03-01</p> <p>More than 25 % of Tanzanian depends on groundwater as the main source of water for drinking, irrigation and industrial activities. The current trend of land use may lead to groundwater contamination and thus increasing risks associated with the usage of contaminated water. <span class="hlt">Nitrate</span> is one of the contaminants resulting largely from anthropogenic activities that may find its way to the aquifers and thus threatening the quality of groundwater. Elevated levels of <span class="hlt">nitrate</span> in groundwater may lead to human health and environmental problems. The current trend of land use in Tanzania associated with high population growth, poor sanitation facilities and fertilizer usage may lead to <span class="hlt">nitrate</span> contamination of groundwater. This paper therefore aimed at providing an overview of to what extent human activities have altered the <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> in groundwater aquifers in Tanzania. The <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> in Tanzanian groundwater is variable with highest values <span class="hlt">observable</span> in Dar es Salaam (up to 477.6 mg/l), Dodoma (up to 441.1 mg/l), Tanga (above 100 mg/l) and Manyara (180 mg/l). Such high values can be attributed to various human activities including onsite sanitation in urban centres and agricultural activities in rural areas. Furthermore, there are some signs of increasing <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> in groundwater with time in some areas in response to increased human activities. However, reports on levels and trends of <span class="hlt">nitrate</span> in groundwater in many regions of the country are lacking. For Tanzania to appropriately address the issue of groundwater contamination, a deliberate move to determine <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater is required, as well as protection of recharge basins and improvement of onsite sanitation systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ACP....18.5293L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ACP....18.5293L"><span><span class="hlt">Nitrate</span>-driven urban haze pollution during summertime over the North China Plain</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Haiyan; Zhang, Qiang; Zheng, Bo; Chen, Chunrong; Wu, Nana; Guo, Hongyu; Zhang, Yuxuan; Zheng, Yixuan; Li, Xin; He, Kebin</p> <p>2018-04-01</p> <p>Compared to the severe winter haze episodes in the North China Plain (NCP), haze pollution during summertime has drawn little public attention. In this study, we present the highly time-resolved chemical composition of submicron particles (PM1) measured in Beijing and Xinxiang in the NCP region during summertime to evaluate the driving factors of aerosol pollution. During the campaign periods (30 June to 27 July 2015, for Beijing and 8 to 25 June 2017, for Xinxiang), the average PM1 <span class="hlt">concentrations</span> were 35.0 and 64.2 µg m-3 in Beijing and Xinxiang. Pollution episodes characterized with largely enhanced <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were <span class="hlt">observed</span> at both sites. In contrast to the slightly decreased mass fractions of sulfate, semivolatile oxygenated organic aerosol (SV-OOA), and low-volatility oxygenated organic aerosol (LV-OOA) in PM1, <span class="hlt">nitrate</span> displayed a significantly enhanced contribution with the aggravation of aerosol pollution, highlighting the importance of <span class="hlt">nitrate</span> formation as the driving force of haze evolution in summer. Rapid <span class="hlt">nitrate</span> production mainly occurred after midnight, with a higher formation rate than that of sulfate, SV-OOA, or LV-OOA. Based on <span class="hlt">observation</span> measurements and thermodynamic modeling, high ammonia emissions in the NCP region favored the high <span class="hlt">nitrate</span> production in summer. Nighttime <span class="hlt">nitrate</span> formation through heterogeneous hydrolysis of dinitrogen pentoxide (N2O5) enhanced with the development of haze pollution. In addition, air masses from surrounding polluted areas during haze episodes led to more <span class="hlt">nitrate</span> production. Finally, atmospheric particulate <span class="hlt">nitrate</span> data acquired by mass spectrometric techniques from various field campaigns in Asia, Europe, and North America uncovered a higher <span class="hlt">concentration</span> and higher fraction of <span class="hlt">nitrate</span> present in China. Although measurements in Beijing during different years demonstrate a decline in the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in recent years, the <span class="hlt">nitrate</span> contribution in PM1 still remains high. To effectively alleviate</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.water.usgs.gov/wri004110/','USGSPUBS'); return false;" href="http://pubs.water.usgs.gov/wri004110/"><span>Estimating the probability of elevated <span class="hlt">nitrate</span> (NO2+NO3-N) <span class="hlt">concentrations</span> in ground water in the Columbia Basin Ground Water Management Area, Washington</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Frans, Lonna M.</p> <p>2000-01-01</p> <p>Logistic regression was used to relate anthropogenic (man-made) and natural factors to the occurrence of elevated <span class="hlt">concentrations</span> of nitrite plus <span class="hlt">nitrate</span> as nitrogen in ground water in the Columbia Basin Ground Water Management Area, eastern Washington. Variables that were analyzed included well depth, depth of well casing, ground-water recharge rates, presence of canals, fertilizer application amounts, soils, surficial geology, and land-use types. The variables that best explain the occurrence of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above 3 milligrams per liter in wells were the amount of fertilizer applied annually within a 2-kilometer radius of a well and the depth of the well casing; the variables that best explain the occurrence of <span class="hlt">nitrate</span> above 10 milligrams per liter included the amount of fertilizer applied annually within a 3-kilometer radius of a well, the depth of the well casing, and the mean soil hydrologic group, which is a measure of soil infiltration rate. Based on the relations between these variables and elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, models were developed using logistic regression that predict the probability that ground water will exceed a <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of either 3 milligrams per liter or 10 milligrams per liter. Maps were produced that illustrate the predicted probability that ground-water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> will exceed 3 milligrams per liter or 10 milligrams per liter for wells cased to 78 feet below land surface (median casing depth) and the predicted depth to which wells would need to be cased in order to have an 80-percent probability of drawing water with a <span class="hlt">nitrate</span> <span class="hlt">concentration</span> below either 3 milligrams per liter or 10 milligrams per liter. Maps showing the predicted probability for the occurrence of elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> indicate that the irrigated agricultural regions are most at risk. The predicted depths to which wells need to be cased in order to have an 80-percent chance of obtaining low <span class="hlt">nitrate</span> ground water exceed 600 feet</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=332780','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=332780"><span>Effects of substrate type on plant growth and nitrogen and <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in spinach</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 effects of three commercial substrates (a mixture of forest residues, composted grape husks, and white peat; black peat; and coir) on plant growth and nitrogen (N) and <span class="hlt">nitrate</span> (NO3) <span class="hlt">concentration</span> and content were evaluated in spinach (Spinacia oleracea L. cv. Tapir). Spinach seedlings were trans...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26134447','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26134447"><span><span class="hlt">Nitrate</span> removal from high strength <span class="hlt">nitrate</span>-bearing wastes in granular sludge sequencing batch reactors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Krishna Mohan, Tulasi Venkata; Renu, Kadali; Nancharaiah, Yarlagadda Venkata; Satya Sai, Pedapati Murali; Venugopalan, Vayalam Purath</p> <p>2016-02-01</p> <p>A 6-L sequencing batch reactor (SBR) was operated for development of granular sludge capable of denitrification of high strength <span class="hlt">nitrates</span>. Complete and stable denitrification of up to 5420 mg L(-1) <span class="hlt">nitrate</span>-N (2710 mg L(-1) <span class="hlt">nitrate</span>-N in reactor) was achieved by feeding simulated <span class="hlt">nitrate</span> waste at a C/N ratio of 3. Compact and dense denitrifying granular sludge with relatively stable microbial community was developed during reactor operation. Accumulation of large amounts of nitrite due to incomplete denitrification occurred when the SBR was fed with 5420 mg L(-1) NO3-N at a C/N ratio of 2. Complete denitrification could not be achieved at this C/N ratio, even after one week of reactor operation as the nitrite levels continued to accumulate. In order to improve denitrification performance, the reactor was fed with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> of 1354 mg L(-1), while keeping C/N ratio at 2. Subsequently, <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the feed was increased in a step-wise manner to establish complete denitrification of 5420 mg L(-1) NO3-N at a C/N ratio of 2. The results show that substrate <span class="hlt">concentration</span> plays an important role in denitrification of high strength <span class="hlt">nitrate</span> by influencing nitrite accumulation. Complete denitrification of high strength <span class="hlt">nitrates</span> can be achieved at lower substrate <span class="hlt">concentrations</span>, by an appropriate acclimatization strategy. Copyright © 2015 The Society for Biotechnology, Japan. Published by Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2011/5018/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2011/5018/"><span>Recent (2008-10) <span class="hlt">concentrations</span> and isotopic compositions of <span class="hlt">nitrate</span> and <span class="hlt">concentrations</span> of wastewater compounds in the Barton Springs zone, south-central Texas, and their potential relation to urban development in the contributing zone</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Mahler, Barbara J.; Musgrove, MaryLynn; Herrington, Chris; Sample, Thomas L.</p> <p>2011-01-01</p> <p>During 2008–10, the U.S. Geological Survey, in cooperation with the City of Austin, the City of Dripping Springs, the Barton Springs/Edwards Aquifer Conservation District, the Lower Colorado River Authority, Hays County, and Travis County, collected and analyzed water samples from five streams (Barton, Williamson, Slaughter, Bear, and Onion Creeks), two groundwater wells (Marbridge well [YD–58–50–704] and Buda well [LR–58–58–403]), and the main orifice of Barton Springs in Austin, Texas, with the objective of characterizing <span class="hlt">concentrations</span> and isotopic compositions of <span class="hlt">nitrate</span> and <span class="hlt">concentrations</span> of wastewater compounds in the Barton Springs zone. The Barton Springs zone is in south-central Texas, an area undergoing rapid growth in population and in land area affected by development, with associated increases in wastewater generation. Over a period of 17 months, during which the hydrologic conditions transitioned from dry to wet, samples were collected routinely from the streams, wells, and spring and, in response to storms, from the streams and spring; some or all samples were analyzed for <span class="hlt">nitrate</span>, nitrogen and oxygen isotopes of <span class="hlt">nitrate</span>, and waste­water compounds. The median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in routine samples from all sites were higher in samples collected during the wet period than in samples collected during the dry period, with the greatest difference for stream samples (0.05 milligram per liter during the dry period to 0.96 milligram per liter for the wet period). <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in recent (2008–10) samples were elevated relative to <span class="hlt">concentrations</span> in historical (1990–2008) samples from streams and from Barton Springs under medium- and high-flow conditions. Recent <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were higher than historical <span class="hlt">concentrations</span> at the Marbridge well but the reverse was true at the Buda well. The elevated <span class="hlt">concentrations</span> likely are related to the cessation of dry conditions coupled with increased nitrogen loading in the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24288757','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24288757"><span>Effect of cation type and <span class="hlt">concentration</span> of <span class="hlt">nitrates</span> on neurological disorders during experimental cerebral ischemia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kuzenkov, V S; Krushinskii, A L; Reutov, V P</p> <p>2013-10-01</p> <p>Experiments were performed on the model of ischemic stroke due to bilateral occlusion of the carotid arteries. <span class="hlt">Nitrates</span> had various effects on the dynamics of neurological disorders and mortality rate of Wistar rats, which depended on the cation type and <span class="hlt">concentration</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26495144','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26495144"><span>Effect of Nitrite and <span class="hlt">Nitrate</span> <span class="hlt">Concentrations</span> on the Performance of AFB-MFC Enriched with High-Strength Synthetic Wastewater.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huang, Jian-Sheng; Yang, Ping; Li, Chong-Ming; Guo, Yong; Lai, Bo; Wang, Ye; Feng, Li; Zhang, Yun</p> <p>2015-01-01</p> <p>In order to study the effect of nitrite and <span class="hlt">nitrate</span> on the performance of microbial fuel cell, a system combining an anaerobic fluidized bed (AFB) and a microbial fuel cell (MFC) was employed for high-strength nitrogen-containing synthetic wastewater treatment. Before this study, the AFB-MFC had been used to treat high-strength organic wastewater for about one year in a continuous flow mode. The results showed that when the <span class="hlt">concentrations</span> of nitrite nitrogen and <span class="hlt">nitrate</span> nitrogen were increased from 1700 mg/L to 4045 mg/L and 545 mg/L to 1427 mg/L, respectively, the nitrite nitrogen and <span class="hlt">nitrate</span> nitrogen removal efficiencies were both above 99%; the COD removal efficiency went up from 60.00% to 88.95%; the voltage was about 375 ± 15 mV while the power density was at 70 ± 5 mW/m(2). However, when the <span class="hlt">concentrations</span> of nitrite nitrogen and <span class="hlt">nitrate</span> nitrogen were above 4045 mg/L and 1427 mg/L, respectively, the removal of nitrite nitrogen, <span class="hlt">nitrate</span> nitrogen, COD, voltage, and power density were decreased to be 86%, 88%, 77%, 180 mV, and 17 mW/m(2) when nitrite nitrogen and <span class="hlt">nitrate</span> nitrogen were increased to 4265 mg/L and 1661 mg/L. In addition, the composition of biogas generated in the anode chamber was analyzed by a gas chromatograph. Nitrogen gas, methane, and carbon dioxide were obtained. The results indicated that denitrification happened in anode chamber.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ACP....17.8635S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ACP....17.8635S"><span><span class="hlt">Nitrate</span> radical oxidation of γ-terpinene: hydroxy <span class="hlt">nitrate</span>, total organic <span class="hlt">nitrate</span>, and secondary organic aerosol yields</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Slade, Jonathan H.; de Perre, Chloé; Lee, Linda; Shepson, Paul B.</p> <p>2017-07-01</p> <p>Polyolefinic monoterpenes represent a potentially important but understudied source of organic <span class="hlt">nitrates</span> (ONs) and secondary organic aerosol (SOA) following oxidation due to their high reactivity and propensity for multi-stage chemistry. Recent modeling work suggests that the oxidation of polyolefinic γ-terpinene can be the dominant source of nighttime ON in a mixed forest environment. However, the ON yields, aerosol partitioning behavior, and SOA yields from γ-terpinene oxidation by the <span class="hlt">nitrate</span> radical (NO3), an important nighttime oxidant, have not been determined experimentally. In this work, we present a comprehensive experimental investigation of the total (gas + particle) ON, hydroxy <span class="hlt">nitrate</span>, and SOA yields following γ-terpinene oxidation by NO3. Under dry conditions, the hydroxy <span class="hlt">nitrate</span> yield = 4(+1/-3) %, total ON yield = 14(+3/-2) %, and SOA yield ≤ 10 % under atmospherically relevant particle mass loadings, similar to those for α-pinene + NO3. Using a chemical box model, we show that the measured <span class="hlt">concentrations</span> of NO2 and γ-terpinene hydroxy <span class="hlt">nitrates</span> can be reliably simulated from α-pinene + NO3 chemistry. This suggests that NO3 addition to either of the two internal double bonds of γ-terpinene primarily decomposes forming a relatively volatile keto-aldehyde, reconciling the small SOA yield <span class="hlt">observed</span> here and for other internal olefinic terpenes. Based on aerosol partitioning analysis and identification of speciated particle-phase ON applying high-resolution liquid chromatography-mass spectrometry, we estimate that a significant fraction of the particle-phase ON has the hydroxy <span class="hlt">nitrate</span> moiety. This work greatly contributes to our understanding of ON and SOA formation from polyolefin monoterpene oxidation, which could be important in the northern continental US and the Midwest, where polyolefinic monoterpene emissions are greatest.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H53G1556L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H53G1556L"><span><span class="hlt">Nitrate</span> <span class="hlt">concentration</span>-drainage flow (C-Q) relationship for a drained agricultural field in Eastern North Carolina Plain</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, W.; Youssef, M.; Birgand, F.; Chescheir, G. M.; Maxwell, B.; Tian, S.</p> <p>2017-12-01</p> <p>Agricultural drainage is a practice used to artificially enhance drainage characteristics of naturally poorly drained soils via subsurface drain tubing or open-ditch systems. Approximately 25% of the U.S. agricultural land requires improved drainage for economic crop production. However, drainage increases the transport of dissolved agricultural chemicals, particularly <span class="hlt">nitrates</span> to downstream surface waters. Nutrient export from artificially drained agricultural landscapes has been identified as the leading source of elevated nutrient levels in major surface water bodies in the U.S. Controlled drainage has long been practiced to reduce nitrogen export from agricultural fields to downstream receiving waters. It has been hypothesized that controlled drainage reduces nitrogen losses by promoting denitrification, reducing drainage outflow from the field, and increasing plant uptake. The documented performance of the practice was widely variable as it depends on several site-specific factors. The goal of this research was to utilize high frequency measurements to investigate the effect of agricultural drainage and related management practices on <span class="hlt">nitrate</span> fate and transport for an artificially drained agricultural field in eastern North Carolina. We deployed a field spectrophotometer to measure <span class="hlt">nitrate</span> <span class="hlt">concentration</span> every 45 minutes and measured drainage flow rate using a V-notch weir every 15 minutes. Furthermore, we measured groundwater level, precipitation, irrigation amount, temperature to characterize antecedent conditions for each event. <span class="hlt">Nitrate</span> <span class="hlt">concentration</span>-drainage flow (C-Q) relationships generated from the high frequency measurements illustrated anti-clockwise hysteresis loops and <span class="hlt">nitrate</span> flushing mechanism in response to most precipitation and irrigation events. Statistical evaluation will be carried out for the C-Q relationships. The results of our analysis, combined with numerical modeling, will provide a better understanding of hydrological and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5386202','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5386202"><span>Disruption of the rice <span class="hlt">nitrate</span> transporter OsNPF2.2 hinders root-to-shoot <span class="hlt">nitrate</span> transport and vascular development</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Li, Yuge; Ouyang, Jie; Wang, Ya-Yun; Hu, Rui; Xia, Kuaifei; Duan, Jun; Wang, Yaqin; Tsay, Yi-Fang; Zhang, Mingyong</p> <p>2015-01-01</p> <p>Plants have evolved to express some members of the <span class="hlt">nitrate</span> transporter 1/peptide transporter family (NPF) to uptake and transport <span class="hlt">nitrate</span>. However, little is known of the physiological and functional roles of this family in rice (Oryza sativa L.). Here, we characterized the vascular specific transporter OsNPF2.2. Functional analysis using cDNA-injected Xenopus laevis oocytes revealed that OsNPF2.2 is a low-affinity, pH-dependent <span class="hlt">nitrate</span> transporter. Use of a green fluorescent protein tagged OsNPF2.2 showed that the transporter is located in the plasma membrane in the rice protoplast. Expression analysis showed that OsNPF2.2 is <span class="hlt">nitrate</span> inducible and is mainly expressed in parenchyma cells around the xylem. Disruption of OsNPF2.2 increased <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the shoot xylem exudate when <span class="hlt">nitrate</span> was supplied after a deprivation period; this result suggests that OsNPF2.2 may participate in unloading <span class="hlt">nitrate</span> from the xylem. Under steady-state <span class="hlt">nitrate</span> supply, the osnpf2.2 mutants maintained high levels of <span class="hlt">nitrate</span> in the roots and low shoot:root <span class="hlt">nitrate</span> ratios; this <span class="hlt">observation</span> suggests that OsNPF2.2 is involved in root-to-shoot <span class="hlt">nitrate</span> transport. Mutation of OsNPF2.2 also caused abnormal vasculature and retarded plant growth and development. Our findings demonstrate that OsNPF2.2 can unload <span class="hlt">nitrate</span> from the xylem to affect the root-to-shoot <span class="hlt">nitrate</span> transport and plant development. PMID:25923512</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://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1586190','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1586190"><span>Does the evidence about health risks associated with <span class="hlt">nitrate</span> ingestion warrant an increase of the <span class="hlt">nitrate</span> standard for drinking water?</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 Grinsven, Hans JM; Ward, Mary H; Benjamin, Nigel; de Kok, Theo M</p> <p>2006-01-01</p> <p>Several authors have suggested that it is safe to raise the health standard for <span class="hlt">nitrate</span> in drinking water, and save money on measures associated with <span class="hlt">nitrate</span> pollution of drinking water resources. The major argument has been that the epidemiologic evidence for acute and chronic health effects related to drinking water <span class="hlt">nitrate</span> at <span class="hlt">concentrations</span> near the health standard is inconclusive. With respect to the chronic effects, the argument was motivated by the absence of evidence for adverse health effects related to ingestion of <span class="hlt">nitrate</span> from dietary sources. An interdisciplinary discussion of these arguments led to three important <span class="hlt">observations</span>. First, there have been only a few well-designed epidemiologic studies that evaluated ingestion of <span class="hlt">nitrate</span> in drinking water and risk of specific cancers or adverse reproductive outcomes among potentially susceptible subgroups likely to have elevated endogenous nitrosation. Positive associations have been <span class="hlt">observed</span> for some but not all health outcomes evaluated. Second, the epidemiologic studies of cancer do not support an association between ingestion of dietary <span class="hlt">nitrate</span> (vegetables) and an increased risk of cancer, because intake of dietary <span class="hlt">nitrate</span> is associated with intake of antioxidants and other beneficial phytochemicals. Third, 2–3 % of the population in Western Europe and the US could be exposed to <span class="hlt">nitrate</span> levels in drinking water exceeding the WHO standard of 50 mg/l <span class="hlt">nitrate</span>, particularly those living in rural areas. The health losses due to this exposure cannot be estimated. Therefore, we conclude that it is not possible to weigh the costs and benefits from changing the <span class="hlt">nitrate</span> standard for drinking water and groundwater resources by considering the potential consequences for human health and by considering the potential savings due to reduced costs for <span class="hlt">nitrate</span> removal and prevention of <span class="hlt">nitrate</span> pollution. PMID:16989661</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.water.usgs.gov/wri004219/','USGSPUBS'); return false;" href="http://pubs.water.usgs.gov/wri004219/"><span>Monthly variability and possible sources of <span class="hlt">nitrate</span> in ground water beneath mixed agricultural land use, Suwannee and Lafayette Counties, Florida</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Katz, Brian G.; Böhlke, J.K.</p> <p>2000-01-01</p> <p>In an area of mixed agricultural land use in Suwannee and Lafayette Counties of northern Florida, water samples were collected monthly from 14 wells tapping the Upper Floridan aquifer during July 1998 through June 1999 to assess hydrologic and land-use factors affecting the variability in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in ground water. Unusually high amounts of rainfall in September and October 1998 (43.5 centimeters total for both months) resulted in an increase in water levels in all wells in October 1998. This was followed by unusually low amounts of rainfall during November 1998 through May 1999, when rainfall was 40.7 centimeters below 30-year mean monthly values. The presence of karst features (sinkholes, springs, solution conduits) and the highly permeable sands that overlie the Upper Floridan aquifer provide for rapid movement of water containing elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> to the aquifer. <span class="hlt">Nitrate</span> was the dominant form of nitrogen in ground water collected at all sites and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> ranged from less than 0.02 to 22 milligrams per liter (mg/L), as nitrogen. Water samples from most wells showed substantial monthly or seasonal fluctuations in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Generally, water samples from wells with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> higher than 10 mg/L showed the greatest amount of monthly fluctuation. For example, water samples from six of eight wells had monthly <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> that varied by at least 5 mg/L during the study period. Water from most wells with lower <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> (less than 6 mg/L) also showed large monthly fluctuations. For instance, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in water from four sites showed monthly variations of more than 50 percent. Large fluctuations in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> likely result from seasonal agricultural practices (fertilizer application and animal waste spreading) at a particular site. For example, an increase in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> <span class="hlt">observed</span> in water samples from seven sites in February or March 1999 most</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.A13F..01C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.A13F..01C"><span>Laboratory Measurements of Isoprene-Derived <span class="hlt">Nitrates</span> Using TD-LIF</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cohen, R.; Perring, A.; Wooldridge, P.; Shepson, P.; Lockwood, A.; Hill, K.; Moffat, C.; Mielke, L.; Cavender, A.; Stevens, P.; Dusanter, S.; Vimal, D.; Wisthaler, A.; Graus, M.</p> <p>2006-12-01</p> <p>Isoprene represents the largest flux of reactive non-methane hydrocarbon to the atmosphere and the production of isoprene-derived <span class="hlt">nitrates</span> is currently one of the major controversies in nitrogen oxide chemistry. Alkyl and multifunctional <span class="hlt">nitrates</span> (ΣANs), measured by Thermal Dissociation Laser Induced Fluorescence (TD-LIF), have been <span class="hlt">observed</span> as a significant NOy component during many ground-based and airborne field experiments. A strong hypothesis is that many of these <span class="hlt">nitrates</span>, especially in forest- impacted environments, are isoprene-derived. We present smog chamber measurements (made at Purdue University in June of 2006) of ΣANs, produced through both NO3 and OH-initiated oxidation of isoprene. Isoprene, OH, HO2, NO, NO2, NOy, PAN, HNO3 and speciated first generation isoprene <span class="hlt">nitrates</span> were also measured simultaneously and chamber chemistry was subsequently modeled. We compare these measurements with previous measurements of isoprene <span class="hlt">nitrate</span> yields and examine the relative contribution of secondary <span class="hlt">nitrates</span> to the measured total organic <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20391599','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20391599"><span>Contribution of atmospheric <span class="hlt">nitrate</span> to stream-water <span class="hlt">nitrate</span> in Japanese coniferous forests revealed by the oxygen isotope ratio of <span class="hlt">nitrate</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tobari, Y; Koba, K; Fukushima, K; Tokuchi, N; Ohte, N; Tateno, R; Toyoda, S; Yoshioka, T; Yoshida, N</p> <p>2010-05-15</p> <p>Evaluation of the openness of the nitrogen (N) cycle in forest ecosystems is important in efforts to improve forest management because the N supply often limits primary production. The use of the oxygen isotope ratio (delta(18)O) of <span class="hlt">nitrate</span> is a promising approach to determine how effectively atmospheric <span class="hlt">nitrate</span> can be retained in a forest ecosystem. We investigated the delta(18)O of <span class="hlt">nitrate</span> in stream water in order to estimate the contribution of atmospheric NO(3) (-) in stream-water NO(3) (-) (f(atm)) from 26 watersheds with different stand ages (1-87 years) in Japan. The stream-water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were high in young forests whereas, in contrast, old forests discharged low-<span class="hlt">nitrate</span> stream water. These results implied a low f(atm) and a closed N cycle in older forests. However, the delta(18)O values of <span class="hlt">nitrate</span> in stream water revealed that f(atm) values were higher in older forests than in younger forests. These results indicated that even in old forests, where the discharged N loss was small, atmospheric <span class="hlt">nitrate</span> was not retained effectively. The steep slopes of the studied watersheds (>40 degrees ) which hinder the capturing of atmospheric <span class="hlt">nitrate</span> by plants and microbes might be responsible for the inefficient utilization of atmospheric <span class="hlt">nitrate</span>. Moreover, the unprocessed fraction of atmospheric <span class="hlt">nitrate</span> in the stream-water <span class="hlt">nitrate</span> in the forest (f(unprocessed)) was high in the young forest (78%), although f(unprocessed) was stable and low for other forests (5-13%). This high f(unprocessed) of the young forest indicated that the young forest retained neither atmospheric NO(3) (-) nor soil NO(3) (-) effectively, engendering high stream-water NO(3) (-) <span class="hlt">concentrations</span>. Copyright (c) 2010 John Wiley & Sons, Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/868323','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/868323"><span>Thermochemical <span class="hlt">nitrate</span> destruction</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Cox, John L.; Hallen, Richard T.; Lilga, Michael A.</p> <p>1992-01-01</p> <p>A method is disclosed for denitrification of <span class="hlt">nitrates</span> and <span class="hlt">nitrates</span> present in aqueous waste streams. The method comprises the steps of (1) identifying the <span class="hlt">concentration</span> <span class="hlt">nitrates</span> and nitrites present in a waste stream, (2) causing formate to be present in the waste stream, (3) heating the mixture to a predetermined reaction temperature from about 200.degree. C. to about 600.degree. C., and (4) holding the mixture and accumulating products at heated and pressurized conditions for a residence time, thereby resulting in nitrogen and carbon dioxide gas, and hydroxides, and reducing the level of <span class="hlt">nitrates</span> and nitrites to below drinking water standards.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19920000492&hterms=wheat&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dwheat','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920000492&hterms=wheat&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dwheat"><span>Efflux Of <span class="hlt">Nitrate</span> From Hydroponically Grown Wheat</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Huffaker, R. C.; Aslam, M.; Ward, M. R.</p> <p>1992-01-01</p> <p>Report describes experiments to measure influx, and efflux of <span class="hlt">nitrate</span> from hydroponically grown wheat seedlings. Ratio between efflux and influx greater in darkness than in light; increased with <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> in nutrient solution. On basis of experiments, authors suggest nutrient solution optimized at lowest possible <span class="hlt">concentration</span> of <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.H53H1512G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.H53H1512G"><span>Spatial and temporal variability of <span class="hlt">nitrate</span> and nitrous oxide <span class="hlt">concentrations</span> in the unsaturated zone at a corn field in the US Midwest</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gopalakrishnan, G.; Negri, C.</p> <p>2011-12-01</p> <p>There has been a significant increase in reactive nitrogen in the environment as a result of human activity. Reactive nitrogen of anthropogenic origin now equals that derived from natural terrestrial nitrogen fixation and is expected to exceed it by the end of the decade. Nitrogen is applied to crops as fertilizer and impacts the environment through water quality impairments (mostly as <span class="hlt">nitrate</span>) and as greenhouse gas emissions (as nitrous oxide). Research on environmental impacts resulting from nitrogen application in the form of fertilizers has focused disproportionately on the degradation of water quality from agricultural non-point sources. The impacts of this degradation are registered both locally, with runoff and percolation of agrochemicals into local surface water and groundwater, and on a larger scale, such as the increase in the anoxic zone in the Gulf of Mexico attributed to <span class="hlt">nitrate</span> from the Mississippi River. Impacts to the global climate from increased production of nitrous oxide as a result of increased fertilization are equally significant. Nitrous oxide is a greenhouse gas with a warming potential that is approximately 300 times greater than carbon dioxide. Direct emissions of nitrous oxide from the soil have been expressed as 1% of the applied nitrogen. Indirect emissions due to runoff, leaching and volatilization of the nitrogen from the field have been expressed as 0.75% of the applied nitrogen. Many studies have focused on processes governing nitrogen fluxes in the soil, surface water and groundwater systems. However, research on the biogeochemical processes regulating nitrogen fluxes in the unsaturated zone and consequent impacts on <span class="hlt">nitrate</span> and nitrous oxide <span class="hlt">concentrations</span> in groundwater are lacking. Our study explores the spatial and temporal variability of <span class="hlt">nitrate</span> and nitrous oxide <span class="hlt">concentrations</span> in the vadose zone at a 15 acre corn field in the US Midwest and links it to the <span class="hlt">concentrations</span> found in the groundwater at the field site. Results</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1092268','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1092268"><span><span class="hlt">Nitrate</span> Utilization by the Diatom Skeletonema costatum</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Serra, Juan L.; Llama, Maria J.; Cadenas, Eduardo</p> <p>1978-01-01</p> <p><span class="hlt">Nitrate</span> uptake has been studied in nitrogen-deficient cells of the marine diatom Skeletonema costatum. When these cells are incubated in the presence of <span class="hlt">nitrate</span>, this ion is quickly taken up from the medium, and nitrite is excreted by the cells. Nitrite is excreted following classical saturation kinetics, its rate being independent of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the incubation medium for <span class="hlt">nitrate</span> <span class="hlt">concentration</span> values higher than 3 micromolar. <span class="hlt">Nitrate</span> uptake shows mixed-transfer kinetics, which can be attributed to the simultaneous contributions of mediated and diffusion transfer. Cycloheximide and p-hydroxymercuribenzoate inhibit the carrier-mediated contribution to <span class="hlt">nitrate</span> uptake, without affecting the diffusion component. When cells are preincubated with <span class="hlt">nitrate</span>, the net nitrogen uptake is increased. PMID:16660652</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/1995/0468/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/1995/0468/report.pdf"><span>A computerized data base of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in Indiana ground water</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Risch, M.R.; Cohen, D.A.</p> <p>1995-01-01</p> <p>The <span class="hlt">nitrate</span> data base was compiled from numerous data sets that were readily accessible in electronic format. The uses of these data may be limited because they were neither comprehensive nor of a single statistical design. Nonetheless, the <span class="hlt">nitrate</span> data can be used in several ways: (1) to identify geographic areas with and without <span class="hlt">nitrate</span> data; (2) to evaluate assumptions, models, and maps of ground-water-contamination potential; and (3) to investigate the relation between environmental factors, land-use types, and the occurrence of <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AtmEn..76...43N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AtmEn..76...43N"><span>Atmospheric <span class="hlt">concentrations</span> of particulate sulfate and <span class="hlt">nitrate</span> in Hong Kong during 1995-2008: Impact of local emission and super-regional transport</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nie, Wei; Wang, Tao; Wang, Wenxing; Wei, Xiaolin; Liu, Qian</p> <p>2013-09-01</p> <p>The release of large amounts of sulfur dioxide (SO2) and nitrogen oxides (NOx) from the burning of fossil fuel leads to regional air pollution phenomena such as haze and acidic deposition. Despite longstanding recognition of the severity of these problems and the numerous studies conducted in China, little is known of long-term trends in particulate sulfate and <span class="hlt">nitrate</span> and their association with changes in precursor emissions. In this study, we analyze records covering a 14-year period (1995-2008) of PM10 composition in the subtropical city of Hong Kong, situated in the rapidly developing Pearl River Delta region of southern China. A linear regression method and a Regional Kendall test are employed for trend calculations. In contrast to the decreased levels of SO2 and NOx emissions in Hong Kong, there are increasing overall trends in ambient <span class="hlt">concentrations</span> of PM10 sulfate and PM10 <span class="hlt">nitrate</span>, with the most obvious rise seen during 2001-2005. The percentages of sulfate and <span class="hlt">nitrate</span> in the PM10 mass and rainwater acidity also increased. Backward trajectories are computed to help identify the origin of large-scale air masses arriving in Hong Kong. In air masses dominated by Hong Kong urban sources and ship emissions, there was no statistically significant trend for PM10 sulfate and a small increase for PM10 <span class="hlt">nitrate</span>; however, the evident increases in PM10 sulfate and PM10 <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were <span class="hlt">observed</span> in air masses originating from eastern China and are generally consistent with changes in emissions of their precursors in eastern China. Examination of PM10 mass data recorded at a pair of upwind-urban sites also indicates that long-range transport makes a large contribution (>80%) to PM10 loadings in Hong Kong. Together with our previous study on the ozone trend, these results demonstrate the important impact exerted by long-distance sources and suggest a need to consider the impact of super-regional transport when formulating air-quality management strategy in Hong</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005JHyd..310..201J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005JHyd..310..201J"><span>Spatial and temporal changes in the structure of groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span> time series (1935 1999) as demonstrated by autoregressive modelling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jones, A. L.; Smart, P. L.</p> <p>2005-08-01</p> <p>Autoregressive modelling is used to investigate the internal structure of long-term (1935-1999) records of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> for five karst springs in the Mendip Hills. There is a significant short term (1-2 months) positive autocorrelation at three of the five springs due to the availability of sufficient <span class="hlt">nitrate</span> within the soil store to maintain <span class="hlt">concentrations</span> in winter recharge for several months. The absence of short term (1-2 months) positive autocorrelation in the other two springs is due to the marked contrast in land use between the limestone and swallet parts of the catchment, rapid <span class="hlt">concentrated</span> recharge from the latter causing short term switching in the dominant water source at the spring and thus fluctuating <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Significant negative autocorrelation is evident at lags varying from 4 to 7 months through to 14-22 months for individual springs, with positive autocorrelation at 19-20 months at one site. This variable timing is explained by moderation of the exhaustion effect in the soil by groundwater storage, which gives longer residence times in large catchments and those with a dominance of diffuse flow. The lags derived from autoregressive modelling may therefore provide an indication of average groundwater residence times. Significant differences in the structure of the autocorrelation function for successive 10-year periods are evident at Cheddar Spring, and are explained by the effect the ploughing up of grasslands during the Second World War and increased fertiliser usage on available nitrogen in the soil store. This effect is moderated by the influence of summer temperatures on rates of mineralization, and of both summer and winter rainfall on the timing and magnitude of <span class="hlt">nitrate</span> leaching. The pattern of <span class="hlt">nitrate</span> leaching also appears to have been perturbed by the 1976 drought.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12022650','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12022650"><span>An accurate and stable <span class="hlt">nitrate</span>-selective electrode for the in situ determination of <span class="hlt">nitrate</span> in agricultural drainage waters.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Le Goff, Thierry; Braven, Jim; Ebdon, Les; Chilcottt, Neil P; Scholefield, David; Wood, John W</p> <p>2002-04-01</p> <p>A field evaluation of a novel <span class="hlt">nitrate</span>-ion selective electrode (ISE) was undertaken by continuous immersion over a period of 5 months in agricultural drainage weirs. The <span class="hlt">nitrate</span> sensor N,N,N-triallyl leucine betaine was covalently attached to polystyrene-block-polybutadiene-block-polystyrene (SBS) using a free radical initiated co-polymerisation, to produce a rubbery membrane which was incorporated into a commercially available electrode body. A measurement unit was constructed comprising the <span class="hlt">nitrate</span>-ISEs, a reference electrode and a temperature probe connected through a pre-amplifier to a data-logger and battery supply. A temperature correction algorithm was developed to accomodate the temperature changes encountered in the drainage weirs. The <span class="hlt">nitrate</span> results obtained with the ISEs at hourly intervals compared very favourably (R2 = 0.99) with those obtained with laboratory automated chemical determinations made on contemporaneous samples of drainage in a <span class="hlt">concentration</span> range 0.47-16 ppm <span class="hlt">nitrate</span>-N. The ISEs did not require re-calibration and no deterioration in performance or fouling of the membrane surface was <span class="hlt">observed</span> over four months of deployment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26439862','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26439862"><span>Impacts of management and climate change on <span class="hlt">nitrate</span> leaching in a forested karst area.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dirnböck, Thomas; Kobler, Johannes; Kraus, David; Grote, Rüdiger; Kiese, Ralf</p> <p>2016-01-01</p> <p>Forest management and climate change, directly or indirectly, affect drinking water resources, both in terms of quality and quantity. In this study in the Northern Limestone Alps in Austria we have chosen model calculations (LandscapeDNDC) in order to resolve the complex long-term interactions of management and climate change and their effect on nitrogen dynamics, and the consequences for <span class="hlt">nitrate</span> leaching from forest soils into the karst groundwater. Our study highlights the dominant role of forest management in controlling <span class="hlt">nitrate</span> leaching. Both clear-cut and shelterwood-cut disrupt the nitrogen cycle to an extent that causes peak <span class="hlt">concentrations</span> and high fluxes into the seepage water. While this effect is well known, our modelling approach has revealed additional positive as well as negative impacts of the expected climatic changes on <span class="hlt">nitrate</span> leaching. First, we show that peak <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> during post-cutting periods were elevated under all climate scenarios. The maximal effects of climatic changes on <span class="hlt">nitrate</span> <span class="hlt">concentration</span> peaks were 20-24 mg L(-1) in 2090 with shelterwood or clear-cut management. Second, climate change significantly decreased the cumulative <span class="hlt">nitrate</span> losses over full forest rotation periods (by 10-20%). The stronger the expected temperature increase and precipitation decrease (in summer), the lesser were the <span class="hlt">observed</span> <span class="hlt">nitrate</span> losses. However, mean annual seepage water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and cumulative <span class="hlt">nitrate</span> leaching were higher under continuous forest cover management than with shelterwood-cut and clear-cut systems. Watershed management can thus be adapted to climate change by either reducing peak <span class="hlt">concentrations</span> or long-term loads of <span class="hlt">nitrate</span> in the karst groundwater. Copyright © 2015 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17932709','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17932709"><span>Toxicity of un-ionized ammonia, nitrite, and <span class="hlt">nitrate</span> to juvenile bay scallops, Argopecten irradians irradians.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Widman, James C; Meseck, Shannon L; Sennefelder, George; Veilleux, David J</p> <p>2008-04-01</p> <p>Juvenile bay scallops (7.2-26.4 mm) were exposed for 72 h to different <span class="hlt">concentrations</span> of un-ionized ammonia, nitrite, or <span class="hlt">nitrate</span>. Using the Trimmed Spearman Karber method, 50% lethal <span class="hlt">concentrations</span> (LC(50)) and 95% confidence limits were calculated individually for each. Un-ionized ammonia <span class="hlt">concentrations</span> above 1.0 mg N-NH(3)/L resulted in 100% scallop mortality within 72 h. The 72-h LC(50) for un-ionized ammonia was calculated at 0.43 mg N/L. At nitrite <span class="hlt">concentrations</span> of 800 mg N/L or higher 100% mortality was <span class="hlt">observed</span>. The 72-h LC(50) for nitrite was calculated at 345 mg N/L. <span class="hlt">Nitrate</span> was the least toxic, with 100% mortality <span class="hlt">observed</span> at a <span class="hlt">concentration</span> of 5000 mg N/L. The calculated <span class="hlt">nitrate</span> 72-h LC(50) was 4453 mg N/L. Our results indicate that un-ionized ammonia is the most lethal nitrogenous waste component to bay scallops.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.C24B..02H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.C24B..02H"><span>The preservation of long-range transported <span class="hlt">nitrate</span> in snow at Summit, Greenland (Invited)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hastings, M. G.</p> <p>2013-12-01</p> <p> insights into the contributions of <span class="hlt">nitrate</span> sources to Summit. There are several important implications of this work including that <span class="hlt">nitrate</span> at Summit appears to be largely preserved in surface snow during photoactive periods, and that <span class="hlt">nitrate</span> in snow at Summit also appears to be representative of long-range transported <span class="hlt">nitrate</span>/NOx. The surface snow work is further substantiated by relationships found between and among seasonally-resolved ice core measurements of the isotopic composition of <span class="hlt">nitrate</span>, <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and a suite of chemical and elemental tracers. The seasonality <span class="hlt">observed</span> in 15N/14N ratios in an ice core representing accumulation since 1760 C.E. cannot be explained by diffusion or other processes occurring in the firn over time. A marked negative trend in 15N/14N since industrialization, parallels a nearly three-fold increase in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> as well as pronounced increases in tracers such as excess lead and non-sea-salt sulfur. This, along with independent estimates of oil burning and transport studies, indicate that North American oil combustion is the primary driver of the modern negative trend in 15N/14N of <span class="hlt">nitrate</span>. The high, positive 15N/14N ratios found in pre-industrial ice link to biomass burning based upon <span class="hlt">concentrations</span> of black carbon and ammonium.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23801341','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23801341"><span>The <span class="hlt">nitrate</span> time bomb: a numerical way to investigate <span class="hlt">nitrate</span> storage and lag time in the unsaturated zone.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, L; Butcher, A S; Stuart, M E; Gooddy, D C; Bloomfield, J P</p> <p>2013-10-01</p> <p><span class="hlt">Nitrate</span> pollution in groundwater, which is mainly from agricultural activities, remains an international problem. It threatens the environment, economics and human health. There is a rising trend in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in many UK groundwater bodies. Research has shown it can take decades for leached <span class="hlt">nitrate</span> from the soil to discharge into groundwater and surface water due to the 'store' of <span class="hlt">nitrate</span> and its potentially long travel time in the unsaturated and saturated zones. However, this time lag is rarely considered in current water <span class="hlt">nitrate</span> management and policy development. The aim of this study was to develop a catchment-scale integrated numerical method to investigate the <span class="hlt">nitrate</span> lag time in the groundwater system, and the Eden Valley, UK, was selected as a case study area. The method involves three models, namely the <span class="hlt">nitrate</span> time bomb-a process-based model to simulate the <span class="hlt">nitrate</span> transport in the unsaturated zone (USZ), GISGroundwater--a GISGroundwater flow model, and N-FM--a model to simulate the <span class="hlt">nitrate</span> transport in the saturated zone. This study answers the scientific questions of when the <span class="hlt">nitrate</span> currently in the groundwater was loaded into the unsaturated zones and eventually reached the water table; is the rising groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the study area caused by historic <span class="hlt">nitrate</span> load; what caused the uneven distribution of groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the study area; and whether the historic peak <span class="hlt">nitrate</span> loading has reached the water table in the area. The groundwater <span class="hlt">nitrate</span> in the area was mainly from the 1980s to 2000s, whilst the groundwater <span class="hlt">nitrate</span> in most of the source protection zones leached into the system during 1940s-1970s; the large and spatially variable thickness of the USZ is one of the major reasons for unevenly distributed groundwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the study area; the peak <span class="hlt">nitrate</span> loading around 1983 has affected most of the study area. For areas around the Bowscar, Beacon Edge, Low Plains, Nord Vue</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22060756','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22060756"><span>Monitoring nitrite and <span class="hlt">nitrate</span> residues in frankfurters during processing and storage.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pérez-Rodríguez, M L; Bosch-Bosch, N; Garciá-Mata, M</p> <p>1996-09-01</p> <p>Frankfurter-type sausages were prepared in a pilot plant with different <span class="hlt">concentrations</span> of NaNO(2) (75, 125 or 250 ppm) combined or not with 200 ppm KNO(3). A meat system, free of curing agents, was also used as control. Nitrite and <span class="hlt">nitrate</span> levels were tested in various processing steps and over 120 days storage at 3 °C of the vacuum-packaged frankfurters. Little influence of the originally added nitrite level on the amount of <span class="hlt">nitrate</span> formed was <span class="hlt">observed</span>. Important losses of nitrite and <span class="hlt">nitrate</span> were due to cooking. Thereafter about 50% of the nitrite added initially remained in this form in all samples (39, 59 and 146 ppm, respectively) and between 10 and 15% as <span class="hlt">nitrate</span>. When only <span class="hlt">nitrate</span> was initially added, formation of nitrite after cooking was <span class="hlt">observed</span> (maximum level 43 ppm NaNO(2)). Formulations prepared with both <span class="hlt">nitrate</span> and nitrite showed no significant differences (p < 0.01) respect to their nitrite or <span class="hlt">nitrate</span> counterparts. A good correlation among nitrite and <span class="hlt">nitrate</span> levels and storage time was showed by multiple linear regression analysis. It is concluded that the use of <span class="hlt">nitrate</span> in combination with nitrite in cooked meat products seems to have little technological significance and adds to the total body burden of nitrite.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23431201','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23431201"><span>Isotopic composition of atmospheric <span class="hlt">nitrate</span> in a tropical marine boundary layer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Savarino, Joel; Morin, Samuel; Erbland, Joseph; Grannec, Francis; Patey, Matthew D; Vicars, William; Alexander, Becky; Achterberg, Eric P</p> <p>2013-10-29</p> <p>Long-term <span class="hlt">observations</span> of the reactive chemical composition of the tropical marine boundary layer (MBL) are rare, despite its crucial role for the chemical stability of the atmosphere. Recent <span class="hlt">observations</span> of reactive bromine species in the tropical MBL showed unexpectedly high levels that could potentially have an impact on the ozone budget. Uncertainties in the ozone budget are amplified by our poor understanding of the fate of NOx (= NO + NO2), particularly the importance of nighttime chemical NOx sinks. Here, we present year-round <span class="hlt">observations</span> of the multiisotopic composition of atmospheric <span class="hlt">nitrate</span> in the tropical MBL at the Cape Verde Atmospheric Observatory. We show that the <span class="hlt">observed</span> oxygen isotope ratios of <span class="hlt">nitrate</span> are compatible with <span class="hlt">nitrate</span> formation chemistry, which includes the BrNO3 sink at a level of ca. 20 ± 10% of <span class="hlt">nitrate</span> formation pathways. The results also suggest that the N2O5 pathway is a negligible NOx sink in this environment. <span class="hlt">Observations</span> further indicate a possible link between the NO2/NOx ratio and the nitrogen isotopic content of <span class="hlt">nitrate</span> in this low NOx environment, possibly reflecting the seasonal change in the photochemical equilibrium among NOx species. This study demonstrates the relevance of using the stable isotopes of oxygen and nitrogen of atmospheric <span class="hlt">nitrate</span> in association with <span class="hlt">concentration</span> measurements to identify and constrain chemical processes occurring in the MBL.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3816420','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3816420"><span>Isotopic composition of atmospheric <span class="hlt">nitrate</span> in a tropical marine boundary layer</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Savarino, Joel; Morin, Samuel; Erbland, Joseph; Grannec, Francis; Patey, Matthew D.; Vicars, William; Alexander, Becky; Achterberg, Eric P.</p> <p>2013-01-01</p> <p>Long-term <span class="hlt">observations</span> of the reactive chemical composition of the tropical marine boundary layer (MBL) are rare, despite its crucial role for the chemical stability of the atmosphere. Recent <span class="hlt">observations</span> of reactive bromine species in the tropical MBL showed unexpectedly high levels that could potentially have an impact on the ozone budget. Uncertainties in the ozone budget are amplified by our poor understanding of the fate of NOx (= NO + NO2), particularly the importance of nighttime chemical NOx sinks. Here, we present year-round <span class="hlt">observations</span> of the multiisotopic composition of atmospheric <span class="hlt">nitrate</span> in the tropical MBL at the Cape Verde Atmospheric Observatory. We show that the <span class="hlt">observed</span> oxygen isotope ratios of <span class="hlt">nitrate</span> are compatible with <span class="hlt">nitrate</span> formation chemistry, which includes the BrNO3 sink at a level of ca. 20 ± 10% of <span class="hlt">nitrate</span> formation pathways. The results also suggest that the N2O5 pathway is a negligible NOx sink in this environment. <span class="hlt">Observations</span> further indicate a possible link between the NO2/NOx ratio and the nitrogen isotopic content of <span class="hlt">nitrate</span> in this low NOx environment, possibly reflecting the seasonal change in the photochemical equilibrium among NOx species. This study demonstrates the relevance of using the stable isotopes of oxygen and nitrogen of atmospheric <span class="hlt">nitrate</span> in association with <span class="hlt">concentration</span> measurements to identify and constrain chemical processes occurring in the MBL. PMID:23431201</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70007513','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70007513"><span><span class="hlt">Nitrate</span> in groundwater of the United States, 1991-2003</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Burow, Karen R.; Nolan, Bernard T.; Rupert, Michael G.; Dubrovsky, Neil M.</p> <p>2010-01-01</p> <p>An assessment of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater in the United States indicates that <span class="hlt">concentrations</span> are highest in shallow, oxic groundwater beneath areas with high N inputs. During 1991-2003, 5101 wells were sampled in 51 study areas throughout the U.S. as part of the U.S. Geological Survey National Water-Quality Assessment (NAWQA) program. The well networks reflect the existing used resource represented by domestic wells in major aquifers (major aquifer studies), and recently recharged groundwater beneath dominant land-surface activities (land-use studies). <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> were highest in shallow groundwater beneath agricultural land use in areas with well-drained soils and oxic geochemical conditions. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> were lowest in deep groundwater where groundwater is reduced, or where groundwater is older and hence <span class="hlt">concentrations</span> reflect historically low N application rates. Classification and regression tree analysis was used to identify the relative importance of N inputs, biogeochemical processes, and physical aquifer properties in explaining <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater. Factors ranked by reduction in sum of squares indicate that dissolved iron <span class="hlt">concentrations</span> explained most of the variation in groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, followed by manganese, calcium, farm N fertilizer inputs, percent well-drained soils, and dissolved oxygen. Overall, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater are most significantly affected by redox conditions, followed by nonpoint-source N inputs. Other water-quality indicators and physical variables had a secondary influence on <span class="hlt">nitrate</span> <span class="hlt">concentrations</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_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://www.ncbi.nlm.nih.gov/pubmed/28131098','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28131098"><span><span class="hlt">Nitrate</span> reduction and its effects on trichloroethylene degradation by granular iron.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lu, Qiong; Jeen, Sung-Wook; Gui, Lai; Gillham, Robert W</p> <p>2017-04-01</p> <p>Laboratory column experiments and reactive transport modeling were performed to evaluate the reduction of <span class="hlt">nitrate</span> and its effects on trichloroethylene (TCE) degradation by granular iron. In addition to determining degradation kinetics of TCE in the presence of <span class="hlt">nitrate</span>, the columns used in this study were equipped with electrodes which allowed for in situ measurements of corrosion potentials of the iron material. Together with Raman spectroscopic measurements the mechanisms of decline in iron reactivity were examined. The experimental results showed that the presence of <span class="hlt">nitrate</span> resulted in an increase in corrosion potential and the formation of thermodynamically stable passive films on the iron surface which impaired iron reactivity. The extent of the decline in iron reactivity was proportional to the <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. Consequently, significant decreases in TCE and <span class="hlt">nitrate</span> degradation rates and migration of degradation profiles for both compounds occurred. Furthermore, the TCE degradation kinetics deviated from the pseudo-first-order model. The results of reactive transport modeling, which related the amount of a passivating iron oxide, hematite (α-Fe 2 O 3 ), to the reactivity of iron, were generally consistent with the patterns of migration of TCE and <span class="hlt">nitrate</span> profiles <span class="hlt">observed</span> in the column experiments. More encouragingly, the simulations successfully demonstrated the differences in performances of three columns without changing model parameters other than <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> in the influent. This study could be valuable in the design of iron permeable reactive barriers (PRBs) or in the development of effective maintenance procedures for PRBs treating TCE-contaminated groundwater with elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Copyright © 2017 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70028392','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70028392"><span>Estimating background and threshold <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> using probability graphs</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Panno, S.V.; Kelly, W.R.; Martinsek, A.T.; Hackley, Keith C.</p> <p>2006-01-01</p> <p>Because of the ubiquitous nature of anthropogenic <span class="hlt">nitrate</span> (NO 3-) in many parts of the world, determining background <span class="hlt">concentrations</span> of NO3- in shallow ground water from natural sources is probably impossible in most environments. Present-day background must now include diffuse sources of NO3- such as disruption of soils and oxidation of organic matter, and atmospheric inputs from products of combustion and evaporation of ammonia from fertilizer and livestock waste. Anomalies can be defined as NO3- derived from nitrogen (N) inputs to the environment from anthropogenic activities, including synthetic fertilizers, livestock waste, and septic effluent. Cumulative probability graphs were used to identify threshold <span class="hlt">concentrations</span> separating background and anomalous NO3-N <span class="hlt">concentrations</span> and to assist in the determination of sources of N contamination for 232 spring water samples and 200 well water samples from karst aquifers. Thresholds were 0.4, 2.5, and 6.7 mg/L for spring water samples, and 0.1, 2.1, and 17 mg/L for well water samples. The 0.4 and 0.1 mg/L values are assumed to represent thresholds for present-day precipitation. Thresholds at 2.5 and 2.1 mg/L are interpreted to represent present-day background <span class="hlt">concentrations</span> of NO3-N. The population of spring water samples with <span class="hlt">concentrations</span> between 2.5 and 6.7 mg/L represents an amalgam of all sources of NO3- in the ground water basins that feed each spring; <span class="hlt">concentrations</span> >6.7 mg/L were typically samples collected soon after springtime application of synthetic fertilizer. The 17 mg/L threshold (adjusted to 15 mg/L) for well water samples is interpreted as the level above which livestock wastes dominate the N sources. Copyright ?? 2006 The Author(s).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.B41J0188L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.B41J0188L"><span><span class="hlt">Nitrate</span> dynamics within a stream-lake network through time and space</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Loken, L. C.; Crawford, J. T.; Childress, E. S.; Casson, N. J.; Stanley, E. H.</p> <p>2014-12-01</p> <p><span class="hlt">Nitrate</span> dynamics in streams are governed by biology, hydrology, and geomorphology, and the ability to parse these drivers apart has improved with the development of accurate high-frequency sensors. By combining a stationary Eulerian and a quasi-Lagrangian sensor platform, we investigated the timing of <span class="hlt">nitrate</span> flushing and identified locations of elevated biogeochemical cycling along a stream-lake network in Northern Wisconsin, USA. Two years of continuous oxygen, carbon dioxide, and discharge measurements were used to compute gross primary production (GPP) and ecosystem respiration (ER) downstream of a wetland reach of Allequash Creek. Metabolic rates and flow patterns were compared with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> measured every 30 minutes using an optical sensor. Additionally, we floated a sensor array from the headwater spring ponds through a heterogeneous stream reach consisting of wetlands, beaver ponds, forested segments, and two lakes. Two distinct temporal patterns of stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were <span class="hlt">observed</span>. During high flow events such as spring snowmelt and summer rain events, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increased from ~5 μM (baseflow) to 12 μM, suggesting flushing from catchment sources. During baseflow conditions, <span class="hlt">nitrate</span> followed a diel cycle with a 0.3-1.0 μM daytime draw down. Daily <span class="hlt">nitrate</span> reduction was positively correlated with GPP calculated from oxygen and carbon dioxide records. Lastly, spatial analyses revealed lowest <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the wetland reach, approximately 2-3 μM lower than the upstream spring ponds, and downstream lakes and forested reaches. This snapshot implies greater <span class="hlt">nitrate</span> removal potential in the wetland reach likely driven by denitrification in organic rich sediments and macrophyte uptake in the open canopy stream segment. Taken together the temporal and spatial results show the dynamics of hydrology, geomorphology, and biology to influence <span class="hlt">nitrate</span> delivery and variability in ecosystem processing through a stream</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70032686','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70032686"><span>Sources, transformations, and hydrological processes that control stream <span class="hlt">nitrate</span> and dissolved organic matter <span class="hlt">concentrations</span> during snowmelt in an upland forest</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Sebestyen, Stephen D.; Boyer, Elizabeth W.; Shanley, James B.; Kendall, Carol; Doctor, Daniel H.; Aiken, George R.; Ohte, Nobuhito</p> <p>2008-01-01</p> <p>We explored catchment processes that control stream nutrient <span class="hlt">concentrations</span> at an upland forest in northeastern Vermont, USA, where inputs of nitrogen via atmospheric deposition are among the highest in the nation and affect ecosystem functioning. We traced sources of water, <span class="hlt">nitrate</span>, and dissolved organic matter (DOM) using stream water samples collected at high frequency during spring snowmelt. Hydrochemistry, isotopic tracers, and end‐member mixing analyses suggested the timing, sources, and source areas from which water and nutrients entered the stream. Although stream‐dissolved organic carbon (DOC) and dissolved organic nitrogen (DON) both originated from leaching of soluble organic matter, flushing responses between these two DOM components varied because of dynamic shifts of hydrological flow paths and sources that supply the highest <span class="hlt">concentrations</span> of DOC and DON. High <span class="hlt">concentrations</span> of stream water <span class="hlt">nitrate</span> originated from atmospheric sources as well as nitrified sources from catchment soils. We detected nitrification in surficial soils during late snowmelt which affected the <span class="hlt">nitrate</span> supply that was available to be transported to streams. However, isotopic tracers showed that the majority of <span class="hlt">nitrate</span> in upslope surficial soil waters after the onset of snowmelt originated from atmospheric sources. A fraction of the atmospheric nitrogen was directly delivered to the stream, and this finding highlights the importance of quick flow pathways during snowmelt events. These findings indicate that interactions among sources, transformations, and hydrologic transport processes must be deciphered to understand why <span class="hlt">concentrations</span> vary over time and over space as well as to elucidate the direct effects of human activities on nutrient dynamics in upland forest streams.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26606093','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26606093"><span>Enhanced removal of <span class="hlt">nitrate</span> from water using nZVI@MWCNTs composite: synthesis, kinetics and mechanism of reduction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Babaei, Ali Akbar; Azari, Ali; Kalantary, Roshanak Rezaei; Kakavandi, Babak</p> <p>2015-01-01</p> <p>Herein, multi-wall carbon nanotubes (MWCNTs) were used as the carrier of nano-zero valent iron (nZVI) particles to fabricate a composite known as nZVI@MWCNTs. The composite was then characterized and applied in the <span class="hlt">nitrate</span> removal process in a batch system under anoxic conditions. The influential parameters such as pH, various <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and composite were investigated within 240 min of the reaction. The mechanism, kinetics and end-products of <span class="hlt">nitrate</span> reduction were also evaluated. Results revealed that the removal <span class="hlt">nitrate</span> percentage for nZVI@MWCNTs composite was higher than that of nZVI and MWCNTs alone. Experimental data from <span class="hlt">nitrate</span> reduction were fitted to the Langmuir-Hinshelwood kinetic model. The values of <span class="hlt">observed</span> rate constant (kobs) decreased with increasing the initial <span class="hlt">concentration</span> of <span class="hlt">nitrate</span>. Our experiments proved that the <span class="hlt">nitrate</span> removal efficiency was favorable once both high amounts of nZVI@MWCNTs and low <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> were applied. The predominant end-products of the <span class="hlt">nitrate</span> reduction were ammonium (84%) and nitrogen gas (15%). Our findings also revealed that ZVI@MWCNTs is potentially a good composite for removal/reduction of <span class="hlt">nitrate</span> from aqueous solutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H21R..03T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H21R..03T"><span>Legacy <span class="hlt">Nitrate</span> Impacts on Groundwater and Streams</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tesoriero, A. J.; Juckem, P. F.; Miller, M. P.</p> <p>2017-12-01</p> <p>Decades of recharge of high-<span class="hlt">nitrate</span> groundwater have created a legacy—a mass of high-<span class="hlt">nitrate</span> groundwater—that has implications for future <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater and in streams. In the United States, inorganic nitrogen fertilizer applications to the land surface have increased ten-fold since 1950, resulting in sharp increases in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in recharging groundwater, which pose a risk to deeper groundwater and streams. This study assesses the factors that control time lags and eventual <span class="hlt">concentrations</span> of legacy <span class="hlt">nitrate</span> in groundwater and streams. Results from the USGS National Water-Quality Assessment Project are presented which elucidate <span class="hlt">nitrate</span> trends in recharging groundwater, delineate redox zones and assess groundwater and stream vulnerability to legacy <span class="hlt">nitrate</span> sources on a regional scale. This study evaluated trends and transformations of agricultural chemicals based on groundwater age and water chemistry data along flow paths from recharge areas to streams at 20 study sites across the United States. Median <span class="hlt">nitrate</span> recharge <span class="hlt">concentrations</span> in these agricultural areas have increased markedly over the last 50 years, from 4 to 7.5 mg N/L. The effect that <span class="hlt">nitrate</span> accumulation in shallow aquifers will have on drinking water quality and stream ecosystems is dependent on the redox zones encountered along flow paths and on the age distribution of <span class="hlt">nitrate</span> discharging to supply wells and streams. Delineating redox zones on a regional scale is complicated by the spatial variability of reaction rates. To overcome this limitation, we applied logistic regression and machine learning techniques to predict the probability of a specific redox condition in groundwater in the Chesapeake Bay watershed and the Fox-Wolf-Peshtigo study area in Wisconsin. By relating redox-active constituent <span class="hlt">concentrations</span> in groundwater samples to indicators of residence time and/or electron donor availability, we were able to delineate redox zones on a regional scale</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3482663','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3482663"><span>Groundwater <span class="hlt">nitrate</span> contamination: Factors and indicators</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wick, Katharina; Heumesser, Christine; Schmid, Erwin</p> <p>2012-01-01</p> <p>Identifying significant determinants of groundwater <span class="hlt">nitrate</span> contamination is critical in order to define sensible agri-environmental indicators that support the design, enforcement, and monitoring of regulatory policies. We use data from approximately 1200 Austrian municipalities to provide a detailed statistical analysis of (1) the factors influencing groundwater <span class="hlt">nitrate</span> contamination and (2) the predictive capacity of the Gross Nitrogen Balance, one of the most commonly used agri-environmental indicators. We find that the percentage of cropland in a given region correlates positively with <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater. Additionally, environmental characteristics such as temperature and precipitation are important co-factors. Higher average temperatures result in lower <span class="hlt">nitrate</span> contamination of groundwater, possibly due to increased evapotranspiration. Higher average precipitation dilutes <span class="hlt">nitrates</span> in the soil, further reducing groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. Finally, we assess whether the Gross Nitrogen Balance is a valid predictor of groundwater <span class="hlt">nitrate</span> contamination. Our regression analysis reveals that the Gross Nitrogen Balance is a statistically significant predictor for <span class="hlt">nitrate</span> contamination. We also show that its predictive power can be improved if we account for average regional precipitation. The Gross Nitrogen Balance predicts <span class="hlt">nitrate</span> contamination in groundwater more precisely in regions with higher average precipitation. PMID:22906701</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H52E..05K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H52E..05K"><span>In what root-zone N <span class="hlt">concentration</span> does <span class="hlt">nitrate</span> start to leach significantly? A reasonable answer from modeling Mediterranean field data and closed root-zone experiments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kurtzman, D.; Kanner, B.; Levy, Y.; Shapira, R. H.; Bar-Tal, A.</p> <p>2017-12-01</p> <p>Closed-root-zone experiments (e.g. pots, lyzimeters) reveal in many cases a mineral-nitrogen (N) <span class="hlt">concentration</span> from which the root-N-uptake efficiency reduces significantly and <span class="hlt">nitrate</span> leaching below the root-zone increases dramatically. A les-direct way to reveal this threshold <span class="hlt">concentration</span> in agricultural fields is to calibrate N-transport models of the unsaturated zone to <span class="hlt">nitrate</span> data of the deep samples (under the root-zone) by fitting the threshold <span class="hlt">concentration</span> of the <span class="hlt">nitrate</span>-uptake function. Independent research efforts of these two types in light soils where <span class="hlt">nitrate</span> problems in underlying aquifers are common reviled: 1) that the threshold exists for most crops (filed, vegetables and orchards); 2) nice agreement on the threshold value between the two very different research methodologies; and 3) the threshold lies within 20-50 mg-N/L. Focusing on being below the threshold is a relatively simple aim in the way to maintain intensive agriculture with limited effects on the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the underlying water resource. Our experience show that in some crops this threshold coincides with the end-of-rise of the N-yield curve (e.g. corn); in this case, it is relatively easy to convince farmers to fertilize below threshold. In other crops, although significant N is lost to leaching the crop can still use higher N <span class="hlt">concentration</span> to increase yield (e.g. potato).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4439352','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4439352"><span>Skeletal muscle as an endogenous <span class="hlt">nitrate</span> reservoir</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Piknova, Barbora; Park, Ji Won; Swanson, Kathryn M.; Dey, Soumyadeep; Noguchi, Constance Tom; Schechter, Alan N</p> <p>2015-01-01</p> <p>The nitric oxide synthase (NOS) family of enzymes form nitric oxide (NO) from arginine in the presence of oxygen. At reduced oxygen availability NO is also generated from <span class="hlt">nitrate</span> in a two step process by bacterial and mammalian molybdopterin proteins, and also directly from nitrite by a variety of five-coordinated ferrous hemoproteins. The mammalian NO cycle also involves direct oxidation of NO to nitrite, and both NO and nitrite to <span class="hlt">nitrate</span> by oxy-ferrous hemoproteins. The liver and blood are considered the sites of active mammalian NO metabolism and nitrite and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the liver and blood of several mammalian species, including human, have been determined. However, the large tissue mass of skeletal muscle had not been generally considered in the analysis of the NO cycle, in spite of its long-known presence of significant levels of active neuronal NOS (nNOS or NOS1). We hypothesized that skeletal muscle participates in the NO cycle and, due to its NO oxidizing heme protein, oxymyoglobin, has high <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> ions. We measured nitrite and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in rat and mouse leg skeletal muscle and found unusually high <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> but similar levels of nitrite, when compared to the liver. The <span class="hlt">nitrate</span> reservoir in muscle is easily accessible via the bloodstream and therefore <span class="hlt">nitrate</span> is available for transport to internal organs where it can be reduced to nitrite and NO. <span class="hlt">Nitrate</span> levels in skeletal muscle and blood in nNOS−/− mice were dramatically lower when compared with controls, which support further our hypothesis. Although the <span class="hlt">nitrate</span> reductase activity of xanthine oxidoreductase in muscle is less than that of liver, the residual activity in muscle could be very important in view of its total mass and the high basal level of <span class="hlt">nitrate</span>. We suggest that skeletal muscle participates in overall NO metabolism, serving as a <span class="hlt">nitrate</span> reservoir, for direct formation of nitrite and NO, and for determining levels of <span class="hlt">nitrate</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1474580','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1474580"><span>Groundwater pollution by <span class="hlt">nitrates</span> from livestock wastes.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Goldberg, V M</p> <p>1989-01-01</p> <p>Utilization of wastes from livestock complexes for irrigation involves the danger of groundwater pollution by <span class="hlt">nitrates</span>. In order to prevent and minimize pollution, it is necessary to apply geological-hydrogeological evidence and concepts to the situation of wastewater irrigation for the purposes of studying natural groundwater protectiveness and predicting changes in groundwater quality as a result of infiltrating wastes. The procedure of protectiveness evaluation and quality prediction is described. With groundwater pollution by <span class="hlt">nitrate</span> nitrogen, the <span class="hlt">concentration</span> of ammonium nitrogen noticeably increases. One of the reasons for this change is the process of denitrification due to changes in the hydrogeochemical conditions in a layer. At representative field sites, it is necessary to collect systematic stationary <span class="hlt">observations</span> of the <span class="hlt">concentrations</span> of nitrogenous compounds in groundwater and changes in redox conditions and temperature. PMID:2620669</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25345876','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25345876"><span>Growing patterns to produce '<span class="hlt">nitrate</span>-free' lettuce (Lactuca sativa).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Croitoru, Mircea Dumitru; Muntean, Daniela-Lucia; Fülöp, Ibolya; Modroiu, Adriana</p> <p>2015-01-01</p> <p>Vegetables can contain significant amounts of <span class="hlt">nitrate</span> and, therefore, may pose health hazards to consumers by exceeding the accepted daily intake for <span class="hlt">nitrate</span>. Different hydroponic growing patterns were examined in this work in order to obtain '<span class="hlt">nitrate</span>-free lettuces'. Growing lettuces on low <span class="hlt">nitrate</span> content nutrient solution resulted in a significant decrease in lettuces' <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> (1741 versus 39 mg kg(-1)), however the beneficial effect was cancelled out by an increase in the ambient temperature. <span class="hlt">Nitrate</span> replacement with ammonium was associated with an important decrease of the lettuces' <span class="hlt">nitrate</span> <span class="hlt">concentration</span> (from 1896 to 14 mg kg(-1)) and survival rate. An economically feasible method to reduce <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> was the removal of all inorganic nitrogen from the nutrient solution before the exponential growth phase. This method led to lettuces almost devoid of <span class="hlt">nitrate</span> (10 mg kg(-1)). The dried mass and calcinated mass of lettuces, used as markers of lettuces' quality, were not influenced by this treatment, but a small reduction (18%, p < 0.05) in the fresh mass was recorded. The <span class="hlt">concentrations</span> of nitrite in the lettuces and their modifications are also discussed in the paper. It is possible to obtain '<span class="hlt">nitrate</span>-free' lettuces in an economically feasible way.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3657194','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3657194"><span>Pseudo-constitutivity of <span class="hlt">nitrate</span>-responsive genes in <span class="hlt">nitrate</span> reductase mutants</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Schinko, Thorsten; Gallmetzer, Andreas; Amillis, Sotiris; Strauss, Joseph</p> <p>2013-01-01</p> <p>In fungi, transcriptional activation of genes involved in NO3- assimilation requires the presence of an inducer (<span class="hlt">nitrate</span> or nitrite) and low intracellular <span class="hlt">concentrations</span> of the pathway products ammonium or glutamine. In Aspergillus nidulans, the two transcription factors NirA and AreA act synergistically to mediate <span class="hlt">nitrate</span>/nitrite induction and nitrogen metabolite derepression, respectively. In all studied fungi and in plants, mutants lacking <span class="hlt">nitrate</span> reductase (NR) activity express <span class="hlt">nitrate</span>-metabolizing enzymes constitutively without the addition of inducer molecules. Based on their work in A. nidulans, Cove and Pateman proposed an “autoregulation control” model for the synthesis of <span class="hlt">nitrate</span> metabolizing enzymes in which the functional <span class="hlt">nitrate</span> reductase molecule would act as co-repressor in the absence and as co-inducer in the presence of <span class="hlt">nitrate</span>. However, NR mutants could simply show “pseudo-constitutivity” due to induction by <span class="hlt">nitrate</span> which accumulates over time in NR-deficient strains. Here we examined this possibility using strains which lack flavohemoglobins (fhbs), and are thus unable to generate <span class="hlt">nitrate</span> internally, in combination with <span class="hlt">nitrate</span> transporter mutations (nrtA, nrtB) and a GFP-labeled NirA protein. Using different combinations of genotypes we demonstrate that <span class="hlt">nitrate</span> transporters are functional also in NR null mutants and show that the constitutive phenotype of NR mutants is not due to <span class="hlt">nitrate</span> accumulation from intracellular sources but depends on the activity of <span class="hlt">nitrate</span> transporters. However, these transporters are not required for <span class="hlt">nitrate</span> signaling because addition of external <span class="hlt">nitrate</span> (10 mM) leads to standard induction of <span class="hlt">nitrate</span> assimilatory genes in the <span class="hlt">nitrate</span> transporter double mutants. We finally show that NR does not regulate NirA localization and activity, and thus the autoregulation model, in which NR would act as a co-repressor of NirA in the absence of <span class="hlt">nitrate</span>, is unlikely to be correct. Results from this study instead suggest</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title49-vol2/pdf/CFR-2011-title49-vol2-sec176-410.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title49-vol2/pdf/CFR-2011-title49-vol2-sec176-410.pdf"><span>49 CFR 176.410 - Division 1.5 materials, ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures.</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>2011-10-01</p> <p>... 49 Transportation 2 2011-10-01 2011-10-01 false Division 1.5 materials, ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures. 176.410 Section 176.410 Transportation Other Regulations Relating to... <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures. (a) This section prescribes requirements to be <span class="hlt">observed</span> with...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title49-vol2/pdf/CFR-2012-title49-vol2-sec176-410.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title49-vol2/pdf/CFR-2012-title49-vol2-sec176-410.pdf"><span>49 CFR 176.410 - Division 1.5 materials, ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures.</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>2012-10-01</p> <p>... 49 Transportation 2 2012-10-01 2012-10-01 false Division 1.5 materials, ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures. 176.410 Section 176.410 Transportation Other Regulations Relating to... <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures. (a) This section prescribes requirements to be <span class="hlt">observed</span> with...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title49-vol2/pdf/CFR-2013-title49-vol2-sec176-410.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title49-vol2/pdf/CFR-2013-title49-vol2-sec176-410.pdf"><span>49 CFR 176.410 - Division 1.5 materials, ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures.</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>2013-10-01</p> <p>... 49 Transportation 2 2013-10-01 2013-10-01 false Division 1.5 materials, ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures. 176.410 Section 176.410 Transportation Other Regulations Relating to... <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures. (a) This section prescribes requirements to be <span class="hlt">observed</span> with...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title49-vol2/pdf/CFR-2014-title49-vol2-sec176-410.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title49-vol2/pdf/CFR-2014-title49-vol2-sec176-410.pdf"><span>49 CFR 176.410 - Division 1.5 materials, ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-10-01</p> <p>... 49 Transportation 2 2014-10-01 2014-10-01 false Division 1.5 materials, ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures. 176.410 Section 176.410 Transportation Other Regulations Relating to... <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> mixtures. (a) This section prescribes requirements to be <span class="hlt">observed</span> with...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020143','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020143"><span>Effect of <span class="hlt">nitrate</span>, organic carbon, and temperature on potential denitrification rates in <span class="hlt">nitrate</span>-rich riverbed sediments</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Pfenning, K.S.; McMahon, P.B.</p> <p>1997-01-01</p> <p>A study conducted in 1994 as part of the US Geological Survey's National Water-Quality Assessment Program, South Platte River Basin investigation, examined the effect of certain environmental factors on potential denitrification rates in <span class="hlt">nitrate</span>-rich riverbed sediments. The acetylene block technique was used to measure nitrous oxide (N2O) production rates in laboratory incubations of riverbed sediments to evaluate the effect of varying <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, organic carbon <span class="hlt">concentrations</span> and type, and water temperature on potential denitrification rates. Sediment incubations amended with <span class="hlt">nitrate</span>, at <span class="hlt">concentrations</span> ranging from 357 to 2142 ??mol l-1 (as measured in the field), produced no significant increase (P > 0.05) in N2O production rates, indicating that the denitrification potential in these sediments was not <span class="hlt">nitrate</span> limited. In contrast, incubations amended with acetate as a source of organic carbon, at <span class="hlt">concentrations</span> ranging from 0 to 624 ??mol l-1, produced significant increases (P < 0.05) in N2O production rates with increased organic carbon <span class="hlt">concentration</span>, indicating that the denitrification potential in these sediments was organic carbon limited. Furthermore, N2O production rates also were affected by the type of organic carbon available as an electron donor. Acetate and surface-water-derived fulvic acid supported higher N2O production rates than groundwater-derived fulvic acid or sedimentary organic carbon. Lowering incubation temperatures from 22 to 4??C resulted in about a 77% decrease in the N2O production rates. These results help to explain findings from previous studies indicating that only 15-30% of <span class="hlt">nitrate</span> in groundwater was denitrified before discharging to the South Platte River and that <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the river generally were higher in winter than in summer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24843910','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24843910"><span>Sensitivity analyses of factors influencing CMAQ performance for fine particulate <span class="hlt">nitrate</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shimadera, Hikari; Hayami, Hiroshi; Chatani, Satoru; Morino, Yu; Mori, Yasuaki; Morikawa, Tazuko; Yamaji, Kazuyo; Ohara, Toshimasa</p> <p>2014-04-01</p> <p>Improvement of air quality models is required so that they can be utilized to design effective control strategies for fine particulate matter (PM2.5). The Community Multiscale Air Quality modeling system was applied to the Greater Tokyo Area of Japan in winter 2010 and summer 2011. The model results were compared with <span class="hlt">observed</span> <span class="hlt">concentrations</span> of PM2.5 sulfate (SO4(2-)), <span class="hlt">nitrate</span> (NO3(-)) and ammonium, and gaseous nitric acid (HNO3) and ammonia (NH3). The model approximately reproduced PM2.5 SO4(2-) <span class="hlt">concentration</span>, but clearly overestimated PM2.5 NO3(-) <span class="hlt">concentration</span>, which was attributed to overestimation of production of ammonium <span class="hlt">nitrate</span> (NH4NO3). This study conducted sensitivity analyses of factors associated with the model performance for PM2.5 NO3(-) <span class="hlt">concentration</span>, including temperature and relative humidity, emission of nitrogen oxides, seasonal variation of NH3 emission, HNO3 and NH3 dry deposition velocities, and heterogeneous reaction probability of dinitrogen pentoxide. Change in NH3 emission directly affected NH3 <span class="hlt">concentration</span>, and substantially affected NH4NO3 <span class="hlt">concentration</span>. Higher dry deposition velocities of HNO3 and NH3 led to substantial reductions of <span class="hlt">concentrations</span> of the gaseous species and NH4NO3. Because uncertainties in NH3 emission and dry deposition processes are probably large, these processes may be key factors for improvement of the model performance for PM2.5 NO3(-). The Community Multiscale Air Quality modeling system clearly overestimated the <span class="hlt">concentration</span> of fine particulate <span class="hlt">nitrate</span> in the Greater Tokyo Area of Japan, which was attributed to overestimation of production of ammonium <span class="hlt">nitrate</span>. Sensitivity analyses were conducted for factors associated with the model performance for <span class="hlt">nitrate</span>. Ammonia emission and dry deposition of nitric acid and ammonia may be key factors for improvement of the model performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70036793','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70036793"><span>Identifying pathways and processes affecting <span class="hlt">nitrate</span> and orthophosphate inputs to streams in agricultural watersheds</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Tesoriero, A.J.; Duff, J.H.; Wolock, D.M.; Spahr, N.E.; Almendinger, J.E.</p> <p>2009-01-01</p> <p>Understanding nutrient pathways to streams will improve nutrient management strategies and estimates of the time lag between when changes in land use practices occur and when water quality effects that result from these changes are <span class="hlt">observed</span>. <span class="hlt">Nitrate</span> and orthophosphate (OP) <span class="hlt">concentrations</span> in several environmental compartments were examined in watersheds having a range of base flow index (BFI) values across the continental United States to determine the dominant pathways for water and nutrient inputs to streams. Estimates of the proportion of stream <span class="hlt">nitrate</span> that was derived from groundwater increased as BFI increased. <span class="hlt">Nitrate</span> <span class="hlt">concentration</span> gradients between groundwater and surface water further supported the groundwater source of <span class="hlt">nitrate</span> in these high BFI streams. However, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in stream-bed pore water in all settings were typically lower than stream or upland groundwater <span class="hlt">concentrations</span>, suggesting that <span class="hlt">nitrate</span> discharge to streams was not uniform through the bed. Rather, preferential pathways (e.g., springs, seeps) may allow high <span class="hlt">nitrate</span> groundwater to bypass sites of high biogeochemical transformation. Rapid pathway compartments (e.g., overland flow, tile drains) had OP <span class="hlt">concentrations</span> that were typically higher than in streams and were important OP conveyers in most of these watersheds. In contrast to <span class="hlt">nitrate</span>, the proportion of stream OP that is derived from ground water did not systematically increase as BFI increased. While typically not the dominant source of OP, groundwater discharge was an important pathway of OP transport to streams when BFI values were very high and when geochemical conditions favored OP mobility in groundwater. Copyright ?? 2009 by the American Society of Agronomy, Crop Science Society of America, and Soil Science Society of America. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22165219','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22165219"><span>[Photodegradation of UV filter PABA in <span class="hlt">nitrate</span> solution].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Meng, Cui; Ji, Yue-Fei; Zeng, Chao; Yang, Xi</p> <p>2011-09-01</p> <p>The aqueous photolysis of a UV filter p-aminobenzoic acid (PABA) using Xe lamp as simulated solar irradiation source was investigated in the presence of <span class="hlt">nitrate</span> ions. The effects of pH, <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> ions and <span class="hlt">concentration</span> of humic substance in natural water on the photodegradation of PABA were studied. The results showed that photodegradation of PABA in <span class="hlt">nitrate</span> solution followed the first order kinetics. The increasing <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> ion increased favored the photodegradaton of PABA, of which the first order constant increased from 0.002 2 min(-10 to 0.017 9 min(-1). The photodegradation of PABA promoted with the increase of pH while the increasing <span class="hlt">concentration</span> of humic substance showed inhibiting effect. Hydroxyl radicals determined by the molecular probe method played a very importnant role in the photolysis process of PABA. Photoproducts upon irradiation of PABA in <span class="hlt">nitrate</span> solution were isolated by means of solid-phase extraction (SPE) and identified by LC-MS techniques. The probable photoinduced degradation pathways in <span class="hlt">nitrate</span> solution were proposed.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.H43K..07H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.H43K..07H"><span>Fabrication and In Situ Testing of Scalable <span class="hlt">Nitrate</span>-Selective Electrodes for Distributed <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Harmon, T. C.; Rat'ko, A.; Dietrich, H.; Park, Y.; Wijsboom, Y. H.; Bendikov, M.</p> <p>2008-12-01</p> <p>Inorganic nitrogen (<span class="hlt">nitrate</span> (NO3-) and ammonium (NH+)) from chemical fertilizer and livestock waste is a major source of pollution in groundwater, surface water and the air. While some sources of these chemicals, such as waste lagoons, are well-defined, their application as fertilizer has the potential to create distributed or non-point source pollution problems. Scalable <span class="hlt">nitrate</span> sensors (small and inexpensive) would enable us to better assess non-point source pollution processes in agronomic soils, groundwater and rivers subject to non-point source inputs. This work describes the fabrication and testing of inexpensive PVC-membrane- based ion selective electrodes (ISEs) for monitoring <span class="hlt">nitrate</span> levels in soil water environments. ISE-based sensors have the advantages of being easy to fabricate and use, but suffer several shortcomings, including limited sensitivity, poor precision, and calibration drift. However, modern materials have begun to yield more robust ISE types in laboratory settings. This work emphasizes the in situ behavior of commercial and fabricated sensors in soils subject to irrigation with dairy manure water. Results are presented in the context of deployment techniques (in situ versus soil lysimeters), temperature compensation, and uncertainty analysis. <span class="hlt">Observed</span> temporal responses of the <span class="hlt">nitrate</span> sensors exhibited diurnal cycling with elevated <span class="hlt">nitrate</span> levels at night and depressed levels during the day. Conventional samples collected via lysimeters validated this response. It is concluded that while modern ISEs are not yet ready for long-term, unattended deployment, short-term installations (on the order of 2 to 4 days) are viable and may provide valuable insights into nitrogen dynamics in complex soil systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.B42D..01G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.B42D..01G"><span>Drivers of inverse DOC-<span class="hlt">nitrate</span> loss patterns in forest soils and streams</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Goodale, C. L.</p> <p>2013-12-01</p> <p><span class="hlt">Nitrate</span> loss from forested catchments varies greatly across sites and over time, with few reliable correlates. One of the few recurring patterns, however, is the negative nonlinear relationship that occurs regularly between surface water <span class="hlt">nitrate</span> and dissolved organic carbon (DOC) <span class="hlt">concentrations</span>: that is, <span class="hlt">nitrate</span> declines sharply as DOC <span class="hlt">concentrations</span> increase, and high <span class="hlt">nitrate</span> levels occur only at low DOC <span class="hlt">concentrations</span>. Several hypotheses have been proposed to explain this pattern, but its cause has remained speculative. It is likely to be driven by C- or N-limitation of biological processes such as assimilation or denitrification, but the identity of which biological process or the main landscape position of their activity are not known. We examined whether DOC and <span class="hlt">nitrate</span> are both driven by soil C content, at scales of both soil blocks and across catchments, by measuring soil, soil extract, and surface water chemistry across nine catchments selected from long-term monitoring networks in the Catskill and Adirondack Mountains. We measured soil C and N status and solution <span class="hlt">nitrate</span>, DOC, bioavailable DOC (bDOC), and isotopic composition (13C-DOC, 15N- and 18O-NO3) to examine whether variation in stocks of soil C partly controls DOC and <span class="hlt">nitrate</span> loss from forested catchments in New York State. These measurements showed that surface soil C and C:N ratio together determine soil production of DOC and <span class="hlt">nitrate</span>, reflecting assimilative demand for N by heterotrophic microbes. Yet, they also show that these processes do not produce the inverse DOC-NO3 curve <span class="hlt">observed</span> at the catchment scale. Rather, catchment-scale DOC-<span class="hlt">nitrate</span> patterns are more likely to be governed by the balance between excess <span class="hlt">nitrate</span> production and its bDOC-mediated loss to denitrification.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/7073164','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/7073164"><span>Thermochemical <span class="hlt">nitrate</span> destruction</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Cox, J.L.; Hallen, R.T.; Lilga, M.A.</p> <p>1992-06-02</p> <p>A method is disclosed for denitrification of <span class="hlt">nitrates</span> and nitrites present in aqueous waste streams. The method comprises the steps of (1) identifying the <span class="hlt">concentration</span> <span class="hlt">nitrates</span> and nitrites present in a waste stream, (2) causing formate to be present in the waste stream, (3) heating the mixture to a predetermined reaction temperature from about 200 C to about 600 C, and (4) holding the mixture and accumulating products at heated and pressurized conditions for a residence time, thereby resulting in nitrogen and carbon dioxide gas, and hydroxides, and reducing the level of <span class="hlt">nitrates</span> and nitrites to below drinking water standards.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.water.usgs.gov/wri004137','USGSPUBS'); return false;" href="http://pubs.water.usgs.gov/wri004137"><span><span class="hlt">Nitrate</span> source indicators in ground water of the Scimitar Subdivision, Peters Creek area, Anchorage, Alaska</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Wang, Bronwen; Strelakos, Pat M.; Jokela, Brett</p> <p>2000-01-01</p> <p>A combination of aqueous chemistry, isotopic measurement, and in situ tracers were used to study the possible <span class="hlt">nitrate</span> sources, the factors contributing to the spatial distribution of <span class="hlt">nitrate</span>, and possible septic system influence in the ground water in the Scimitar Subdivision, Municipality of Anchorage, Alaska. Two water types were distinguished on the basis of the major ion chemistry: (1) a calcium sodium carbonate water, which was associated with isotopically heavier boron and with chlorofluorocarbons (CFC's) that were in the range expected from equilibration with the atmosphere (group A water) and (2) a calcium magnesium carbonate water, which was associated with elevated <span class="hlt">nitrate</span>, chloride, and magnesium <span class="hlt">concentrations</span>, generally isotopically lighter boron, and CFC's <span class="hlt">concentrations</span> that were generally in excess of that expected from equilibration with the atmosphere (group B water). Water from wells in group B had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> that were greater than 3 milligrams per liter, whereas those in group A had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> of 0.2 milligram per liter or less. <span class="hlt">Nitrate</span> does not appear to be undergoing extensive transformation in the ground-water system and behaves as a conservative ion. The major ion chemistry trends and the presence of CFC's in excess of an atmospheric source for group B wells are consistent with waste-water influences. The spatial distribution of the <span class="hlt">nitrate</span> among wells is likely due to the magnitude of this influence on any given well. Using an expanded data set composed of 16 wells sampled only for <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, a significant difference in the static water level relative to bedrock was found. Well water samples with less than 1 milligram per liter <span class="hlt">nitrate</span> had static water levels within the bedrock, whereas those samples with greater than 1 milligram per liter <span class="hlt">nitrate</span> had static water levels near or above the top of the bedrock. This <span class="hlt">observation</span> would be consistent with a conceptual model of a low-<span class="hlt">nitrate</span> fractured bedrock</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.B11O..05A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.B11O..05A"><span>An unexpected truth: increasing <span class="hlt">nitrate</span> loading can decrease <span class="hlt">nitrate</span> export from watersheds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Askarizadeh Bardsiri, A.; Grant, S. B.; Rippy, M.</p> <p>2015-12-01</p> <p>The discharge of anthropogenic <span class="hlt">nitrate</span> (e.g., from partially treated sewage, return flows from agricultural irrigation, and runoff from animal feeding operations) to streams can negatively impact both human and ecosystem health. Managing these many point and non-point sources to achieve some specific end-point—for example, reducing the annual mass of <span class="hlt">nitrate</span> exported from a watershed—can be a challenge, particularly in rapidly growing urban areas. Adding to this complexity is the fact that streams are not inert: they too can add or remove <span class="hlt">nitrate</span> through assimilation (e.g., by stream-associated plants and animals) and microbially-mediated biogeochemical reactions that occur in streambed sediments (e.g., respiration, ammonification, nitrification, denitrification). By coupling a previously published correlation for in-stream processing of <span class="hlt">nitrate</span> [Mulholland et al., Nature, 2008, 452, 202-205] with a stream network model of the Jacksons Creek watershed (Victoria, Australia) I demonstrate that managing anthropogenic sources of stream <span class="hlt">nitrate</span> without consideration of in-stream processing can result in a number of non-intuitive "surprises"; for example, wastewater effluent discharges that increase <span class="hlt">nitrate</span> loading but decrease in-stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> can reduce the mass of <span class="hlt">nitrate</span> exported from a watershed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24211765','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24211765"><span>Total salivary <span class="hlt">nitrates</span> and nitrites in oral health and periodontal disease.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sánchez, Gabriel A; Miozza, Valeria A; Delgado, Alejandra; Busch, Lucila</p> <p>2014-01-30</p> <p>It is well known that nitrites are increased in saliva from patients with periodontal disease. In the oral cavity, nitrites may derive partly from the reduction of <span class="hlt">nitrates</span> by oral bacteria. <span class="hlt">Nitrates</span> have been reported as a defence-related mechanism. Thus, the aim of the present study was to determine the salivary levels of total <span class="hlt">nitrate</span> and nitrite and their relationship, in unstimulated and stimulated saliva from periodontal healthy subjects, and from patients with chronic periodontal disease. <span class="hlt">Nitrates</span> and nitrites were determined in saliva from thirty healthy subjects and forty-four patients with periodontal disease. A significant increase in salivary <span class="hlt">nitrates</span> and nitrites was <span class="hlt">observed</span>. <span class="hlt">Nitrates</span> and nitrites <span class="hlt">concentration</span> was related to clinical attachment level (CAL). A positive and significant Pearson's correlation was found between salivary total <span class="hlt">nitrates</span> and nitrites. Periodontal treatment induced clinical improvement and decreased <span class="hlt">nitrates</span> and nitrites. It is concluded that salivary <span class="hlt">nitrates</span> and nitrites increase, in patients with periodontal disease, could be related to defence mechanisms. The possibility that the salivary glands respond to oral infectious diseases by increasing <span class="hlt">nitrate</span> secretion should be explored further. Copyright © 2013 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014NatCC...4..477J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014NatCC...4..477J"><span><span class="hlt">Nitrate</span> assimilation is inhibited by elevated CO2 in field-grown wheat</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>J. Bloom, Arnold; Burger, Martin; A. Kimball, Bruce; J. Pinter, Paul, Jr.</p> <p>2014-06-01</p> <p>Total protein and nitrogen <span class="hlt">concentrations</span> in plants generally decline under elevated CO2 atmospheres. Explanations for this decline include that plants under elevated CO2 grow larger, diluting the protein within their tissues; that carbohydrates accumulate within leaves, downregulating the amount of the most prevalent protein Rubisco; that carbon enrichment of the rhizosphere leads to progressively greater limitations of the nitrogen available to plants; and that elevated CO2 directly inhibits plant nitrogen metabolism, especially the assimilation of <span class="hlt">nitrate</span> into proteins in leaves of C3 plants. Recently, several meta-analyses have indicated that CO2 inhibition of <span class="hlt">nitrate</span> assimilation is the explanation most consistent with <span class="hlt">observations</span>. Here, we present the first direct field test of this explanation. We analysed wheat (Triticum aestivum L.) grown under elevated and ambient CO2 <span class="hlt">concentrations</span> in the free-air CO2 enrichment experiment at Maricopa, Arizona. In leaf tissue, the ratio of <span class="hlt">nitrate</span> to total nitrogen <span class="hlt">concentration</span> and the stable isotope ratios of organic nitrogen and free <span class="hlt">nitrate</span> showed that <span class="hlt">nitrate</span> assimilation was slower under elevated than ambient CO2. These findings imply that food quality will suffer under the CO2 levels anticipated during this century unless more sophisticated approaches to nitrogen fertilization are employed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27325522','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27325522"><span>Thermal Storage Properties of Molten <span class="hlt">Nitrate</span> Salt-Based Nanofluids with Graphene Nanoplatelets.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xie, Qiangzhi; Zhu, Qunzhi; Li, Yan</p> <p>2016-12-01</p> <p>In this study, the effect of <span class="hlt">concentration</span> of nanoparticles on the thermal storage properties of molten <span class="hlt">nitrate</span> salt-based nanofluids with graphene nanoplatelets (GNPs) was investigated. Solar salt consisting of sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span> was utilized as the base material for the nanofluids. Homogeneous dispersion of GNPs within the solar salt was <span class="hlt">observed</span> through scanning electron microscopy analysis. For both solar salt and resultant nanofluids, differential scanning calorimetry was employed to measure the thermal storage properties, including characteristic temperatures of phase change, startup heat, and specific heat capacity (SHC). A maximum increase of 16.7 % in SHC at the liquid phase was found at an optimal <span class="hlt">concentration</span> of 1 wt% of GNPs. At the same <span class="hlt">concentration</span>, the onset temperature decreased by 10.4 °C, the endset temperature decreased by 4.7 °C, and the startup heat decreased by 9 %.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1572527','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1572527"><span>Impaired vasodilator response to organic <span class="hlt">nitrates</span> in isolated basilar arteries</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Martens, Dorothee; Kojda, Georg</p> <p>2001-01-01</p> <p>The differential responsiveness of various sections and regions in the vascular system to the vasodilator activity of organic <span class="hlt">nitrates</span> is important for the beneficial antiischaemic effects of these drugs. In this study we examined the vasodilator activity of organic <span class="hlt">nitrates</span> in cerebral arteries, where vasodilation causes substantial <span class="hlt">nitrate</span> induced headache. Isolated porcine basilar and coronary arteries were subjected to increasing <span class="hlt">concentrations</span> of glyceryl trinitrate (GTN), isosorbide-5-<span class="hlt">nitrate</span> (ISMN) and pentaerythritol tetranitrate (PETN). S-nitroso-N-acetyl-D,L-penicillamine (SNAP) and endothelium-dependent vasodilation was investigated for comparison purpose. The vasodilator potency (halfmaximal effective <span class="hlt">concentration</span> in −logM) of GTN (4.33±0.1, n=8), ISMN (1.61±0.07, n=7) and PETN (>10 μM, n=7) in basilar arteries was more than 100 fold lower than that of GTN (6.52±0.06, n=12), ISMN (3.66±0.08, n=10) and PETN (6.3±0.13, n=8) <span class="hlt">observed</span> in coronary arteries. In striking contrast, the vasodilator potency of SNAP (halfmaximal effective <span class="hlt">concentration</span> in −logM) was almost similar in basilar (7.76±0.05, n=7) and coronary arteries (7.59±0.05, n=9). Likewise, no difference in endothelium dependent relaxation was <span class="hlt">observed</span>. Denudation of the endothelium resulted in a small increase of the vasodilator potency (halfmaximal effective <span class="hlt">concentration</span> in −logM) of GTN (4.84±0.09, n=7, P<0.03) in basilar arteries and similar results were obtained in the presence of the NO-synthase inhibitor Nω-nitro-L-arginine (4.59±0.05, n=9, P<0.03). These results suggest that cerebral conductance blood vessels such as porcine basilar arteries seems to have a reduced expression and/or activity of certain cellular enzymatic electron transport systems such as cytochrome P450 enzymes, which are necessary to bioconvert organic <span class="hlt">nitrates</span> to NO. PMID:11156558</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..16.2094S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..16.2094S"><span>Identification of groundwater <span class="hlt">nitrate</span> sources in pre-alpine catchments: a multi-tracer approach</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stoewer, Myriam; Stumpp, Christine</p> <p>2014-05-01</p> <p>Porous aquifers in pre-alpine areas are often used as drinking water resources due to their good water quality status and water yield. Maintaining these resources requires knowledge about possible sources of pollutants and a sustainable management practice in groundwater catchment areas. Of particular interest in agricultural areas, like in pre-alpine regions, is limiting <span class="hlt">nitrate</span> input as main groundwater pollutant. Therefore, the objective of the presented study is i) to identify main <span class="hlt">nitrate</span> sources in a pre-alpine groundwater catchment with current low <span class="hlt">nitrate</span> <span class="hlt">concentration</span> using stable isotopes of <span class="hlt">nitrate</span> (d18O and d15N) and ii) to investigate seasonal dynamics of nitrogen compounds. The groundwater catchment areas of four porous aquifers are located in Southern Germany. Most of the land use is organic grassland farming as well as forestry and residential area. Thus, potential sources of <span class="hlt">nitrate</span> mainly are mineral fertilizer, manure/slurry, leaking sewage system and atmospheric deposition of nitrogen compounds. Monthly freshwater samples (precipitation, river water and groundwater) are analysed for stable isotope of water (d2H, d18O), the <span class="hlt">concentration</span> of major anions and cations, electrical conductivity, water temperature, pH and oxygen. In addition, isotopic analysis of d18O-NO3- and d15N-NO3- for selected samples is carried out using the denitrifier method. In general, all groundwater samples were oxic (10.0±2.6mg/L) and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were low (0.2 - 14.6mg/L). The <span class="hlt">observed</span> <span class="hlt">nitrate</span> isotope values in the <span class="hlt">observation</span> area compared to values from local precipitation, sewage, manure and mineral fertilizer as well as to data from literature shows that the <span class="hlt">nitrate</span> in freshwater samples is of microbial origin. <span class="hlt">Nitrate</span> derived from ammonium in fertilizers and precipitation as well as from soil nitrogen. It is suggested that a major potential threat to the groundwater quality is ammonia and ammonium at a constant level mainly from agriculture activities as</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?direntryid=151143&sitype=pr&','PESTICIDES'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?direntryid=151143&sitype=pr&"><span>CARBON-BASED REACTIVE BARRIER FOR <span class="hlt">NITRATE</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><span class="hlt">Nitrate</span> (NO3-) is a common ground water contaminant related to agricultural activity, waste water disposal, leachate from landfills, septic systems, and industrial processes. This study reports on the performance of a carbon-based permeable reactive barrier (PRB) that was constructed for in-situ bioremediation of a ground water <span class="hlt">nitrate</span> plume caused by leakage from a swine CAFO (<span class="hlt">concentrated</span> animal feeding operation) lagoon. The swine CAFO, located in Logan County, Oklahoma, was in operation from 1992-1999. The overall site remediation strategy includes an ammonia recovery trench to intercept ammonia-contaminated ground water and a hay straw PRB which is used to intercept a <span class="hlt">nitrate</span> plume caused by nitrification of sorbed ammonia. The PRB extends approximately 260 m to intercept the <span class="hlt">nitrate</span> plume. The depth of the trench averages 6 m and corresponds to the thickness of the surficial saturated zone; the width of the trench is 1.2 m. Detailed quarterly monitoring of the PRB began in March, 2004, about 1 year after construction activities ended. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> hydraulically upgradient of the PRB have ranged from 23 to 77 mg/L N, from 0 to 3.2 mg/L N in the PRB, and from 0 to 65 mg/L N hydraulically downgradient of the PRB. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> have generally decreased in downgradient locations with successive monitoring events. Mass balance considerations indicate that <span class="hlt">nitrate</span> attenuation is dominantly from denitrification but with some component of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1914306M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1914306M"><span>Identification of <span class="hlt">nitrate</span> sources and discharge-depending <span class="hlt">nitrate</span> dynamics in a mesoscale catchment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mueller, Christin; Strachauer, Ulrike; Brauns, Mario; Musolff, Andreas; Kunz, Julia Vanessa; Brase, Lisa; Tarasova, Larisa; Merz, Ralf; Knöller, Kay</p> <p>2017-04-01</p> <p>During the last decades, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in surface and groundwater have increased due to land use change and accompanying application of fertilizer in agriculture as well as increased atmospheric deposition. To mitigate nutrient impacts on downstream aquatic ecosystems, it is important to quantify potential <span class="hlt">nitrate</span> sources, instream <span class="hlt">nitrate</span> processing and its controls in a river system. The objective of this project is to characterize and quantify (regional) scale dynamics and trends in water and nitrogen fluxes of the entire Holtemme river catchment in central Germany making use of isotopic fingerprinting methods. Here we compare two key date sampling campaigns in 2014 and 2015, with spatially highly resolved measurements of discharge at 23 sampling locations including 11 major tributaries and 12 locations at the main river. Additionally, we have data from continuous runoff measurements at 10 locations operated by the local water authorities. Two waste water treatment plants contribute nitrogen to the Holtemme stream. This contribution impacts <span class="hlt">nitrate</span> loads and <span class="hlt">nitrate</span> isotopic signatures depending on the prevailing hydrological conditions. Nitrogen isotopic signatures in the catchment are mainly controlled by different sources (nitrified soil nitrogen in the headwater and manure/ effluents from WWTPs in the lowlands) and increase with raising <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> along the main river. <span class="hlt">Nitrate</span> loads at the outlet of the catchment are extremely different between both sampling campaigns (2014: NO3- = 97 t a-1, 2015: NO3- = 5 t a-1) which is associated with various runoff (2014: 0.8 m3 s-1, 2015: 0.2 m3 s-1). In 2015, the inflow from WWTP's raises the NO3- loads and enriches δ18O-NO3 values. Generally, oxygen isotope signatures from <span class="hlt">nitrate</span> are more variable and are controlled by biogeochemical processes in concert with the oxygen isotopic composition of the ambient water. Elevated δ18O-NO3 in 2015 are most likely due to higher temperatures and lower</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1710824B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1710824B"><span>Variability of pesticides and <span class="hlt">nitrates</span> <span class="hlt">concentrations</span> along a river transect: chemical and isotopic evidence of groundwater - surface water interconnections</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Baran, Nicole; Petelet-Giraud, Emmanuelle; Saplairoles, Maritxu</p> <p>2015-04-01</p> <p>Groundwater quality is increasingly monitored in Europe where various levels of <span class="hlt">nitrate</span> and pesticide and/or metabolite contamination have been demonstrated (Loos et al., 2010, Stuart et al., 2012). The Groundwater Daughter Directive (2006/118/EC) to Water Framework Directive (WFD) particularly requires measures to prevent or limit inputs of pollutants into groundwater and compliance with good chemical status criteria (based on EU standards of <span class="hlt">nitrate</span> and pesticides). The WFD mentioned the need to protect groundwater but also to have a particular regard to its impact and interrelationship with associated surface waters and directly dependent terrestrial Ecosystems. The Ariège river basin (SW France - 538 km²) is an alluvial plain under high agricultural pressure leading to a contamination of the aquifer by several pesticides and metabolites (Amalric et al., 2013). The Crieu is an allochtone river, crossing the plain (~ 10 km length) before joining the Ariège River. The Crieu is often dry in its middle section suggesting water leakage from surface water towards groundwater. At the opposite, the permanent flow <span class="hlt">observed</span> downstream suggests an input of groundwater into surface water. In May 2014, while the Crieu flow was continuous through the plain, 7 river samples were collected and analyzed for pesticides, major ions, strontium <span class="hlt">concentration</span> and isotopes. In situ measurements of electric conductivity were also performed as well as flow gauging. Two groundwaters close to the river were also sampled. The flow gauging measurements show a decreasing river discharge in the central area of the Crieu River, suggesting surface water leakage towards groundwater. Nevertheless, the electric conductivity increases along the river flow as well as some pesticides and <span class="hlt">nitrates</span> <span class="hlt">concentrations</span>. This chemical evolution of the river water is thus inconsistent with a simple water infiltration and another source of dissolved solutes is required to explain the increased of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27486968','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27486968"><span>Association between dietary <span class="hlt">nitrate</span> and nitrite intake and sitespecific cancer risk: evidence from <span class="hlt">observational</span> studies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xie, Li; Mo, Miao; Jia, Hui-Xun; Liang, Fei; Yuan, Jing; Zhu, Ji</p> <p>2016-08-30</p> <p>Epidemiological studies have reported inconsistent findings on the association between dietary <span class="hlt">nitrate</span> and nitrite intake and cancer risk. We performed a meta-analysis of epidemiological studies to summarize available evidence on the association between dietary <span class="hlt">nitrate</span> and nitrite intake and cancer risk from published prospective and case-control studies. PubMed database was searched to identify eligible publications through April 30th, 2016. Study-specific relative risks (RRs) with corresponding 95% confidence interval (CI) from individual studies were pooled by using random- or fixed- model, and heterogeneity and publication bias analyses were conducted. Data from 62 <span class="hlt">observational</span> studies, 49 studies for <span class="hlt">nitrates</span> and 51 studies for nitrites, including a total of 60,627 cancer cases were analyzed. Comparing the highest vs. lowest levels, dietary <span class="hlt">nitrate</span> intake was inversely associated with gastric cancer risk (RR = 0.78; 95%CI = 0.67-0.91) with moderate heterogeneity (I2 = 42.3%). In contrast, dietary nitrite intake was positively associated with adult glioma and thyroid cancer risk with pooled RR of 1.21 (95%CI = 1.03-1.42) and 1.52 (95%CI = 1.12-2.05), respectively. No significant associations were found between dietary <span class="hlt">nitrate</span>/nitrite and cancers of the breast, bladder, colorectal, esophagus, renal cell, non-Hodgkin lymphoma, ovarian, and pancreas. The present meta-analysis provided modest evidence that positive associations of dietary <span class="hlt">nitrate</span> and negative associations of dietary nitrite with certain cancers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C11D..02D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C11D..02D"><span><span class="hlt">Nitrate</span> photolysis in salty snow</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Donaldson, D. J.; Morenz, K.; Shi, Q.; Murphy, J. G.</p> <p>2016-12-01</p> <p><span class="hlt">Nitrate</span> photolysis from snow can have a significant impact on the oxidative capacity of the local atmosphere, but the factors affecting the release of gas phase products are not well understood. Here, we report the first systematic study of the amounts of NO, NO2, and total nitrogen oxides (NOy) emitted from illuminated snow samples as a function of both <span class="hlt">nitrate</span> and total salt (NaCl and Instant Ocean) <span class="hlt">concentration</span>. We show that the release of nitrogen oxides to the gas phase is directly related to the expected <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the brine at the surface of the snow crystals, increasing to a plateau value with increasing <span class="hlt">nitrate</span>, and generally decreasing with increasing NaCl or Instant Ocean (I.O.). In frozen mixed <span class="hlt">nitrate</span> (25 mM) - salt (0-500 mM) solutions, there is an increase in gas phase NO2 seen at low added salt amounts: NO2 production is enhanced by 35% at low prefreezing [NaCl] and by 70% at similar prefreezing [I.O.]. Raman microscopy of frozen <span class="hlt">nitrate</span>-salt solutions shows evidence of stronger <span class="hlt">nitrate</span> exclusion to the air interface in the presence of I.O. than with added NaCl. The enhancement in nitrogen oxides emission in the presence of salts may prove to be important to the atmospheric oxidative capacity in polar regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=307558','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=307558"><span>Modeling <span class="hlt">nitrate</span> removal in a denitrification bed</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>Denitrification beds are being promoted to reduce <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in agricultural drainage water to alleviate the adverse environmental effects associated with <span class="hlt">nitrate</span> pollution in surface water. In this system, water flows through a trench filled with a carbon media where <span class="hlt">nitrate</span> is transfor...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2013/5150/pdf/sir2013-5150.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2013/5150/pdf/sir2013-5150.pdf"><span>Estimating <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater at selected wells and springs in the surficial aquifer system and Upper Floridan aquifer, Dougherty Plain and Marianna Lowlands, Georgia, Florida, and Alabama, 2002-50</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Crandall, Christy A.; Katz, Brian G.; Berndt, Marian P.</p> <p>2013-01-01</p> <p>Groundwater from the surficial aquifer system and Upper Floridan aquifer in the Dougherty Plain and Marianna Lowlands in southwestern Georgia, northwestern Florida, and southeastern Alabama is affected by elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> as a result of the vulnerability of the aquifer, irrigation water-supply development, and intensive agricultural land use. The region relies primarily on groundwater from the Upper Floridan aquifer for drinking-water and irrigation supply. Elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in drinking water are a concern because infants under 6 months of age who drink water containing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above the U.S. Environmental Protection Agency maximum contaminant level of 10 milligrams per liter as nitrogen can become seriously ill with blue baby syndrome. In response to concerns about water quality in domestic wells and in springs in the lower Apalachicola–Chattahoochee–Flint River Basin, the Florida Department of Environmental Protection funded a study in cooperation with the U.S. Geological Survey to examine water quality in groundwater and springs that provide base flow to the Chipola River. A three-dimensional, steady-state, regional-scale groundwater-flow model and two local-scale models were used in conjunction with particle tracking to identify travel times and areas contributing recharge to six groundwater sites—three long-term monitor wells (CP-18A, CP-21A, and RF-41) and three springs (Jackson Blue Spring, Baltzell Springs Group, and Sandbag Spring) in the lower Apalachicola–Chattahoochee–Flint River Basin. Estimated <span class="hlt">nitrate</span> input to groundwater at land surface, based on previous studies of nitrogen fertilizer sales and atmospheric <span class="hlt">nitrate</span> deposition data, were used in the advective transport models for the period 2002 to 2050. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in groundwater samples collected from the six sites during 1993 to 2007 and groundwater age tracer data were used to calibrate the transport aspect of the simulations</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1917257K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1917257K"><span><span class="hlt">Nitrate</span> Contamination in the groundwater of the Lake Acıgöl Basin, SW Turkey</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Karaman, Muhittin; Budakoǧlu, Murat; Taşdelen, Suat</p> <p>2017-04-01</p> <p>The lacustrine Acıgöl basin, formed as an extensional half-graben, hosts various bodies of water, such as cold-hot springs, lakes, streams, and wells. The hydrologically closed basin contains a hypersaline lake (Lake Acıgöl) located in the southern part of the basin. The brackish springs and deep waters discharged along the Acıgöl fault zone in the southern part of the basin feed the hypersaline lake. Groundwater is used as drinking, irrigation, and domestic water in the closed Acıgöl Basin. Groundwater flows into the hypersaline lake from the highland. The Acıgöl basin hosts large plains (Hambat, Başmakçı, and Evciler). Waters in agricultural areas contain high amounts of <span class="hlt">nitrate</span>; groundwater samples in agricultural areas contain <span class="hlt">nitrate</span> levels higher than 10 mg/L. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in the groundwater samples varied from 0 to 487 mg/L (n=165); 25.4 % of the groundwater samples from the basin had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above 10 mg/L (the WHO drinking guideline) and 52.2% of the groundwater samples from the basin had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above 3.0 mg/L, and these high values were regarded as the result of human activity. The highest <span class="hlt">nitrate</span> values were measured in the Hambat plain (480 and 100 mg/L) and Yirce Pinari spring (447 mg/L), which discharges along the Acıgöl fault zone in the southern part of the basin. The average multi-temporal <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of the Yirce Pınarı spring was 3.3 mg/L. Extreme <span class="hlt">nitrate</span> values were measured in the Yirce Pınarı spring during periods when sheep wool was washed (human activity). The lowest <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were <span class="hlt">observed</span> in some springs that discharged along the Acıgöl fault zone in the southern part of the basin. <span class="hlt">Nitrate</span> was not detected in deep groundwater discharged along the Acıgöl fault zone. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in deep groundwater and some springs discharged along the Acıgöl fault zone and those feeding the hypersaline lake were significantly affected by redox conditions</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5298580','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5298580"><span>Characteristics of the Fiber Laser Sensor System Based on Etched-Bragg Grating Sensing Probe for Determination of the Low <span class="hlt">Nitrate</span> <span class="hlt">Concentration</span> in Water</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Pham, Thanh Binh; Bui, Huy; Le, Huu Thang; Pham, Van Hoi</p> <p>2016-01-01</p> <p>The necessity of environmental protection has stimulated the development of many kinds of methods allowing the determination of different pollutants in the natural environment, including methods for determining <span class="hlt">nitrate</span> in source water. In this paper, the characteristics of an etched fiber Bragg grating (e-FBG) sensing probe—which integrated in fiber laser structure—are studied by numerical simulation and experiment. The proposed sensor is demonstrated for determination of the low <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in a water environment. Experimental results show that this sensor could determine <span class="hlt">nitrate</span> in water samples at a low <span class="hlt">concentration</span> range of 0–80 ppm with good repeatability, rapid response, and average sensitivity of 3.5 × 10−3 nm/ppm with the detection limit of 3 ppm. The e-FBG sensing probe integrated in fiber laser demonstrates many advantages, such as a high resolution for wavelength shift identification, high optical signal-to-noise ratio (OSNR of 40 dB), narrow bandwidth of 0.02 nm that enhanced accuracy and precision of wavelength peak measurement, and capability for optical remote sensing. The obtained results suggested that the proposed e-FBG sensor has a large potential for the determination of low <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in water in outdoor field work. PMID:28025512</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28025512','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28025512"><span>Characteristics of the Fiber Laser Sensor System Based on Etched-Bragg Grating Sensing Probe for Determination of the Low <span class="hlt">Nitrate</span> <span class="hlt">Concentration</span> in Water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pham, Thanh Binh; Bui, Huy; Le, Huu Thang; Pham, Van Hoi</p> <p>2016-12-22</p> <p>The necessity of environmental protection has stimulated the development of many kinds of methods allowing the determination of different pollutants in the natural environment, including methods for determining <span class="hlt">nitrate</span> in source water. In this paper, the characteristics of an etched fiber Bragg grating (e-FBG) sensing probe-which integrated in fiber laser structure-are studied by numerical simulation and experiment. The proposed sensor is demonstrated for determination of the low <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in a water environment. Experimental results show that this sensor could determine <span class="hlt">nitrate</span> in water samples at a low <span class="hlt">concentration</span> range of 0-80 ppm with good repeatability, rapid response, and average sensitivity of 3.5 × 10 -3 nm/ppm with the detection limit of 3 ppm. The e-FBG sensing probe integrated in fiber laser demonstrates many advantages, such as a high resolution for wavelength shift identification, high optical signal-to-noise ratio (OSNR of 40 dB), narrow bandwidth of 0.02 nm that enhanced accuracy and precision of wavelength peak measurement, and capability for optical remote sensing. The obtained results suggested that the proposed e-FBG sensor has a large potential for the determination of low <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in water in outdoor field work.</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/2017AGUFM.A53F2338P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A53F2338P"><span>Investigating a Sulphate-<span class="hlt">Nitrate</span> Chemical Indirect Effect over Europe from 1980-2010</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pearce, H.; Mann, G. W.; Arnold, S.; O'Connor, F.; Conibear, L.; Turnock, S.; Rumbold, S.; Benduhn, F.</p> <p>2017-12-01</p> <p>Sulphur dioxide emission reductions have been successful in reducing surface sulphate <span class="hlt">concentrations</span> over Europe between 1980 and 2010, with positive implications for air quality and human health. However the response of <span class="hlt">nitrate</span> aerosol <span class="hlt">concentrations</span> to declining NOx emissions has been non-linear. Previous studies have indicated that decreasing ammonium sulphate formation, as a result of SO2 emission reduction, may be partly responsible for this non-linearity by increasing the availability of ammonia and, hence, indirectly increasing ammonium <span class="hlt">nitrate</span> aerosol formation. We use the UM-UKCA composition-climate model, including the GLOMAP interactive aerosol microphysics module and a recently developed `hybrid' dissolution solver (HyDis), to investigate the size-resolved partitioning of ammonia and nitric acid to the particle phase over Europe in the period 1980 to 2010. Anthropogenic emissions of SO2, NOx and NH3 are included from the MACCity inventory and change by approximately -79%, -33% and +30% respectively over Europe in this time. We evaluate the UM-UKCA simulated 1980-2010 variability in <span class="hlt">nitrate</span>, ammonium and sulphate aerosol mass <span class="hlt">concentrations</span> and aerosol pH, with comparison to EMEP <span class="hlt">observations</span>, and isolate the indirect influence of reduced SO2 emissions on <span class="hlt">nitrate</span> formation. Preliminary sensitivity tests indicate that simulated <span class="hlt">nitrate</span> aerosol <span class="hlt">concentrations</span> over Europe were 8% higher in 2009 than they would have been if SO2 emissions had not been reduced. The implications of this change for air quality, aerosol acidity and regional climate will be presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.8401V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.8401V"><span>Oxygen isotope dynamics of atmospheric <span class="hlt">nitrate</span> over the Antarctic plateau: First combined measurements of ozone and <span class="hlt">nitrate</span> 17O-excess (Δ17O)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vicars, William; Savarino, Joël; Erbland, Joseph; Preunkert, Susanne; Jourdain, Bruno; Frey, Markus; Gil, Jaime; Legrand, Michel</p> <p>2013-04-01</p> <p>Variations in the isotopic composition of atmospheric <span class="hlt">nitrate</span> (NO3-) provide novel indicators for important processes in boundary layer chemistry, often acting as source markers for reactive nitrogen (NOx = NO + NO2) and providing both qualitative and quantitative constraints on the pathways that determine its fate. Stable isotope ratios of <span class="hlt">nitrate</span> (δ15N, δ17O, δ18O) offer direct insight into the nature and magnitude of the fluxes associated with different processes, thus providing unique information regarding phenomena that are often difficult to quantify from <span class="hlt">concentration</span> measurements alone. The unique and distinctive 17O-excess (Δ17O = δ17O - 0.52 × δ18O ) of ozone (O3), which is transferred to NOx via oxidation reactions in the atmosphere, has been found to be a particularly useful isotopic fingerprint in studies of NOx transformations. Constraining the propagation of 17O-excess within the NOx cycle is critical in polar areas where there exists the possibility of extending atmospheric interpretations to the glacial/interglacial time scale using deep ice core records of <span class="hlt">nitrate</span>. Here we present measurements of the comprehensive isotopic composition of atmospheric <span class="hlt">nitrate</span> collected at Dome C, Antarctica during December 2011 to January 2012. Sampling was conducted within the framework of the OPALE (Oxidant Production over Antarctic Land and its Export) project, thus providing an opportunity to combine our isotopic <span class="hlt">observations</span> with a wealth of meteorological and chemical data, including in-situ <span class="hlt">concentration</span> measurements of the gas-phase precursors involved in <span class="hlt">nitrate</span> production (NOx, O3, OH, HO2, etc.). Furthermore, <span class="hlt">nitrate</span> isotope analysis has been combined in this study for the first time with parallel <span class="hlt">observations</span> of the transferrable Δ17O of surface ozone, which was measured concurrently at Dome C using our recently developed analytical approach. This unique dataset has allowed for a direct comparison of <span class="hlt">observed</span> Δ17O(NO3-) values to those that are</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26802346','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26802346"><span>Environmental land use conflicts in catchments: A major cause of amplified <span class="hlt">nitrate</span> in river water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pacheco, F A L; Sanches Fernandes, L F</p> <p>2016-04-01</p> <p>Environmental land use conflicts are uses of the land that ignore soil capability. In this study, environmental land use conflicts were investigated in mainland Portugal, using Partial Least Squares (PLS) regression combined with GIS modeling and a group of 85 agricultural watersheds (with >50% occupation by agriculture) as work sample. The results indicate a dominance of conflicts in a region where vineyards systematically invaded steep hillsides (the River Douro basin), where forests would be the most appropriate use. As a consequence of the conflicts, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in rivers and lakes from these areas have increased, sometimes beyond the legal limit of 50mg/L imposed by the European and Portuguese laws. Excessive <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were also <span class="hlt">observed</span> along the Atlantic coast of continental Portugal, but associated to a combination of other factors: large population densities, and incomplete coverage by sewage systems and inadequate functioning of wastewater treatment plants. Before this study, environmental land use conflicts were never recognized as possible boost of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in surface water. Bearing in mind the consequences of drinking water <span class="hlt">nitrate</span> for human health, a number of land use change scenarios were investigated to forecast their impact on freshwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. It was seen that an aggravation of the conflicts would duplicate the number of watersheds with maximum <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above 50mg/L (from 11 to 20 watersheds), while the elimination of the conflicts would greatly reduce that number (to 3 watersheds). Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/865454','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/865454"><span>Purification of alkali metal <span class="hlt">nitrates</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Fiorucci, Louis C.; Gregory, Kevin M.</p> <p>1985-05-14</p> <p>A process is disclosed for removing heavy metal contaminants from impure alkali metal <span class="hlt">nitrates</span> containing them. The process comprises mixing the impure <span class="hlt">nitrates</span> with sufficient water to form a <span class="hlt">concentrated</span> aqueous solution of the impure <span class="hlt">nitrates</span>, adjusting the pH of the resulting solution to within the range of between about 2 and about 7, adding sufficient reducing agent to react with heavy metal contaminants within said solution, adjusting the pH of the solution containing reducing agent to effect precipitation of heavy metal impurities and separating the solid impurities from the resulting purified aqueous solution of alkali metal <span class="hlt">nitrates</span>. The resulting purified solution of alkali metal <span class="hlt">nitrates</span> may be heated to evaporate water therefrom to produce purified molten alkali metal <span class="hlt">nitrate</span> suitable for use as a heat transfer medium. If desired, the purified molten form may be granulated and cooled to form discrete solid particles of alkali metal <span class="hlt">nitrates</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.H12E..02W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.H12E..02W"><span>The UK <span class="hlt">Nitrate</span> Time Bomb (Invited)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ward, R.; Wang, L.; Stuart, M.; Bloomfield, J.; Gooddy, D.; Lewis, M.; McKenzie, A.</p> <p>2013-12-01</p> <p>The developed world has benefitted enormously from the intensification of agriculture and the increased availability and use of synthetic fertilizers during the last century. However there has also been unintended adverse impact on the natural environment (water and ecosystems) with <span class="hlt">nitrate</span> the most significant cause of water pollution and ecosystem damage . Many countries have introduced controls on <span class="hlt">nitrate</span>, e.g. the European Union's Water Framework and <span class="hlt">Nitrate</span> Directives, but despite this are continuing to see a serious decline in water quality. The purpose of our research is to investigate and quantify the importance of the unsaturated (vadose) zone pathway and groundwater in contributing to the decline. Understanding nutrient behaviour in the sub-surface environment and, in particular, the time lag between action and improvement is critical to effective management and remediation of nutrient pollution. A readily-transferable process-based model has been used to predict temporal loading of <span class="hlt">nitrate</span> at the water table across the UK. A time-varying <span class="hlt">nitrate</span> input function has been developed based on <span class="hlt">nitrate</span> usage since 1925. Depth to the water table has been calculated from groundwater levels based on regional-scale <span class="hlt">observations</span> in-filled by interpolated river base levels and vertical unsaturated zone velocities estimated from hydrogeological properties and mapping. The model has been validated using the results of more than 300 unsaturated zone <span class="hlt">nitrate</span> profiles. Results show that for about 60% of the Chalk - the principal aquifer in the UK - peak <span class="hlt">nitrate</span> input has yet to reach the water table and <span class="hlt">concentrations</span> will continue to rise over the next 60 years. The implications are hugely significant especially where environmental objectives must be achieved in much shorter timescales. Current environmental and regulatory management strategies rarely take lag times into account and as a result will be poorly informed, leading to inappropriate controls and conflicts</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.H21E1449R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.H21E1449R"><span>The <span class="hlt">Nitrate</span> App: Enhancing nutrient best management practice adoption and targeting via instantaneous, on-farm <span class="hlt">nitrate</span> data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rozemeijer, J.; De Geus, D.; Ekkelenkamp, R.</p> <p>2016-12-01</p> <p>Sociological surveys suggest that farmers understand that agriculture contributes to nutrient pollution but the same surveys also indicate that in the absence of on-farm <span class="hlt">nitrate</span> data, farmers assume someone else is causing the problem. This tendency to overestimate our own abilities is common to all of us and often described as "Lake Wobegon Syndrome" after the mythical town where "where all the women are strong, all the men are good-looking, and all the children are above average." We developed the <span class="hlt">Nitrate</span> App for smartphones to enable farmers and citizens to collect and share <span class="hlt">nitrate</span> <span class="hlt">concentration</span> measurements. The app accurately reads and interprets <span class="hlt">nitrate</span> test strips, directly displays the measured <span class="hlt">concentration</span>, and gives the option to share the result. The shared results are immediately visualised in the online Delta Data Viewer. Within this viewer, user group specific combinations of background maps, monitoring data, and study area characteristics can be configured. Through the <span class="hlt">Nitrate</span> App's mapping function project managers can more accurately target conservation practices to areas with the highest <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and loads. Furthermore, we expect that the actual on-farm data helps to overcome the "Lake Wobegon Effect" and will encourage farmers to talk to specialists about the right nutrient best management practices (BMP's) for their farm. After implementing these BMP's, the farmers can keep monitoring to evaluate the reduction in <span class="hlt">nitrate</span> losses. In this presentation, we explain the <span class="hlt">Nitrate</span> App technology and present the results of the first field applications in The Netherlands. We expect this free to download app to have wide transferability across watershed projects worldwide focusing on <span class="hlt">nitrate</span> contamination of groundwater or surface water. Its simple design requires no special equipment outside of the <span class="hlt">nitrate</span> test strips, a reference card, and a smartphone. The technology is also transferable to other relevant solutes for which test strips</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29282476','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29282476"><span>Effects of Environmental Contamination and Acute Toxicity of N-<span class="hlt">Nitrate</span> on Early Life Stages of Endemic Arboreal Frog, Polypedates cruciger (Blyth, 1852).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Balangoda, Anusha; Deepananda, K H M Ashoka; Wegiriya, H C E</p> <p>2018-02-01</p> <p>This study investigated the potential toxic effects of environmentally relevant <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> on development, growth, and mortality of early life stages of common hour-glass tree frog, Polypedates cruciger. Tadpoles from hatchlings through pre-adult were exposed to environmentally relevant <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> detected in Mirissa, Sri Lanka. Newly hatched, external gill stage, and internal gill stage tadpoles were exposed to potassium <span class="hlt">nitrate</span> for bioassay tests. No behavioral changes or abnormalities were <span class="hlt">observed</span> in control and <span class="hlt">nitrate</span>-induced group. However, detected environmental <span class="hlt">nitrate</span> <span class="hlt">concentration</span> significantly increased (p < 0.05) the growth of the tadpoles up to 25 days old. Results revealed that newly hatched and external gill stage was more susceptible to the <span class="hlt">nitrate</span> pollution than internal gill stage. The results suggest that environmentally relevant <span class="hlt">nitrate</span> can cause mortality on the amphibian population in ecosystems associated with agro-pastoral activities through altering the growth and direct toxicological effects on the survivorship.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JChPh.147x4503Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JChPh.147x4503Y"><span>Preferential solvation, ion pairing, and dynamics of <span class="hlt">concentrated</span> aqueous solutions of divalent metal <span class="hlt">nitrate</span> salts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yadav, Sushma; Chandra, Amalendu</p> <p>2017-12-01</p> <p>We have investigated the characteristics of preferential solvation of ions, structure of solvation shells, ion pairing, and dynamics of aqueous solutions of divalent alkaline-earth metal <span class="hlt">nitrate</span> salts at varying <span class="hlt">concentration</span> by means of molecular dynamics simulations. Hydration shell structures and the extent of preferential solvation of the metal and <span class="hlt">nitrate</span> ions in the solutions are investigated through calculations of radial distribution functions, tetrahedral ordering, and also spatial distribution functions. The Mg2+ ions are found to form solvent separated ion-pairs while the Ca2+ and Sr2+ ions form contact ion pairs with the <span class="hlt">nitrate</span> ions. These findings are further corroborated by excess coordination numbers calculated through Kirkwood-Buff G factors for different ion-ion and ion-water pairs. The ion-pairing propensity is found to be in the order of Mg(NO3) 2 < C a (NO3) 2 < S r (NO3) 2, and it follows the trend given by experimental activity coefficients. It is found that proper modeling of these solutions requires the inclusion of electronic polarization of the ions which is achieved in the current study through electronic continuum correction force fields. A detailed analysis of the effects of ion-pairs on the structure and dynamics of water around the hydrated ions is done through classification of water into different subspecies based on their locations around the cations or anions only or bridged between them. We have looked at the diffusion coefficients, relaxation of orientational correlation functions, and also the residence times of different subspecies of water to explore the dynamics of water in different structural environments in the solutions. The current results show that the water molecules are incorporated into fairly well-structured hydration shells of the ions, thus decreasing the single-particle diffusivities and increasing the orientational relaxation times of water with an increase in salt <span class="hlt">concentration</span>. The different structural</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25592012','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25592012"><span>Effects of polymorphisms in endothelial nitric oxide synthase and folate metabolizing genes on the <span class="hlt">concentration</span> of serum <span class="hlt">nitrate</span>, folate, and plasma total homocysteine after folic acid supplementation: a double-blind crossover study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cabo, Rona; Hernes, Sigrunn; Slettan, Audun; Haugen, Margaretha; Ye, Shu; Blomhoff, Rune; Mansoor, M Azam</p> <p>2015-02-01</p> <p>A number of studies have explored the effects of dietary <span class="hlt">nitrate</span> on human health. <span class="hlt">Nitrate</span> in the blood can be recycled to nitric oxide, which is an essential mediator involved in many important biochemical mechanisms. Nitric oxide is also formed in the body from l-arginine by nitric oxide synthase. The aim of this study was to investigate whether genetic polymorphisms in endothelial nitric oxide synthase (eNOS) and genes involved in folate metabolism affect the <span class="hlt">concentration</span> of serum <span class="hlt">nitrate</span>, serum folate, and plasma total homocysteine in healthy individuals after folic acid supplementation. In a randomized double-blind, crossover study, participants were given either folic acid 800 μg/d (n = 52) or placebo (n = 51) for 2 wk. Wash-out period was 2 wk. Fasting blood samples were collected, DNA was extracted by salting-out method and the polymorphisms in eNOS synthase and folate genes were genotyped by polymerase chain reaction methods. Measurement of serum <span class="hlt">nitrate</span> and plasma total homocysteine (p-tHcy) <span class="hlt">concentration</span> was done by high-performance liquid chromatography. The <span class="hlt">concentration</span> of serum <span class="hlt">nitrate</span> did not change in individuals after folic acid supplements (trial 1); however, the <span class="hlt">concentration</span> of serum <span class="hlt">nitrate</span> increased in the same individuals after placebo (P = 0.01) (trial 2). The individuals with three polymorphisms in eNOS gene had increased <span class="hlt">concentration</span> of serum folate and decreased <span class="hlt">concentration</span> of p-tHcy after folic acid supplementation. Among the seven polymorphisms tested in folate metabolizing genes, serum <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was significantly decreased only in DHFR del 19 gene variant. A significant difference in the <span class="hlt">concentration</span> of serum <span class="hlt">nitrate</span> was detected among individuals with MTHFR C > T677 polymorphisms. Polymorphisms in eNOS and folate genes affect the <span class="hlt">concentration</span> of serum folate and p-tHcy but do not have any effect on the <span class="hlt">concentration</span> of NO3 in healthy individuals after folic acid supplementation. Copyright © 2015 Elsevier Inc. All</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AtmEn..62..228Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AtmEn..62..228Z"><span>Source apportionment of PM2.5 <span class="hlt">nitrate</span> and sulfate in China using a source-oriented chemical transport model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Hongliang; Li, Jingyi; Ying, Qi; Yu, Jian Zhen; Wu, Dui; Cheng, Yuan; He, Kebin; Jiang, Jingkun</p> <p>2012-12-01</p> <p><span class="hlt">Nitrate</span> and sulfate account for a significant fraction of PM2.5 mass and are generally secondary in nature. Contributions to these two inorganic aerosol components from major sources need to be identified for policy makers to develop cost effective regional emission control strategies. In this work, a source-oriented version of the Community Multiscale Air Quality (CMAQ) model that directly tracks the contributions from multiple emission sources to secondary PM2.5 is developed to determine the regional contributions of power, industry, transportation and residential sectors as well as biogenic sources to <span class="hlt">nitrate</span> and sulfate <span class="hlt">concentrations</span> in China in January and August 2009.The source-oriented CMAQ model is capable of reproducing most of the available PM10 and PM2.5 mass, and PM2.5 <span class="hlt">nitrate</span> and sulfate <span class="hlt">observations</span>. Model prediction suggests that monthly average PM2.5 inorganic components (<span class="hlt">nitrate</span> + sulfate + ammonium ion) can be as high as 60 μg m-3 in January and 45 μg m-3 in August, accounting for 20-40% and 50-60% of total PM2.5 mass. The model simulations also indicate significant spatial and temporal variation of the <span class="hlt">nitrate</span> and sulfate <span class="hlt">concentrations</span> as well as source contributions in the country. In January, <span class="hlt">nitrate</span> is high over Central and East China with a maximum of 30 μg m-3 in the Sichuan Basin. In August, <span class="hlt">nitrate</span> is lower and the maximum <span class="hlt">concentration</span> of 16 μg m-3 occurs in North China. In January, highest sulfate occurs in the Sichuan Basin with a maximum <span class="hlt">concentration</span> of 18 μg m-3 while in August high sulfate <span class="hlt">concentration</span> occurs in North and East China with a similar maximum <span class="hlt">concentration</span>. Power sector is the dominating source of <span class="hlt">nitrate</span> and sulfate in both January and August. Transportation sector is an important source of <span class="hlt">nitrate</span> (20-30%) in both months. Industry sector contributes to both <span class="hlt">nitrate</span> and sulfate <span class="hlt">concentrations</span> by approximately 20-30%. Residential sector contributes to approximately 10-20% of <span class="hlt">nitrate</span> and sulfate in January but</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28954516','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28954516"><span>High-Resolution in Situ Measurement of <span class="hlt">Nitrate</span> in Runoff from the Greenland Ice Sheet.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Beaton, Alexander D; Wadham, Jemma L; Hawkings, Jon; Bagshaw, Elizabeth A; Lamarche-Gagnon, Guillaume; Mowlem, Matthew C; Tranter, Martyn</p> <p>2017-11-07</p> <p>We report the first in situ high-resolution <span class="hlt">nitrate</span> time series from two proglacial meltwater rivers draining the Greenland Ice Sheet, using a recently developed submersible analyzer based on lab-on-chip (LOC) technology. The low sample volume (320 μL) required by the LOC analyzer meant that low <span class="hlt">concentration</span> (few micromolar to submicromolar), highly turbid subglacial meltwater could be filtered and colorimetrically analyzed in situ. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in rivers draining Leverett Glacier in southwest Greenland and Kiattuut Sermiat in southern Greenland exhibited a clear diurnal signal and a gradual decline at the commencement of the melt season, displaying trends that would not be discernible using traditional daily manual sampling. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> varied by 4.4 μM (±0.2 μM) over a 10 day period at Kiattuut Sermiat and 3.0 μM (±0.2 μM) over a 14 day period at Leverett Glacier. Marked changes in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were <span class="hlt">observed</span> when discharge began to increase. High-resolution in situ measurements such as these have the potential to significantly advance the understanding of nutrient cycling in remote systems, where the dynamics of nutrient release are complex but are important for downstream biogeochemical cycles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1850g0002A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1850g0002A"><span>Aluminide slurry coatings for protection of ferritic steel in molten <span class="hlt">nitrate</span> corrosion for <span class="hlt">concentrated</span> solar power technology</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Audigié, Pauline; Bizien, Nicolas; Baráibar, Ignacio; Rodríguez, Sergio; Pastor, Ana; Hernández, Marta; Agüero, Alina</p> <p>2017-06-01</p> <p>Molten <span class="hlt">nitrates</span> can be employed as heat storage fluids in solar <span class="hlt">concentration</span> power plants. However molten <span class="hlt">nitrates</span> are corrosive and if operating temperatures are raised to increase efficiencies, the corrosion rates will also increase. High temperature corrosion resistant coatings based on Al have demonstrated excellent results in other sectors such as gas turbines. Aluminide slurry coated and uncoated P92 steel specimens were exposed to the so called Solar Salt (industrial grade), a binary eutectic mixture of 60 % NaNO3 - 40 % KNO3, in air for 2000 hours at 550°C and 580°C in order to analyze their behavior as candidates to be used in future solar <span class="hlt">concentration</span> power plants employing molten <span class="hlt">nitrates</span> as heat transfer fluids. Coated ferritic steels constitute a lower cost technology than Ni based alloy. Two different coating morphologies resulting from two heat treatment performed at 700 and 1050°C after slurry application were tested. The coated systems exhibited excellent corrosion resistance at both temperatures, whereas uncoated P92 showed significant mass loss from the beginning of the test. The coatings showed very slow reaction with the molten Solar Salt. In contrast, uncoated P92 developed a stratified, unprotected Fe, Cr oxide with low adherence which shows oscillating Cr content as a function of coating depth. NaFeO2 was also found at the oxide surface as well as within the Fe, Cr oxide.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/4310773','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/4310773"><span>TREATMENT OF AMMONIUM <span class="hlt">NITRATE</span> SOLUTIONS</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Boyer, T.W.; MacHutchin, J.G.; Yaffe, L.</p> <p>1958-06-10</p> <p>The treatment of waste solutions obtained in the processing of neutron- irradiated uranium containing fission products and ammonium <span class="hlt">nitrate</span> is described. The object of this process is to provide a method whereby the ammonium <span class="hlt">nitrate</span> is destroyed and removed from the solution so as to permit subsequent <span class="hlt">concentration</span> of the solution.. In accordance with the process the residual <span class="hlt">nitrate</span> solutions are treated with an excess of alkyl acid anhydride, such as acetic anhydride. Preferably, the residual <span class="hlt">nitrate</span> solution is added to an excess of the acetic anhydride at such a rate that external heat is not required. The result of this operation is that the ammonium <span class="hlt">nitrate</span> and acetic anhydride react to form N/sub 2/ O and acetic acid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850014976','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850014976"><span>Geographic variation in the relationships of temperature, salinity or sigma sub t versus plant nutrient <span class="hlt">concentrations</span> in the world ocean. [silicic acid, <span class="hlt">nitrate</span>, and phosphate <span class="hlt">concentration</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kamykowski, D.; Zentara, S. J.</p> <p>1985-01-01</p> <p>A NODC data set representing all regions of the world ocean was analyzed for temperature and sigma-t relationships with <span class="hlt">nitrate</span>, phosphate or silicic acid. Six cubic regressions were for each ten degree square of latitude and longitude containing adequate data. World maps display the locations that allow the prediction of plant nutrient <span class="hlt">concentrations</span> from temperature or sigma-t. Geographic coverage improves along the sequence: <span class="hlt">nitrate</span>, phosphate, and silicic acid and is better for sigma-t than for temperature. Contour maps of the approximate temperature of sigma-t at which these nitrients are no longer measurable in a parcel of water are generated, based on a percentile analysis of the temperature or sigma-t at which less than a selected amount of plant nutrient occurs. Results are stored on magnetic tape in tabular form. The global potential to predict plant nutrient <span class="hlt">concentrations</span> from remotely sensed temperature of sigma-t and to emphasize the latitudinally and longitudinally changing phytoplankton growth environment in present and past oceans is demonstrated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4059197','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4059197"><span>Active secretion and protective effect of salivary <span class="hlt">nitrate</span> against stress in human volunteers and rats</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jin, Luyuan; Qin, Lizheng; Xia, Dengsheng; Liu, Xibao; Fan, Zhipeng; Zhang, Chunmei; Gu, Liankun; He, Junqi; Ambudkar, Indu S.; Deng, Dajun; Wang, Songlin</p> <p>2014-01-01</p> <p>Up to 25% of the circulating <span class="hlt">nitrate</span> in blood is actively taken up, <span class="hlt">concentrated</span>, and secreted into saliva by the salivary glands. Salivary <span class="hlt">nitrate</span> can be reduced to nitrite by the commensal bacteria in the oral cavity or stomach and then further converted to nitric oxide (NO) in vivo, which may play a role in gastric protection. However, whether salivary <span class="hlt">nitrate</span> is actively secreted in human beings has not yet been determined. This study was designed to determine whether salivary <span class="hlt">nitrate</span> is actively secreted in human beings as an acute stress response and what role salivary <span class="hlt">nitrate</span> plays in stress-induced gastric injury. To <span class="hlt">observe</span> salivary <span class="hlt">nitrate</span> function under stress conditions, alteration of salivary <span class="hlt">nitrate</span> and nitrite was analyzed among 22 healthy volunteers before and after a strong stress activity, jumping down from a platform at the height of 68m. A series of stress indexes was analyzed to monitor the stress situation. We found that both the <span class="hlt">concentration</span> and the total amount of <span class="hlt">nitrate</span> in mixed saliva were significantly increased in the human volunteers immediately after the jump, with an additional increase 1 h later (p < 0.01). Saliva nitrite reached a maximum immediately after the jump and was maintained 1 h later. To study the biological functions of salivary <span class="hlt">nitrate</span> and nitrite in stress protection, we further carried out a water-immersion-restraint stress (WIRS) assay in male adult rats with bilateral parotid and submandibular duct ligature (BPSDL). Intragastric <span class="hlt">nitrate</span>, nitrite, and NO; gastric mucosal blood flow; and gastric ulcer index (UI) were monitored and <span class="hlt">nitrate</span> was administrated in drinking water to compensate for <span class="hlt">nitrate</span> secretion in BPSDL animals. Significantly decreased levels of intragastric <span class="hlt">nitrate</span>, nitrite, and NO and gastricmucosal blood flow were measured in BPSDL rats during the WIRS assay compared to sham control rats (p < 0.05). Recovery was <span class="hlt">observed</span> in the BPSDL rats upon <span class="hlt">nitrate</span> administration. The WIRS-induced UI was</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.7852S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.7852S"><span>Fate of <span class="hlt">nitrate</span> and origin of ammonium during infiltration of treated wastewater investigated through stable isotopes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Silver, Matthew; Schlögl, Johanna; Knöller, Kay; Schüth, Christoph</p> <p>2017-04-01</p> <p> <span class="hlt">concentration</span> is not clearly logarithmic, so processes other than denitrification are not ruled out for explaining the fate of <span class="hlt">nitrate</span>. The δ15N of ammonium in the water samples and of nitrogen in the soil were also measured. With increasing depth and time, the δ15N-NH4+ (mean 4.3‰) decreases and approaches the δ15N of the pre-experimental soil of 2.4‰. This suggests that ammonium is formed at least in part from the soil organic matter, likely through a combination of leaching and microbial processes. Although most <span class="hlt">nitrate</span> attenuates by 15 cm depth and very little ammonium is <span class="hlt">observed</span> here, some <span class="hlt">nitrate</span> (usually <0.5 mg-N/L) was <span class="hlt">observed</span> at depths of 30 cm and below, especially early in the experiments. Starting at 30 cm depth, organic carbon <span class="hlt">concentrations</span> and thereby also C:NO3-ratios become high (>10), which are conditions sometimes found to be favorable to dissimilatory <span class="hlt">nitrate</span> reduction to ammonium. Rayleigh enrichment factors also suggest that <span class="hlt">nitrate</span> may be the source of some of the ammonium. Measurements of additional samples and organic nitrogen isotopes are planned, in order to further evaluate the fate of <span class="hlt">nitrate</span> and the source(s) of the ammonium.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/fs/2010/3077/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/fs/2010/3077/"><span>Sustainability of natural attenuation of <span class="hlt">nitrate</span> in agricultural aquifers</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Green, Christopher T.; Bekins, Barbara A.</p> <p>2010-01-01</p> <p>Increased <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> in groundwater in agricultural areas, coinciding with increased use of chemical and organic fertilizers, have raised concern because of risks to environmental and human health. At some sites, these problems are mitigated by natural attenuation of <span class="hlt">nitrate</span> as a result of microbially mediated reactions. Results from U.S. Geological Survey (USGS) research under the National Water-Quality Assessment (NAWQA) program show that reactions of dissolved <span class="hlt">nitrate</span> with solid aquifer minerals and organic carbon help lower <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater beneath agricultural fields. However, increased fluxes of <span class="hlt">nitrate</span> cause ongoing depletion of the finite pool of solid reactants. Consumption of the solid reactants diminishes the capacity of the aquifer to remove <span class="hlt">nitrate</span>, calling into question the long-term sustainability of these natural attenuation processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15382874','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15382874"><span>Precipitation of <span class="hlt">nitrate</span>-cancrinite in Hanford Tank Sludge.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Buck, E C; McNamara, B K</p> <p>2004-08-15</p> <p>The chemistry of underground storage tanks containing high-level waste at the Hanford Site in Washington State is an area of continued research interest. Thermodynamic models have predicted the formation of analcime and clinoptilolite in Hanford tanks, rather than cancrinite; however, these predictions were based on carbonate-cancrinite. We report the first <span class="hlt">observation</span> of a <span class="hlt">nitrate</span>-cancrinite [possibly Na8(K,Cs)(AlSiO4)6(NO3)2 x nH2O] extracted from a Hanford tank 241-AP-101 sample that was evaporated to 6, 8, and 10 M NaOH <span class="hlt">concentrations</span>. The <span class="hlt">nitrate</span>-cancrinite phase formed spherical aggregates (4 microm in diameter) that consisted of platy hexagonal crystals (approximately 0.2 microm thick). Cesium-137 was <span class="hlt">concentrated</span> in these aluminosilicate structures. These phases possessed a morphology identical to that of <span class="hlt">nitrate</span>-cancrinite synthesized using simulant tests of nonradioactive tank waste, supporting the contention that it is possible to develop nonradioactive artificial sludges. This investigation points to the continued importance of understanding the solubility of NO3-cancrinite and related phases. Knowledge of the detailed structure of actual phases in the tank waste helps with thermodynamic modeling of tank conditions and waste processing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5302962','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5302962"><span>Association between dietary <span class="hlt">nitrate</span> and nitrite intake and site-specific cancer risk: evidence from <span class="hlt">observational</span> studies</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jia, Hui-Xun; Liang, Fei; Yuan, Jing; Zhu, Ji</p> <p>2016-01-01</p> <p>Epidemiological studies have reported inconsistent findings on the association between dietary <span class="hlt">nitrate</span> and nitrite intake and cancer risk. We performed a meta-analysis of epidemiological studies to summarize available evidence on the association between dietary <span class="hlt">nitrate</span> and nitrite intake and cancer risk from published prospective and case-control studies. PubMed database was searched to identify eligible publications through April 30th, 2016. Study-specific relative risks (RRs) with corresponding 95% confidence interval (CI) from individual studies were pooled by using random- or fixed- model, and heterogeneity and publication bias analyses were conducted. Data from 62 <span class="hlt">observational</span> studies, 49 studies for <span class="hlt">nitrates</span> and 51 studies for nitrites, including a total of 60,627 cancer cases were analyzed. Comparing the highest vs. lowest levels, dietary <span class="hlt">nitrate</span> intake was inversely associated with gastric cancer risk (RR = 0.78; 95%CI = 0.67-0.91) with moderate heterogeneity (I2 = 42.3%). In contrast, dietary nitrite intake was positively associated with adult glioma and thyroid cancer risk with pooled RR of 1.21 (95%CI = 1.03-1.42) and 1.52 (95%CI = 1.12-2.05), respectively. No significant associations were found between dietary <span class="hlt">nitrate</span>/nitrite and cancers of the breast, bladder, colorectal, esophagus, renal cell, non-Hodgkin lymphoma, ovarian, and pancreas. The present meta-analysis provided modest evidence that positive associations of dietary <span class="hlt">nitrate</span> and negative associations of dietary nitrite with certain cancers. PMID:27486968</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5078692','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5078692"><span>Knock-Down of a Tonoplast Localized Low-Affinity <span class="hlt">Nitrate</span> Transporter OsNPF7.2 Affects Rice Growth under High <span class="hlt">Nitrate</span> Supply</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hu, Rui; Qiu, Diyang; Chen, Yi; Miller, Anthony J.; Fan, Xiaorong; Pan, Xiaoping; Zhang, Mingyong</p> <p>2016-01-01</p> <p>The large <span class="hlt">nitrate</span> transporter 1/peptide transporter family (NPF) has been shown to transport diverse substrates, including <span class="hlt">nitrate</span>, amino acids, peptides, phytohormones, and glucosinolates. However, the rice (Oryza sativa) root-specific family member OsNPF7.2 has not been functionally characterized. Here, our data show that OsNPF7.2 is a tonoplast localized low-affinity <span class="hlt">nitrate</span> transporter, that affects rice growth under high <span class="hlt">nitrate</span> supply. Expression analysis showed that OsNPF7.2 was mainly expressed in the elongation and maturation zones of roots, especially in the root sclerenchyma, cortex and stele. It was also induced by high <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>. Subcellular localization analysis showed that OsNPF7.2 was localized on the tonoplast of large and small vacuoles. Heterologous expression in Xenopus laevis oocytes suggested that OsNPF7.2 was a low-affinity <span class="hlt">nitrate</span> transporter. Knock-down of OsNPF7.2 retarded rice growth under high <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>. Therefore, we deduce that OsNPF7.2 plays a role in intracellular allocation of <span class="hlt">nitrate</span> in roots, and thus influences rice growth under high <span class="hlt">nitrate</span> supply. PMID:27826301</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/biblio/7189347-development-accelerated-net-nitrate-uptake-zea-mays','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/7189347-development-accelerated-net-nitrate-uptake-zea-mays"><span>Development of accelerated net <span class="hlt">nitrate</span> uptake. [Zea mays L</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>MacKown, C.T.; McClure, P.R.</p> <p>1988-05-01</p> <p>Upon initial <span class="hlt">nitrate</span> exposure, net <span class="hlt">nitrate</span> uptake rates in roots of a wide variety of plants accelerate within 6 to 8 hours to substantially greater rates. Effects of solution <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and short pulses of <span class="hlt">nitrate</span> ({le}1 hour) upon <span class="hlt">nitrate</span>-induced acceleration of <span class="hlt">nitrate</span> uptake in maize (Zea mays L.) were determined. Root cultures of dark-grown seedlings, grown without <span class="hlt">nitrate</span>, were exposed to 250 micromolar <span class="hlt">nitrate</span> for 0.25 to 1 hour or to various solution <span class="hlt">nitrate</span> <span class="hlt">concentration</span> (10-250 micromolar) for 1 hour before returning them to a <span class="hlt">nitrate</span>-free solution. Net <span class="hlt">nitrate</span> uptake rates were assayed at various periods following <span class="hlt">nitrate</span> exposuremore » and compared to rates of roots grown either in the absence of <span class="hlt">nitrate</span> (CaSO{sub 4}-grown) or with continuous <span class="hlt">nitrate</span> for at least 20 hours. Three hours after initial <span class="hlt">nitrate</span> exposure, <span class="hlt">nitrate</span> pulse treatments increased <span class="hlt">nitrate</span> uptake rates three- to four-fold compared to the rates of CaSO{sub 4}-grown roots. When cycloheximide (5 micrograms per milliliter) was included during a 1-hour pulse with 250 micromolar <span class="hlt">nitrate</span>, development of the accelerated <span class="hlt">nitrate</span> uptake state was delayed. Otherwise, <span class="hlt">nitrate</span> uptake rates reached maximum values within 6 hours before declining. Maximum rates, however, were significantly less than those of roots exposed continuously for 20, 32, or 44 hours. Pulsing for only 0.25 hour with 250 micromolar <span class="hlt">nitrate</span> and for 1 hour with 10 micromolar caused acceleration of <span class="hlt">nitrate</span> uptake, but the rates attained were either less than or not sustained for a duration comparable to those of roots pulsed for 1 hour with 250 micromolar <span class="hlt">nitrate</span>. These results indicate that substantial development of <span class="hlt">nitrate</span>-induced accelerated <span class="hlt">nitrate</span> uptake state can be achieved by small endogenous accumulations of <span class="hlt">nitrate</span>, which appear to moderate the activity or level of root <span class="hlt">nitrate</span> uptake.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27871751','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27871751"><span>Regional assessment of <span class="hlt">concentrations</span> and sources of pharmaceutically active compounds, pesticides, <span class="hlt">nitrate</span>, and E. coli in post-glacial aquifer environments (Canada).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Saby, Marion; Larocque, Marie; Pinti, Daniele L; Barbecot, Florent; Gagné, Sylvain; Barnetche, Diogo; Cabana, Hubert</p> <p>2017-02-01</p> <p>There is growing concern worldwide about the exposure of groundwater resources to pharmaceutically active compounds (PhACs) and agricultural contaminants, such as pesticides, <span class="hlt">nitrate</span>, and Escherichia coli. For regions with a low population density and an abundance of water, regional contamination assessments are not carried out systematically due to the typically low <span class="hlt">concentrations</span> and high costs of analyses. The objectives of this study were to evaluate regional-scale contaminant distributions in untreated groundwater in a rural region of Quebec (Canada). The geological and hydrogeological settings of this region are typical of post-glacial regions around the world, where groundwater flow can be complex due to heterogeneous geological conditions. A new spatially distributed Anthropogenic Footprint Index (AFI), based on land use data, was developed to assess surface pollution risks. The Hydrogeochemical Vulnerability Index (HVI) was computed to estimate aquifer vulnerability. Nine wells had detectable <span class="hlt">concentrations</span> of one to four of the 13 tested PhACs, with a maximum <span class="hlt">concentration</span> of 116ng·L -1 for benzafibrate. A total of 34 of the 47 tested pesticides were detected in <span class="hlt">concentrations</span> equal to or greater than the detection limit, with a maximum total pesticide <span class="hlt">concentration</span> of 692ng·L -1 . <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> exceeded 1mg·L -1 N-NO 3 in 15.3% of the wells, and the Canadian drinking water standard was exceeded in one well. Overall, 13.5% of the samples had detectable E. coli. Including regional-scale sources of pollutants to the assessment of aquifer vulnerability with the AFI did not lead to the identification of contaminated wells, due to the short groundwater flow paths between recharge and the sampled wells. Given the occurrence of contaminants, the public health concerns stemming from these new data on regional-scale PhAC and pesticide <span class="hlt">concentrations</span>, and the local flow conditions <span class="hlt">observed</span> in post-glacial terrains, there is a clear need to investigate</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AtmEn..99...24X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AtmEn..99...24X"><span>Effect of <span class="hlt">nitrate</span> and sulfate relative abundance in PM2.5 on liquid water content explored through half-hourly <span class="hlt">observations</span> of inorganic soluble aerosols at a polluted receptor site</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xue, Jian; Griffith, Stephen M.; Yu, Xin; Lau, Alexis K. H.; Yu, Jian Zhen</p> <p>2014-12-01</p> <p>Liquid water content (LWC) is the amount of liquid water on aerosols. It contributes to visibility degradation, provides a surface for gas condensation, and acts as a medium for heterogeneous gas/particle reactions. In this study, 520 half-hourly measurements of ionic chemical composition in PM2.5 at a receptor site in Hong Kong are used to investigate the dependence of LWC on ionic chemical composition, particularly on the relative abundance of sulfate and <span class="hlt">nitrate</span>. LWC was estimated using a thermodynamic model (AIM-III). Within this data set of PM2.5 ionic compositions, LWC was highly correlated with the multivariate combination of sulfate and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and RH (R2 = 0.90). The empirical linear regression result indicates that LWC is more sensitive to <span class="hlt">nitrate</span> mass than sulfate. During a <span class="hlt">nitrate</span> episode, the highest LWC (80.6 ± 17.9 μg m-3) was <span class="hlt">observed</span> and the level was 70% higher than that during a sulfate episode despite a similar ionic PM2.5 mass <span class="hlt">concentration</span>. A series of sensitivity tests were conducted to study LWC change as a function of the relative <span class="hlt">nitrate</span> and sulfate abundance, the trend of which is expected to shift to more <span class="hlt">nitrate</span> in China as a result of SO2 reduction and increase in NOx emission. Starting from a base case that uses the average of measured PM2.5 ionic chemical composition (63% SO42-, 11% NO3-, 19% NH4+, and 7% other ions) and an ionic equivalence ratio, [NH4+]/(2[SO42-] + [NO3-]), set constant to 0.72, the results show LWC would increase by 204% at RH = 40% when 50% of the SO42- is replaced by NO3- mass <span class="hlt">concentration</span>. This is largely due to inhibition of (NH4)3H(SO4)2 crystallization while PM2.5 ionic species persist in the aqueous phase. At RH = 90%, LWC would increase by 12% when 50% of the SO42- is replaced by NO3- mass <span class="hlt">concentration</span>. The results of this study highlight the important implications to aerosol chemistry and visibility degradation associated with LWC as a result of a shift in PM2.5 ionic chemical</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.pca.state.mn.us/index.php/view-document.html?gid=19844','USGSPUBS'); return false;" href="http://www.pca.state.mn.us/index.php/view-document.html?gid=19844"><span><span class="hlt">Nitrate</span> Trends in Minnesota Rivers</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Wall, Dave; Christopherson, Dave; Lorenz, Dave; Martin, Gary</p> <p>2013-01-01</p> <p>The objective of this study was to assess long-term trends (30 to 35 years) of flow-adjusted <span class="hlt">concentrations</span> of nitrite+<span class="hlt">nitrate</span>-N (hereinafter referred to as <span class="hlt">nitrate</span>) in a way that would allow us to discern changing trends. Recognizing that these trends are commonly different from one river to another river and from one part of the state to another, our objective was to examine as many river monitoring sites across the state as possible for which sufficient long term streamflow and <span class="hlt">concentration</span> data were available.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002AGUFM.H11F..11D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002AGUFM.H11F..11D"><span>Identifying the Source of High-<span class="hlt">Nitrate</span> Ground Water Related to Artificial Recharge in a Desert Basin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Densmore, J. N.; Nishikawa, T.; Bohlke, J. K.; Martin, P.</p> <p>2002-12-01</p> <p>Ground water has been the sole source of water supply for the community of Yucca Valley in the Mojave Desert, California. Domestic wastewater from the community is treated using septic tanks. An imbalance between ground-water recharge and pumpage caused ground-water levels in the ground-water basin to decline by as much as 300 feet from the late 1940s through 1994. In response to this decline, the local water district, Hi-Desert Water District, began an artificial recharge program in February 1995 to replenish the ground water in the basin using imported surface water. The artificial recharge program resulted in water-level recovery of about 250 feet between February 1995 and December 2001; however, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in some wells also increased from a background <span class="hlt">concentration</span> of 10 mg/L as NO3 to more than the U.S. Environmental Protection Agency maximum contaminant level of 45 mg/L as NO3, limiting water use for some wells. The largest increase in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> occurred adjacent to the artificial recharge sites where the largest increase in water levels occurred even though the recharge water had low <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. The source of high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> <span class="hlt">observed</span> in ground water during aquifer recovery was identified by compiling historical water-quality data; monitoring changes in water quality since artificial recharge began; and analyzing selected samples for major-ion chemistry, stable isotopes of H,O, and N, caffeine, and pharmaceuticals. The major-ions and H and O stable-isotope data indicate that ground-water samples that had the highest <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were mixtures of imported water and native ground water. <span class="hlt">Nitrate</span>-to-chloride ratios, N isotopes and caffeine and pharmaceutical data indicate septic-tank seepage (septage) is the primary source of increases in <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. The rapid rise in water levels entrained the large volume of high-<span class="hlt">nitrate</span> septage stored in the unsaturated zone, resulting in the rapid increase</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24205754','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24205754"><span>Effect of lead <span class="hlt">nitrate</span> on the liver of the cichlid fish (Oreochromis niloticus): a light microscope study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Khidr, Bothaina M; Mekkawy, Imam A A; Harabawy, Ahmed S A; Ohaida, Abdel Salam M I</p> <p>2012-09-15</p> <p>The adverse impacts of heavy metals on fish liver were evident with great variability among organs and species. The present study deals with the histological changes of the hepatocytes of the Nile tilapia, Oreochromis niloticus, following exposure to 2.5, 5, 10 ppm of lead <span class="hlt">nitrate</span> for 1, 2, 3, 4 weeks. The present results revealed that lead <span class="hlt">nitrate</span> exerts some histological effects on the hepatic tissue after exposure to the first <span class="hlt">concentration</span> in the form of dilatation and congestion of the blood vessels, vacuolation of hepatic cells, proliferation of connective tissue and hepatic necrosis. Leucocyte aggregation-mostly lymphatic in nature-was seen infiltrating hepatic tissue. These alterations became more pronounced in liver of fishes exposed to second <span class="hlt">concentrations</span> indicating more progressive signs of necrosis. The presence of eosinophilic oedematous areas surrounding some blood vessels was also <span class="hlt">observed</span>. Finally, at the third <span class="hlt">concentration</span>, in addition to the above alterations, melanomacrophages, which store lipofuscin at the site of necrosis, were <span class="hlt">observed</span>. These histological results imply that the fish liver may serve as a target organ for the toxicity of sublethal <span class="hlt">concentrations</span> of lead <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25149457','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25149457"><span>Effects of <span class="hlt">nitrate</span> injection on microbial enhanced oil recovery and oilfield reservoir souring.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>da Silva, Marcio Luis Busi; Soares, Hugo Moreira; Furigo, Agenor; Schmidell, Willibaldo; Corseuil, Henry Xavier</p> <p>2014-11-01</p> <p>Column experiments were utilized to investigate the effects of <span class="hlt">nitrate</span> injection on sulfate-reducing bacteria (SRB) inhibition and microbial enhanced oil recovery (MEOR). An indigenous microbial consortium collected from the produced water of a Brazilian offshore field was used as inoculum. The presence of 150 mg/L volatile fatty acids (VFA´s) in the injection water contributed to a high biological electron acceptors demand and the establishment of anaerobic sulfate-reducing conditions. Continuous injection of <span class="hlt">nitrate</span> (up to 25 mg/L) for 90 days did not inhibit souring. Contrariwise, in nitrogen-limiting conditions, the addition of <span class="hlt">nitrate</span> stimulated the proliferation of δ-Proteobacteria (including SRB) and the associated sulfide <span class="hlt">concentration</span>. Denitrification-specific nirK or nirS genes were not detected. A sharp decrease in water interfacial tension (from 20.8 to 14.5 mN/m) <span class="hlt">observed</span> concomitantly with <span class="hlt">nitrate</span> consumption and increased oil recovery (4.3 % v/v) demonstrated the benefits of <span class="hlt">nitrate</span> injection on MEOR. Overall, the results support the notion that the addition of <span class="hlt">nitrate</span>, at this particular oil reservoir, can benefit MEOR by stimulating the proliferation of fortuitous biosurfactant-producing bacteria. Higher <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> exceeding the stoichiometric volatile fatty acid (VFA) biodegradation demands and/or the use of alternative biogenic souring control strategies may be necessary to warrant effective SRB inhibition down gradient from the injection wells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2012/5287/support/sir2012-5287.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2012/5287/support/sir2012-5287.pdf"><span>Nutrient <span class="hlt">concentrations</span> in surface water and groundwater, and <span class="hlt">nitrate</span> source identification using stable isotope analysis, in the Barnegat Bay-Little Egg Harbor watershed, New Jersey, 2010–11</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Wieben, Christine M.; Baker, Ronald J.; Nicholson, Robert S.</p> <p>2013-01-01</p> <p>Five streams in the Barnegat Bay-Little Egg Harbor (BB-LEH) watershed in southern New Jersey were sampled for nutrient <span class="hlt">concentrations</span> and stable isotope composition under base-flow and stormflow conditions, and during the growing and nongrowing seasons, to help quantify and identify sources of nutrient loading. Samples were analyzed for <span class="hlt">concentrations</span> of total nitrogen, ammonia, <span class="hlt">nitrate</span> plus nitrite, organic nitrogen, total phosphorus, and orthophosphate, and for nitrogen and oxygen stable isotope ratios. <span class="hlt">Concentrations</span> of total nitrogen in the five streams appear to be related to land use, such that streams in subbasins characterized by extensive urban development (and historical agricultural land use)—North Branch Metedeconk and Toms Rivers—exhibited the highest total nitrogen <span class="hlt">concentrations</span> (0.84–1.36 milligrams per liter (mg/L) in base flow). Base-flow total nitrogen <span class="hlt">concentrations</span> in these two streams were dominated by <span class="hlt">nitrate</span>; <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> decreased during storm events as a result of dilution by storm runoff. The two streams in subbasins with the least development—Cedar Creek and Westecunk Creek—exhibited the lowest total nitrogen <span class="hlt">concentrations</span> (0.16–0.26 mg/L in base flow), with organic nitrogen as the dominant species in both base flow and stormflow. A large proportion of these subbasins lies within forested parts of the Pinelands Area, indicating the likelihood of natural inputs of organic nitrogen to the streams that increase during periods of storm runoff. Base-flow total nitrogen <span class="hlt">concentrations</span> in Mill Creek, in a moderately developed basin, were 0.43 to 0.62 mg/L and were dominated by ammonia, likely associated with leachate from a landfill located upstream. Total phosphorus and orthophosphate were not found at detectable <span class="hlt">concentrations</span> in most of the surface-water samples, with the exception of samples collected from the North Branch Metedeconk River, where <span class="hlt">concentrations</span> ranged from 0.02 to 0.09 mg/L for total phosphorus and 0</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70023438','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70023438"><span>Catchment-scale variation in the <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> of groundwater seeps in the Catskill Mountains, New York, U.S.A.</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>West, A.J.; Findlay, S.E.G.; Burns, Douglas A.; Weathers, K.C.; Lovett, Gary M.</p> <p>2001-01-01</p> <p>Forested headwater streams in the Catskill Mountains of New York show significant among-catchment variability in mean annual <span class="hlt">nitrate</span> (NO3-) <span class="hlt">concentrations</span>. Large contributions from deep groundwater with high NO3- <span class="hlt">concentrations</span> have been invoked to explain high NO3- <span class="hlt">concentrations</span> in stream water during the growing season. To determine whether variable contributions of groundwater could explain among-catchment differences in streamwater, we measured NO3- <span class="hlt">concentrations</span> in 58 groundwater seeps distributed across six catchments known to have different annual average streamwater <span class="hlt">concentrations</span>. Seeps were identified based on release from bedrock fractures and bedding planes and had consistently lower temperatures than adjacent streamwaters. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in seeps ranged from near detection limits (0.005 mg NO3--N/L) to 0.75 mg NO3--N/L. Within individual catchments, groundwater residence time does not seem to strongly affect NO3- <span class="hlt">concentrations</span> because in three out of four catchments there were non-significant correlations between seep silica (SiO2) <span class="hlt">concentrations</span>, a proxy for residence time, and seep NO3- <span class="hlt">concentrations</span>. Across catchments, there was a significant but weak negative relationship between NO3- and SiO2 <span class="hlt">concentrations</span>. The large range in NO3- <span class="hlt">concentrations</span> of seeps across catchments suggests: 1) the principal process generating among-catchment differences in streamwater NO3- <span class="hlt">concentrations</span> must influence water before it enters the groundwater flow system and 2) this process must act at large spatial scales because among-catchment variability is much greater than intra-catchment variability. Differences in the quantity of groundwater contribution to stream baseflow are not sufficient to account for differences in streamwater NO3- <span class="hlt">concentrations</span> among catchments in the Catskill Mountains.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024539','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024539"><span><span class="hlt">Nitrate</span> in aquifers beneath agricultural systems</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Burkart, M.R.; Stoner, J.D.</p> <p>2002-01-01</p> <p>Research from several regions of the world provides spatially anecdotal evidence to hypothesize which hydrologic and agricultural factors contribute to groundwater vulnerability to <span class="hlt">nitrate</span> contamination. Analysis of nationally consistent measurements from the U.S. Geological Survey's NAWOA program confirms these hypotheses for a substantial range of agricultural systems. Shallow unconfined aquifers are most susceptible to <span class="hlt">nitrate</span> contamination associated with agricultural systems. Alluvial and other unconsolidated aquifers are the most vulnerable and shallow carbonate aquifers provide a substantial but smaller contamination risk. Where any of these aquifers are overlain by permeable soils the risk of contamination is larger. Irrigated systems can compound this vulnerability by increasing leaching facilitated by additional recharge and additional nutrient applications. The agricultural system of corn, soybeans, and hogs produced significantly larger <span class="hlt">concentrations</span> of groundwater <span class="hlt">nitrate</span> than all other agricultural systems, although mean <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in counties with dairy, poultry, cattle and grains, and horticulture systems were similar. If trends in the relation between increased fertilizer use and groundwater <span class="hlt">nitrate</span> in the United States are repeated in other regions of the world, Asia may experience increasing problems because of recent increases in fertilizer use. Groundwater monitoring in Western and Eastern Europe as well as Russia over the next decade may provide data to determine if the trend in increased <span class="hlt">nitrate</span> contamination can be reversed. If the <span class="hlt">concentrated</span> livestock trend in the United States is global, it may be accompanied by increasing nitrogen contamination in groundwater. <span class="hlt">Concentrated</span> livestock provide both point sources in the confinement area and intense non-point sources as fields close to facilities are used for manure disposal. Regions where irrigated cropland is expanding, such as in Asia, may experience the greatest impact of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27566937','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27566937"><span>Context-dependent environmental quality standards of soil <span class="hlt">nitrate</span> for terrestrial plant communities.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>van Goethem, Thomas M W J; Schipper, Aafke M; Wamelink, G W Wieger; Huijbregts, Mark A J</p> <p>2016-10-01</p> <p>Environmental quality standards (EQS) specify the maximum permissible <span class="hlt">concentration</span> or level of a specific environmental stressor. Here, a procedure is proposed to derive EQS that are specific to a representative species pool and conditional on confounding environmental factors. To illustrate the procedure, a dataset was used with plant species richness <span class="hlt">observations</span> of grasslands and forests and accompanying soil <span class="hlt">nitrate</span>-N and pH measurements collected from 981 sampling sites in the Netherlands. Species richness was related to soil <span class="hlt">nitrate</span>-N and pH with quantile regression allowing for interaction effects. The resulting regression models were used to derive EQS for <span class="hlt">nitrate</span> conditional on pH, quantified as the <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> at a specific pH level corresponding with a species richness equal to 95% of the species pool, for both grasslands and forest communities. The EQS varied between 1.8 mg/kg <span class="hlt">nitrate</span>-N at pH 9-65 mg/kg <span class="hlt">nitrate</span>-N at pH 4. EQS for forests and grasslands were similar, but EQS based on Red List species richness were considerably lower (more stringent) than those based on overall species richness, particularly at high pH levels. The results indicate that both natural background pH conditions and Red List species are important factors to consider in the derivation of EQS for soil <span class="hlt">nitrate</span>-N for terrestrial ecosystems. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23959335','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23959335"><span>Gallium <span class="hlt">nitrate</span> induces fibrinogen flocculation: an explanation for its hemostatic effect?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bauters, A; Holt, D J; Zerbib, P; Rogosnitzky, M</p> <p>2013-12-01</p> <p>A novel hemostatic effect of gallium <span class="hlt">nitrate</span> has recently been discovered. Our aim was to perform a preliminary investigation into its mode of action. Thromboelastography® showed no effect on coagulation but pointed instead to changes in fibrinogen <span class="hlt">concentration</span>. We measured functional fibrinogen in whole blood after addition of gallium <span class="hlt">nitrate</span> and nitric acid. We found that gallium <span class="hlt">nitrate</span> induces fibrinogen precipitation in whole blood to a significantly higher degree than solutions of nitric acid alone. This precipitate is not primarily pH driven, and appears to occur via flocculation. This behavior is in line with the generally <span class="hlt">observed</span> ability of metals to induce fibrinogen precipitation. Further investigation is required into this novel phenomenon.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4134218','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4134218"><span>The Fate of <span class="hlt">Nitrate</span> in Intertidal Permeable Sediments</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Marchant, Hannah K.; Lavik, Gaute; Holtappels, Moritz; Kuypers, Marcel M. M.</p> <p>2014-01-01</p> <p>Coastal zones act as a sink for riverine and atmospheric nitrogen inputs and thereby buffer the open ocean from the effects of anthropogenic activity. Recently, microbial activity in sandy permeable sediments has been identified as a dominant source of N-loss in coastal zones, namely through denitrification. Some of the highest coastal denitrification rates measured so far occur within the intertidal permeable sediments of the eutrophied Wadden Sea. Still, denitrification alone can often account for only half of the substantial <span class="hlt">nitrate</span> (NO3 −) consumption. Therefore, to investigate alternative NO3 − sinks such as dissimilatory <span class="hlt">nitrate</span> reduction to ammonium (DNRA), intracellular <span class="hlt">nitrate</span> storage by eukaryotes and isotope equilibration effects we carried out 15NO3 − amendment experiments. By considering all of these sinks in combination, we could quantify the fate of the 15NO3 − added to the sediment. Denitrification was the dominant <span class="hlt">nitrate</span> sink (50–75%), while DNRA, which recycles N to the environment accounted for 10–20% of NO3 − consumption. Intriguingly, we also <span class="hlt">observed</span> that between 20 and 40% of 15NO3 − added to the incubations entered an intracellular pool of NO3 − and was subsequently respired when <span class="hlt">nitrate</span> became limiting. Eukaryotes were responsible for a large proportion of intracellular <span class="hlt">nitrate</span> storage, and it could be shown through inhibition experiments that at least a third of the stored <span class="hlt">nitrate</span> was subsequently also respired by eukaryotes. The environmental significance of the intracellular <span class="hlt">nitrate</span> pool was confirmed by in situ measurements which revealed that intracellular storage can accumulate <span class="hlt">nitrate</span> at <span class="hlt">concentrations</span> six fold higher than the surrounding porewater. This intracellular pool is so far not considered when modeling N-loss from intertidal permeable sediments; however it can act as a reservoir for <span class="hlt">nitrate</span> during low tide. Consequently, <span class="hlt">nitrate</span> respiration supported by intracellular <span class="hlt">nitrate</span> storage can add an additional</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25127459','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25127459"><span>The fate of <span class="hlt">nitrate</span> in intertidal permeable sediments.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Marchant, Hannah K; Lavik, Gaute; Holtappels, Moritz; Kuypers, Marcel M M</p> <p>2014-01-01</p> <p>Coastal zones act as a sink for riverine and atmospheric nitrogen inputs and thereby buffer the open ocean from the effects of anthropogenic activity. Recently, microbial activity in sandy permeable sediments has been identified as a dominant source of N-loss in coastal zones, namely through denitrification. Some of the highest coastal denitrification rates measured so far occur within the intertidal permeable sediments of the eutrophied Wadden Sea. Still, denitrification alone can often account for only half of the substantial <span class="hlt">nitrate</span> (NO3-) consumption. Therefore, to investigate alternative NO3- sinks such as dissimilatory <span class="hlt">nitrate</span> reduction to ammonium (DNRA), intracellular <span class="hlt">nitrate</span> storage by eukaryotes and isotope equilibration effects we carried out 15NO3- amendment experiments. By considering all of these sinks in combination, we could quantify the fate of the 15NO3- added to the sediment. Denitrification was the dominant <span class="hlt">nitrate</span> sink (50-75%), while DNRA, which recycles N to the environment accounted for 10-20% of NO3- consumption. Intriguingly, we also <span class="hlt">observed</span> that between 20 and 40% of 15NO3- added to the incubations entered an intracellular pool of NO3- and was subsequently respired when <span class="hlt">nitrate</span> became limiting. Eukaryotes were responsible for a large proportion of intracellular <span class="hlt">nitrate</span> storage, and it could be shown through inhibition experiments that at least a third of the stored <span class="hlt">nitrate</span> was subsequently also respired by eukaryotes. The environmental significance of the intracellular <span class="hlt">nitrate</span> pool was confirmed by in situ measurements which revealed that intracellular storage can accumulate <span class="hlt">nitrate</span> at <span class="hlt">concentrations</span> six fold higher than the surrounding porewater. This intracellular pool is so far not considered when modeling N-loss from intertidal permeable sediments; however it can act as a reservoir for <span class="hlt">nitrate</span> during low tide. Consequently, <span class="hlt">nitrate</span> respiration supported by intracellular <span class="hlt">nitrate</span> storage can add an additional 20% to previous <span class="hlt">nitrate</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3319660','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3319660"><span>Plasma <span class="hlt">nitrate</span> and nitrite are increased by a high <span class="hlt">nitrate</span> supplement, but not by high <span class="hlt">nitrate</span> foods in older adults</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Miller, Gary D.; Marsh, Anthony P.; Dove, Robin W.; Beavers, Daniel; Presley, Tennille; Helms, Christine; Bechtold, Erika; King, S. Bruce; Kim-Shapiro, Daniel</p> <p>2012-01-01</p> <p>Little is known about the effect of dietary <span class="hlt">nitrate</span> on the <span class="hlt">nitrate</span>/nitrite/NO (nitric oxide) cycle in older adults. We examined the effect of a 3-day control diet vs. high <span class="hlt">nitrate</span> diet, with and without a high <span class="hlt">nitrate</span> supplement (beetroot juice), on plasma <span class="hlt">nitrate</span> and nitrite kinetics, and blood pressure using a randomized four period cross-over controlled design. We hypothesized that the high <span class="hlt">nitrate</span> diet would show higher levels of plasma <span class="hlt">nitrate</span>/nitrite and blood pressure compared to the control diet, which would be potentiated by the supplement. Participants were eight normotensive older men and women (5 female, 3 male, 72.5±4.7 yrs) with no overt disease or medications that affect NO metabolism. Plasma <span class="hlt">nitrate</span> and nitrite levels and blood pressure were measured prior to and hourly for 3 hours after each meal. The mean daily changes in plasma <span class="hlt">nitrate</span> and nitrite were significantly different from baseline for both control diet+supplement (p<0.001 and =0.017 for <span class="hlt">nitrate</span> and nitrite, respectively) and high <span class="hlt">nitrate</span> diet+supplement (p=0.001 and 0.002), but not for control diet (p=0.713 and 0.741) or high <span class="hlt">nitrate</span> diet (p=0.852 and 0.500). Blood pressure decreased from the morning baseline measure to the three 2 hr post-meal follow-up time-points for all treatments, but there was no main effect for treatment. In healthy older adults, a high <span class="hlt">nitrate</span> supplement consumed at breakfast elevated plasma <span class="hlt">nitrate</span> and nitrite levels throughout the day. This <span class="hlt">observation</span> may have practical utility for the timing of intake of a <span class="hlt">nitrate</span> supplement with physical activity for older adults with vascular dysfunction. PMID:22464802</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4080430','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4080430"><span><span class="hlt">Nitrate</span> and periplasmic <span class="hlt">nitrate</span> reductases</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Sparacino-Watkins, Courtney; Stolz, John F.; Basu, Partha</p> <p>2014-01-01</p> <p>The <span class="hlt">nitrate</span> anion is a simple, abundant and relatively stable species, yet plays a significant role in global cycling of nitrogen, global climate change, and human health. Although it has been known for quite some time that <span class="hlt">nitrate</span> is an important species environmentally, recent studies have identified potential medical applications. In this respect the <span class="hlt">nitrate</span> anion remains an enigmatic species that promises to offer exciting science in years to come. Many bacteria readily reduce <span class="hlt">nitrate</span> to nitrite via <span class="hlt">nitrate</span> reductases. Classified into three distinct types – periplasmic <span class="hlt">nitrate</span> reductase (Nap), respiratory <span class="hlt">nitrate</span> reductase (Nar) and assimilatory <span class="hlt">nitrate</span> reductase (Nas), they are defined by their cellular location, operon organization and active site structure. Of these, Nap proteins are the focus of this review. Despite similarities in the catalytic and spectroscopic properties Nap from different Proteobacteria are phylogenetically distinct. This review has two major sections: in the first section, <span class="hlt">nitrate</span> in the nitrogen cycle and human health, taxonomy of <span class="hlt">nitrate</span> reductases, assimilatory and dissimilatory <span class="hlt">nitrate</span> reduction, cellular locations of <span class="hlt">nitrate</span> reductases, structural and redox chemistry are discussed. The second section focuses on the features of periplasmic <span class="hlt">nitrate</span> reductase where the catalytic subunit of the Nap and its kinetic properties, auxiliary Nap proteins, operon structure and phylogenetic relationships are discussed. PMID:24141308</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24095993','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24095993"><span>Sustainable <span class="hlt">nitrate</span>-contaminated water treatment using multi cycle ion-exchange/bioregeneration of <span class="hlt">nitrate</span> selective resin.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ebrahimi, Shelir; Roberts, Deborah J</p> <p>2013-11-15</p> <p>The sustainability of ion-exchange treatment processes using high capacity single use resins to remove <span class="hlt">nitrate</span> from contaminated drinking water can be achieved by regenerating the exhausted resin and reusing it multiple times. In this study, multi cycle loading and bioregeneration of tributylamine strong base anion (SBA) exchange resin was studied. After each cycle of exhaustion, biological regeneration of the resin was performed using a salt-tolerant, <span class="hlt">nitrate</span>-perchlorate-reducing culture for 48 h. The resin was enclosed in a membrane to avoid direct contact of the resin with the culture. The results show that the culture was capable of regenerating the resin and allowing the resin to be used in multiple cycles. The <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> in the samples reached a peak in first 0.5-1h after placing the resin in medium because of desorption of <span class="hlt">nitrate</span> from resin with desorption rate of 0.099 ± 0.003 hr(-1). After this time, since microorganisms began to degrade the <span class="hlt">nitrate</span> in the aqueous phase, the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was generally non-detectable after 10h. The average of calculated specific degradation rate of <span class="hlt">nitrate</span> was -0.015 mg NO3(-)/mg VSS h. Applying 6 cycles of resin exhaustion/regeneration shows resin can be used for 4 cycles without a loss of capacity, after 6 cycles only 6% of the capacity was lost. This is the first published research to examine the direct regeneration of a resin enclosed in a membrane, to allow reuse without any disinfection or cleaning procedures. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4280893','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4280893"><span>Nebraska's groundwater legacy: <span class="hlt">Nitrate</span> contamination beneath irrigated cropland</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Exner, Mary E; Hirsh, Aaron J; Spalding, Roy F</p> <p>2014-01-01</p> <p>A 31 year record of ∼44,000 <span class="hlt">nitrate</span> analyses in ∼11,500 irrigation wells was utilized to depict the decadal expansion of groundwater <span class="hlt">nitrate</span> contamination (N ≥ 10 mg/L) in the irrigated corn-growing areas of eastern and central Nebraska and analyze long-term <span class="hlt">nitrate</span> <span class="hlt">concentration</span> trends in 17 management areas (MAs) subject to N fertilizer and budgeting requirements. The 1.3 M contaminated hectares were characterized by irrigation method, soil drainage, and vadose zone thickness and lithology. The areal extent and growth of contaminated groundwater in two predominately sprinkler-irrigated areas was only ∼20% smaller beneath well-drained silt loams with thick clayey-silt unsaturated layers and unsaturated thicknesses >15 m (400,000 ha and 15,000 ha/yr) than beneath well and excessively well-drained soils with very sandy vadose zones (511,000 ha and 18,600 ha/yr). Much slower expansion (3700 ha/yr) occurred in the 220,000 contaminated hectares in the central Platte valley characterized by predominately gravity irrigation on thick, well-drained silt loams above a thin (∼5.3 m), sandy unsaturated zone. The only reversals in long-term <span class="hlt">concentration</span> trends occurred in two MAs (120,500 ha) within this contaminated area. <span class="hlt">Concentrations</span> declined 0.14 and 0.20 mg N/L/yr (p < 0.02) to ∼18.3 and 18.8 mg N/L, respectively, during >20 years of management. Average annual <span class="hlt">concentrations</span> in 10 MAs are increasing (p < 0.05) and indicate that average <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in leachates below the root zone and groundwater <span class="hlt">concentrations</span> have not yet reached steady state. While management practices likely have slowed increases in groundwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, irrigation and nutrient applications must be more effectively controlled to retain <span class="hlt">nitrate</span> in the root zone. PMID:25558112</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4495186','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4495186"><span>Molybdenum Availability Is Key to <span class="hlt">Nitrate</span> Removal in Contaminated Groundwater Environments</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Thorgersen, Michael P.; Lancaster, W. Andrew; Vaccaro, Brian J.; Poole, Farris L.; Rocha, Andrea M.; Mehlhorn, Tonia; Pettenato, Angelica; Ray, Jayashree; Waters, R. Jordan; Melnyk, Ryan A.; Chakraborty, Romy; Deutschbauer, Adam M.; Arkin, Adam P.</p> <p>2015-01-01</p> <p>The <span class="hlt">concentrations</span> of molybdenum (Mo) and 25 other metals were measured in groundwater samples from 80 wells on the Oak Ridge Reservation (ORR) (Oak Ridge, TN), many of which are contaminated with <span class="hlt">nitrate</span>, as well as uranium and various other metals. The <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and uranium were in the ranges of 0.1 μM to 230 mM and <0.2 nM to 580 μM, respectively. Almost all metals examined had significantly greater median <span class="hlt">concentrations</span> in a subset of wells that were highly contaminated with uranium (≥126 nM). They included cadmium, manganese, and cobalt, which were 1,300- to 2,700-fold higher. A notable exception, however, was Mo, which had a lower median <span class="hlt">concentration</span> in the uranium-contaminated wells. This is significant, because Mo is essential in the dissimilatory <span class="hlt">nitrate</span> reduction branch of the global nitrogen cycle. It is required at the catalytic site of <span class="hlt">nitrate</span> reductase, the enzyme that reduces <span class="hlt">nitrate</span> to nitrite. Moreover, more than 85% of the groundwater samples contained less than 10 nM Mo, whereas <span class="hlt">concentrations</span> of 10 to 100 nM Mo were required for efficient growth by <span class="hlt">nitrate</span> reduction for two Pseudomonas strains isolated from ORR wells and by a model denitrifier, Pseudomonas stutzeri RCH2. Higher <span class="hlt">concentrations</span> of Mo tended to inhibit the growth of these strains due to the accumulation of toxic <span class="hlt">concentrations</span> of nitrite, and this effect was exacerbated at high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. The relevance of these results to a Mo-based <span class="hlt">nitrate</span> removal strategy and the potential community-driving role that Mo plays in contaminated environments are discussed. PMID:25979890</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1334444','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1334444"><span>Molybdenum Availability Is Key to <span class="hlt">Nitrate</span> Removal in Contaminated Groundwater Environments</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>Thorgersen, Michael P.; Lancaster, W. Andrew; Vaccaro, Brian J.</p> <p>2015-05-15</p> <p>The <span class="hlt">concentrations</span> of molybdenum (Mo) and 25 other metals were measured in groundwater samples from 80 wells on the Oak Ridge Reservation (ORR) (Oak Ridge, TN), many of which are contaminated with <span class="hlt">nitrate</span>, as well as uranium and various other metals. Moreover, the <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and uranium were in the ranges of 0.1 μM to 230 mM and <0.2 nM to 580 μM, respectively. Most metals examined had significantly greater median <span class="hlt">concentrations</span> in a subset of wells that were highly contaminated with uranium (≥126 nM). They included cadmium, manganese, and cobalt, which were 1,300- to 2,700-fold higher. A notablemore » exception, however, was Mo, which had a lower median <span class="hlt">concentration</span> in the uranium-contaminated wells. This is significant, because Mo is essential in the dissimilatory <span class="hlt">nitrate</span> reduction branch of the global nitrogen cycle. It is required at the catalytic site of <span class="hlt">nitrate</span> reductase, the enzyme that reduces <span class="hlt">nitrate</span> to nitrite. Furthermore, more than 85% of the groundwater samples contained less than 10 nM Mo, whereas <span class="hlt">concentrations</span> of 10 to 100 nM Mo were required for efficient growth by <span class="hlt">nitrate</span> reduction for twoPseudomonasstrains isolated from ORR wells and by a model denitrifier,Pseudomonas stutzeriRCH2. Higher <span class="hlt">concentrations</span> of Mo tended to inhibit the growth of these strains due to the accumulation of toxic <span class="hlt">concentrations</span> of nitrite, and this effect was exacerbated at high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. The relevance of these results to a Mo-based <span class="hlt">nitrate</span> removal strategy and the potential community-driving role that Mo plays in contaminated environments are discussed.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_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('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=542443','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=542443"><span>Differential Effect of Irradiance and Nutrient <span class="hlt">Nitrate</span> on the Relationship of in Vivo and in Vitro <span class="hlt">Nitrate</span> Reductase Assay in Chlorophyllous Tissues 1</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jones, Richard Wyn; Sheard, Robert W.</p> <p>1977-01-01</p> <p>Growth at increasing continuous irradiance (at high nutrient <span class="hlt">nitrate</span>) and nutrient <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> (at high continuous irradiance) furnished increases in the in vivo and in vitro <span class="hlt">nitrate</span> reductase activities of corn (Zea mays L.), field peas (Pisum arvense L.), wheat (Triticum aestivum L.), barley (Hordeum vulgare L.), and globe amaranth (Gomphrena globosa L.) leaves and of marrow (Cucurbita pepo L.) cotyledons. Ratios of in vivo to in vitro activity declined exponentially in all species with increasing <span class="hlt">nitrate</span> reductase levels promoted by nutrient <span class="hlt">nitrate</span>. The ratios were more nearly independent of <span class="hlt">nitrate</span> reductase levels generated by adjusting the irradiance; major exceptions were marrow and wheat at low (1.5 klux and less) irradiances and peas throughout the irradiance range, where decreases in the ratio were accompanied by increases in in situ <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. The ratio also increased at the highest irradiance (39.2 klux) in wheat and barley, associated with a decline of in vitro <span class="hlt">nitrate</span> reductase. These differences in response to irradiance and nutrient <span class="hlt">nitrate</span> indicate that the in vivo assay does not provide a simple measure of <span class="hlt">nitrate</span> reductase but rather yields a more composite measure of <span class="hlt">nitrate</span> reduction, possibly related both to <span class="hlt">nitrate</span> reductase level and to the supply of reductant for in vivo activity. PMID:16659888</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/fs/1997/0085/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/fs/1997/0085/report.pdf"><span>Herbicides and <span class="hlt">nitrates</span> in the Iowa River alluvial aquifer prior to changing land use, Iowa County, Iowa, 1996</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Savoca, Mark E.; Tobias, Jennifer L.; Sadorf, Eric M.; Birkenholtz, Trevor L.</p> <p>1997-01-01</p> <p>Four herbicides (alachlor, atrazine, cyanazine, and metolachlor) and one nutrient (<span class="hlt">nitrate</span>) were selected for study on the basis of frequent usage in Iowa and high detection rates in ground water (Detroy and Kuzniar, 1988). Alachlor was not detected at <span class="hlt">concentrations</span> greater than the method detection limit (MDL). Atrazine was detected at <span class="hlt">concentrations</span> greater than the MDL in samples from 48 percent of the 23 wells, cyanazine from 13 percent, metolachlor from 26 percent, and <span class="hlt">nitrate</span> from 91 percent. None of the four herbicides were detected at <span class="hlt">concentrations</span> greater than the respective U.S. Environmental Protection Agency's (USEPA) Maximum Contaminant Level (MCL) for drinking water. Thirteen percent of the samples had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above the USEPA's MCL of 10 mg/L (milligrams per liter). Relations between constituent <span class="hlt">concentration</span> and well depth were <span class="hlt">observed</span> for specific constituents at individual well nests.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19186792','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19186792"><span>[<span class="hlt">Nitrate</span> pollution in groundwater for drinking and its affecting factors in Hailun, northeast China].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, Xin-Feng; Yang, Li-Rong; Shi, Qian; Ma, Yan; Zhang, Yan-Yan; Chen, Li-Ding; Zheng, Hai-Feng</p> <p>2008-11-01</p> <p><span class="hlt">Nitrate</span> pollution in groundwater has become a worldwide problem. It may affect the water quality for daily use and thus the health of people. The temporal and spatial characteristics of <span class="hlt">nitrate</span> pollution in the groundwater were addressed by sample analysis of the drinkable water from 157 wells in Hailun, Heilongjiang, northeastern China. It was found that the mean value of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in all wells was 14.01 mg x L(-1). Of all the samples, the <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> of 26.11% wells exceeded the standard of drinkable water (10.00 mg x L(-1)). A significant difference was found on the spatial distribution of <span class="hlt">nitrate</span> pollution in the study area. The pollution degree in term of <span class="hlt">nitrate</span> pollution was in the order: the central rolling hills and flooding plain > the northeastern mountain area > the southwest rolling hills and plain. Based on the results, the factors causing the pollution we analyzed from the well properties and pollution sources. As for well properties, the type of the pipe material plays a critical role in the groundwater <span class="hlt">nitrate</span> pollution. It was found that the wells with seamless pipe have less pollution than those with multiple-sections pipe. The <span class="hlt">concentrations</span> of seamless pipe wells and multiple ones were respectively 5.08 mg x L(-1) and 32.57 mg x L(-1), 12.26% and 82.35% of these two kinds wells exceeded 10.00 mg x L(-1), the state drinking water standard. In the whole Hailun, there is no statistically relationship between <span class="hlt">nitrate</span>-N levels of wells and the well depth. However, a statistically lower <span class="hlt">nitrate</span>-N was <span class="hlt">observed</span> in the deep wells than that in the shallower ones. The mean values of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of the seamless-pipe deep wells, seamless-pipe shallow wells, multiple-section-pipe deep wells and multiple-section-pipe shallow wells were 1.84, 12.02, 25.14 and 45.61 mg x L(-1). Analysis of pollution source shows that the heavily polluted regions are usually associated with large use of nitrogen fertilizer and household livestock</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29605158','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29605158"><span>Three-dimensional modeling of <span class="hlt">nitrate</span>-N transport in vadose zone: Roles of soil heterogeneity and groundwater flux.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Akbariyeh, Simin; Bartelt-Hunt, Shannon; Snow, Daniel; Li, Xu; Tang, Zhenghong; Li, Yusong</p> <p>2018-04-01</p> <p>Contamination of groundwater from nitrogen fertilizers in agricultural lands is an important environmental and water quality management issue. It is well recognized that in agriculturally intensive areas, fertilizers and pesticides may leach through the vadose zone and eventually reach groundwater. While numerical models are commonly used to simulate fate and transport of agricultural contaminants, few models have considered a controlled field work to investigate the influence of soil heterogeneity and groundwater flow on <span class="hlt">nitrate</span>-N distribution in both root zone and deep vadose zone. In this work, a numerical model was developed to simulate <span class="hlt">nitrate</span>-N transport and transformation beneath a center pivot-irrigated corn field on Nebraska Management System Evaluation area over a three-year period. The model was based on a realistic three-dimensional sediment lithology, as well as carefully controlled irrigation and fertilizer application plans. In parallel, a homogeneous soil domain, containing the major sediment type of the site (i.e. sandy loam), was developed to conduct the same water flow and <span class="hlt">nitrate</span>-N leaching simulations. Simulated <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> were compared with the monitored <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> in 10 multi-level sampling wells over a three-year period. Although soil heterogeneity was mainly <span class="hlt">observed</span> from top soil to 3 m below the surface, heterogeneity controlled the spatial distribution of <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span>. Soil heterogeneity, however, has minimal impact on the total mass of <span class="hlt">nitrate</span>-N in the domain. In the deeper saturated zone, short-term variations of <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span> correlated with the groundwater level fluctuations. Copyright © 2018 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JCHyd.211...15A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JCHyd.211...15A"><span>Three-dimensional modeling of <span class="hlt">nitrate</span>-N transport in vadose zone: Roles of soil heterogeneity and groundwater flux</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Akbariyeh, Simin; Bartelt-Hunt, Shannon; Snow, Daniel; Li, Xu; Tang, Zhenghong; Li, Yusong</p> <p>2018-04-01</p> <p>Contamination of groundwater from nitrogen fertilizers in agricultural lands is an important environmental and water quality management issue. It is well recognized that in agriculturally intensive areas, fertilizers and pesticides may leach through the vadose zone and eventually reach groundwater. While numerical models are commonly used to simulate fate and transport of agricultural contaminants, few models have considered a controlled field work to investigate the influence of soil heterogeneity and groundwater flow on <span class="hlt">nitrate</span>-N distribution in both root zone and deep vadose zone. In this work, a numerical model was developed to simulate <span class="hlt">nitrate</span>-N transport and transformation beneath a center pivot-irrigated corn field on Nebraska Management System Evaluation area over a three-year period. The model was based on a realistic three-dimensional sediment lithology, as well as carefully controlled irrigation and fertilizer application plans. In parallel, a homogeneous soil domain, containing the major sediment type of the site (i.e. sandy loam), was developed to conduct the same water flow and <span class="hlt">nitrate</span>-N leaching simulations. Simulated <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> were compared with the monitored <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> in 10 multi-level sampling wells over a three-year period. Although soil heterogeneity was mainly <span class="hlt">observed</span> from top soil to 3 m below the surface, heterogeneity controlled the spatial distribution of <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span>. Soil heterogeneity, however, has minimal impact on the total mass of <span class="hlt">nitrate</span>-N in the domain. In the deeper saturated zone, short-term variations of <span class="hlt">nitrate</span>-N <span class="hlt">concentration</span> correlated with the groundwater level fluctuations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29572447','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29572447"><span>Coupling of oceanic carbon and nitrogen facilitates spatially resolved quantitative reconstruction of <span class="hlt">nitrate</span> inventories.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Glock, Nicolaas; Erdem, Zeynep; Wallmann, Klaus; Somes, Christopher J; Liebetrau, Volker; Schönfeld, Joachim; Gorb, Stanislav; Eisenhauer, Anton</p> <p>2018-03-23</p> <p>Anthropogenic impacts are perturbing the global nitrogen cycle via warming effects and pollutant sources such as chemical fertilizers and burning of fossil fuels. Understanding controls on past nitrogen inventories might improve predictions for future global biogeochemical cycling. Here we show the quantitative reconstruction of deglacial bottom water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> from intermediate depths of the Peruvian upwelling region, using foraminiferal pore density. Deglacial <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> correlate strongly with downcore δ 13 C, consistent with modern water column <span class="hlt">observations</span> in the intermediate Pacific, facilitating the use of δ 13 C records as a paleo-<span class="hlt">nitrate</span>-proxy at intermediate depths and suggesting that the carbon and nitrogen cycles were closely coupled throughout the last deglaciation in the Peruvian upwelling region. Combining the pore density and intermediate Pacific δ 13 C records shows an elevated <span class="hlt">nitrate</span> inventory of >10% during the Last Glacial Maximum relative to the Holocene, consistent with a δ 13 C-based and δ 15 N-based 3D ocean biogeochemical model and previous box modeling studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26318690','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26318690"><span>Residence times of groundwater and <span class="hlt">nitrate</span> transport in coastal aquifer systems: Daweijia area, northeastern China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Han, Dongmei; Cao, Guoliang; McCallum, James; Song, Xianfang</p> <p>2015-12-15</p> <p>Groundwater within the coastal aquifer systems of the Daweijia area in northeastern China is characterized by a large of variations (33-521mg/L) in NO3(-) <span class="hlt">concentrations</span>. Elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, in addition to seawater intrusion in the Daweijia well field, both attributable to anthropogenic activities, may impact future water-management practices. Chemical and stable isotopic (δ(18)O, δ(2)H) analysis, (3)H and CFCs methods were applied to provide a better understanding of the relationship between the distribution of groundwater mean residence time (MRT) and <span class="hlt">nitrate</span> transport, and to identify sources of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the complex coastal aquifer systems. There is a relatively narrow range of isotopic composition (ranging from -8.5 to -7.0‰) in most groundwater. Generally higher tritium contents <span class="hlt">observed</span> in the wet season relative to the dry season may result from rapid groundwater circulation in response to the rainfall through the preferential flow paths. In the well field, the relatively increased <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> of groundwater, accompanied by the higher tritium contents in the wet season, indicate the <span class="hlt">nitrate</span> pollution can be attributed to domestic wastes. The binary exponential and piston-flow mixing model (BEP) yielded feasible age distributions based on the conceptual model. The good inverse relationship between groundwater MRTs (92-467years) and the NO3(-) <span class="hlt">concentrations</span> in the shallow Quaternary aquifers indicates that elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are attributable to more recent recharge for shallow groundwater. However, there is no significant relationship between the MRTs (8-411years) and the NO3(-) <span class="hlt">concentrations</span> existing in the carbonate aquifer system, due to the complex hydrogeological conditions, groundwater age distributions and the range of contaminant source areas. <span class="hlt">Nitrate</span> in the groundwater system without denitrification effects could accumulate and be transported for tens of years, through the complex carbonate</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70031037','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70031037"><span><span class="hlt">Nitrate</span> in aquifers beneath agricultural systems</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Burkart, M.R.; Stoner, J.D.; ,</p> <p>2007-01-01</p> <p>Research from several regions of the world provides spatially anecdotal evidence to hypothesize which hydrologic and agricultural factors contribute to groundwater vulnerability to <span class="hlt">nitrate</span> contamination. Analysis of nationally consistent measurements from the U.S. Geological Survey's NAWQA program confirms these hypotheses for a substantial range of agricultural systems. Shallow unconfined aquifers are most susceptible to <span class="hlt">nitrate</span> contamination associated with agricultural systems. Alluvial and other unconsolidated aquifers are the most vulnerable and also shallow carbonate aquifers that provide a substantial but smaller contamination risk. Where any of these aquifers are overlain by permeable soils the risk of contamination is larger. Irrigated systems can compound this vulnerability by increasing leaching facilitated by additional recharge and additional nutrient applications. The system of corn, soybean, and hogs produced significantly larger <span class="hlt">concentrations</span> of groundwater <span class="hlt">nitrate</span> than all other agricultural systems because this system imports the largest amount of N-fertilizer per unit production area. Mean <span class="hlt">nitrate</span> under dairy, poultry, horticulture, and cattle and grains systems were similar. If trends in the relation between increased fertilizer use and groundwater <span class="hlt">nitrate</span> in the United States are repeated in other regions of the world, Asia may experience increasing problems because of recent increases in fertilizer use. Groundwater monitoring in Western and Eastern Europe as well as Russia over the next decade may provide data to determine if the trend in increased <span class="hlt">nitrate</span> contamination can be reversed. If the <span class="hlt">concentrated</span> livestock trend in the United States is global, it may be accompanied by increasing nitrogen contamination in groundwater. <span class="hlt">Concentrated</span> livestock provide both point sources in the confinement area and intense non-point sources as fields close to facilities are used for manure disposal. Regions where irrigated cropland is expanding, such as</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.H44A..05M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.H44A..05M"><span><span class="hlt">Nitrate</span> Contamination of Deep Aquifers in the Salinas Valley, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Moran, J. E.; Esser, B. K.; Hillegonds, D. J.; Holtz, M.; Roberts, S. K.; Singleton, M. J.; Visser, A.; Kulongoski, J. T.; Belitz, K.</p> <p>2011-12-01</p> <p>The Salinas Valley, known as 'the salad bowl of the world', has been an agricultural center for more than 100 years. Irrigated row crops such as lettuce and strawberries dominate both land use and water use. Groundwater is the exclusive supply for both irrigation and drinking water. Some irrigation wells and most public water supply wells in the Salinas Valley are constructed to draw water from deep portions of the aquifer system, where contamination by <span class="hlt">nitrate</span> is less likely than in the shallow portions of the aquifer system. However, a number of wells with top perforations greater than 75 m deep, screened below confining or semi-confining units, have <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> greater than the Maximum Contaminant Limit (MCL) of 45 mg/L as NO3-. This study uses <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> from several hundred irrigation, drinking water, and monitoring wells (Monterey County Water Resources Agency, 1997), along with tritium-helium groundwater ages acquired at Lawrence Livermore National Laboratory through the State of California Groundwater Monitoring and Assessment (GAMA) program (reported in Kulongoski et al., 2007 and in Moran et al., in press), to identify <span class="hlt">nitrate</span> 'hot spots' in the deep aquifer and to examine possible modes of <span class="hlt">nitrate</span> transport to the deep aquifer. In addition, <span class="hlt">observed</span> apparent groundwater ages are compared with the results of transport simulations that use particle tracking and a stochastic-geostatistical framework to incorporate aquifer heterogeneity to determine the distribution of travel times from the water table to each well (Fogg et al., 1999). The combined evidence from <span class="hlt">nitrate</span>, tritium, tritiogenic 3He, and radiogenic 4He <span class="hlt">concentrations</span>, reveals complex recharge and flow to the capture zone of the deep drinking water wells. Widespread groundwater pumping for irrigation accelerates vertical groundwater flow such that high <span class="hlt">nitrate</span> groundwater reaches some deep drinking water wells. Deeper portions of the wells often draw in water that recharged</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/fs/2013/3109/pdf/fs2013-3109.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/fs/2013/3109/pdf/fs2013-3109.pdf"><span>Real-time continuous <span class="hlt">nitrate</span> monitoring in Illinois in 2013</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Warner, Kelly L.; Terrio, Paul J.; Straub, Timothy D.; Roseboom, Donald; Johnson, Gary P.</p> <p>2013-01-01</p> <p>Many sources contribute to the nitrogen found in surface water in Illinois. Illinois is located in the most productive agricultural area in the country, and nitrogen fertilizer is commonly used to maximize corn production in this area. Additionally, septic/wastewater systems, industrial emissions, and lawn fertilizer are common sources of nitrogen in urban areas of Illinois. In agricultural areas, the use of fertilizer has increased grain production to meet the needs of a growing population, but also has resulted in increases in nitrogen <span class="hlt">concentrations</span> in many streams and aquifers (Dubrovsky and others, 2010). The urban sources can increase nitrogen <span class="hlt">concentrations</span>, too. The Federal limit for <span class="hlt">nitrate</span> nitrogen in water that is safe to drink is 10 milligrams per liter (mg/L) (http://water.epa.gov/drink/contaminants/basicinformation/<span class="hlt">nitrate</span>.cfm, accessed on May 24, 2013). In addition to the concern with <span class="hlt">nitrate</span> nitrogen in drinking water, nitrogen, along with phosphorus, is an aquatic concern because it feeds the intensive growth of algae that are responsible for the hypoxic zone in the Gulf of Mexico. The largest nitrogen flux to the waters feeding the Gulf of Mexico is from Illinois (Alexander and others, 2008). Most studies of nitrogen in surface water and groundwater include samples for <span class="hlt">nitrate</span> nitrogen collected weekly or monthly, but <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> can change rapidly and these discrete samples may not capture rapid changes in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> that can affect human and aquatic health. Continuous monitoring for <span class="hlt">nitrate</span> could inform scientists and water-resource managers of these changes and provide information on the transport of <span class="hlt">nitrate</span> in surface water and groundwater.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024706','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024706"><span>Probability of <span class="hlt">nitrate</span> contamination of recently recharged groundwaters in the conterminous United States</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Nolan, B.T.; Hitt, K.J.; Ruddy, B.C.</p> <p>2002-01-01</p> <p>A new logistic regression (LR) model was used to predict the probability of <span class="hlt">nitrate</span> contamination exceeding 4 mg/L in predominantly shallow, recently recharged ground waters of the United States. The new model contains variables representing (1) N fertilizer loading (p 2 = 0.875), indicating that the LR model fits the data well. The likelihood of <span class="hlt">nitrate</span> contamination is greater in areas with high N loading and well-drained surficial soils over unconsolidated sand and gravels. The LR model correctly predicted the status of <span class="hlt">nitrate</span> contamination in 75% of wells in a validation data set. Considering all wells used in both calibration and validation, <span class="hlt">observed</span> median <span class="hlt">nitrate</span> <span class="hlt">concentration</span> increased from 0.24 to 8.30 mg/L as the mapped probability of <span class="hlt">nitrate</span> exceeding 4 mg/L increased from less than or equal to 0.17 to > 0.83.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70010011','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70010011"><span>Decadal-scale changes of <span class="hlt">nitrate</span> in ground water of the United States, 1988-2004</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Rupert, Michael G.</p> <p>2008-01-01</p> <p>This study evaluated decadal-scale changes of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater samples collected by the USGS National Water-Quality Assessment Program from 495 wells in 24 well networks across the USA in predominantly agricultural areas. Each well network was sampled once during 1988-1995 and resampled once during 2000-2004. Statistical tests of decadal-scale changes of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in water from all 495 wells combined indicate there is a significant increase in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the data set as a whole. Eight out of the 24 well networks, or about 33%, had significant changes of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Of the eight well networks with significant decadal-scale changes of <span class="hlt">nitrate</span>, all except one, the Willamette Valley of Oregon, had increasing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> of three of those eight well networks increased above the USEPA maximum contaminant level of 10 mg L-1. <span class="hlt">Nitrate</span> in water from wells with reduced conditions had significantly smaller decadal-scale changes in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> than oxidized and mixed waters. A subset of wells had data on ground water recharge date; <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increased in response to the increase of N fertilizer use since about 1950. Determining ground water recharge dates is an important component of a ground water trends investigation because recharge dates provide a link between changes in ground water quality and changes in land-use practices. Copyright ?? 2008 by the American Society of Agronomy, Crop Science Society of America, and Soil Science Society of America. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMPP51C1087J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMPP51C1087J"><span>First Results of <span class="hlt">Nitrate</span> and its Stable Isotopic Composition in an Ice Core from Dome A, East Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jiang, S.</p> <p>2017-12-01</p> <p>During the 21st Chinese Antarctic Research Expedition in 2004/2005 austral summer, a 109.91 m ice core (hereafter DA2005 core) was recovered at the site about 300 m away from the summit of Dome A. The top 100.42 m was analyzed for major chemical impurities and isotopic composition of <span class="hlt">nitrate</span>. Dating was based on the volcanic stratigraphy and average annual accumulation rate. Results showed that the analyzed 100.42 m part of the core covers the last 2840 years before present, from 840 BC to AD 1998. <span class="hlt">Nitrate</span> <span class="hlt">concentration</span> in the DA2005 core varies between 2.86 μg kg-1 and 30.75 μg kg-1 throughout the 2840 years, with the mean <span class="hlt">concentration</span> of 11.84 µg kg-1. Comparisons with previous Antarctic ice core <span class="hlt">nitrate</span> records show that the DA2005 core has the lowest mean <span class="hlt">concentration</span> of <span class="hlt">nitrate</span>, which is consistent with the lowest accumulation rate at Dome A among these sampling sites. Decreased <span class="hlt">nitrate</span> <span class="hlt">concentration</span> during the period of Little Ice Age (AD 1500-1900) is <span class="hlt">observed</span> in the DA2005 core. The δ15N(NO3-) values vary between 235.4 ‰ and 279.4 ‰, which suggest strong 15N enrichment in the DA2005 core. The sample covering the most recent time period (AD 1695-1838) has the lowest δ15N(NO3-) value. The Δ17O(NO3-) values span from 28.9 ‰ to 31.4 ‰, which is among the range ever <span class="hlt">observed</span>. An increasing trend is seen during the period of AD 1225-1838, which corresponds to the time period when <span class="hlt">nitrate</span> <span class="hlt">concentration</span> remains low. The maximum Δ17O(NO3-) value occurs in the period AD 1695-1838, and the minimum value occurs in the period AD 62-166.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26261901','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26261901"><span>Organic <span class="hlt">Nitrate</span> Therapy, <span class="hlt">Nitrate</span> Tolerance, and <span class="hlt">Nitrate</span>-Induced Endothelial Dysfunction: Emphasis on Redox Biology and Oxidative Stress.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Daiber, Andreas; Münzel, Thomas</p> <p>2015-10-10</p> <p>Organic <span class="hlt">nitrates</span>, such as nitroglycerin (GTN), isosorbide-5-mononitrate and isosorbide dinitrate, and pentaerithrityl tetranitrate (PETN), when given acutely, have potent vasodilator effects improving symptoms in patients with acute and chronic congestive heart failure, stable coronary artery disease, acute coronary syndromes, or arterial hypertension. The mechanisms underlying vasodilation include the release of •NO or a related compound in response to intracellular bioactivation (for GTN, the mitochondrial aldehyde dehydrogenase [ALDH-2]) and activation of the enzyme, soluble guanylyl cyclase. Increasing cyclic guanosine-3',-5'-monophosphate (cGMP) levels lead to an activation of the cGMP-dependent kinase I, thereby causing the relaxation of the vascular smooth muscle by decreasing intracellular calcium <span class="hlt">concentrations</span>. The hemodynamic and anti-ischemic effects of organic <span class="hlt">nitrates</span> are rapidly lost upon long-term (low-dose) administration due to the rapid development of tolerance and endothelial dysfunction, which is in most cases linked to increased intracellular oxidative stress. Enzymatic sources of reactive oxygen species under <span class="hlt">nitrate</span> therapy include mitochondria, NADPH oxidases, and an uncoupled •NO synthase. Acute high-dose challenges with organic <span class="hlt">nitrates</span> cause a similar loss of potency (tachyphylaxis), but with distinct pathomechanism. The differences among organic <span class="hlt">nitrates</span> are highlighted regarding their potency to induce oxidative stress and subsequent tolerance and endothelial dysfunction. We also address pleiotropic effects of organic <span class="hlt">nitrates</span>, for example, their capacity to stimulate antioxidant pathways like those demonstrated for PETN, all of which may prevent adverse effects in response to long-term therapy. Based on these considerations, we will discuss and present some preclinical data on how the <span class="hlt">nitrate</span> of the future should be designed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4752190','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4752190"><span>Organic <span class="hlt">Nitrate</span> Therapy, <span class="hlt">Nitrate</span> Tolerance, and <span class="hlt">Nitrate</span>-Induced Endothelial Dysfunction: Emphasis on Redox Biology and Oxidative Stress</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>2015-01-01</p> <p>Abstract Organic <span class="hlt">nitrates</span>, such as nitroglycerin (GTN), isosorbide-5-mononitrate and isosorbide dinitrate, and pentaerithrityl tetranitrate (PETN), when given acutely, have potent vasodilator effects improving symptoms in patients with acute and chronic congestive heart failure, stable coronary artery disease, acute coronary syndromes, or arterial hypertension. The mechanisms underlying vasodilation include the release of •NO or a related compound in response to intracellular bioactivation (for GTN, the mitochondrial aldehyde dehydrogenase [ALDH-2]) and activation of the enzyme, soluble guanylyl cyclase. Increasing cyclic guanosine-3′,-5′-monophosphate (cGMP) levels lead to an activation of the cGMP-dependent kinase I, thereby causing the relaxation of the vascular smooth muscle by decreasing intracellular calcium <span class="hlt">concentrations</span>. The hemodynamic and anti-ischemic effects of organic <span class="hlt">nitrates</span> are rapidly lost upon long-term (low-dose) administration due to the rapid development of tolerance and endothelial dysfunction, which is in most cases linked to increased intracellular oxidative stress. Enzymatic sources of reactive oxygen species under <span class="hlt">nitrate</span> therapy include mitochondria, NADPH oxidases, and an uncoupled •NO synthase. Acute high-dose challenges with organic <span class="hlt">nitrates</span> cause a similar loss of potency (tachyphylaxis), but with distinct pathomechanism. The differences among organic <span class="hlt">nitrates</span> are highlighted regarding their potency to induce oxidative stress and subsequent tolerance and endothelial dysfunction. We also address pleiotropic effects of organic <span class="hlt">nitrates</span>, for example, their capacity to stimulate antioxidant pathways like those demonstrated for PETN, all of which may prevent adverse effects in response to long-term therapy. Based on these considerations, we will discuss and present some preclinical data on how the <span class="hlt">nitrate</span> of the future should be designed. Antioxid. Redox Signal. 23, 899–942. PMID:26261901</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.H53H1524P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.H53H1524P"><span>In situ <span class="hlt">nitrate</span> measurements capture short-term variability and seasonal transitions during a drought - flood year in the Mississippi River Basin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pellerin, B. A.; Bergamaschi, B. A.; Saraceno, J.; Downing, B. D.; Crawford, C.; Gilliom, R.; Frederick, P.</p> <p>2013-12-01</p> <p>Nitrogen flux from the Mississippi River to the Gulf of Mexico has received considerable attention because it fuels primary production on the continental shelf and can contribute to the summer hypoxia <span class="hlt">observed</span> in the Gulf. Accurately quantifying the load of nitrogen - particularly as <span class="hlt">nitrate</span> - to the Gulf is critical for both predicting the size of the oxygen-depleted dead zone and establishing targets for N load reduction from the basin. Fluxes have been historically calculated with load estimation models using 5-10 years of discrete <span class="hlt">nitrate</span> data collected approximately 12-18 times per year. These traditional monthly to biweekly sampling intervals often fail to adequately capture hydrologic pulses ranging from early snowmelt periods to short-duration rainfall events in small streams, but the ability to adequately resolve patterns in water quality in large rivers has received much less attention. The recent commercial availability of in situ optical sensors for <span class="hlt">nitrate</span>, together with new techniques for data collection and analysis, provides an opportunity to measure <span class="hlt">nitrate</span> <span class="hlt">concentration</span> on time scales in which environmental conditions actually change. Data have been collected and analyzed from a USGS optical <span class="hlt">nitrate</span> sensor deployed in the Mississippi River at Baton Rouge, Louisiana, since November 2011. Our <span class="hlt">nitrate</span> data, collected at three hour intervals, shows a strong relationship to depth- and width-integrated discrete <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> measured on 20 dates (r2=0.99, slope=1) after correcting for a consistent, small positive bias (0.10 mg/L). The close relationship between the in situ data measured on edge of the channel and the depth- and width-integrated sample suggests that the fixed sensor measurements provide a robust proxy for cross-sectional averaged <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> at Baton Rouge under a range of flow conditions. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> ranged from a low of 0.19 mg/L as N on September 11, 2012 to a high of 3.09 mg/L as N on July 12, 2013</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ACP....1711819M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ACP....1711819M"><span>Light-induced protein <span class="hlt">nitration</span> and degradation with HONO emission</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Meusel, Hannah; Elshorbany, Yasin; Kuhn, Uwe; Bartels-Rausch, Thorsten; Reinmuth-Selzle, Kathrin; Kampf, Christopher J.; Li, Guo; Wang, Xiaoxiang; Lelieveld, Jos; Pöschl, Ulrich; Hoffmann, Thorsten; Su, Hang; Ammann, Markus; Cheng, Yafang</p> <p>2017-10-01</p> <p>Proteins can be <span class="hlt">nitrated</span> by air pollutants (NO2), enhancing their allergenic potential. This work provides insight into protein <span class="hlt">nitration</span> and subsequent decomposition in the presence of solar radiation. We also investigated light-induced formation of nitrous acid (HONO) from protein surfaces that were <span class="hlt">nitrated</span> either online with instantaneous gas-phase exposure to NO2 or offline by an efficient <span class="hlt">nitration</span> agent (tetranitromethane, TNM). Bovine serum albumin (BSA) and ovalbumin (OVA) were used as model substances for proteins. <span class="hlt">Nitration</span> degrees of about 1 % were derived applying NO2 <span class="hlt">concentrations</span> of 100 ppb under VIS/UV illuminated conditions, while simultaneous decomposition of (<span class="hlt">nitrated</span>) proteins was also found during long-term (20 h) irradiation exposure. Measurements of gas exchange on TNM-<span class="hlt">nitrated</span> proteins revealed that HONO can be formed and released even without contribution of instantaneous heterogeneous NO2 conversion. NO2 exposure was found to increase HONO emissions substantially. In particular, a strong dependence of HONO emissions on light intensity, relative humidity, NO2 <span class="hlt">concentrations</span> and the applied coating thickness was found. The 20 h long-term studies revealed sustained HONO formation, even when <span class="hlt">concentrations</span> of the intact (<span class="hlt">nitrated</span>) proteins were too low to be detected after the gas exchange measurements. A reaction mechanism for the NO2 conversion based on the Langmuir-Hinshelwood kinetics is proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1138542','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1138542"><span>Determination of intracellular <span class="hlt">nitrate</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>Romero, J M; Lara, C; Guerrero, M G</p> <p>1989-01-01</p> <p>A sensitive procedure has been developed for the determination of intracellular <span class="hlt">nitrate</span>. The method includes: (i) preparation of cell lysates in 2 M-H3PO4 after separation of cells from the outer medium by rapid centrifugation through a layer of silicone oil, and (ii) subsequent <span class="hlt">nitrate</span> analysis by ion-exchange h.p.l.c. with, as mobile phase, a solution containing 50 mM-H3PO4 and 2% (v/v) tetrahydrofuran, adjusted to pH 1.9 with NaOH. The determination of <span class="hlt">nitrate</span> is subjected to interference by chloride and sulphate when present in the samples at high <span class="hlt">concentrations</span>. Nitrite also interferes, but it is easily eliminated by treatment of the samples with sulphamic acid. The method has been successfully applied to the study of <span class="hlt">nitrate</span> transport in the unicellular cyanobacterium Anacystis nidulans. PMID:2497740</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.B11F0489B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.B11F0489B"><span>Patterns of Diel Variation in <span class="hlt">Nitrate</span> <span class="hlt">Concentrations</span> in the Potomac River</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burns, D. A.; Miller, M. P.; Pellerin, B. A.; Capel, P. D.</p> <p>2015-12-01</p> <p>The Potomac River is the second largest source of nitrogen to Chesapeake Bay, where reducing nutrient loads has been a focus of efforts to improve estuarine trophic status. Two years of high frequency sensor measurements of <span class="hlt">nitrate</span> (NO3-) <span class="hlt">concentrations</span> in the Upper Potomac River at the Little Falls gage were analyzed to quantify seasonal variation in the magnitude and timing of the apparent loss of NO3- from the water column that results from diel-driven processes. In addition to broad seasonal and flow-driven variation in NO3- <span class="hlt">concentrations</span>, clear diel patterns were evident in the river, especially during low flow conditions that follow stormflow by several days. Diel variation was about 0.01 mg N/L in winter and 0.02 to 0.03 mg N/L in summer with intermediate values during spring and fall. This variation was equivalent to <1% of the mean daily NO3- <span class="hlt">concentration</span> in winter and about 4% in summer; however, variation >10% occurred during some summer days. Maximum diel <span class="hlt">concentrations</span> occurred during mid- to late-morning in most seasons, with the most repeatable patterns in summer and wider variation in timing during fall and winter. Diel NO3- loss diminished loads by about 0.6% in winter and 1.3% in summer, and diel-driven processes were minor compared to estimates of total in-stream NO3- loss that averaged about one-third of the inferred groundwater NO3- contribution to the river network. The magnitude of diel NO3- variation was more strongly related to metrics based on water temperature and discharge than to metrics based on photosynthetically active radiation. Despite the fairly low diminishment of NO3- loads attributable to diel variation, estimates of diel NO3- uptake were fairly high compared to published values from smaller streams and rivers. The diel NO3- patterns <span class="hlt">observed</span> in the Potomac River are consistent with photosynthesis of periphyton as a principal driver which may be linked to denitrification through the release of labile carbon. The extent to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70162084','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70162084"><span>Long-term changes in <span class="hlt">nitrate</span> conditions over the 20th century in two Midwestern Corn Belt streams</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kelly, Valerie J.; Stets, Edward G.; Crawford, Charles G.</p> <p>2015-01-01</p> <p>Long-term changes in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and flux between the middle of the 20th century and the first decade of the 21st century were estimated for the Des Moines River and the Middle Illinois River, two Midwestern Corn Belt streams, using a novel weighted regression approach that is able to detect subtle changes in solute transport behavior over time. The results show that the largest changes in flow-normalized <span class="hlt">concentration</span> and flux occurred between 1960 and 1980 in both streams, with smaller or negligible changes between 1980 and 2004. Contrasting patterns were <span class="hlt">observed</span> between (1) <span class="hlt">nitrate</span> export linked to non-point sources, explicitly runoff of synthetic fertilizer or other surface sources and (2) <span class="hlt">nitrate</span> export presumably associated with point sources such as urban wastewater or confined livestock feeding facilities, with each of these modes of transport important under different domains of streamflow. Surface runoff was estimated to be consistently most important under high-flow conditions during the spring in both rivers. <span class="hlt">Nitrate</span> export may also have been considerable in the Des Moines River even under some conditions during the winter when flows are generally lower, suggesting the influence of point sources during this time. Similar results were shown for the Middle Illinois River, which is subject to significant influence of wastewater from the Chicago area, where elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were associated with at the lowest flows during the winter and fall. By modeling <span class="hlt">concentration</span> directly, this study highlights the complex relationship between <span class="hlt">concentration</span> and streamflow that has evolved in these two basins over the last 50 years. This approach provides insights about changing conditions that only become <span class="hlt">observable</span> when stationarity in the relationship between <span class="hlt">concentration</span> and streamflow is not assumed.</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('https://www.ncbi.nlm.nih.gov/pubmed/28593809','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28593809"><span><span class="hlt">Nitrate</span> analogs as attractants for soybean cyst nematode.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hosoi, Akito; Katsuyama, Tsutomu; Sasaki, Yasuyuki; Kondo, Tatsuhiko; Yajima, Shunsuke; Ito, Shinsaku</p> <p>2017-08-01</p> <p>Soybean cyst nematode (SCN) Heterodera glycines Ichinohe, a plant parasite, is one of the most serious pests of soybean. In this paper, we report that SCN is attracted to <span class="hlt">nitrate</span> and its analogs. We performed attraction assays to screen for novel attractants for SCN and found that <span class="hlt">nitrates</span> were attractants for SCN and SCN recognized <span class="hlt">nitrate</span> gradients. However, attraction of SCN to <span class="hlt">nitrates</span> was not <span class="hlt">observed</span> on agar containing <span class="hlt">nitrate</span>. To further elucidate the attraction mechanism in SCN, we performed attraction assays using <span class="hlt">nitrate</span> analogs ([Formula: see text], [Formula: see text], [Formula: see text]). SCN was attracted to all <span class="hlt">nitrate</span> analogs; however, attraction of SCN to <span class="hlt">nitrate</span> analogs was not <span class="hlt">observed</span> on agar containing <span class="hlt">nitrate</span>. In contrast, SCN was attracted to azuki root, irrespective of presence or absence of <span class="hlt">nitrate</span> in agar media. Our results suggest that the attraction mechanisms differ between plant-derived attractant and <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950029610&hterms=Phytoplankton&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DPhytoplankton','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950029610&hterms=Phytoplankton&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DPhytoplankton"><span>Annual and interannual variations of phytoplankton pigment <span class="hlt">concentration</span> and upwelling along the Pacific equator</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Halpern, David; Feldman, Gene C.</p> <p>1994-01-01</p> <p>The following variables along the Pacific equator from 145 deg E to 95 deg W were employed: surface layer phytoplankton pigment <span class="hlt">concentrations</span> derived from Nimbus 7 coastal zone color scanner (CZCS) measurements of ocean color radiances; vertical velocities simulated at the 90-m bottom of the euphotic layer from a wind-driven ocean general circulation model; and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> estimated from model-simulated temperature. The upward flux of <span class="hlt">nitrate</span> into the euphotic layer was calculated from the simulated vertical motion and <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. The CZCS-derived phytoplankton pigment <span class="hlt">concentration</span> was uniform from 175 deg to 95 deg W. Longitudinal profiles of upwelling, phytoplankton biomass, and 90-m <span class="hlt">nitrate</span> flux were of different shapes. The small annual cycles of the phytoplankton pigment and <span class="hlt">nitrate</span> flux were in phase: increased phytoplankton biomass was associated with increased upward <span class="hlt">nitrate</span> flux, but the phase was not consistent with the annual cycles of the easterly wind or of the upwelling intensity. Variation of phytoplankton pigment <span class="hlt">concentration</span> was greater during El Nino than during the annual cycle. The substantially reduced phytoplankton pigment <span class="hlt">concentration</span> <span class="hlt">observed</span> during El Nino was associated with smaller upward <span class="hlt">nitrate</span> flux. Phytoplankton biomass during non-El Nino conditions was not related to <span class="hlt">nitrate</span> flux into the euphotic layer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29571890','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29571890"><span>Response of humic acid formation to elevated <span class="hlt">nitrate</span> during chicken manure composting.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shi, Mingzi; Wei, Zimin; Wang, Liqin; Wu, Junqiu; Zhang, Duoying; Wei, Dan; Tang, Yu; Zhao, Yue</p> <p>2018-06-01</p> <p><span class="hlt">Nitrate</span> can stimulate microbes to degrade aromatic compounds, whereas humic acid (HA) as a high molecular weight aromatic compound, its formation may be affected by elevated <span class="hlt">nitrate</span> during composting. Therefore, this study is conducted to determine the effect of elevated <span class="hlt">nitrate</span> on HA formation. Five tests were executed by adding different <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> to chicken manure composting. Results demonstrate that the <span class="hlt">concentration</span> of HA in treatment group is significantly decreased compared with control group (p < 0.05), especially in the highest <span class="hlt">nitrate</span> <span class="hlt">concentration</span> group. RDA indicates that the microbes associated with HA and environmental parameters are influenced by elevated <span class="hlt">nitrate</span>. Furthermore, structural equation model reveals that elevated <span class="hlt">nitrate</span> reduces HA formation by mediating microbes directly, or by affecting ammonia and pH as the indirect drivers to regulate microbial community structure. Copyright © 2018 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26443731','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26443731"><span>Respiration of <span class="hlt">Nitrate</span> and Nitrite.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cole, Jeffrey A; Richardson, David J</p> <p>2008-09-01</p> <p><span class="hlt">Nitrate</span> reduction to ammonia via nitrite occurs widely as an anabolic process through which bacteria, archaea, and plants can assimilate <span class="hlt">nitrate</span> into cellular biomass. Escherichia coli and related enteric bacteria can couple the eight-electron reduction of <span class="hlt">nitrate</span> to ammonium to growth by coupling the <span class="hlt">nitrate</span> and nitrite reductases involved to energy-conserving respiratory electron transport systems. In global terms, the respiratory reduction of <span class="hlt">nitrate</span> to ammonium dominates <span class="hlt">nitrate</span> and nitrite reduction in many electron-rich environments such as anoxic marine sediments and sulfide-rich thermal vents, the human gastrointestinal tract, and the bodies of warm-blooded animals. This review reviews the regulation and enzymology of this process in E. coli and, where relevant detail is available, also in Salmonella and draws comparisons with and implications for the process in other bacteria where it is pertinent to do so. Fatty acids may be present in high levels in many of the natural environments of E. coli and Salmonella in which oxygen is limited but <span class="hlt">nitrate</span> is available to support respiration. In E. coli, <span class="hlt">nitrate</span> reduction in the periplasm involves the products of two seven-gene operons, napFDAGHBC, encoding the periplasmic <span class="hlt">nitrate</span> reductase, and nrfABCDEFG, encoding the periplasmic nitrite reductase. No bacterium has yet been shown to couple a periplasmic <span class="hlt">nitrate</span> reductase solely to the cytoplasmic nitrite reductase NirB. The cytoplasmic pathway for <span class="hlt">nitrate</span> reduction to ammonia is restricted almost exclusively to a few groups of facultative anaerobic bacteria that encounter high <span class="hlt">concentrations</span> of environmental <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20012780-sensitivity-nitrate-nitrite-pond-breeding-amphibians-from-pacific-northwest-usa','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20012780-sensitivity-nitrate-nitrite-pond-breeding-amphibians-from-pacific-northwest-usa"><span>Sensitivity to <span class="hlt">nitrate</span> and nitrite in pond-breeding amphibians from the Pacific Northwest, USA</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>Marco, A.; Quilchano, C.; Blaustein, A.R.</p> <p>1999-12-01</p> <p>In static experiments, the authors studied the effects of <span class="hlt">nitrate</span> and <span class="hlt">nitrate</span> solutions on newly hatched larvae of five species of amphibians, namely Rana pretiosa, Rana aurora, Bufo boreas, Hyla regilla, and Ambystoma gracile. When <span class="hlt">nitrate</span> or nitrite ions were added to the water, some larvae of some species reduced feeding activity, swam less vigorously, showed disequilibrium and paralysis, suffered abnormalities and edemas, and eventually died. The <span class="hlt">observed</span> effects increased with both <span class="hlt">concentration</span> and time, and there were significant differences in sensitivity among species. Ambrystoma gracile displayed the highest acute effect in water with <span class="hlt">nitrate</span> and nitrite. The three ranidmore » species had acute effects in water with nitrite. In chronic exposures, R. pretiosa was the most sensitive species to <span class="hlt">nitrates</span> and nitrites. All species showed 15-d LC50s lower than 2 mg N-NO{sub 2{sup {minus}}}/L. For both N ions, B. boreas was the least sensitive amphibian. All species showed a high morality at the US Environmental Protection Agency-recommended limits of nitrite for warm-water fishes and a significant larval mortality at the recommended limits of nitrite <span class="hlt">concentration</span> for drinking water. The recommended levels of <span class="hlt">nitrate</span> for warm-water fishes were highly toxic for R. pretiosa and A. gracile larvae.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70017118','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70017118"><span>Effects of nutrient management on <span class="hlt">nitrate</span> levels in ground water near Ephrata Pennsylvania</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, David W.</p> <p>1992-01-01</p> <p>Effects of the implementation of nutrient management practices on ground-water quality were studied at a 55-acre farm in Lancaster County, Pennsylvania, from 1985-90. After nutrient management practices were implemented at the site in October 1986, statistically significant decreases (Wilcoxon Mann-Whitney test) in median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in ground-water samples occurred at four of the five wells monitored. The largest decreases in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> occurred in samples collected at the wells that had the largest <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> prior to nutrient management. The decreases in median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in ground-water samples ranged from 8 to 32 percent of the median <span class="hlt">concentrations</span> prior to nutrient management and corresponded to nitrogen application decreases of 39 to 67 percent in contributing areas that were defined upgradient of these wells. Changes in nitrogen applications to the contributing areas of five water wells were correlated (Spearman rank-sum test) with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> of the well water. Changes in ground-water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> lagged behind the changes in applied-nitrogen fertilizers (primarily manure) by approximately 4 to 19 months.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AtmEn..42.2720L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AtmEn..42.2720L"><span><span class="hlt">Observations</span> of fine and coarse particle <span class="hlt">nitrate</span> at several rural locations in the United States</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Taehyoung; Yu, Xiao-Ying; Ayres, Benjamin; Kreidenweis, Sonia M.; Malm, William C.; Collett, Jeffrey L.</p> <p></p> <p><span class="hlt">Nitrate</span> comprises an important part of aerosol mass at many non-urban locations during some times of the year. Little is known, however, about the chemical form and size distribution of particulate <span class="hlt">nitrate</span> in these environments. While submicron ammonium <span class="hlt">nitrate</span> is often assumed to be the dominant species, this assumption is rarely tested. Properties of aerosol <span class="hlt">nitrate</span> were characterized at several IMPROVE monitoring sites during a series of field studies. Study sites included Bondville, Illinois (February 2003), San Gorgonio Wilderness Area, California (April and July 2003), Grand Canyon National Park, Arizona (May 2003), Brigantine National Wildlife Refuge, New Jersey (November 2003), and Great Smoky Mountains National Park, Tennessee (July/August 2004). <span class="hlt">Nitrate</span> was found predominantly in submicron ammonium <span class="hlt">nitrate</span> particles during the Bondville and San Gorgonio (April) campaigns. Coarse mode <span class="hlt">nitrate</span> particles, resulting from reactions of nitric acid or its precursors with sea salt or soil dust, were more important at Grand Canyon and Great Smoky Mountains. Both fine and coarse mode <span class="hlt">nitrate</span> were important during the studies at Brigantine and San Gorgonio (July). These results, which complement earlier findings about the importance of coarse particle <span class="hlt">nitrate</span> at Yosemite and Big Bend National Parks, suggest a need to more closely examine common assumptions regarding the importance of ammonium <span class="hlt">nitrate</span> at non-urban sites, to include pathways for coarse mode <span class="hlt">nitrate</span> formation in regional models, and to consider impacts of coarse particle <span class="hlt">nitrate</span> on visibility. Because coarse particle <span class="hlt">nitrate</span> modes often extend well below 2.5 μm aerodynamic diameter, measurements of PM 2.5 <span class="hlt">nitrate</span> in these environments should not automatically be assumed to contain only ammonium <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014GeCoA.125..528H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014GeCoA.125..528H"><span>Trace <span class="hlt">concentration</span> - Huge impact: <span class="hlt">Nitrate</span> in the calcite/Eu(III) system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hofmann, Sascha; Voïtchovsky, Kislon; Schmidt, Moritz; Stumpf, Thorsten</p> <p>2014-01-01</p> <p>The interactions of trivalent lanthanides and actinides with secondary mineral phases such as calcite is of high importance for the safety assessment of deep geological repositories for high level nuclear waste (HLW). Due to similar ionic radii, calcium-bearing mineral phases are suitable host minerals for Ln(III) and An(III) ions. Especially calcite has been proven to retain these metal ions effectively by both surface complexation and bulk incorporation. Since anionic ligands (e.g., <span class="hlt">nitrate</span>) are omnipresent in the geological environment and due to their coordinating properties, their influence on retentive processes should not be underestimated. <span class="hlt">Nitrate</span> is a common contaminant in most HLW forms as a result of using nitric acid in fuel reprocessing. It is also formed by microbial activity under aerobic conditions. In this study, atomic force microscopy investigations revealed a major influence of <span class="hlt">nitrate</span> upon the surface of calcite crystals. NaNO3 causes serious modifications even in trace amounts (<10-7 M) and forms a soft surface layer of low crystallinity on top of the calcite crystal. Time-resolved laser fluorescence spectroscopy of Eu(III) showed that, within this layer, Eu(III) ions are incorporated, while losing most of their hydration shell. The results show that solid solution modelling for actinides in calcite must take into account the presence of <span class="hlt">nitrate</span> in pore and ground waters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997GeCoA..61.1819O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997GeCoA..61.1819O"><span>Chemical catalysis of <span class="hlt">nitrate</span> reduction by iron (II)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ottley, C. J.; Davison, W.; Edmunds, W. M.</p> <p>1997-05-01</p> <p>Experiments have been conducted to investigate the chemical reduction of <span class="hlt">nitrate</span> under conditions relevant to the often low organic carbon environment of groundwaters. At pH 8 and 20 ± 2°C, in the presence of Cu(II), NO 3- was chemically reduced by Fe(II) to NH 4+ with an average stoichiometric liberation of 8 protons. The rate of the reaction systematically increased with pH in the range pH 7-8.5. The half-life for <span class="hlt">nitrate</span> reduction, t 1/2, was inversely related to the total molar copper <span class="hlt">concentration</span>, [Cu T], by the equation log t 1/2 = -1.35 log [Cu T] -2.616, for all measured values of t 1/2 from 23 min to 15 days. At the Cu(II) <span class="hlt">concentrations</span> used of 7 × 10 -6 -10 -3 M, Cu was present mainly as a solid phase, either adsorbed to the surfaces of precipitated iron oxides or as a saturated solid. It is this solid phase copper rather than CU 2+ in solution which is catalytically active. Neither magnetite, which was formed as a product of the reaction, nor freshly prepared lepidocrocite catalysed the reaction, but goethite did. Although traces of oxygen accelerated the reaction, at higher partial pressures (>0.01 atm) the reduction of <span class="hlt">nitrate</span> was inhibited, probably due to competition between NO 3- and O 2 for Fe(II). Appreciable catalytic effects were also <span class="hlt">observed</span> for solid phase forms of Ag(I), Cd(H), Ni(H), Hg(II), and Pb(II). Mn(II) enhanced the rate slightly, and there was evidence for slow abiotic reduction in the absence of any added metal catalysts. These results suggest that the chemical reduction of <span class="hlt">nitrate</span> at catalytic <span class="hlt">concentrations</span> and temperatures appropriate to groundwater conditions is feasible on a timescale of months to years.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18359993','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18359993"><span>Effect of didecyl dimethyl ammonium chloride on <span class="hlt">nitrate</span> reduction in a mixed methanogenic culture.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tezel, U; Pierson, J A; Pavlostathis, S G</p> <p>2008-01-01</p> <p>The effect of the quaternary ammonium compound, didecyl dimethyl ammonium chloride (DDAC), on <span class="hlt">nitrate</span> reduction was investigated at <span class="hlt">concentrations</span> up to 100 mg/L in a batch assay using a mixed, mesophilic (35 degrees C) methanogenic culture. Glucose was used as the carbon and energy source and the initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was 70 mg N/L. Dissimilatory <span class="hlt">nitrate</span> reduction to ammonia (DNRA) and to dinitrogen (denitrification) were <span class="hlt">observed</span> at DDAC <span class="hlt">concentrations</span> up to 25 mg/L. At and above 50 mg DDAC/L, DNRA was inhibited and denitrification was incomplete resulting in accumulation of nitrous oxide. At DDAC <span class="hlt">concentrations</span> above 10 mg/L, production of nitrous oxide, even transiently, resulted in complete, long-term inhibition of methanogenesis and accumulation of volatile fatty acids. Fermentation was inhibited at and above 75 mg DDAC/L. DDAC suppressed microbial growth and caused cell lysis at a <span class="hlt">concentration</span> 50 mg/L or higher. Most of the added DDAC was adsorbed on the biomass. Over 96% of the added DDAC was recovered from all cultures at the end of the 100-days incubation period, indicating that DDAC did not degrade in the mixed methanogenic culture under the conditions of this study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2012/5049/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2012/5049/"><span>Methods for evaluating temporal groundwater quality data and results of decadal-scale changes in chloride, dissolved solids, and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater in the United States, 1988-2010</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lindsey, Bruce D.; Rupert, Michael G.</p> <p>2012-01-01</p> <p>Decadal-scale changes in groundwater quality were evaluated by the U.S. Geological Survey National Water-Quality Assessment (NAWQA) Program. Samples of groundwater collected from wells during 1988-2000 - a first sampling event representing the decade ending the 20th century - were compared on a pair-wise basis to samples from the same wells collected during 2001-2010 - a second sampling event representing the decade beginning the 21st century. The data set consists of samples from 1,236 wells in 56 well networks, representing major aquifers and urban and agricultural land-use areas, with analytical results for chloride, dissolved solids, and <span class="hlt">nitrate</span>. Statistical analysis was done on a network basis rather than by individual wells. Although spanning slightly more or less than a 10-year period, the two-sample comparison between the first and second sampling events is referred to as an analysis of decadal-scale change based on a step-trend analysis. The 22 principal aquifers represented by these 56 networks account for nearly 80 percent of the estimated withdrawals of groundwater used for drinking-water supply in the Nation. Well networks where decadal-scale changes in <span class="hlt">concentrations</span> were statistically significant were identified using the Wilcoxon-Pratt signed-rank test. For the statistical analysis of chloride, dissolved solids, and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> at the network level, more than half revealed no statistically significant change over the decadal period. However, for networks that had statistically significant changes, increased <span class="hlt">concentrations</span> outnumbered decreased <span class="hlt">concentrations</span> by a large margin. Statistically significant increases of chloride <span class="hlt">concentrations</span> were identified for 43 percent of 56 networks. Dissolved solids <span class="hlt">concentrations</span> increased significantly in 41 percent of the 54 networks with dissolved solids data, and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increased significantly in 23 percent of 56 networks. At least one of the three - chloride, dissolved solids, or</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990111645&hterms=Organic+fertilizers&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DOrganic%2Bfertilizers','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990111645&hterms=Organic+fertilizers&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DOrganic%2Bfertilizers"><span>Techniques for Measurement of <span class="hlt">Nitrate</span> Movement in Soils</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Broadbent, F. E.</p> <p>1971-01-01</p> <p>Contamination of surface and ground waters with <span class="hlt">nitrate</span> usually involves leaching through soil of <span class="hlt">nitrate</span> produced by mineralization of soil organic matter, decomposition of animal wastes or plant residues, or derived from fertilizers. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in the soil solution may be measured by several chemical procedures or by the <span class="hlt">nitrate</span> electrode. since <span class="hlt">nitrate</span> is produced throughout the soil mass it is difficult to identify a source of <span class="hlt">nitrate</span> contamination by conventional means. This problem can be solved by use of N-15-enriched or N-15-depleted materials as tracers. The latter is particularly attractive because of the negligible possibility of the tracer hazardous to health.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H41C1458S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H41C1458S"><span>Source Areas of Water and <span class="hlt">Nitrate</span> in a Peatland Catchment, Minnesota, USA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sebestyen, S. D.</p> <p>2017-12-01</p> <p>In nitrogen polluted forests, stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increase and some unprocessed atmospheric <span class="hlt">nitrate</span> may be transported to streams during stormflow events. This understanding has emerged from forests with upland mineral soils. In contrast, catchments with northern peatlands may have both upland soils and lowlands with deep organic soils, each with unique effects on <span class="hlt">nitrate</span> transport and processing. While annual budgets show <span class="hlt">nitrate</span> yields to be relatively lower from peatland than upland-dominated catchments, little is known about particular runoff events when stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> have been higher (despite long periods with little or no <span class="hlt">nitrate</span> in outlet streams) or the reasons why. I used site knowledge and expansive/extensive monitoring at the Marcell Experimental Forest in Minnesota, along with a targeted 2-year study to determine landscape areas, water sources, and <span class="hlt">nitrate</span> sources that affected stream <span class="hlt">nitrate</span> variation in a peatland catchment. I combined streamflow, upland runoff, snow amount, and frost depth data from long-term monitoring with <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, yield, and isotopic data to show that up to 65% of stream <span class="hlt">nitrate</span> during snowmelt of 2009 and 2010 was unprocessed atmospheric <span class="hlt">nitrate</span>. Up to 46% of subsurface runoff from upland soils during 2009 was unprocessed atmospheric <span class="hlt">nitrate</span>, which shows the uplands to be a stream <span class="hlt">nitrate</span> source during 2009, but not during 2010 when upland runoff <span class="hlt">concentrations</span> were below the detection limit. Differences are attributable to variations in water and <span class="hlt">nitrate</span> sources. Little snow (a <span class="hlt">nitrate</span> source), less upland runoff relative to peatland runoff, and deeper soil frost in the peatland caused a relatively larger input of <span class="hlt">nitrate</span> from the uplands to the stream during 2009 and the peatland to the stream during 2010. Despite the near-absence of stream <span class="hlt">nitrate</span> during much of rest of the year, these findings show an important time when <span class="hlt">nitrate</span> transport affected downstream aquatic ecosystems, reasons</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/2000/4163/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/2000/4163/report.pdf"><span>Probability of detecting atrazine/desethyl-atrazine and elevated <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> plus <span class="hlt">nitrate</span> as nitrogen in ground water in the Idaho part of the western Snake River Plain</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Donato, Mary M.</p> <p>2000-01-01</p> <p>. In areas where this map overlaps the 1998 map of the upper Snake River Basin, the two maps show broadly similar probabilities of detecting atrazine. Logistic regression also was used to develop a preliminary statistical model that predicts the probability of detecting elevated <span class="hlt">nitrate</span> in the western Snake River Plain. A <span class="hlt">nitrate</span> probability map was produced from this model. Results showed that elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were correlated with land use, soil organic content, well depth, and water level. Detailed information on <span class="hlt">nitrate</span> input, specifically fertilizer application, might have improved the effectiveness of this model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1322494-simulations-sulfate-nitrate-ammonium-sna-aerosols-during-extreme-haze-events-over-northern-china','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1322494-simulations-sulfate-nitrate-ammonium-sna-aerosols-during-extreme-haze-events-over-northern-china"><span>Simulations of Sulfate-<span class="hlt">Nitrate</span>-Ammonium (SNA) aerosols during the extreme haze events over Northern China in 2014</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Chen, Dan; Liu, Zhiquan; Fast, Jerome D.</p> <p></p> <p>Extreme haze events have occurred frequently over China in recent years. Although many studies have investigated the formation mechanisms associated with PM2.5 for heavily polluted regions in China based on <span class="hlt">observational</span> data, adequately predicting peak PM2.5 <span class="hlt">concentrations</span> is still challenging for regional air quality models. In this study, we evaluate the performance of one configuration of the Weather Research and Forecasting model coupled with chemistry (WRF-Chem) and use the model to investigate the sensitivity of heterogeneous reactions on simulated peak sulfate, <span class="hlt">nitrate</span>, and ammonium <span class="hlt">concentrations</span> in the vicinity of Beijing during four extreme haze episodes in October 2014 over themore » North China Plain. The highest <span class="hlt">observed</span> PM2.5 <span class="hlt">concentration</span> of 469 μg m-3 occurred in Beijing. Comparisons with <span class="hlt">observations</span> show that the model reproduced the temporal variability in PM2.5 with the highest PM2.5 values on polluted days (defined as days in which <span class="hlt">observed</span> PM2.5 is greater than 75 μg m-3), but predictions of sulfate, <span class="hlt">nitrate</span>, and ammonium were too low on days with the highest <span class="hlt">observed</span> <span class="hlt">concentrations</span>. <span class="hlt">Observational</span> data indicate that the sulfur/nitric oxidation rates are strongly correlated with relative humidity during periods of peak PM2.5; however, the model failed to reproduce the highest PM2.5 <span class="hlt">concentrations</span> due to missing heterogeneous reactions. As the parameterizations of those reactions is not well established yet, estimates of SO2-to-H2SO4 and NO2/NO3-to-HNO3 reaction rates that depend on relative humidity were applied which improved the simulation of sulfate, <span class="hlt">nitrate</span>, and ammonium enhancement on polluted days in terms of both <span class="hlt">concentrations</span> and partitioning among those species. Sensitivity simulations showed that the extremely high heterogeneous reaction rates and also higher emission rates than those reported in the emission inventory« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23505772','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23505772"><span><span class="hlt">Nitrate</span> and ammonia contaminations in drinking water and the affecting factors in Hailun, northeast China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, Xinfeng; Chen, Liding; Zhang, Haiping</p> <p>2013-03-01</p> <p>Drinking water samples (N = 228) from domestic tube wells (DTWs) and seven samples from public water supply wells (PWSWs) were collected and tested in Hailun, northeast China. The percentage of samples with <span class="hlt">nitrate</span> and ammonia <span class="hlt">concentrations</span> above the maximum acceptable <span class="hlt">concentration</span> of <span class="hlt">nitrate</span>, 10 mg N/L, and the maximum ensure <span class="hlt">concentration</span> of ammonia, 1.5 mg/L, for the DTWs were significantly higher than for the PWSWs. Of the DTWs, an important <span class="hlt">observation</span> was that the occurrence of groundwater <span class="hlt">nitrate</span> contamination was directly related to well tube material with different joint pathways. <span class="hlt">Nitrate</span> in seamless-tube wells was lower statistically significantly than those in multiple-section-tube wells (p < .001). Furthermore, well depth and hydrogeological setting might have some impacts on nitrogen contamination and the major sources of inorganic nitrogen contamination may be nitrogenous chemical fertilizer. Therefore, PWSWs built for all families are the best way to ensure the drinking water safety in villages. For DTWs it is necessary to use seamless tubes and to dig deep enough according to the depth of groundwater level. Improving the efficiency of chemical fertilizer use would also reduce the risk of groundwater contamination.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AIPC.1251...41M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AIPC.1251...41M"><span>Photodegradation of Paracetamol in <span class="hlt">Nitrate</span> Solution</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Meng, Cui; Qu, Ruijuan; Liang, Jinyan; Yang, Xi</p> <p>2010-11-01</p> <p>The photodegradation of paracetamol in <span class="hlt">nitrate</span> solution under simulated solar irradiation has been investigated. The degradation rates were compared by varying environmental parameters including <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> ion, humic substance and pH values. The quantifications of paracetamol were conducted by HPLC method. The results demonstrate that the photodegradation of paracetamol followed first-order kinetics. The photoproducts and intermediates of paracetamol in the presence of <span class="hlt">nitrate</span> ions were identified by extensive GC-MS method. The photodegradation pathways involving. OH radicals as reactive species were proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21506740-photodegradation-paracetamol-nitrate-solution','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21506740-photodegradation-paracetamol-nitrate-solution"><span>Photodegradation of Paracetamol in <span class="hlt">Nitrate</span> Solution</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Meng Cui; Qu Ruijuan; Liang Jinyan</p> <p>2010-11-24</p> <p>The photodegradation of paracetamol in <span class="hlt">nitrate</span> solution under simulated solar irradiation has been investigated. The degradation rates were compared by varying environmental parameters including <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> ion, humic substance and pH values. The quantifications of paracetamol were conducted by HPLC method. The results demonstrate that the photodegradation of paracetamol followed first-order kinetics. The photoproducts and intermediates of paracetamol in the presence of <span class="hlt">nitrate</span> ions were identified by extensive GC-MS method. The photodegradation pathways involving. OH radicals as reactive species were proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.H34C..02B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.H34C..02B"><span><span class="hlt">NITRATE</span> POLLUTION IN SHALLOW GROUNDWATER OF A HARD ROCK REGION IN SOUTH CENTRAL INDIA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brindha, K.; Rajesh, R.; Murugan, R.; Elango, L.</p> <p>2009-12-01</p> <p>Groundwater forms a major source of drinking water in most parts of the world. Due to the lack of piped drinking water supply, the population in rural areas depend on the groundwater resources for domestic purposes. Hence, the quality of groundwater in such regions needs to be monitored regularly. Presence of high <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> in groundwater used for drinking is a major problem in many countries as it causes health related problems. Most often infants are affected by the intake of high <span class="hlt">nitrate</span> in drinking water and food. The present study was carried out with the objective of assessing the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater and determining the causes for <span class="hlt">nitrate</span> in groundwater in parts of Nalgonda district in India which is located at a distance of about 135 km towards ESE direction from Hyderabad. <span class="hlt">Nitrate</span> <span class="hlt">concentration</span> in groundwater of this area was analysed by collecting groundwater samples from forty six representative wells. Samples were collected once in two months from March 2008 to March 2009. A total of 244 groundwater samples were collected during the study. Soil samples were collected from fifteen locations during May 2009 and the denitrifying bacteria were isolated from the soil using spread plate method. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater samples were analysed in the laboratory using Metrohm 861 advanced compact ion chromatograph using appropriate standards. The highest <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> recorded during the sampling period was 879.65mg/l and the lowest <span class="hlt">concentration</span> was below detection limit. The maximum permissible limit of <span class="hlt">nitrate</span> for drinking water as per Bureau of Indian Standards is 45mg/l. About 13% of the groundwater samples collected from this study area possessed <span class="hlt">nitrate</span> <span class="hlt">concentration</span> beyond this limit. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was high in the southeastern part of the study area. This implies that the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater tends to increase along the flow direction. Application of fertilizers is one</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003HyPr...17.1197M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003HyPr...17.1197M"><span>Geologic controls on the chemical behaviour of <span class="hlt">nitrate</span> in riverside alluvial aquifers, Korea</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Min, Joong-Hyuk; Yun, Seong-Taek; Kim, Kangjoo; Kim, Hyoung-Soo; Kim, Dong-Ju</p> <p>2003-04-01</p> <p>To investigate the origin and behaviour of <span class="hlt">nitrate</span> in alluvial aquifers adjacent to Nakdong River, Korea, we chose two representative sites (Wolha and Yongdang) having similar land-use characteristics but different geology. A total of 96 shallow groundwater samples were collected from irrigation and domestic wells tapping alluvial aquifers.About 63% of the samples analysed had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> that exceeded the Korean drinking water limit (44·3 mg l-1 NO3-), and about 35% of the samples had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> that exceeded the Korean groundwater quality standard for agricultural use (88·6 mg l-1 NO3-). Based on nitrogen isotope analysis, two major <span class="hlt">nitrate</span> sources were identified: synthetic fertilizer (about 4 15N) applied to farmland, and animal manure and sewage (15-20 15N) originating from upstream residential areas. Shallow groundwater in the farmland generally had higher <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> than those in residential areas, due to the influence of synthetic fertilizer. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> at both study sites were highest near the water table and then progressively decreased with depth. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> are also closely related to the geologic characteristics of the aquifer. In Yongdang, denitrification is important in regulating <span class="hlt">nitrate</span> chemistry because of the availability of organic carbon from a silt layer (about 20 m thick) below a thin, sandy surface aquifer. In Wolha, however, conservative mixing between farmland-recharged water and water coming from a village is suggested as the dominant process. Mixing ratios estimated based on the <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and the 15N values indicate that water originating from the village affects the <span class="hlt">nitrate</span> chemistry of the shallow groundwater underneath the farmland to a large extent. </TR</TABLE</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://pubs.er.usgs.gov/publication/70018077','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018077"><span>Denitrification and mixing in a stream-aquifer system: Effects on <span class="hlt">nitrate</span> loading to surface water</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>McMahon, P.B.; Böhlke, J.K.</p> <p>1996-01-01</p> <p>Ground water in terrace deposits of the South Platte River alluvial aquifer near Greeley, Colorado, USA, had a median <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of 1857 ??mol l-1. Median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in ground water from adjacent floodplain deposits (468 ??mol l-1) and riverbed sediments (461 ??mol l-1), both of which are downgradient from the terrace deposits, were lower than the median <span class="hlt">concentration</span> in the terrace deposits. The <span class="hlt">concentrations</span> and ??15N values of <span class="hlt">nitrate</span> and N2 in ground water indicated that denitrifying activity in the floodplain deposits and riverbed sediments accounted for 15- 30% of the difference in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. <span class="hlt">Concentrations</span> of Cl- and SiO2 indicated that mixing between river water and ground water in the floodplain deposits and riverbed sediments accounted for the remainder of the difference in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. River flux measurements indicated that ground-water discharge in a 7.5 km segment of river had a <span class="hlt">nitrate</span> load of 1718 kg N day-1 and accounted for about 18% of the total <span class="hlt">nitrate</span> load in the river at the downstream end of that segment. This <span class="hlt">nitrate</span> load was 70% less than the load predicted on the basis of the median <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the terrace deposits and assuming no denitrification or mixing in the aquifer. Water exchange between the river and aquifer caused ground water that originally discharged to the river to reenter denitrifying sediments in the riverbed and floodplain, thereby further decreasing the <span class="hlt">nitrate</span> load in this stream-aquifer system. Results from this study indicated that denitrification and mixing within alluvial aquifer sediments may substantially decrease the <span class="hlt">nitrate</span> load added to rivers by discharging ground water.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27371928','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27371928"><span>Does <span class="hlt">nitrate</span> co-pollution affect biological responses of an aquatic plant to two common herbicides?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nuttens, A; Chatellier, S; Devin, S; Guignard, C; Lenouvel, A; Gross, E M</p> <p>2016-08-01</p> <p>Aquatic systems in agricultural landscapes are subjected to multiple stressors, among them pesticide and <span class="hlt">nitrate</span> run-off, but effects of both together have rarely been studied. We investigated possible stress-specific and interaction effects using the new OECD test organism, Myriophyllum spicatum, a widespread aquatic plant. In a fully factorial design, we used two widely applied herbicides, isoproturon and mesosulfuron-methyl, in <span class="hlt">concentration</span>-response curves at two <span class="hlt">nitrate</span> levels (219.63 and 878.52mg N-NO3). We applied different endpoints reflecting plant performance such as growth, pigment content, content in phenolic compounds, and plant stoichiometry. Relative growth rates based on length (RGR-L) were affected strongly by both herbicides, while effects on relative growth rate based on dry weight (RGR-DW) were apparent for isoproturon but hardly visible for mesosulfuron-methyl due to an increase in dry matter content. The higher <span class="hlt">nitrate</span> level further reduced growth rates, specifically with mesosulfuron-methyl. Effects were visible between 50 and 500μgL(-1) for isoproturon and 0.5-5μgL(-1) for mesosulfuron-methyl, with some differences between endpoints. The two herbicides had opposite effects on chlorophyll, carotenoid and nitrogen contents in plants, with values increasing with increasing <span class="hlt">concentrations</span> of isoproturon and decreasing for mesosulfuron-methyl. Herbicides and <span class="hlt">nitrate</span> level exhibited distinct effects on the content in phenolic compounds, with higher <span class="hlt">nitrate</span> levels reducing total phenolic compounds in controls and with isoproturon, but not with mesosulfuron-methyl. Increasing <span class="hlt">concentrations</span> of mesosulfuron-methyl lead to a decline of total phenolic compounds, while isoproturon had little effect. Contents of carbon, nitrogen and phosphorus changed depending on the stressor combination. We <span class="hlt">observed</span> higher phosphorus levels in plants exposed to certain <span class="hlt">concentrations</span> of herbicides, potentially indicating a metabolic response. The C:N molar ratio</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21671306','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21671306"><span>Toxic effects of lead and nickel <span class="hlt">nitrate</span> on rat liver chromatin components.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rabbani-Chadegani Iii, Azra; Fani, Nesa; Abdossamadi, Sayeh; Shahmir, Nosrat</p> <p>2011-01-01</p> <p>The biological activity of heavy metals is related to their physicochemical interaction with biological receptors. In the present study, the effect of low <span class="hlt">concentrations</span> of nickel <span class="hlt">nitrate</span> and lead <span class="hlt">nitrate</span> (<0.3 mM) on rat liver soluble chromatin and histone proteins was examined. The results showed that addition of various <span class="hlt">concentrations</span> of metals to chromatin solution preceded the chromatin into aggregation and precipitation in a dose-dependant manner; however, the extent of absorbance changes at 260 and 400 nm was different between two metals. Gel electrophoresis of histone proteins and DNA of the supernatants obtained from the metal-treated chromatin and the controls revealed higher affinity of lead <span class="hlt">nitrate</span> to chromatin compared to nickel <span class="hlt">nitrate</span>. Also, the binding affinity of lead <span class="hlt">nitrate</span> to histone proteins free in solution was higher than nickel. On the basis of the results, it is concluded that lead reacts with chromatin components even at very low <span class="hlt">concentrations</span> and induce chromatin aggregation through histone-DNA cross-links. Whereas, nickel <span class="hlt">nitrate</span> is less effective on chromatin at low <span class="hlt">concentrations</span>, suggesting higher toxicity of lead <span class="hlt">nitrate</span> on chromatin compared to nickel. Copyright © 2010 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18765775','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18765775"><span>Temporal trends in <span class="hlt">nitrate</span> and selected pesticides in Mid-Atlantic ground water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Debrewer, Linda M; Ator, Scott W; Denver, Judith M</p> <p>2008-01-01</p> <p>Evaluating long-term temporal trends in regional ground-water quality is complicated by variable hydrogeologic conditions and typically slow flow, and such trends have rarely been directly measured. Ground-water samples were collected over near-decadal and annual intervals from unconfined aquifers in agricultural areas of the Mid-Atlantic region, including fractured carbonate rocks in the Great Valley, Potomac River Basin, and unconsolidated sediments on the Delmarva Peninsula. <span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> and selected pesticides and degradates were compared among sampling events and to apparent recharge dates. <span class="hlt">Observed</span> temporal trends are related to changes in land use and chemical applications, and to hydrogeology and climate. Insignificant differences in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the Great Valley between 1993 and 2002 are consistent with relatively steady fertilizer application during respective recharge periods and are likely related to drought conditions in the later sampling period. Detecting trends in Great Valley ground water is complicated by long open boreholes characteristic of wells sampled in this setting which facilitate significant ground-water mixing. Decreasing atrazine and prometon <span class="hlt">concentrations</span>, however, reflect reported changes in usage. On the Delmarva Peninsula between 1988 and 2001, median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increased 2 mg per liter in aerobic ground water, reflecting increasing fertilizer applications. Correlations between selected pesticide compounds and apparent recharge date are similarly related to changing land use and chemical application. <span class="hlt">Observed</span> trends in the two settings demonstrate the importance of considering hydrogeology and recharge date along with changing land and chemical uses when interpreting trends in regional ground-water quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28500889','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28500889"><span>Ammonium stimulates <span class="hlt">nitrate</span> reduction during simultaneous nitrification and denitrification process by Arthrobacter arilaitensis Y-10.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>He, Tengxia; Xie, Deti; Li, Zhenlun; Ni, Jiupai; Sun, Quan</p> <p>2017-09-01</p> <p>The ability of Arthrobacter arilaitensis Y-10 for nitrogen removal from simulated wastewater was studied. Results showed that ammonium was the best inorganic nitrogen for strain Y-10's cell growth, which could also promote <span class="hlt">nitrate</span> reduction. Approximately 100.0% of ammonium was removed in the nitrogen removal experiments. The <span class="hlt">nitrate</span> removal efficiency was 73.3% with <span class="hlt">nitrate</span> as sole nitrogen source, and then the <span class="hlt">nitrate</span> efficiency was increased to 85.3% and 100.0% with ammonium and <span class="hlt">nitrate</span> (both about 5 or 100mg/L) as the mixed nitrogen sources. Nitrite accumulation was <span class="hlt">observed</span> in presence of ammonium and <span class="hlt">nitrate</span>. When the <span class="hlt">concentration</span> of sole nitrite nitrogen was 10.31mg/L, the nitrite removal efficiency was 100.0%. Neither ammonium nor <span class="hlt">nitrate</span> was accumulated during the whole experimental process. All experimental results indicated that A. arilaitensis Y-10 could remove ammonium, <span class="hlt">nitrate</span> and nitrite at 15°C from wastewater, and could also perform simultaneous nitrification and denitrification under aerobic condition. Copyright © 2017. Published by Elsevier Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2013/5136/pdf/sir2013-5136.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2013/5136/pdf/sir2013-5136.pdf"><span>The distribution and modeling of <span class="hlt">nitrate</span> transport in the Carson Valley alluvial aquifer, Douglas County, Nevada</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Naranjo, Ramon C.; Welborn, Toby L.; Rosen, Michael R.</p> <p>2013-01-01</p> <p>The distribution of <span class="hlt">nitrate</span> as nitrogen (referred to herein as <span class="hlt">nitrate</span>-N) <span class="hlt">concentrations</span> in groundwater was determined by collecting more than 200 samples from 8 land-use categories: single family residential, multifamily residential, rural (including land use for agriculture), vacant land, commercial, industrial, utilities, and unclassified. <span class="hlt">Nitrate</span>-N <span class="hlt">concentrations</span> ranged from below detection (less than 0.05 milligrams per liter) to 18 milligrams per liter. The results of <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> that were sampled from three wells equalled or exceeded the maximum contaminant level of 10 milligrams per liter set by the U.S. Environmental Protection Agency. <span class="hlt">Nitrate</span>-N <span class="hlt">concentrations</span> in sampled wells showed a positive correlation between elevated <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> and the percentage of single-family land use and septic-system density. Wells sampled in other land-use categories did not have any correlation to <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span>. In areas with greater than 50-percent single-family land use, <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> were two times greater than in areas with less than 50 percent single-family land use. <span class="hlt">Nitrate</span>-N <span class="hlt">concentrations</span> in groundwater near septic systems that had been used more than 20 years were more than two times greater than in areas where septic systems had been used less than 20 years. Lower <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> in the areas where septic systems were less than 20 years old probably result from temporary storage of nitrogen leaching from septic systems into the unsaturated zone. In areas where septic systems are abundant, <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> were predicted to 2059 by using numerical models within the Ruhenstroth and Johnson Lane subdivisions in the Carson Valley. Model results indicated that <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> will continue to increase and could exceed the maximum contaminant level over extended areas inside and outside the subdivisions. Two modeling scenarios were used to simulate future transport as a result of removal of septic</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUSM.H23B..03S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUSM.H23B..03S"><span>Evaluating <span class="hlt">Nitrate</span> Contributions From Different Land Use Types Across a Regional Watershed Using Flow and Transport Models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Spansky, M. C.; Hyndman, D. W.; Long, D. T.; Pijanowski, B. C.</p> <p>2004-05-01</p> <p>Regional inputs of non-point source pollutants to groundwater, such as agriculturally-derived <span class="hlt">nitrate</span>, have typically proven difficult to model due to sparse <span class="hlt">concentration</span> data and complex system dynamics. We present an approach to evaluate the relative contribution of various land use types to groundwater <span class="hlt">nitrate</span> across a regional Michigan watershed using groundwater flow and transport models. The models were parameterized based on land use data, and calibrated to a 20 year database of <span class="hlt">nitrate</span> measured in drinking water wells. The database spans 1983-2003 and contains approximately 27,000 <span class="hlt">nitrate</span> records for the five major counties encompassed by the watershed. The Grand Traverse Bay Watershed (GTBW), located in the northwest Lower Peninsula of Michigan, was chosen for this research. Groundwater flow and <span class="hlt">nitrate</span> transport models were developed for the GTBW using MODFLOW2000 and RT3D, respectively. In a preliminary transport model, agricultural land uses were defined as the sole source of groundwater <span class="hlt">nitrate</span>. <span class="hlt">Nitrate</span> inputs were then refined to reflect variations in nitrogen loading rates for different agriculture types, including orchards, row crops, and pastureland. The calibration dataset was created by assigning spatial coordinates to each water well sample using address matching from a geographic information system (GIS). Preliminary results show that there is a significant link between agricultural sources and measured groundwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. In cases where <span class="hlt">observed</span> <span class="hlt">concentrations</span> remain significantly higher than simulated values, other sources of <span class="hlt">nitrate</span> (e.g. septic tanks or abandoned agricultural fields) will be evaluated. This research will eventually incorporate temporal variations in fertilizer application rates and changing land use patterns to better represent fluid and solute fluxes at a regional scale.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/20216','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/20216"><span>Changes in stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> due to land management practices, ecological succession, and climate: Developing a system approach to integrated catchment response</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>F. Worrall; Wayne T. Swank; T. P. Burt</p> <p>2003-01-01</p> <p>This study uses time series analysis to examine long-term stream water <span class="hlt">nitrate</span> <span class="hlt">concentration</span> records from a pair of forested catchments at the Coweeta Hydrologic Laboratory, North Carolina, USA. Monthly average <span class="hlt">concentrations</span> were available from 1970 through 1997 for two forested catchments, one of which was clear-felled in 1977 and the other maintained as a control....</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4108036','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4108036"><span><span class="hlt">Nitrate</span> dynamics in natural plants: insights based on the <span class="hlt">concentration</span> and natural isotope abundances of tissue <span class="hlt">nitrate</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>Liu, Xue-Yan; Koba, Keisuke; Makabe, Akiko; Liu, Cong-Qiang</p> <p>2014-01-01</p> <p>The dynamics of <span class="hlt">nitrate</span> (NO−3), a major nitrogen (N) source for natural plants, has been studied mostly through experimental N addition, enzymatic assay, isotope labeling, and genetic expression. However, artificial N supply may not reasonably reflect the N strategies in natural plants because NO−3 uptake and reduction may vary with external N availability. Due to abrupt application and short operation time, field N addition, and isotopic labeling hinder the elucidation of in situ NO−3-use mechanisms. The <span class="hlt">concentration</span> and natural isotopes of tissue NO−3 can offer insights into the plant NO−3 sources and dynamics in a natural context. Furthermore, they facilitate the exploration of plant NO−3 utilization and its interaction with N pollution and ecosystem N cycles without disturbing the N pools. The present study was conducted to review the application of the denitrifier method for <span class="hlt">concentration</span> and isotope analyses of NO−3 in plants. Moreover, this study highlights the utility and advantages of these parameters in interpreting NO−3 sources and dynamics in natural plants. We summarize the major sources and reduction processes of NO−3 in plants, and discuss the implications of NO−3 <span class="hlt">concentration</span> in plant tissues based on existing data. Particular emphasis was laid on the regulation of soil NO−3 and plant ecophysiological functions in interspecific and intra-plant NO−3 variations. We introduce N and O isotope systematics of NO−3 in plants and discuss the principles and feasibilities of using isotopic enrichment and fractionation factors; the correlation between <span class="hlt">concentration</span> and isotopes (N and O isotopes: δ18O and Δ17O); and isotope mass-balance calculations to constrain sources and reduction of NO−3 in possible scenarios for natural plants are deliberated. Finally, we offer a preliminary framework of intraplant δ18O-NO−3 variation, and summarize the uncertainties in using tissue NO−3 parameters to interpret plant NO−3 utilization</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25101106','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25101106"><span><span class="hlt">Nitrate</span> dynamics in natural plants: insights based on the <span class="hlt">concentration</span> and natural isotope abundances of tissue <span class="hlt">nitrate</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, Xue-Yan; Koba, Keisuke; Makabe, Akiko; Liu, Cong-Qiang</p> <p>2014-01-01</p> <p>The dynamics of <span class="hlt">nitrate</span> (NO(-) 3), a major nitrogen (N) source for natural plants, has been studied mostly through experimental N addition, enzymatic assay, isotope labeling, and genetic expression. However, artificial N supply may not reasonably reflect the N strategies in natural plants because NO(-) 3 uptake and reduction may vary with external N availability. Due to abrupt application and short operation time, field N addition, and isotopic labeling hinder the elucidation of in situ NO(-) 3-use mechanisms. The <span class="hlt">concentration</span> and natural isotopes of tissue NO(-) 3 can offer insights into the plant NO(-) 3 sources and dynamics in a natural context. Furthermore, they facilitate the exploration of plant NO(-) 3 utilization and its interaction with N pollution and ecosystem N cycles without disturbing the N pools. The present study was conducted to review the application of the denitrifier method for <span class="hlt">concentration</span> and isotope analyses of NO(-) 3 in plants. Moreover, this study highlights the utility and advantages of these parameters in interpreting NO(-) 3 sources and dynamics in natural plants. We summarize the major sources and reduction processes of NO(-) 3 in plants, and discuss the implications of NO(-) 3 <span class="hlt">concentration</span> in plant tissues based on existing data. Particular emphasis was laid on the regulation of soil NO(-) 3 and plant ecophysiological functions in interspecific and intra-plant NO(-) 3 variations. We introduce N and O isotope systematics of NO(-) 3 in plants and discuss the principles and feasibilities of using isotopic enrichment and fractionation factors; the correlation between <span class="hlt">concentration</span> and isotopes (N and O isotopes: δ(18)O and Δ(17)O); and isotope mass-balance calculations to constrain sources and reduction of NO(-) 3 in possible scenarios for natural plants are deliberated. Finally, we offer a preliminary framework of intraplant δ(18)O-NO(-) 3 variation, and summarize the uncertainties in using tissue NO(-) 3 parameters to interpret</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.H13B0925L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.H13B0925L"><span>Transformation of <span class="hlt">Nitrate</span> and Toluene in Groundwater by Sulfur Modified Iron(SMI-III)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, W.; Park, S.; Lim, J.; Hong, U.; Kwon, S.; Kim, Y.</p> <p>2009-12-01</p> <p>In Korea, <span class="hlt">nitrate</span> and benzene, toluene, ethylbenzene, and xylene isomers (BTEX) are frequently detected together as ground water contaminants. Therefore, a system simultaneously treating both <span class="hlt">nitrate</span> (inorganic compound) and BTEX (organic compounds) is required to utilize groundwater as a water resource. In this study, we investigated the efficiency of Sulfur Modified Iron (SMI-III) in treating both <span class="hlt">nitrate</span> and BTEX contaminated groundwater. Based on XRD (X-Ray Diffraction) analysis, the SMI-III is mainly composed of Fe3O4, S, and Fe. A series of column tests were conducted at three different empty bed contact times (EBCTs) for each compound. During the experiments, removal efficiency for both <span class="hlt">nitrate</span> and toluene were linearly correlated with EBCT, suggesting that SMI-III have an ability to transform both <span class="hlt">nitrate</span> and toluene. The <span class="hlt">concentration</span> of SO42- and oxidation/reduction potential (ORP) were also measured. After exposed to <span class="hlt">nitrate</span> contaminated groundwater, the composition of SMI-III was changed to Fe2O3, Fe3O4, Fe, and Fe0.95S1.05. The trends of effluent sulfate <span class="hlt">concentrations</span> were inversely correlated with effluent <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, while the trends of ORP values, having the minimum values of -480 mV, were highly correlated with effluent <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. XRD analysis before and after exposed to <span class="hlt">nitrate</span> contaminated groundwater, sulfate production, and nitrite detection as a reductive transformation by-product of <span class="hlt">nitrate</span> suggest that <span class="hlt">nitrate</span> is reductively transformed by SMI-III. Interestingly, in the toluene experiments, the trends of ORP values were inversely correlated with effluent toluene <span class="hlt">concentrations</span>, suggesting that probably degrade through oxidation reaction. Consequently, <span class="hlt">nitrate</span> and toluene probably degrade through reduction and oxidation reaction, respectively and SMI-III could serve as both electron donor and acceptor.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EurSS..50.1450K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EurSS..50.1450K"><span>Viscoelastic Properties of Soil with Different Ammonium <span class="hlt">Nitrate</span> Addition</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kawecka-Radomska, M.; Tomczyńska-Mleko, M.; Muszyńskic, S.; Wesołowska-Trojanowska, M.; Mleko, S.</p> <p>2017-12-01</p> <p>Four different soils samples were taken from not cultivated recreational places. Particle-size distribution and pH (in water and in 1 M KCl) of the soil samples were measured. Soil samples were saturated with deionized water and solution of ammonium <span class="hlt">nitrate</span> with the <span class="hlt">concentration</span> of 5, 50 or 500 mM for 3 days. The samples were analyzed using dynamic oscillatory rheometer by frequency and strain sweeps. Soil samples were similar to physical gels, as they presented rheological properties between those of a <span class="hlt">concentrated</span> biopolymer and a true gel. 50 mM <span class="hlt">concentration</span> of the salt was enough to make changes in the elasticity of the soils. Small <span class="hlt">concentration</span> of the fertilizer caused weakening of the soil samples structure. Higher <span class="hlt">concentration</span> of ammonium <span class="hlt">nitrate</span> caused the increase in the moduli crossover strain value. For the loam sample taken from a playground, with the highest content of the particles <0.002 mm (clay aluminosilicates), the lowest value of strain was <span class="hlt">observed</span> at the moduli intersection. Lower strain value was necessary for the sliding shear effect of soil A sample effecting transgression to the "flowing" state. Strain sweep moduli crossover point can be used as a determinant of the rheological properties of soil.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007JGRC..112.1012D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007JGRC..112.1012D"><span>Alkyl <span class="hlt">nitrate</span> (C1-C3) depth profiles in the tropical Pacific Ocean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dahl, E. E.; Yvon-Lewis, S. A.; Saltzman, E. S.</p> <p>2007-01-01</p> <p>This paper reports the first depth profile measurements of methyl, ethyl, isopropyl and n-propyl <span class="hlt">nitrates</span> in the tropical Pacific Ocean. Depth profile measurements were made at 22 stations during the Project Halocarbon Air Sea Exchange cruise, in warm pool, equatorial, subequatorial, and gyre waters. The highest <span class="hlt">concentrations</span>, up to several hundred pM of methyl <span class="hlt">nitrate</span>, were <span class="hlt">observed</span> in the central Pacific within 8 degrees of the equator. In general, alkyl <span class="hlt">nitrate</span> levels were highest in the surface mixed layer, and decreased with depth below the mixed layer. The spatial distribution of the alkyl <span class="hlt">nitrates</span> suggests that there is a strong source associated with biologically productive ocean regions, that is characterized by high ratios of methyl:ethyl <span class="hlt">nitrate</span>. However, the data do not allow discrimination between direct biological emissions and photochemistry as production mechanisms. Alkyl <span class="hlt">nitrates</span> were consistently detectable at several hundred meters depth. On the basis of the estimated chemical loss rate of these compounds, we conclude that deep water alkyl <span class="hlt">nitrates</span> must be produced in situ. Possible sources include free radical processes initiated by radioactive decay or cosmic rays, enzymatically mediated reactions involving bacteria, or unidentified chemical mechanisms involving dissolved organic matter.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.B41H..05G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.B41H..05G"><span>Isotopic signals from precipitation and denitrification in <span class="hlt">nitrate</span> in a northern hardwood forest</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Goodale, C. L.; Wexller, S.</p> <p>2012-12-01</p> <p>Denitrification can represent an important term in the nitrogen budget of small catchments; however, this process varies greatly over space and time and is notoriously difficult to quantify. Measurements of the natural abundance of stable isotopes of nitrogen and oxygen in dissolved <span class="hlt">nitrate</span> in stream- and river water can sometimes provide evidence of denitrification, particularly in large river basins or agriculturally impacted catchments. To date, however, this approach has provided little to no evidence of denitrification in catchments in temperate forests. Here, we examined d15N and d18O of <span class="hlt">nitrate</span> in water samples collected during summer 2011 not only from streams and precipitation, but also from groundwater from the hydrologic reference watershed (W3) drained by Paradise Brook, at the Hubbard Brook Experimental Forest, in the White Mountains, New Hampshire. Despite low <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> (< 0.5 to 8.8 uM <span class="hlt">nitrate</span>) dual-isotopic signals of <span class="hlt">nitrate</span> sources and nitrogen cycle processes were clearly distinguishable, including sources from atmospheric deposition, and from nitrification of atmospheric ammonium and from or soil organic nitrogen, as well as <span class="hlt">nitrate</span> affected by soil denitrification. An atmospheric signal from <span class="hlt">nitrate</span> in precipitation (enriched with 18O) was <span class="hlt">observed</span> immediately following a precipitation event in mid-July contributing roughly 22% of stream <span class="hlt">nitrate</span> export on this date. Stream samples the day following this and other storms showed this export of event <span class="hlt">nitrate</span> to be short-lived. Hillslope piezometers showed low <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and high d15N- and d18O-<span class="hlt">nitrate</span> values (averaging 12 and 18 per mil, repectively) indicating denitrification, which preferentially removes isotopically light N and O in N gases and leaves isotopically heavy <span class="hlt">nitrate</span> behind. These samples showed a positive relationship between nitrogen and oxygen isotopic composition with a regression line slope of 0.76 (R2 = 0.68), and an isotope enrichment factor -12.7 per</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014WRR....50.8694P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014WRR....50.8694P"><span>Relationships between stream <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and spatially distributed snowmelt in high-elevation catchments of the western U.S.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Perrot, Danielle; Molotch, Noah P.; Williams, Mark W.; Jepsen, Steven M.; Sickman, James O.</p> <p>2014-11-01</p> <p>This study compares stream <span class="hlt">nitrate</span> (NO3-) <span class="hlt">concentrations</span> to spatially distributed snowmelt in two alpine catchments, the Green Lakes Valley, Colorado (GLV4) and Tokopah Basin, California (TOK). A snow water equivalent reconstruction model and Landsat 5 and 7 snow cover data were used to estimate daily snowmelt at 30 m spatial resolution in order to derive indices of new snowmelt areas (NSAs). Estimates of NSA were then used to explain the NO3- flushing behavior for each basin over a 12 year period (1996-2007). To identify the optimal method for defining NSAs and elucidate mechanisms underlying catchment NO3- flushing, we conducted a series of regression analyses using multiple thresholds of snowmelt based on temporal and volumetric metrics. NSA indices defined by volume of snowmelt (e.g., snowmelt ≤ 30 cm) rather than snowmelt duration (e.g., snowmelt ≤ 9 days) were the best predictors of stream NO3- <span class="hlt">concentrations</span>. The NSA indices were better correlated with stream NO3- <span class="hlt">concentration</span> in TOK (average R2= 0.68) versus GLV4 (average R2= 0.44). Positive relationships between NSA and stream NO3- <span class="hlt">concentration</span> were <span class="hlt">observed</span> in TOK with peak stream NO3- <span class="hlt">concentration</span> occurring on the rising limb of snowmelt. Positive and negative relationships between NSA and stream NO3- <span class="hlt">concentration</span> were found in GLV4 with peak stream NO3- <span class="hlt">concentration</span> occurring as NSA expands. Consistent with previous works, the contrasting NO3- flushing behavior suggests that streamflow in TOK was primarily influenced by overland flow and shallow subsurface flow, whereas GLV4 appeared to be more strongly influenced by deeper subsurface flow paths.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29950688','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29950688"><span>Metabolites Re-programming and Physiological Changes Induced in Scenedesmus regularis under <span class="hlt">Nitrate</span> Treatment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ma, Nyuk-Ling; Aziz, Ahmad; Teh, Kit-Yinn; Lam, Su Shiung; Cha, Thye-San</p> <p>2018-06-27</p> <p><span class="hlt">Nitrate</span> is required to maintain the growth and metabolism of plant and animals. Nevertheless, in excess amount such as polluted water, its <span class="hlt">concentration</span> can be harmful to living organisms such as microalgae. Recently, studies on microalgae response towards nutrient fluctuation are usually limited to lipid accumulation for the production of biofuels, disregarding the other potential of microalgae to be used in wastewater treatments and as source of important metabolites. Our study therefore captures the need to investigate overall metabolite changes via NMR spectroscopy approach coupled with multivariate data to understand the complex molecular process under high (4X) and low (1/4X) <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> ([Formula: see text]). NMR spectra with the aid of chemometric analysis revealed contrasting metabolites makeup under abundance and limited <span class="hlt">nitrate</span> treatment. By using NMR technique, 43 types of metabolites and 8 types of fatty acid chains were detected. Nevertheless, only 20 key changes were <span class="hlt">observed</span> and 16 were down regulated in limited <span class="hlt">nitrate</span> condition. This paper has demonstrated the feasibility of NMR-based metabolomics approach to study the physiological impact of changing environment such as pollution to the implications for growth and productivity of microalgae population.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H53H1570N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H53H1570N"><span><span class="hlt">Nitrate</span> pollution and surface water chemistry in Shimabara, Nagasaki Prefecture, Japan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakagawa, K.; Amano, H.</p> <p>2017-12-01</p> <p>Shimabara city has been experiencing serious <span class="hlt">nitrate</span> pollution in groundwater. To evaluate <span class="hlt">nitrate</span> pollution and water chemistry in surface water, water samples were collected at 42 sampling points in 15 rivers in Shimabara including a part of Unzen city from January to February 2017. Firstly, spatial distribution of water chemistry was assessed by describing stiff and piper-trilinear diagrams using major ions <span class="hlt">concentrations</span>. Most of the samples showed Ca-HCO3 or Ca-(NO3+SO4) water types. It corresponds to groundwater chemistry. Some samples were classified into characteristic water types such as Na-Cl, (Na+K)-HCO3, and Ca-Cl. These results indicate sea water mixing and anthropogenic pollution. At the upstream of Nishi-river, although water chemistry showed Ca-HCO3, ions <span class="hlt">concentrations</span> were higher than that of the other rivers. It indicates that this site was affected by the peripheral anthropogenic activities. Secondly, <span class="hlt">nitrate</span>-pollution assessment was performed by using NO3-, NO2-, coprostanol (5β(H)-Cholestan-3β-ol), and cholestanol (5α(H)-Cholestan-3β-ol). NO2-N was detected at the 2 sampling points and exceeded drinking standard 0.9 mg L-1 for bottle-fed infants (WHO, 2011). NO3-N + NO2-N <span class="hlt">concentrations</span> exceeded Japanese drinking standard 10 mg L-1 at 18 sampling points. The highest <span class="hlt">concentration</span> was 27.5 mg L-1. Higher NO3-N levels were <span class="hlt">observed</span> in the rivers in the northern parts of the study area. Coprostanol has been used as a fecal contamination indicator, since it can be found in only feces of higher animals. Coprostanol <span class="hlt">concentrations</span> at 8 sampling points exceeded 700 ng L-1 (Australian drinking water standard). Coprostanol has a potential to distinguish the <span class="hlt">nitrate</span> pollution sources between chemical fertilizer or livestock wastes, since water samples with similar NO3-N + NO2-N <span class="hlt">concentration</span> showed distinct coprostanol <span class="hlt">concentration</span>. The sterols ratio (5β/ (5β+5α)) exceeded 0.5 at 18 sampling points. This reveals that fecal pollution has occurred.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=248316&keyword=tile&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=248316&keyword=tile&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>Impacts of drainage water management on subsurface drain flow, <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, and <span class="hlt">nitrate</span> loads in Indiana</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>Drainage water management is a conservation practice that has the potential to reduce drainage outflow and <span class="hlt">nitrate</span> (NO3) loss from agricultural fields while maintaining or improving crop yields. The goal of this study was to quantify the impact of drainage water management on dra...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUSM.B32A..01S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUSM.B32A..01S"><span>Effects of Atmospheric <span class="hlt">Nitrate</span> on an Upland Stream of the Northeastern USA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sebestyen, S. D.; Shanley, J. B.; Boyer, E. W.; Kendall, C.</p> <p>2009-05-01</p> <p>Excess nitrogen cascades through terrestrial biogeochemical cycles and affects stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in upland forests where atmospheric deposition is an important source of anthropogenic nitrogen. We will discuss approaches including high-frequency sampling, isotopic tracers, and end-member mixing analysis that can be used to decipher the sources, transformations, and hydrological processes that affect <span class="hlt">nitrate</span> transport through forested upland catchments to streams. We present results of studies at the Sleepers River Research Watershed in Vermont, USA, a site where we have intensively measured stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> during baseflow and stormflow. Stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are typically low and nearly 75% of annual inorganic N inputs from atmospheric deposition are retained within the catchment. However, high <span class="hlt">concentrations</span> and stream loadings of <span class="hlt">nitrate</span> occur during storm events due to source variation and hydrological flushing of <span class="hlt">nitrate</span> from catchment soils. Using isotopic tracers and end-member mixing analysis, we have quantified source inputs of unprocessed atmospheric <span class="hlt">nitrate</span> and show that this stream is directly affected by nitrogen pollution. Using a long-term record of stream hydrochemistry and our findings on event- scale <span class="hlt">nitrate</span> flushing dynamics, we then explore how stream <span class="hlt">nitrate</span> loading may respond to anthropogenic climate forcing during the next century. Results suggest that stream runoff and <span class="hlt">nitrate</span> loadings will change during future emission scenarios (i.e. longer growing seasons and higher winter precipitation rates). Understanding the timing and magnitude of hydrological and hydrochemical responses is important because climate change effects on catchment hydrology may alter how <span class="hlt">nitrate</span> is retained, produced, and hydrologically flushed in headwater ecosystems with implications for aquatic metabolism, nutrient export from catchments, and downstream eutrophication.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27460024','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27460024"><span>The use of multilevel sampling techniques for determining shallow aquifer <span class="hlt">nitrate</span> profiles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lasagna, Manuela; De Luca, Domenico Antonio</p> <p>2016-10-01</p> <p><span class="hlt">Nitrate</span> is a worldwide pollutant in aquifers. Shallow aquifer <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> generally display vertical stratification, with a maximum <span class="hlt">concentration</span> immediately below the water level. The <span class="hlt">concentration</span> then gradually decreases with depth. Different techniques can be used to highlight this stratification. The paper aims at comparing the advantages and limitations of three open hole multilevel sampling techniques (packer system, dialysis membrane samplers and bailer), chosen on the base of a literary review, to highlight a <span class="hlt">nitrate</span> vertical stratification under the assumption of (sub)horizontal flow in the aquifer. The sampling systems were employed at three different times of the year in a shallow aquifer piezometer in northern Italy. The optimal purge time, equilibration time and water volume losses during the time in the piezometer were evaluated. Multilevel techniques highlighted a similar vertical <span class="hlt">nitrate</span> stratification, present throughout the year. Indeed, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> generally decreased with depth downwards, but with significantly different levels in the sampling campaigns. Moreover, the sampling techniques produced different degrees of accuracy. More specifically, the dialysis membrane samplers provided the most accurate hydrochemical profile of the shallow aquifer and they appear to be necessary when the objective is to detect the discontinuities in the <span class="hlt">nitrate</span> profile. Bailer and packer system showed the same <span class="hlt">nitrate</span> profile with little differences of <span class="hlt">concentration</span>. However, the bailer resulted much more easier to use.</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.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3562778','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3562778"><span>Biogeochemistry of beetle-killed forests: Explaining a weak <span class="hlt">nitrate</span> response</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Rhoades, Charles C.; McCutchan, James H.; Cooper, Leigh A.; Clow, David; Detmer, Thomas M.; Briggs, Jennifer S.; Stednick, John D.; Veblen, Thomas T.; Ertz, Rachel M.; Likens, Gene E.; Lewis, William M.</p> <p>2013-01-01</p> <p>A current pine beetle infestation has caused extensive mortality of lodgepole pine (Pinus contorta) in forests of Colorado and Wyoming; it is part of an unprecedented multispecies beetle outbreak extending from Mexico to Canada. In United States and European watersheds, where atmospheric deposition of inorganic N is moderate to low (<10 kg⋅ha⋅y), disturbance of forests by timber harvest or violent storms causes an increase in stream <span class="hlt">nitrate</span> <span class="hlt">concentration</span> that typically is close to 400% of predisturbance <span class="hlt">concentrations</span>. In contrast, no significant increase in streamwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> has occurred following extensive tree mortality caused by the mountain pine beetle in Colorado. A model of <span class="hlt">nitrate</span> release from Colorado watersheds calibrated with field data indicates that stimulation of <span class="hlt">nitrate</span> uptake by vegetation components unaffected by beetles accounts for significant <span class="hlt">nitrate</span> retention in beetle-infested watersheds. The combination of low atmospheric N deposition (<10 kg⋅ha⋅y), tree mortality spread over multiple years, and high compensatory capacity associated with undisturbed residual vegetation and soils explains the ability of these beetle-infested watersheds to retain <span class="hlt">nitrate</span> despite catastrophic mortality of the dominant canopy tree species. PMID:23319612</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890005156','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890005156"><span>Antarctic polar stratospheric aerosols: The roles of <span class="hlt">nitrates</span>, chlorides and sulfates</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Pueschel, R. F.; Snetsinger, K. G.; Goodman, J. K.; Ferry, G. V.; Oberbeck, V. R.; Verma, S.; Fong, W.</p> <p>1988-01-01</p> <p>Nitric and hydrochloric acids have been postulated to condense in the winter polar stratosphere to become an important component of polar stratospheric clouds. One implication is that the removal of NO(y) from the gas phase by this mechanism allows high Cl(x) <span class="hlt">concentrations</span> to react with O3, because the formation of ClNO3 is inhibited. Contributions of NO3 and Cl to the stratospheric aerosol were determined during the 1987 Airborne Antarctic Ozone Experiment by testing for the presence of <span class="hlt">nitrates</span> and chlorides in the condensed phase. Aerosol particles were collected on four 500 micron diameter gold wires, each pretreated differently to give results that were specific to certain physical and chemical aerosol properties. One wire was carbon-coated for <span class="hlt">concentration</span> and size analyses by scanning electron microscopy; X-ray energy dispersive analyses permitted the detection of S and Cl in individual particles. Three more wires were coated with Nitron, barium chloride and silver <span class="hlt">nitrate</span>, respectively, to detect <span class="hlt">nitrate</span>, sulfate and chloride in aerosol particles. All three ions, viz., sulfates, <span class="hlt">nitrates</span> and chlorides were detected in the Antarctic stratospheric aerosol. In terms of number <span class="hlt">concentrations</span>, the aerosol was dominated by sulfates, followed by chlorides and <span class="hlt">nitrates</span>. An inverse linear regression can be established between <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and ozone mixing ratio, and between temperature and <span class="hlt">nitrates</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=227907&keyword=principal+AND+management&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=227907&keyword=principal+AND+management&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>Test/QA Plan for Verification of <span class="hlt">Nitrate</span> Sensors for Groundwater Remediation Monitoring</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 submersible <span class="hlt">nitrate</span> sensor is capable of collecting in-situ measurements of dissolved <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater. Although several types of <span class="hlt">nitrate</span> sensors currently exist, this verification test will focus on submersible sensors equipped with a <span class="hlt">nitrate</span>-specific ion...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70031954','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70031954"><span>Ecohydrological factors affecting <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in a phreatic desert aquifer in northwestern China</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Gates, J.B.; Böhlke, J.K.; Edmunds, W.M.</p> <p>2008-01-01</p> <p>Aerobic conditions in desert aquifers commonly allow high <span class="hlt">nitrate</span> (NO 3-) <span class="hlt">concentrations</span> in recharge to persist for long periods of time, an important consideration for N-cycling and water quality. In this study, stable isotopes of NO3- (??15N NO3 and ??18ONO3) were used to trace NO3- cycling processes which affect <span class="hlt">concentrations</span> in groundwater and unsaturated zone moisture in the arid Badain Jaran Oesert in northwestern China. Most groundwater NO3- appears to be depleted relative to Cl- in rainfall <span class="hlt">concentrated</span> by evapotranspiration, indicating net N losses. Unsaturated zone NO 3- is generally higher than groundwater NO 3- in terms of both <span class="hlt">concentration</span> (up to 15 476 ??M, corresponding to 3.6 mg NO3--N per kg sediment) and ratios with Cl-. Isotopic data indicate that the NO3- derives primarily from nitrification, with a minor direct contribution of atmospheric NO3- inferred for some samples, particularly in the unsaturated zone. Localized denitrification in the saturated zone is suggested by isotopic and geochemical indicators in some areas. Anthropogenic inputs appear to be minimal, and variability is attributed to environmental factors. In comparison to other arid regions, the sparseness of vegetation in the study area appears to play an important role in moderating unsaturated zone NO3- accumulation by allowing solute flushing and deterring extensive N2 fixation. ?? 2008 American Chemical Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1619346','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1619346"><span><span class="hlt">Nitrates</span>, chlorates and trihalomethanes in swimming pool water.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Beech, J A; Diaz, R; Ordaz, C; Palomeque, B</p> <p>1980-01-01</p> <p>Water from swimming pools in the Miami area was analyzed for <span class="hlt">nitrates</span>, chlorates and trihalomethanes. The average <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and chlorate found in freshwater pools were 8.6 mg/liter and 16 mg/liter respectively, with the highest <span class="hlt">concentrations</span> being 54.9 mg/liter and 124 mg/liter, respectively. The average <span class="hlt">concentration</span> of total trihalomethanes found in freshwater pools was 125 micrograms/liter (mainly chloroform) and in saline pools was 657 micrograms/liter (mainly bromoform); the highest <span class="hlt">concentration</span> was 430 micrograms/liter (freshwater) and 1287 micrograms/liter (saltwater). The possible public health significance of these results is briefly discussed. PMID:7350831</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..1615236N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..1615236N"><span>Thermophilic <span class="hlt">nitrate</span>-reducing microorganisms prevent sulfate reduction in cold marine sediments incubated at high temperature</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nepomnyashchaya, Yana; Rezende, Julia; Hubert, Casey</p> <p>2014-05-01</p> <p>Hydrogen sulphide produced during metabolism of sulphate-reducing microorganisms (SRM) is toxic, corrosive and causes detrimental oil reservoir souring. During secondary oil recovery, injecting oil reservoirs with seawater that is rich in sulphate and that also cools high temperature formations provides favourable growth conditions for SRM. <span class="hlt">Nitrate</span> addition can prevent metabolism of SRM by stimulating <span class="hlt">nitrate</span>-reducing microorganisms (NRM). The investigations of thermophilic NRM are needed to develop mechanisms to control the metabolism of SRM in high temperature oil field ecosystems. We therefore established a model system consisting of enrichment cultures of cold surface marine sediments from the Baltic Sea (Aarhus Bay) that were incubated at 60°C. Enrichments contained 25 mM <span class="hlt">nitrate</span> and 40 mM sulphate as potential electron acceptors, and a mixture of the organic substrates acetate, lactate, propionate, butyrate (5 mM each) and yeast extract (0.01%) as potential carbon sources and electron donors. Slurries were incubated at 60°C both with and without initial pasteurization at 80°C for 2 hours. In the enrichments containing both <span class="hlt">nitrate</span> and sulphate, the <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> decreased indicating metabolic activity of NRM. After a four-hour lag phase the rate of <span class="hlt">nitrate</span> reduction increased and the <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> dropped to zero after 10 hours of incubation. The <span class="hlt">concentration</span> of nitrite increased as the reduction of <span class="hlt">nitrate</span> progressed and reached 16.3 mM after 12 hours, before being consumed and falling to 4.4 mM after 19-day of incubation. No evidence for sulphate reduction was <span class="hlt">observed</span> in these cultures during the 19-day incubation period. In contrast, the <span class="hlt">concentration</span> of sulphate decreased up to 50% after one week incubation in controls containing only sulphate but no <span class="hlt">nitrate</span>. Similar sulfate reduction rates were seen in the pasteurized controls suggesting the presence of heat resistant SRM, whereas <span class="hlt">nitrate</span> reduction rates were lower in the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1992JHyd..137..181B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1992JHyd..137..181B"><span>The origin of high-<span class="hlt">nitrate</span> ground waters in the Australian arid zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barnes, C. J.; Jacobson, G.; Smith, G. D.</p> <p>1992-08-01</p> <p><span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> beyond the drinking-water limit of 10 mg1 -1 NO 3-N, are common in Australian arid-zone ground waters and are often associated with otherwise potable waters. In some aquifers <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> of up to 80 mg1 -1 have been found, and this is a severe constraint on water supply development for small settlements. Water-bore data indicate a correlation of high-<span class="hlt">nitrate</span> ground waters with shallow unconfined aquifers. Aguifer hydrochemistry indicats that these ground waters were emplaced by episodic Holocene recharge events in an otherwise arid climate regime. <span class="hlt">Nitrate</span> has been flushed through the unsaturated zone which apparently lacks denitrification activity. The <span class="hlt">nitrate</span> originates by near-surface biological fixation and contributing organisms include cyanobacteria in soil crusts and bacteria in termite mounds with the highest soil <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> found in the outer skin of termite mounds. Bacteria associated with the termites appear to fix nitrogen, which eventually appears in an inorganic form, principally as ammonia. <span class="hlt">Nitrate</span> is produced by bacterial oxidation of the ammonia, and is leached to the outside of the termite mound by capillary action. Diffuse recharge from extreme rainfall events then flushes this <span class="hlt">nitrate</span> to the water table.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18172806','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18172806"><span>Reduction of trichloroethylene and <span class="hlt">nitrate</span> by zero-valent iron with peat.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Min, Jee-Eun; Kim, Meejeong; Pardue, John H; Park, Jae-Woo</p> <p>2008-02-01</p> <p>The feasibility of using zero-valent iron (ZVI) and peat mixture as in situ barriers for contaminated sediments and groundwater was investigated. Trichloroethylene (TCE) and <span class="hlt">nitrate</span> (NO(3)(-)), redox sensitive contaminants were reduced by ZVI and peat soil mixture under anaerobic condition. Peat was used to support the sorption of TCE, microbial activity for biodegradation of TCE and denitrification while TCE and <span class="hlt">nitrate</span> were reduced by ZVI. Decreases in TCE <span class="hlt">concentrations</span> were mainly due to ZVI, while peat supported denitrifying microbes and further affected the sorption of TCE. Due to the competition of electrons, <span class="hlt">nitrate</span> reduction was inhibited by TCE, while TCE reduction was not affected by <span class="hlt">nitrate</span>. From the results of peat and sterilized peat, it can be concluded that peat was involved in both dechlorination and denitrification but biological reduction of TCE was negligible compared to that of <span class="hlt">nitrate</span>. The results from hydrogen and methane gas analyses confirmed that hydrogen utilization by microbes and methanogenic process had occurred in the ZVI-peat system. Even though effect of the peat on TCE reduction were quantitatively small, ZVI and peat contributed to the removal of TCE and <span class="hlt">nitrate</span> independently. The 16S rRNA analysis revealed that viable bacterial diversity was narrow and the most frequently <span class="hlt">observed</span> genera were Bacillus and Staphylococcus spp.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3984109','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3984109"><span><span class="hlt">Nitrate</span> Reduction Functional Genes and <span class="hlt">Nitrate</span> Reduction Potentials Persist in Deeper Estuarine Sediments. Why?</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Papaspyrou, Sokratis; Smith, Cindy J.; Dong, Liang F.; Whitby, Corinne; Dumbrell, Alex J.; Nedwell, David B.</p> <p>2014-01-01</p> <p>Denitrification and dissimilatory <span class="hlt">nitrate</span> reduction to ammonium (DNRA) are processes occurring simultaneously under oxygen-limited or anaerobic conditions, where both compete for <span class="hlt">nitrate</span> and organic carbon. Despite their ecological importance, there has been little investigation of how denitrification and DNRA potentials and related functional genes vary vertically with sediment depth. <span class="hlt">Nitrate</span> reduction potentials measured in sediment depth profiles along the Colne estuary were in the upper range of <span class="hlt">nitrate</span> reduction rates reported from other sediments and showed the existence of strong decreasing trends both with increasing depth and along the estuary. Denitrification potential decreased along the estuary, decreasing more rapidly with depth towards the estuary mouth. In contrast, DNRA potential increased along the estuary. Significant decreases in copy numbers of 16S rRNA and <span class="hlt">nitrate</span> reducing genes were <span class="hlt">observed</span> along the estuary and from surface to deeper sediments. Both metabolic potentials and functional genes persisted at sediment depths where porewater <span class="hlt">nitrate</span> was absent. Transport of <span class="hlt">nitrate</span> by bioturbation, based on macrofauna distributions, could only account for the upper 10 cm depth of sediment. A several fold higher combined freeze-lysable KCl-extractable <span class="hlt">nitrate</span> pool compared to porewater <span class="hlt">nitrate</span> was detected. We hypothesised that his could be attributed to intracellular <span class="hlt">nitrate</span> pools from <span class="hlt">nitrate</span> accumulating microorganisms like Thioploca or Beggiatoa. However, pyrosequencing analysis did not detect any such organisms, leaving other bacteria, microbenthic algae, or foraminiferans which have also been shown to accumulate <span class="hlt">nitrate</span>, as possible candidates. The importance and bioavailability of a KCl-extractable <span class="hlt">nitrate</span> sediment pool remains to be tested. The significant variation in the vertical pattern and abundance of the various <span class="hlt">nitrate</span> reducing genes phylotypes reasonably suggests differences in their activity throughout the sediment column. This</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H42E..01A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H42E..01A"><span><span class="hlt">Nitrate</span> and Moisture Content of Broad Permafrost Landscape Features in the Barrow Peninsula: Predicting Evolving NO3 <span class="hlt">Concentrations</span> in a Changing Arctic</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arendt, C. A.; Heikoop, J. M.; Newman, B. D.; Wales, N. A.; McCaully, R. E.; Wilson, C. J.; Wullschleger, S.</p> <p>2017-12-01</p> <p>The geochemical evolution of Arctic regions as permafrost degrades, significantly impacts nutrient availability. The release of nitrogen compounds from permafrost degradation fertilizes both microbial decomposition and plant productivity. Arctic warming promotes permafrost degradation, causing geomorphic and hydrologic transitions that have the potential to convert saturated zones to unsaturated zones and subsequently alter the <span class="hlt">nitrate</span> production capacity of permafrost regions. Changes in <span class="hlt">Nitrate</span> (NO3-) content associated with shifting moisture regimes are a primary factor determining Arctic fertilization and subsequent primary productivity, and have direct feedbacks to carbon cycling. We have documented a broad survey of co-located soil moisture and <span class="hlt">nitrate</span> <span class="hlt">concentration</span> measurements in shallow active layer regions across a variety of topographic features in the expansive continuous permafrost region encompassing the Barrow Peninsula of Alaska. Topographic features of interest are slightly higher relative to surrounding landscapes with drier soils and elevated <span class="hlt">nitrate</span>, including the rims of low centered polygons, the centers of flat and high centered polygons, the rims of young, old and ancient drain thaw lake basins and drainage slopes that exist across the landscape. With this information, we model the <span class="hlt">nitrate</span> inventory of the Barrow Peninsula using multiple geospatial approaches to estimate total area cover by unsaturated features of interest and further predict how various drying scenarios increase the magnitude of <span class="hlt">nitrate</span> produced in degrading permafrost regions across the Arctic. This work is supported by the US Department of Energy Next Generation Ecosystem Experiment, NGEE-Arctic.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title33-vol2/pdf/CFR-2010-title33-vol2-sec126-28.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title33-vol2/pdf/CFR-2010-title33-vol2-sec126-28.pdf"><span>33 CFR 126.28 - Ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>...</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-07-01</p> <p>... <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. 126.28 Section 126..., ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. (a) When any item of ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title33-vol2/pdf/CFR-2014-title33-vol2-sec126-28.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title33-vol2/pdf/CFR-2014-title33-vol2-sec126-28.pdf"><span>33 CFR 126.28 - Ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>...</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-07-01</p> <p>... <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. 126.28 Section 126..., ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. (a) When any item of ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title33-vol2/pdf/CFR-2013-title33-vol2-sec126-28.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title33-vol2/pdf/CFR-2013-title33-vol2-sec126-28.pdf"><span>33 CFR 126.28 - Ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</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>... <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. 126.28 Section 126..., ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. (a) When any item of ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title33-vol2/pdf/CFR-2011-title33-vol2-sec126-28.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title33-vol2/pdf/CFR-2011-title33-vol2-sec126-28.pdf"><span>33 CFR 126.28 - Ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</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>... <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. 126.28 Section 126..., ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. (a) When any item of ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title33-vol2/pdf/CFR-2012-title33-vol2-sec126-28.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title33-vol2/pdf/CFR-2012-title33-vol2-sec126-28.pdf"><span>33 CFR 126.28 - Ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</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>... <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. 126.28 Section 126..., ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>; general provisions. (a) When any item of ammonium <span class="hlt">nitrate</span>, ammonium <span class="hlt">nitrate</span> fertilizers, fertilizer mixtures, or nitro carbo <span class="hlt">nitrate</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21870765','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21870765"><span>[<span class="hlt">Nitrates</span> and nitrites in meat products--nitrosamines precursors].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Avasilcăi, Liliana; Cuciureanu, Rodica</p> <p>2011-01-01</p> <p>To determine the content in <span class="hlt">nitrates</span> and nitrites and the formation of two nitrosamines (N-nitrosodimethylamine--NDMA, and N-nitrosodiethylaamine--NDEA) in samples of chicken ham, dry Banat salami, dry French salami, traditional Romania sausages, and pork pastrami. Nitrites were determined by spectrophotometry with Peter-Griess reagent, and <span class="hlt">nitrates</span> by the same method after reduction to nitrites with cadmium powder. High performance liquid chromatography with UV detection was used to determine nitrosamines. The initial <span class="hlt">concentration</span> of <span class="hlt">nitrates</span>, nitrites, NDMA and NDEA in the samples ranged as follows: 14.10-60.40 mg NO3/kg, 2.70-26.70 mg NO2/kg, from non-detectable to 0.90 microg NDMA/kg, and from non-detectable to 0.27 microg NDEA/kg, respectively. After 28 days the <span class="hlt">concentrations</span> were: 3.24-17.1 mg NO3/kg, 0.04 -1.87 mg NO2/kg, 0.8-29 microg NDMA/kg, and 11.6-61.9 microg NDEA/kg, respectively. The decreased <span class="hlt">nitrate</span> and nitrite and increased NDMA and NDEA <span class="hlt">concentrations</span> prove that in food products nitrosamines are formed due to residual nitrite during their preservation. The determination of nitrasamines revealed levels much above the admitted maximal <span class="hlt">concentration</span> for these food products.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70019648','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70019648"><span>Risk of <span class="hlt">nitrate</span> in groundwaters of the United States - A national perspective</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Nolan, B.T.; Ruddy, B.C.; Hitt, K.J.; Helsel, D.R.</p> <p>1997-01-01</p> <p><span class="hlt">Nitrate</span> contamination of groundwater occurs in predictable patterns, based on findings of the U.S. Geological Survey's (USGS) National Water Quality Assessment (NAWQA) Program. The NAWQA Program was begun in 1991 to describe the quality of the Nation's water resources, using nationally consistent methods. Variables affecting <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater were grouped as 'input' factors (population density end the amount of nitrogen contributed by fertilizer, manure, and atmospheric sources) and 'aquifer vulnerability' factors (soil drainage characteristic and the ratio of woodland acres to cropland acres in agricultural areas) and compiled in a national map that shows patterns of risk for <span class="hlt">nitrate</span> contamination of groundwater. Areas with high nitrogen input, well-drained soils, and low woodland to cropland ratio have the highest potential for contamination of shallow groundwater by <span class="hlt">nitrate</span>. Groundwater <span class="hlt">nitrate</span> data collected through 1992 from wells less than 100 ft deep generally verified the risk patterns shown on the national map. Median <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was 0.2 mg/L in wells representing the low-risk group, and the maximum contaminant level (MCL) was exceeded in 3% of the wells. In contrast, median <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was 4.8 mg/L in wells representing the high-risk group, and the MCL was exceeded in 25% of the wells.<span class="hlt">Nitrate</span> contamination of groundwater occurs in predictable patterns, based on findings of the U.S. Geological Survey's (USGS) National Water Quality Assessment (NAWQA) Program. The NAWQA Program was begun in 1991 to describe the quality of the Nation's water resources, using nationally consistent methods. Variables affecting <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater were grouped as `input' factors (population density and the amount of nitrogen contributed by fertilizer, manure, and atmospheric sources) and `aquifer vulnerability' factors (soil drainage characteristic and the ratio of woodland acres to cropland acres in agricultural areas</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20864207','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20864207"><span>Regional transport modelling for <span class="hlt">nitrate</span> trend assessment and forecasting in a chalk aquifer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Orban, Philippe; Brouyère, Serge; Batlle-Aguilar, Jordi; Couturier, Julie; Goderniaux, Pascal; Leroy, Mathieu; Maloszewski, Piotr; Dassargues, Alain</p> <p>2010-10-21</p> <p>Regional degradation of groundwater resources by <span class="hlt">nitrate</span> has become one of the main challenges for water managers worldwide. Regulations have been defined to reverse <span class="hlt">observed</span> <span class="hlt">nitrate</span> trends in groundwater bodies, such as the Water Framework Directive and the Groundwater Daughter Directive in the European Union. In such a context, one of the main challenges remains to develop efficient approaches for groundwater quality assessment at regional scale, including quantitative numerical modelling, as a decision support for groundwater management. A new approach combining the use of environmental tracers and the innovative 'Hybrid Finite Element Mixing Cell' (HFEMC) modelling technique is developed to study and forecast the groundwater quality at the regional scale, with an application to a regional chalk aquifer in the Geer basin in Belgium. Tritium data and <span class="hlt">nitrate</span> time series are used to produce a conceptual model for regional groundwater flow and contaminant transport in the combined unsaturated and saturated zones of the chalk aquifer. This shows that the spatial distribution of the contamination in the Geer basin is essentially linked to the hydrodynamic conditions prevailing in the basin, more precisely to groundwater age and mixing and not to the spatial patterns of land use or local hydrodispersive processes. A three-dimensional regional scale groundwater flow and solute transport model is developed. It is able to reproduce the spatial patterns of tritium and <span class="hlt">nitrate</span> and the <span class="hlt">observed</span> <span class="hlt">nitrate</span> trends in the chalk aquifer and it is used to predict the evolution of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the basin. The modelling application shows that the global inertia of groundwater quality is strong in the basin and trend reversal is not expected to occur before the 2015 deadline fixed by the European Water Framework Directive. The expected time required for trend reversal ranges between 5 and more than 50 years, depending on the location in the basin and the expected reduction</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.springerlink.com/content/xk5242515098g188/','USGSPUBS'); return false;" href="http://www.springerlink.com/content/xk5242515098g188/"><span>A caveat regarding diatom-inferred nitrogen <span class="hlt">concentrations</span> in oligotrophic lakes</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Arnett, Heather A.; Saros, Jasmine E.; Mast, M. Alisa</p> <p>2012-01-01</p> <p>Atmospheric deposition of reactive nitrogen (Nr) has enriched oligotrophic lakes with nitrogen (N) in many regions of the world and elicited dramatic changes in diatom community structure. The lakewater <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> that cause these community changes remain unclear, raising interest in the development of diatom-based transfer functions to infer <span class="hlt">nitrate</span>. We developed a diatom calibration set using surface sediment samples from 46 high-elevation lakes across the Rocky Mountains of the western US, a region spanning an N deposition gradient from very low to moderate levels (<1 to 3.2 kg Nr ha−1 year−1 in wet deposition). Out of the fourteen measured environmental variables for these 46 lakes, ordination analysis identified that <span class="hlt">nitrate</span>, specific conductance, total phosphorus, and hypolimnetic water temperature were related to diatom distributions. A transfer function was developed for <span class="hlt">nitrate</span> and applied to a sedimentary diatom profile from Heart Lake in the central Rockies. The model coefficient of determination (bootstrapping validation) of 0.61 suggested potential for diatom-inferred reconstructions of lakewater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> over time, but a comparison of <span class="hlt">observed</span> versus diatom-inferred <span class="hlt">nitrate</span> values revealed the poor performance of this model at low <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Resource physiology experiments revealed that nitrogen requirements of two key taxa were opposite to <span class="hlt">nitrate</span> optima defined in the transfer function. Our data set reveals two underlying ecological constraints that impede the development of <span class="hlt">nitrate</span> transfer functions in oligotrophic lakes: (1) even in lakes with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> below quantification (<1 μg L−1), diatom assemblages were already dominated by species indicative of moderate N enrichment; (2) N-limited oligotrophic lakes switch to P limitation after receiving only modest inputs of reactive N, shifting the controls on diatom species changes along the length of the <span class="hlt">nitrate</span> gradient. These</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21885082','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21885082"><span><span class="hlt">Nitrate</span> reduction in a simulated free-water surface wetland system.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Misiti, Teresa M; Hajaya, Malek G; Pavlostathis, Spyros G</p> <p>2011-11-01</p> <p>The feasibility of using a constructed wetland for treatment of <span class="hlt">nitrate</span>-contaminated groundwater resulting from the land application of biosolids was investigated for a site in the southeastern United States. Biosolids degradation led to the release of ammonia, which upon oxidation resulted in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the upper aquifer in the range of 65-400 mg N/L. A laboratory-scale system was constructed in support of a pilot-scale project to investigate the effect of temperature, hydraulic retention time (HRT) and <span class="hlt">nitrate</span> and carbon loading on denitrification using soil and groundwater from the biosolids application site. The maximum specific reduction rates (MSRR), measured in batch assays conducted with an open to the atmosphere reactor at four initial <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> from 70 to 400 mg N/L, showed that the <span class="hlt">nitrate</span> reduction rate was not affected by the initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. The MSRR values at 22 °C for <span class="hlt">nitrate</span> and nitrite were 1.2 ± 0.2 and 0.7 ± 0.1 mg N/mg VSS(COD)-day, respectively. MSRR values were also measured at 5, 10, 15 and 22 °C and the temperature coefficient for <span class="hlt">nitrate</span> reduction was estimated at 1.13. Based on the performance of laboratory-scale continuous-flow reactors and model simulations, wetland performance can be maintained at high nitrogen removal efficiency (>90%) with an HRT of 3 days or higher and at temperature values as low as 5 °C, as long as there is sufficient biodegradable carbon available to achieve complete denitrification. The results of this study show that based on the climate in the southeastern United States, a constructed wetland can be used for the treatment of <span class="hlt">nitrate</span>-contaminated groundwater to low, acceptable <span class="hlt">nitrate</span> levels. Copyright © 2011 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_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.fs.usda.gov/treesearch/pubs/19547','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/19547"><span>Sources, transformations, and hydrological processes that control stream <span class="hlt">nitrate</span> and dissolved organic matter <span class="hlt">concentrations</span> during snowmelt in an upland forest</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Stephen D. Sebestyen; Elizabeth W. Boyer; James B. Shanley; Carol Kendall; Daniel H. Doctor; George R. Aiken; Nobuhito Ohte</p> <p>2008-01-01</p> <p>We explored catchment processes that control stream nutrient <span class="hlt">concentrations</span> at an upland forest in northeastern Vermont, USA, where inputs of nitrogen via atmospheric deposition are among the highest in the nation and affect ecosystem functioning. We traced sources of water, <span class="hlt">nitrate</span>, and dissolved organic matter (DOM) using stream water samples collected at high...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JHyd..555..760J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JHyd..555..760J"><span>Vadose zone processes delay groundwater <span class="hlt">nitrate</span> reduction response to BMP implementation as <span class="hlt">observed</span> in paired cultivated vs. uncultivated potato rotation fields</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jiang, Y.; Nyiraneza, J.; Murray, B. J.; Chapman, S.; Malenica, A.; Parker, B.</p> <p>2017-12-01</p> <p><span class="hlt">Nitrate</span> leaching from crop production contributes to groundwater contamination and subsequent eutrophication of the receiving surface water. A study was conducted in a 7-ha potato-grain-forages rotation field in Prince Edward Island (PEI), Canada during 2011-2016 to link potato rotation practices and groundwater quality. The field consists of fine sandy loam soil and is underlain by 7-9 m of glacial till, which overlies the regional fractured ;red-bed; sandstone aquifer. The water table is generally located in overburden close to the bedrock interface. Field treatments included one field zone taken out of production in 2011 with the remaining zones kept under a conventional potato rotation. Agronomy data including crop tissue, soil, and tile-drain water quality were collected. Hydrogeology data including multilevel monitoring of groundwater <span class="hlt">nitrate</span> and hydraulic head and data from rock coring for <span class="hlt">nitrate</span> distribution in overburden and bedrock matrix were also collected. A significant amount of <span class="hlt">nitrate</span> leached below the soil profile after potato plant kill (referred to as topkill) in 2011, most of it from fertilizer N. A high level of <span class="hlt">nitrate</span> was also detected in the till vadose zone through coring in December 2012 and through multilevel groundwater sampling from January to May 2014 in both cultivated and uncultivated field zones. Groundwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increased for about 2.5 years after the overlying potato field was removed from production. Pressure-driven uniform flow processes dominate water and <span class="hlt">nitrate</span> transport in the vadose zone, producing an apparently instant water table response but a delayed groundwater quality response to <span class="hlt">nitrate</span> leaching events. These data suggest that the uniform flow dominated vadose zone in agricultural landscapes can cause the accumulation of a significant amount of <span class="hlt">nitrate</span> originated from previous farming activities, and the long travel time of this legacy <span class="hlt">nitrate</span> in the vadose zone can result in substantially delayed</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27612180','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27612180"><span>Investigating <span class="hlt">nitrate</span> dynamics in a fine-textured soil affected by feedlot effluents.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Veizaga, E A; Rodríguez, L; Ocampo, C J</p> <p>2016-10-01</p> <p>Feedlots <span class="hlt">concentrate</span> large volumes of manure and effluents that contain high <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>, among other constituents. If not managed properly, pen surfaces run-off and lagoons overflows may spread those effluents to surrounding land, infiltrating into the soil. Soil <span class="hlt">nitrate</span> mobilization and distribution are of great concern due to its potential migration towards groundwater resources. This work aimed at evaluating the migration of <span class="hlt">nitrate</span> originated on feedlots effluents in a fine-textured soil under field conditions. Soil water constituents were measured during a three-year period at three distinct locations adjacent to feedlot retention lagoons representing different degrees of exposure to water flow and manure accumulation. A simple statistical analysis was undertaken to identify patterns of <span class="hlt">observed</span> <span class="hlt">nitrate</span> and chloride <span class="hlt">concentrations</span> and electrical conductivity and their differences with depth. HYDRUS-1D was used to simulate water flow and solute transport of Cl - , NO 4 + N, NO 3 - N and electrical conductivity to complement field data interpretation. Results indicated that patterns of NO 3 - N <span class="hlt">concentrations</span> were not only notoriously different from electrical conductivity and Cl - but also ranges and distribution with depth differed among locations. A combination of dilution, transport, reactions such as nitrification/denitrification and vegetation water and solute uptake took place at each plots denoting the complexity of soil-solution behavior under extreme polluting conditions. Simulations using the concept of single porosity-mobile/immobile water (SP-MIM) managed structural controls and correctly simulated - all species <span class="hlt">concentrations</span> under field data constrains. The opposite was true for the other two locations experiencing near-saturation conditions, absence of vegetation and frequent manure accumulation and runoff from feedlot lagoons. Although the results are site specific, findings are relevant to advance the understanding of NO 3 - N</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70010004','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70010004"><span>Temporal trends in <span class="hlt">nitrate</span> and selected pesticides in mid-atlantic ground water</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Debrewer, L.M.; Ator, S.W.; Denver, J.M.</p> <p>2008-01-01</p> <p>Evaluating long-term temporal trends in regional ground-water quality is complicated by variable hydrogeologic conditions and typically slow flow, and such trends have rarely been directly measured. Ground-water samples were collected over near-decadal and annual intervals from unconfined aquifers in agricultural areas of the Mid-Atlantic region, including fractured carbonate rocks in the Great Valley, Potomac River Basin, and unconsolidated sediments on the Delmarva Peninsula. <span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> and selected pesticides and degradates were compared among sampling events and to apparent recharge dates. <span class="hlt">Observed</span> temporal trends are related to changes in land use and chemical applications, and to hydrogeology and climate. Insignificant differences in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the Great Valley between 1993 and 2002 are consistent with relatively steady fertilizer application during respective recharge periods and are likely related to drought conditions in the later sampling period. Detecting trends in Great Valley ground water is complicated by long open boreholes characteristic of wells sampled in this setting which facilitate significant ground-water mixing. Decreasing atrazine and prometon <span class="hlt">concentrations</span>, however, reflect reported changes in usage. On the Delmarva Peninsula between 1988 and 2001, median <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increased 2 mg per liter in aerobic ground water, reflecting increasing fertilizer applications. Correlations between selected pesticide compounds and apparent recharge date are similarly related to changing land use and chemical application. <span class="hlt">Observed</span> trends in the two settings demonstrate the importance of considering hydrogeology and recharge date along with, changing land and chemical uses when interpreting trends in regional ground-water quality. Copyright ?? 2008 by the American Society of Agronomy, Crop Science Society of America, and Soil Science Society of America. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.acs.org/doi/abs/10.1021/es201221s','USGSPUBS'); return false;" href="http://pubs.acs.org/doi/abs/10.1021/es201221s"><span><span class="hlt">Nitrate</span> in the Mississippi River and its tributaries, 1980 to 2008: Are we making progress?</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Sprague, Lori A.; Hirsch, Robert M.; Aulenbach, Brent T.</p> <p>2011-01-01</p> <p>Changes in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and flux between 1980 and 2008 at eight sites in the Mississippi River basin were determined using a new statistical method that accommodates evolving <span class="hlt">nitrate</span> behavior over time and produces flow-normalized estimates of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and flux that are independent of random variations in streamflow. The results show that little consistent progress has been made in reducing riverine <span class="hlt">nitrate</span> since 1980, and that flow-normalized <span class="hlt">concentration</span> and flux are increasing in some areas. Flow-normalized <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and flux increased between 9 and 76% at four sites on the Mississippi River and a tributary site on the Missouri River, but changed very little at tributary sites on the Ohio, Iowa, and Illinois Rivers. Increases in flow-normalized <span class="hlt">concentration</span> and flux at the Mississippi River at Clinton and Missouri River at Hermann were more than three times larger than at any other site. The increases at these two sites contributed much of the 9% increase in flow-normalized <span class="hlt">nitrate</span> flux leaving the Mississippi River basin. At most sites, <span class="hlt">concentrations</span> increased more at low and moderate streamflows than at high streamflows, suggesting that increasing groundwater <span class="hlt">concentrations</span> are having an effect on river <span class="hlt">concentrations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.B13G1832J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.B13G1832J"><span>Comparative Analysis of <span class="hlt">Nitrate</span> Levels in Pensacola Area Rain Water</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jacobs, J.; Caffrey, J. M.; Maestre, A.; Landing, W. M.</p> <p>2017-12-01</p> <p><span class="hlt">Nitrate</span> is an important constituent of acid rain and often correlated with atmospheric NOx levels. This link between air and water quality was tested over a course of summer 2017 and compared to data from 2005-2012. Rain water samples collected from late May through early July of 2017 were tested for pH and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. These months were among the stormiest on record for the Northwest Florida region with a total rainfall of 648 mm. The data analyzed from these rain events was compared to previous data to show the trends of <span class="hlt">nitrate</span> and pH levels in the rainwater. Median pH for this study was 5.2, higher than the medians between 2015-2012 which ranged from 4.2 to 5.0, while <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> for this study were 15.2 µM. This contrasts with a significant drop in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> from 41 µM in 2005 and 2006 to around 12 µM between 2007 and 2012. The drop between 2006-7 was suspected to be a result of implementation of NOx controls at Plant Crist coal fired power plant and other Clean Air Act requirements. These inputs of <span class="hlt">nitrate</span> and H+ ions from rainwater can have a significant influence water quality throughout the region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21284299','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21284299"><span>Multiscale effects of management, environmental conditions, and land use on <span class="hlt">nitrate</span> leaching in dairy farms.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Oenema, Jouke; Burgers, Saskia; Verloop, Koos; Hooijboer, Arno; Boumans, Leo; ten Berge, Hein</p> <p>2010-01-01</p> <p><span class="hlt">Nitrate</span> leaching in intensive grassland- and silage maize-based dairy farming systems on sandy soil is a main environmental concern. Here, statistical relationships are presented between management practices and environmental conditions and <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in shallow groundwater (0.8 m depth) at farm, field, and point scales in The Netherlands, based on data collected in a participatory approach over a 7-yr period at one experimental and eight pilot commercial dairy farms on sandy soil. Farm milk production ranged from 10 to 24 Mg ha(-1). Soil and hydrological characteristics were derived from surveys and weather conditions from meteorological stations. Statistical analyses were performed with multiple regression models. Mean <span class="hlt">nitrate</span> <span class="hlt">concentration</span> at farm scale decreased from 79 mg L(-1) in 1999 to 63 in 2006, with average <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater decreasing under grassland but increasing under maize land over the monitoring period. The effects of management practices on <span class="hlt">nitrate</span> <span class="hlt">concentration</span> varied with spatial scale. At farm scale, nitrogen surplus, grazing intensity, and the relative areas of grassland and maize land significantly contributed to explaining the variance in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater. Mean <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was negatively correlated to the <span class="hlt">concentration</span> of dissolved organic carbon in the shallow groundwater. At field scale, management practices and soil, hydrological, and climatic conditions significantly contributed to explaining the variance in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater under grassland and maize land. We conclude that, on these intensive dairy farms, additional measures are needed to comply with the European Union water quality standard in groundwater of 50 mg <span class="hlt">nitrate</span> L(-1). The most promising measures are omitting fertilization of catch crops and reducing fertilization levels of first-year maize in the rotation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3241578','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3241578"><span>Totomatix: a novel automatic set-up to control diurnal, diel and long-term plant <span class="hlt">nitrate</span> nutrition</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Adamowicz, Stéphane; Le Bot, Jacques; Huanosto Magaña, Ruth; Fabre, José</p> <p>2012-01-01</p> <p>Background Stand-alone nutritional set-ups are useful tools to grow plants at defined nutrient availabilities and to measure nutrient uptake rates continuously, in particular that for <span class="hlt">nitrate</span>. Their use is essential when the measurements are meant to cover long time periods. These complex systems have, however, important drawbacks, including poor long-term reliability and low precision at high <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. This explains why the information dealing with diel dynamics of <span class="hlt">nitrate</span> uptake rate is scarce and concerns mainly young plants grown at low <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. Scope The novel system detailed in this paper has been developed to allow versatile use in growth rooms, greenhouses or open fields at <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> ranging from a few micro- to several millimoles per litres. The system controls, at set frequencies, the solution <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, pH and volumes. <span class="hlt">Nitrate</span> <span class="hlt">concentration</span> is measured by spectral deconvolution of UV spectra. The main advantages of the set-up are its low maintenance (weekly basis), an ability to diagnose interference or erroneous analyses and high precision of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> measurements (0·025 % at 3 mm). The paper details the precision of diurnal <span class="hlt">nitrate</span> uptake rate measurements, which reveals sensitivity to solution volume at low <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, whereas at high <span class="hlt">concentration</span>, it is mostly sensitive to the precision of volume estimates. Conclusions This novel set-up allows us to measure and characterize the dynamics of plant <span class="hlt">nitrate</span> nutrition at high temporal resolution (minutes to hours) over long-term experiments (up to 1 year). It is reliable and also offers a novel method to regulate up to seven N treatments by adjusting the daily uptake of test plants relative to controls, in variable environments such as open fields and glasshouses. PMID:21985796</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26995268','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26995268"><span>Chronic <span class="hlt">nitrate</span> enrichment decreases severity and induces protection against an infectious disease.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Smallbone, Willow; Cable, Jo; Maceda-Veiga, Alberto</p> <p>2016-05-01</p> <p>Excessive fertilisation is one of the most pernicious forms of global change resulting in eutrophication. It has major implications for disease control and the conservation of biodiversity. Yet, the direct link between nutrient enrichment and disease remains largely unexplored. Here, we present the first experimental evidence that chronic <span class="hlt">nitrate</span> enrichment decreases severity and induces protection against an infectious disease. Specifically, this study shows that <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> ranging between 50 and 250mgNO3(-)/l reduce Gyrodactylus turnbulli infection intensity in two populations of Trinidadian guppies Poecilia reticulata, and that the highest <span class="hlt">nitrate</span> <span class="hlt">concentration</span> can even clean the parasites from the fish. This added to the fact that host <span class="hlt">nitrate</span> pre-exposure altered the fish epidermal structure and reduced parasite intensity, suggests that <span class="hlt">nitrate</span> protected the host against the disease. <span class="hlt">Nitrate</span> treatments also caused fish mortality. As we used ecologically-relevant <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, and guppies are top-consumers widely used for mosquito bio-control in tropical and often nutrient-enriched waters, our results can have major ecological and social implications. In conclusion, this study advocates reducing <span class="hlt">nitrate</span> level including the legislative threshold to protect the aquatic biota, even though this may control an ectoparasitic disease. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20585108','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20585108"><span>Inorganic <span class="hlt">nitrate</span> supplementation lowers blood pressure in humans: role for nitrite-derived NO.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kapil, Vikas; Milsom, Alexandra B; Okorie, Michael; Maleki-Toyserkani, Sheiva; Akram, Farihah; Rehman, Farkhanda; Arghandawi, Shah; Pearl, Vanessa; Benjamin, Nigel; Loukogeorgakis, Stavros; Macallister, Raymond; Hobbs, Adrian J; Webb, Andrew J; Ahluwalia, Amrita</p> <p>2010-08-01</p> <p>Ingestion of dietary (inorganic) <span class="hlt">nitrate</span> elevates circulating and tissue levels of nitrite via bioconversion in the entero-salivary circulation. In addition, nitrite is a potent vasodilator in humans, an effect thought to underlie the blood pressure-lowering effects of dietary <span class="hlt">nitrate</span> (in the form of beetroot juice) ingestion. Whether inorganic <span class="hlt">nitrate</span> underlies these effects and whether the effects of either naturally occurring dietary <span class="hlt">nitrate</span> or inorganic <span class="hlt">nitrate</span> supplementation are dose dependent remain uncertain. Using a randomized crossover study design, we show that <span class="hlt">nitrate</span> supplementation (KNO(3) capsules: 4 versus 12 mmol [n=6] or 24 mmol of KNO(3) (1488 mg of <span class="hlt">nitrate</span>) versus 24 mmol of KCl [n=20]) or vegetable intake (250 mL of beetroot juice [5.5 mmol <span class="hlt">nitrate</span>] versus 250 mL of water [n=9]) causes dose-dependent elevation in plasma nitrite <span class="hlt">concentration</span> and elevation of cGMP <span class="hlt">concentration</span> with a consequent decrease in blood pressure in healthy volunteers. In addition, post hoc analysis demonstrates a sex difference in sensitivity to <span class="hlt">nitrate</span> supplementation dependent on resting baseline blood pressure and plasma nitrite <span class="hlt">concentration</span>, whereby blood pressure is decreased in male volunteers, with higher baseline blood pressure and lower plasma nitrite <span class="hlt">concentration</span> but not in female volunteers. Our findings demonstrate dose-dependent decreases in blood pressure and vasoprotection after inorganic <span class="hlt">nitrate</span> ingestion in the form of either supplementation or by dietary elevation. In addition, our post hoc analyses intimate sex differences in <span class="hlt">nitrate</span> processing involving the entero-salivary circulation that are likely to be major contributing factors to the lower blood pressures and the vasoprotective phenotype of premenopausal women.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28013474','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28013474"><span>Use of continuous monitoring to assess stream <span class="hlt">nitrate</span> flux and transformation patterns.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jones, Christopher; Kim, Sea-Won; Schilling, Keith</p> <p>2017-01-01</p> <p>Delivery of nitrogen from farmed fields to the stream network is an ongoing water quality issue in central North America and other parts of the world. Although fertilization and other farming practices have been refined to produce environmental improvements, stemming loss of nitrogen, especially in the soluble <span class="hlt">nitrate</span> form, is a problem that has seemingly defied solution. The Iowa Nutrient Reduction Strategy is a policy initiative designed to implement conservation and other farm management practices to produce reductions in <span class="hlt">nitrate</span> loading. The strategy does not focus on how the streams themselves may or may not be processing nitrogen and reducing downstream loading. We used continuous high-frequency <span class="hlt">nitrate</span> and discharge monitoring over 3 years at two sites separated by 18 km in a low-order, agricultural stream in eastern Iowa to estimate how nitrogen is processed, and whether or not these processes are reducing downstream loading. We conclude that the upstream to downstream <span class="hlt">nitrate</span> <span class="hlt">concentration</span> decline between the two sites was not driven by denitrification. These data also show that <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are closely coupled to discharge during periods of adequate moisture, but decoupling of <span class="hlt">concentration</span> from discharge occurs during dry periods. This decoupling is a possible indicator of in-stream <span class="hlt">nitrate</span> processing. Finally, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> are likely diluted by water sourced from non-row crop land covers in the lower reaches of the watershed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1130835','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1130835"><span><span class="hlt">Nitrate</span> transport in the cyanobacterium Anacystis nidulans R2. Kinetic and energetic aspects.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Rodríguez, R; Lara, C; Guerrero, M G</p> <p>1992-01-01</p> <p><span class="hlt">Nitrate</span> transport has been studied in the cyanobacterium Anacystis nidulans R2 by monitoring intracellular <span class="hlt">nitrate</span> accumulation in intact cells of the mutant strain FM6, which lacks <span class="hlt">nitrate</span> reductase activity and is therefore unable to reduce the transported <span class="hlt">nitrate</span>. Kinetic analysis of <span class="hlt">nitrate</span> transport as a function of external <span class="hlt">nitrate</span> <span class="hlt">concentration</span> revealed apparent substrate inhibition, with a peak velocity at 20-25 microM-<span class="hlt">nitrate</span>. A Ks (NO3-) of 1 microM was calculated. <span class="hlt">Nitrate</span> transport exhibited a stringent requirement for Na+. Neither Li+ nor K+ could substitute for Na+. Monensin depressed <span class="hlt">nitrate</span> transport in a <span class="hlt">concentration</span>-dependent manner, inhibition being more than 60% at 2 microM, indicating that the Na(+)-dependence of active <span class="hlt">nitrate</span> transport relies on the maintenance of a Na+ electrochemical gradient. The operation of an Na+/NO3- symport system is suggested. Nitrite behaved as an effective competitive inhibitor of <span class="hlt">nitrate</span> transport, with a Ki (NO2-) of 3 microM. The time course of nitrite inhibition of <span class="hlt">nitrate</span> transport was consistent with competitive inhibition by mixed alternative substrates. <span class="hlt">Nitrate</span> and nitrite might be transported by the same carrier. PMID:1554347</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25237352','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25237352"><span>Synergistic Activity of Econazole-<span class="hlt">Nitrate</span> and Chelerythrine against Clinical Isolates of Candida albicans.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chen, Zhibao; Li, Xinran; Wu, Xiuping; Wang, Wei; Wang, Wendong; Xin, Mingxun; Shen, Fengge; Liu, Lihui; Liang, Junchao; Li, Lei; Yu, Lu</p> <p>2014-01-01</p> <p>The aim of this investigation was to assess the in-vitro interaction of two antifungal agents, econazole-<span class="hlt">nitrate</span> and chelerythrine, against ten fluconazole-resistant clinical isolates and one ATCC type strain 10231 of Candida albicans. The checkerboard microdilution method was performed according to the recommendations of the National Committee for Clinical Laboratory Standards, and the results were determined by visual examination. The interaction intensity was tested in all isolates using the fractional inhibitory <span class="hlt">concentration</span> index (FICI). These experiments showed synergism between econazole-<span class="hlt">nitrate</span> and chelerythrine in antifungal activity against C. albicans, and no antagonistic activity was <span class="hlt">observed</span> in any of the strains tested. Moreover, time-kill curves were performed with selected strains to confirm the positive interactions. The similarity between the results of the FICI values and the time-kill curves revealed that chelerythrine greatly enhances the antifungal effects of econazole-<span class="hlt">nitrate</span> against isolates of C. albicans. This synergistic effect may markedly reduce the dose of econazole-<span class="hlt">nitrate</span> required to treat candidiasis, thereby decreasing the econazole-<span class="hlt">nitrate</span> toxic side effects. This novel synergism might provide a potential combination treatment against fungal infections.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1911b0012K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1911b0012K"><span>Verification of spectrophotometric method for <span class="hlt">nitrate</span> analysis in water samples</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kurniawati, Puji; Gusrianti, Reny; Dwisiwi, Bledug Bernanti; Purbaningtias, Tri Esti; Wiyantoko, Bayu</p> <p>2017-12-01</p> <p>The aim of this research was to verify the spectrophotometric method to analyze <span class="hlt">nitrate</span> in water samples using APHA 2012 Section 4500 NO3-B method. The verification parameters used were: linearity, method detection limit, level of quantitation, level of linearity, accuracy and precision. Linearity was obtained by using 0 to 50 mg/L <span class="hlt">nitrate</span> standard solution and the correlation coefficient of standard calibration linear regression equation was 0.9981. The method detection limit (MDL) was defined as 0,1294 mg/L and limit of quantitation (LOQ) was 0,4117 mg/L. The result of a level of linearity (LOL) was 50 mg/L and <span class="hlt">nitrate</span> <span class="hlt">concentration</span> 10 to 50 mg/L was linear with a level of confidence was 99%. The accuracy was determined through recovery value was 109.1907%. The precision value was <span class="hlt">observed</span> using % relative standard deviation (%RSD) from repeatability and its result was 1.0886%. The tested performance criteria showed that the methodology was verified under the laboratory conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11572282','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11572282"><span>Recurrent diarrhea in children living in areas with high levels of <span class="hlt">nitrate</span> in drinking water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gupta, S K; Gupta, R C; Gupta, A B; Seth, A K; Bassin, J K; Gupta, A; Sharma, M L</p> <p>2001-01-01</p> <p>Given that there was documented evidence of an association between diarrhea and high <span class="hlt">nitrate</span> ingestion, the authors examined drinking water <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and its possible correlation(s) with methemoglobin levels, cytochrome b5 reductase activity, and recurrent diarrhea. In addition, the authors studied histopathological changes in the intestines of rabbits in an animal model. Five village areas were studied, and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> (expressed in mg of <span class="hlt">nitrate</span> per liter of water) of 26, 45, 95, 220, and 459 existed in the respective villages. The study included 88 randomly selected children who were 8 yr of age or younger; they represented 10% of the total population of each of the areas. Detailed histories of recurrent diarrhea were noted, and medical examinations were conducted. Cytochrome b5 reductase activity and methemoglobin levels were estimated biochemically. Collected data were analyzed statistically with Microsoft Excel software. In addition, the authors exposed rabbits to various levels of <span class="hlt">nitrate</span>, and histopathological changes of the stomach and intestine (small and large) were evaluated. There was a strong relationship between <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and recurrent diarrhea; 80% of the recurrent diarrhea cases were explained by <span class="hlt">nitrate</span> <span class="hlt">concentration</span> alone. In the rabbit intestines, lymphocytic infiltration and hyperplasia characterized the submucosa as <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increased.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70170058','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70170058"><span>Changes between early development (1930–60) and recent (2005–15) groundwater-level altitudes and dissolved-solids and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> In and near Gaines, Terry, and Yoakum Counties, Texas</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thomas, Jonathan V.; Teeple, Andrew; Payne, Jason; Ikard, Scott</p> <p>2016-06-21</p> <p>During the recent period, median dissolved-solids <span class="hlt">concentrations</span> of less than 1,000 milligrams per liter (mg/L) were predominantly measured in the western part of the study area, and median <span class="hlt">concentrations</span> of more than 1,000 mg/L were predominantly measured in the eastern part of the study area. A general pattern of increasing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> from west to the northeast was evident in the study area. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> measured in samples collected from 16 wells completed in the Ogallala aquifer for the recent period were equal to or greater than 10 mg/L, the primary drinking water standard for finished drinking water.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21282331','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21282331"><span>Regulation of <span class="hlt">nitrate</span> assimilation in cyanobacteria.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ohashi, Yoshitake; Shi, Wei; Takatani, Nobuyuki; Aichi, Makiko; Maeda, Shin-ichi; Watanabe, Satoru; Yoshikawa, Hirofumi; Omata, Tatsuo</p> <p>2011-02-01</p> <p><span class="hlt">Nitrate</span> assimilation by cyanobacteria is inhibited by the presence of ammonium in the growth medium. Both <span class="hlt">nitrate</span> uptake and transcription of the <span class="hlt">nitrate</span> assimilatory genes are regulated. The major intracellular signal for the regulation is, however, not ammonium or glutamine, but 2-oxoglutarate (2-OG), whose <span class="hlt">concentration</span> changes according to the change in cellular C/N balance. When nitrogen is limiting growth, accumulation of 2-OG activates the transcription factor NtcA to induce transcription of the <span class="hlt">nitrate</span> assimilation genes. Ammonium inhibits transcription by quickly depleting the 2-OG pool through its metabolism via the glutamine synthetase/glutamate synthase cycle. The P(II) protein inhibits the ABC-type <span class="hlt">nitrate</span> transporter, and also <span class="hlt">nitrate</span> reductase in some strains, by an unknown mechanism(s) when the cellular 2-OG level is low. Upon nitrogen limitation, 2-OG binds to P(II) to prevent the protein from inhibiting <span class="hlt">nitrate</span> assimilation. A pathway-specific transcriptional regulator NtcB activates the <span class="hlt">nitrate</span> assimilation genes in response to nitrite, either added to the medium or generated intracellularly by <span class="hlt">nitrate</span> reduction. It plays an important role in selective activation of the <span class="hlt">nitrate</span> assimilation pathway during growth under a limited supply of <span class="hlt">nitrate</span>. P(II) was recently shown to regulate the activity of NtcA negatively by binding to PipX, a small coactivator protein of NtcA. On the basis of accumulating genome information from a variety of cyanobacteria and the molecular genetic data obtained from the representative strains, common features and group- or species-specific characteristics of the response of cyanobacteria to nitrogen is summarized and discussed in terms of ecophysiological significance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A23B2337C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A23B2337C"><span>Simulations of Sulfate-<span class="hlt">Nitrate</span>-Ammonium (SNA) aerosols during the extreme haze events over Northern China in October 2014</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, D.; Liu, Z.; Fast, J. D.; Ban, J.</p> <p>2017-12-01</p> <p>Extreme haze events have occurred frequently over China in recent years. Although many studies have investigated the formation mechanisms associated with PM2.5 for heavily polluted regions in China based on <span class="hlt">observational</span> data, adequately predicting peak PM2.5 <span class="hlt">concentrations</span> is still challenging for regional air quality models. In this study, we evaluate the performance of one configuration of the Weather Research and Forecasting model coupled with chemistry (WRF-Chem) and use the model to investigate the sensitivity of heterogeneous reactions on simulated peak sulfate, <span class="hlt">nitrate</span>, and ammonium <span class="hlt">concentrations</span> in the vicinity of Beijing during four extreme haze episodes in October 2014 over the North China Plain. The highest <span class="hlt">observed</span> PM2.5 <span class="hlt">concentration</span> of 469 μg m-3 occurred in Beijing. Comparisons with <span class="hlt">observations</span> show that the model reproduced the temporal variability in PM2.5 with the highest PM2.5 values on polluted days (defined as days in which <span class="hlt">observed</span> PM2.5 is greater than 75 μg m-3), but predictions of sulfate, <span class="hlt">nitrate</span>, and ammonium were too low on days with the highest <span class="hlt">observed</span> <span class="hlt">concentrations</span>. <span class="hlt">Observational</span> data indicate that the sulfur/nitric oxidation rates are strongly correlated with relative humidity during periods of peak PM2.5; however, the model failed to reproduce the highest PM2.5 <span class="hlt">concentrations</span> due to missing heterogeneous/aqueous reactions. As the parameterizations of those heterogeneous reactions are not well established yet, estimates of SO2-to-H2SO4 and NO2/NO3-to-HNO3 reaction rates that depend on relative humidity were applied which improved the simulation of sulfate, <span class="hlt">nitrate</span>, and ammonium enhancement on polluted days in terms of both <span class="hlt">concentrations</span> and partitioning among those species. Sensitivity simulations showed that the extremely high heterogeneous reaction rates and also higher emission rates than those reported in the emission inventory were likely important factors contributing to those peak PM2.5 <span class="hlt">concentrations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70032658','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70032658"><span>Verifiable metamodels for <span class="hlt">nitrate</span> losses to drains and groundwater in the Corn Belt, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Nolan, Bernard T.; Malone, Robert W.; Gronberg, Jo Ann M.; Thorp, K.R.; Ma, Liwang</p> <p>2012-01-01</p> <p><span class="hlt">Nitrate</span> leaching in the unsaturated zone poses a risk to groundwater, whereas <span class="hlt">nitrate</span> in tile drainage is conveyed directly to streams. We developed metamodels (MMs) consisting of artificial neural networks to simplify and upscale mechanistic fate and transport models for prediction of <span class="hlt">nitrate</span> losses by drains and leaching in the Corn Belt, USA. The two final MMs predicted <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and flux, respectively, in the shallow subsurface. Because each MM considered both tile drainage and leaching, they represent an integrated approach to vulnerability assessment. The MMs used readily available data comprising farm fertilizer nitrogen (N), weather data, and soil properties as inputs; therefore, they were well suited for regional extrapolation. The MMs effectively related the outputs of the underlying mechanistic model (Root Zone Water Quality Model) to the inputs (R2 = 0.986 for the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> MM). Predicted <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was compared with measured <span class="hlt">nitrate</span> in 38 samples of recently recharged groundwater, yielding a Pearson’s r of 0.466 (p = 0.003). Predicted <span class="hlt">nitrate</span> generally was higher than that measured in groundwater, possibly as a result of the time-lag for modern recharge to reach well screens, denitrification in groundwater, or interception of recharge by tile drains. In a qualitative comparison, predicted <span class="hlt">nitrate</span> <span class="hlt">concentration</span> also compared favorably with results from a previous regression model that predicted total N in streams.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=292398','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=292398"><span>Effects of dietary protein <span class="hlt">concentration</span> on ammonia volatilization, <span class="hlt">nitrate</span> leaching, and plant nitrogen uptake from dairy manure applied to lysimeters</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>This lysimeter experiment was designed to investigate the effects of dietary crude protein (CP) <span class="hlt">concentration</span> on <span class="hlt">nitrate</span>-N (NO3-N) and ammonia (NH3) losses from dairy manure applied to soil and manure N use for plant growth. Lactating dairy cows were fed diets with 16.7 (HighCP) or 14.8% (LowCP) cru...</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.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=144217','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=144217"><span>CHL1 is a dual-affinity <span class="hlt">nitrate</span> transporter of Arabidopsis involved in multiple phases of <span class="hlt">nitrate</span> uptake.</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Liu, K H; Huang, C Y; Tsay, Y F</p> <p>1999-01-01</p> <p>Higher plants have both high- and low-affinity <span class="hlt">nitrate</span> uptake systems. These systems are generally thought to be genetically distinct. Here, we demonstrate that a well-known low-affinity <span class="hlt">nitrate</span> uptake mutant of Arabidopsis, chl1, is also defective in high-affinity <span class="hlt">nitrate</span> uptake. Two to 3 hr after <span class="hlt">nitrate</span> induction, uptake activities of various chl1 mutants at 250 microM <span class="hlt">nitrate</span> (a high-affinity <span class="hlt">concentration</span>) were only 18 to 30% of those of wild-type plants. In these mutants, both the inducible phase and the constitutive phase of high-affinity <span class="hlt">nitrate</span> uptake activities were reduced, with the inducible phase being severely reduced. Expressing a CHL1 cDNA driven by the cauliflower mosaic virus 35S promoter in a transgenic chl1 plant effectively recovered the defect in high-affinity uptake for the constitutive phase but not for the induced phase, which is consistent with the constitutive level of CHL1 expression in the transgenic plant. Kinetic analysis of <span class="hlt">nitrate</span> uptake by CHL1-injected Xenopus oocytes displayed a biphasic pattern with a Michaelis-Menten Km value of approximately 50 microM for the high-affinity phase and approximately 4 mM for the low-affinity phase. These results indicate that in addition to being a low-affinity <span class="hlt">nitrate</span> transporter, as previously recognized, CHL1 is also involved in both the inducible and constitutive phases of high-affinity <span class="hlt">nitrate</span> uptake in Arabidopsis. PMID:10330471</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/864729','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/864729"><span>Method for improved decomposition of metal <span class="hlt">nitrate</span> solutions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Haas, Paul A.; Stines, William B.</p> <p>1983-10-11</p> <p>A method for co-conversion of aqueous solutions of one or more heavy metal <span class="hlt">nitrates</span> wherein thermal decomposition within a temperature range of about 300.degree. to 800.degree. C. is carried out in the presence of about 50 to 500% molar <span class="hlt">concentration</span> of ammonium <span class="hlt">nitrate</span> to total metal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=62022&keyword=copd&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=62022&keyword=copd&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>PERSONAL, INDOOR, AND OUTDOOR <span class="hlt">CONCENTRATIONS</span> OF PM2.5, PARTICULATE <span class="hlt">NITRATE</span>, AND ELEMENTAL CARBON FOR INDIVIDUALS WITH COPD IN LOS ANGELES, CA</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 study characterizes the personal, indoor, and outdoor <span class="hlt">concentrations</span> of PM2.5 and the major components of PM2.5, including <span class="hlt">nitrate</span> (NO3-), elemental carbon (EC), and the elements for individuals with chronic obstructive pulmonary disease (COPD) living in Los Angeles, CA. ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.water.usgs.gov/sir20045221/','USGSPUBS'); return false;" href="http://pubs.water.usgs.gov/sir20045221/"><span>Water quality and possible sources of <span class="hlt">nitrate</span> in the Cimarron Terrace Aquifer, Oklahoma, 2003</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Masoner, Jason R.; Mashburn, Shana L.</p> <p>2004-01-01</p> <p>Water from the Cimarron terrace aquifer in northwest Oklahoma commonly has <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> that exceed the maximum contaminant level of 10 milligrams per liter of nitrite plus <span class="hlt">nitrate</span> as nitrogen (referred to as <span class="hlt">nitrate</span>) set by the U.S. Environmental Protection Agency for public drinking water supplies. Starting in July 2003, the U.S. Geological Survey, in cooperation with the Oklahoma Department of Environmental Quality, conducted a study in the Cimarron terrace aquifer to assess the water quality and possible sources of <span class="hlt">nitrate</span>. A qualitative and quantitative approach based on multiple lines of evidence from chemical analysis of <span class="hlt">nitrate</span>, nitrogen isotopes in <span class="hlt">nitrate</span>, pesticides (indicative of cropland fertilizer application), and wastewater compounds (indicative of animal or human wastewater) were used to indicate possible sources of <span class="hlt">nitrate</span> in the Cimarron terrace aquifer. <span class="hlt">Nitrate</span> was detected in 44 of 45 ground-water samples and had the greatest median <span class="hlt">concentration</span> (8.03 milligrams per liter) of any nutrient analyzed. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> ranged from <0.06 to 31.8 milligrams per liter. Seventeen samples had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> exceeding the maximum contaminant level of 10 milligrams per liter. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in agricultural areas were significantly greater than <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in grassland areas. Pesticides were detected in 15 of 45 ground-water samples. Atrazine and deethylatrazine, a metabolite of atrazine, were detected most frequently. Deethylatrazine was detected in water samples from 9 wells and atrazine was detected in samples from 8 wells. Tebuthiuron was detected in water samples from 5 wells; metolachlor was detected in samples from 4 wells; prometon was detected in samples from 4 wells; and alachlor was detected in 1 well. None of the detected pesticide <span class="hlt">concentrations</span> exceeded the maximum contaminant level or health advisory level set by the U.S. Environmental Protection Agency. Wastewater compounds were detected in 28 of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018WRR....54.1312C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018WRR....54.1312C"><span>Contextualizing Wetlands Within a River Network to Assess <span class="hlt">Nitrate</span> Removal and Inform Watershed Management</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Czuba, Jonathan A.; Hansen, Amy T.; Foufoula-Georgiou, Efi; Finlay, Jacques C.</p> <p>2018-02-01</p> <p>Aquatic <span class="hlt">nitrate</span> removal depends on interactions throughout an interconnected network of lakes, wetlands, and river channels. Herein, we present a network-based model that quantifies <span class="hlt">nitrate</span>-nitrogen and organic carbon <span class="hlt">concentrations</span> through a wetland-river network and estimates <span class="hlt">nitrate</span> export from the watershed. This model dynamically accounts for multiple competing limitations on <span class="hlt">nitrate</span> removal, explicitly incorporates wetlands in the network, and captures hierarchical network effects and spatial interactions. We apply the model to the Le Sueur Basin, a data-rich 2,880 km2 agricultural landscape in southern Minnesota and validate the model using synoptic field measurements during June for years 2013-2015. Using the model, we show that the overall limits to <span class="hlt">nitrate</span> removal rate via denitrification shift between <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, organic carbon availability, and residence time depending on discharge, characteristics of the waterbody, and location in the network. Our model results show that the spatial context of wetland restorations is an important but often overlooked factor because nonlinearities in the system, e.g., deriving from switching of resource limitation on denitrification rate, can lead to unexpected changes in downstream biogeochemistry. Our results demonstrate that reduction of watershed-scale <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and downstream loads in the Le Sueur Basin can be most effectively achieved by increasing water residence time (by slowing the flow) rather than by increasing organic carbon <span class="hlt">concentrations</span> (which may limit denitrification). This framework can be used toward assessing where and how to restore wetlands for reducing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and loads from agricultural watersheds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1310926','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1310926"><span>Workgroup Report: Drinking-Water <span class="hlt">Nitrate</span> and Health—Recent Findings and Research Needs</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ward, Mary H.; deKok, Theo M.; Levallois, Patrick; Brender, Jean; Gulis, Gabriel; Nolan, Bernard T.; VanDerslice, James</p> <p>2005-01-01</p> <p>Human alteration of the nitrogen cycle has resulted in steadily accumulating <span class="hlt">nitrate</span> in our water resources. The U.S. maximum contaminant level and World Health Organization guidelines for <span class="hlt">nitrate</span> in drinking water were promulgated to protect infants from developing methemoglobinemia, an acute condition. Some scientists have recently suggested that the regulatory limit for <span class="hlt">nitrate</span> is overly conservative; however, they have not thoroughly considered chronic health outcomes. In August 2004, a symposium on drinking-water <span class="hlt">nitrate</span> and health was held at the International Society for Environmental Epidemiology meeting to evaluate <span class="hlt">nitrate</span> exposures and associated health effects in relation to the current regulatory limit. The contribution of drinking-water <span class="hlt">nitrate</span> toward endogenous formation of N-nitroso compounds was evaluated with a focus toward identifying subpopulations with increased rates of nitrosation. Adverse health effects may be the result of a complex interaction of the amount of <span class="hlt">nitrate</span> ingested, the concomitant ingestion of nitrosation cofactors and precursors, and specific medical conditions that increase nitrosation. Workshop participants concluded that more experimental studies are needed and that a particularly fruitful approach may be to conduct epidemiologic studies among susceptible subgroups with increased endogenous nitrosation. The few epidemiologic studies that have evaluated intake of nitrosation precursors and/or nitrosation inhibitors have <span class="hlt">observed</span> elevated risks for colon cancer and neural tube defects associated with drinking-water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> below the regulatory limit. The role of drinking-water <span class="hlt">nitrate</span> exposure as a risk factor for specific cancers, reproductive outcomes, and other chronic health effects must be studied more thoroughly before changes to the regulatory level for <span class="hlt">nitrate</span> in drinking water can be considered. PMID:16263519</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70027698','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70027698"><span>Workgroup report: Drinking-water <span class="hlt">nitrate</span> and health - Recent findings and research needs</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ward, M.H.; deKok, T.M.; Levallois, P.; Brender, J.; Gulis, G.; Nolan, B.T.; VanDerslice, J.</p> <p>2005-01-01</p> <p>Human alteration of the nitrogen cycle has resulted in steadily accumulating <span class="hlt">nitrate</span> in our water resources. The U.S. maximum contaminant level and World Health Organization guidelines for <span class="hlt">nitrate</span> in drinking water were promulgated to protect infants from developing methemoglobinemia, an acute condition. Some scientists have recently suggested that the regulatory limit for <span class="hlt">nitrate</span> is overly conservative; however, they have not thoroughly considered chronic health outcomes. In August 2004, a symposium on drinking-water <span class="hlt">nitrate</span> and health was held at the International Society for Environmental Epidemiology meeting to evaluate <span class="hlt">nitrate</span> exposures and associated health effects in relation to the current regulatory limit. The contribution of drinking-water <span class="hlt">nitrate</span> toward endogenous formation of N-nitroso compounds was evaluated with a focus toward identifying subpopulations with increased rates of nitrosation. Adverse health effects may be the result of a complex interaction of the amount of <span class="hlt">nitrate</span> ingested, the concomitant ingestion of nitrosation cofactors and precursors, and specific medical conditions that increase nitrosation. Workshop participants concluded that more experimental studies are needed and that a particularly fruitful approach may be to conduct epidemiologic studies among susceptible subgroups with increased endogenous nitrosation. The few epidemiologic studies that have evaluated intake of nitrosation precursors and/or nitrosation inhibitors have <span class="hlt">observed</span> elevated risks for colon cancer and neural tube defects associated with drinking-water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> below the regulatory limit. The role of drinking-water <span class="hlt">nitrate</span> exposure as a risk factor for specific cancers, reproductive outcomes, and other chronic health effects must be studied more thoroughly before changes to the regulatory level for <span class="hlt">nitrate</span> in drinking water can be considered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16263519','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16263519"><span>Workgroup report: Drinking-water <span class="hlt">nitrate</span> and health--recent findings and research needs.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ward, Mary H; deKok, Theo M; Levallois, Patrick; Brender, Jean; Gulis, Gabriel; Nolan, Bernard T; VanDerslice, James</p> <p>2005-11-01</p> <p>Human alteration of the nitrogen cycle has resulted in steadily accumulating <span class="hlt">nitrate</span> in our water resources. The U.S. maximum contaminant level and World Health Organization guidelines for <span class="hlt">nitrate</span> in drinking water were promulgated to protect infants from developing methemoglobinemia, an acute condition. Some scientists have recently suggested that the regulatory limit for <span class="hlt">nitrate</span> is overly conservative; however, they have not thoroughly considered chronic health outcomes. In August 2004, a symposium on drinking-water <span class="hlt">nitrate</span> and health was held at the International Society for Environmental Epidemiology meeting to evaluate <span class="hlt">nitrate</span> exposures and associated health effects in relation to the current regulatory limit. The contribution of drinking-water <span class="hlt">nitrate</span> toward endogenous formation of N-nitroso compounds was evaluated with a focus toward identifying subpopulations with increased rates of nitrosation. Adverse health effects may be the result of a complex interaction of the amount of <span class="hlt">nitrate</span> ingested, the concomitant ingestion of nitrosation cofactors and precursors, and specific medical conditions that increase nitrosation. Workshop participants concluded that more experimental studies are needed and that a particularly fruitful approach may be to conduct epidemiologic studies among susceptible subgroups with increased endogenous nitrosation. The few epidemiologic studies that have evaluated intake of nitrosation precursors and/or nitrosation inhibitors have <span class="hlt">observed</span> elevated risks for colon cancer and neural tube defects associated with drinking-water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> below the regulatory limit. The role of drinking-water <span class="hlt">nitrate</span> exposure as a risk factor for specific cancers, reproductive outcomes, and other chronic health effects must be studied more thoroughly before changes to the regulatory level for <span class="hlt">nitrate</span> in drinking water can be considered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/10795767','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/10795767"><span>Estimation of <span class="hlt">nitrate</span> metabolism in intestinal tract by measuring breath nitrous oxide <span class="hlt">concentration</span> in Chinese and Japanese.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mitsui, T; Kato, N; Kondo, T</p> <p>2000-05-01</p> <p>We have previously reported that ingestion of vegetables containing high <span class="hlt">nitrate</span> (NO3-) increases breath nitrous oxide (N2O) <span class="hlt">concentration</span>, probably due to denitrification. In the present study, we estimated NO3- metabolism in the intestine by determining exhaled breath N2O <span class="hlt">concentration</span> after the ingestion of vegetables by 16 healthy Chinese and Japanese. Breath samples were collected at 15-min intervals for 4 hr after subjects ingested 180 g of vegetable juice or 50 g of lettuce. Breath N2O <span class="hlt">concentration</span> was measured by an IR-PAS analyzer. Lettuce but not vegetable juice increased N2O <span class="hlt">concentrations</span> similarly in Japanese and Chinese. In control subjects who ingested nothing, there were significant differences between Chinese and Japanese in peak N2O <span class="hlt">concentrations</span> (334 +/- 91 vs 140 +/- 24 ppb, P = 0.027) and total excretions (535 +/- 143 vs 214 +/- 36 microg, P = 0.036). Although the reason for this difference is unclear, Chinese subjects could produce breath N2O from other metabolic pathways than denitrification of dietary NO3.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.agu.org/pubs/crossref/pip/2012JG001990.shtml','USGSPUBS'); return false;" href="http://www.agu.org/pubs/crossref/pip/2012JG001990.shtml"><span><span class="hlt">Nitrate</span> removal in deep sediments of a nitrogen-rich river network: A test of a conceptual model</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Stelzer, Robert S.; Bartsch, Lynn</p> <p>2012-01-01</p> <p>Many estimates of nitrogen removal in streams and watersheds do not include or account for <span class="hlt">nitrate</span> removal in deep sediments, particularly in gaining streams. We developed and tested a conceptual model for <span class="hlt">nitrate</span> removal in deep sediments in a nitrogen-rich river network. The model predicts that oxic, <span class="hlt">nitrate</span>-rich groundwater will become depleted in <span class="hlt">nitrate</span> as groundwater upwelling through sediments encounters a zone that contains buried particulate organic carbon, which promotes redox conditions favorable for <span class="hlt">nitrate</span> removal. We tested the model at eight sites in upwelling reaches of lotic ecosystems in the Waupaca River Watershed that varied by three orders of magnitude in groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. We measured denitrification potential in sediment core sections to 30 cm and developed vertical <span class="hlt">nitrate</span> profiles to a depth of about 1 m with peepers and piezometer nests. Denitrification potential was higher, on average, in shallower core sections. However, core sections deeper than 5 cm accounted for 70%, on average, of the depth-integrated denitrification potential. Denitrification potential increased linearly with groundwater <span class="hlt">nitrate</span> <span class="hlt">concentration</span> up to 2 mg NO3-N/L but the relationship broke down at higher <span class="hlt">concentrations</span> (> 5 mg NO3-N/L), a pattern that suggests <span class="hlt">nitrate</span> saturation. At most sites groundwater <span class="hlt">nitrate</span> declined from high <span class="hlt">concentrations</span> at depth to much lower <span class="hlt">concentrations</span> prior to discharge into the surface water. The profiles suggested that <span class="hlt">nitrate</span> removal occurred at sediment depths between 20 and 40 cm. Dissolved oxygen <span class="hlt">concentrations</span> were much higher in deep sediments than in pore water at 5 cm sediment depth at most locations. The substantial denitrification potential in deep sediments coupled with the declines in <span class="hlt">nitrate</span> and dissolved oxygen <span class="hlt">concentrations</span> in upwelling groundwater suggest that our conceptual model for <span class="hlt">nitrate</span> removal in deep sediments is applicable to this river network. Our results suggest that <span class="hlt">nitrate</span> removal rates</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29775933','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29775933"><span><span class="hlt">Observing</span> HNO3 release dependent upon metal complexes in malonic acid/<span class="hlt">nitrate</span> droplets.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shao, Xu; Wu, Feng-Min; Yang, Hui; Pang, Shu-Feng; Zhang, Yun-Hong</p> <p>2018-05-09</p> <p>Although the dicarboxylic acid has been reported to react with <span class="hlt">nitrate</span> for aged internally mixed aerosols in atmosphere, the quantitative <span class="hlt">nitrate</span> depletion dependent upon composition in particles is still not well constrained. The chemical composition evolutions for malonic acid/sodium <span class="hlt">nitrate</span> (MA/SN), malonic acid/magnesium <span class="hlt">nitrate</span> (MA/MN) and malonic acid/calcium <span class="hlt">nitrate</span> (MA/CN) particles with the organic to inorganic molar ratio (OIR) of 1:1 are investigated by vacuum Fourier transform infrared spectroscopy (FTIR). Upon dehydration, the intensity of the asymmetric stretching mode of COO - group (ν as -COO - ) increases, accompanying the decrease in OH feather band and COOH band and NO 3 - band. These band changes suggest malonate salts formation and HNO 3 release. The quantitative NO 3 - depletion data shows that the reactivity of MA-MN is most and that of MA-SN is least. Analysis of the stretching mode of COO - indicates the different bond type between metal cation and carboxylate anion. In addition, water content in particles decreases at the constant RH, implying water loss with the chemical reaction. When the RH changes very quickly, water uptake delay during the humidification process reveals that water mass transport is limited below 37% RH. Copyright © 2018 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26829247','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26829247"><span>Fiber Type-Specific Effects of Dietary <span class="hlt">Nitrate</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jones, Andrew M; Ferguson, Scott K; Bailey, Stephen J; Vanhatalo, Anni; Poole, David C</p> <p>2016-04-01</p> <p>Dietary <span class="hlt">nitrate</span> supplementation increases circulating nitrite <span class="hlt">concentration</span>, and the subsequent reduction of nitrite to nitric oxide is promoted in hypoxic environments. Given that PO2 is lower in Type II compared with Type I muscle, this article examines the hypothesis that the ergogenicity of <span class="hlt">nitrate</span> supplementation is linked to specific effects on vascular, metabolic, and contractile function in Type II muscle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5906813','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5906813"><span>Organic <span class="hlt">nitrate</span> chemistry and its implications for nitrogen budgets in an isoprene- and monoterpene-rich atmosphere: constraints from aircraft (SEAC4RS) and ground-based (SOAS) <span class="hlt">observations</span> in the Southeast US</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Fisher, J. A.; Jacob, D. J.; Travis, K. R.; Kim, P. S.; Marais, E. A.; Miller, C. Chan; Yu, K.; Zhu, L.; Yantosca, R. M.; Sulprizio, M. P.; Mao, J.; Wennberg, P. O.; Crounse, J. D.; Teng, A. P.; Nguyen, T. B.; St. Clair, J. M.; Cohen, R. C.; Romer, P.; Nault, B. A.; Wooldridge, P. J.; Jimenez, J. L.; Campuzano-Jost, P.; Day, D. A.; Hu, W.; Shepson, P. B.; Xiong, F.; Blake, D. R.; Goldstein, A. H.; Misztal, P. K.; Hanisco, T. F.; Wolfe, G. M.; Ryerson, T. B.; Wisthaler, A.; Mikoviny, T.</p> <p>2018-01-01</p> <p>Formation of organic <span class="hlt">nitrates</span> (RONO2) during oxidation of biogenic volatile organic compounds (BVOCs: isoprene, monoterpenes) is a significant loss pathway for atmospheric nitrogen oxide radicals (NOx), but the chemistry of RONO2 formation and degradation remains uncertain. Here we implement a new BVOC oxidation mechanism (including updated isoprene chemistry, new monoterpene chemistry, and particle uptake of RONO2) in the GEOS-Chem global chemical transport model with ∼25 × 25 km2 resolution over North America. We evaluate the model using aircraft (SEAC4RS) and ground-based (SOAS) <span class="hlt">observations</span> of NOx, BVOCs, and RONO2 from the Southeast US in summer 2013. The updated simulation successfully reproduces the <span class="hlt">concentrations</span> of individual gas- and particle-phase RONO2 species measured during the campaigns. Gas-phase isoprene <span class="hlt">nitrates</span> account for 25-50% of <span class="hlt">observed</span> RONO2 in surface air, and we find that another 10% is contributed by gas-phase monoterpene <span class="hlt">nitrates</span>. <span class="hlt">Observations</span> in the free troposphere show an important contribution from long-lived <span class="hlt">nitrates</span> derived from anthropogenic VOCs. During both campaigns, at least 10% of <span class="hlt">observed</span> boundary layer RONO2 were in the particle phase. We find that aerosol uptake followed by hydrolysis to HNO3 accounts for 60% of simulated gas-phase RONO2 loss in the boundary layer. Other losses are 20% by photolysis to recycle NOx and 15% by dry deposition. RONO2 production accounts for 20% of the net regional NOx sink in the Southeast US in summer, limited by the spatial segregation between BVOC and NOx emissions. This segregation implies that RONO2 production will remain a minor sink for NOx in the Southeast US in the future even as NOx emissions continue to decline. PMID:29681921</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5367448','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/5367448"><span>Method for improved decomposition of metal <span class="hlt">nitrate</span> solutions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Haas, P.A.; Stines, W.B.</p> <p>1981-01-21</p> <p>A method for co-conversion of aqueous solutions of one or more heavy metal <span class="hlt">nitrates</span> is described, wherein thermal decomposition within a temperature range of about 300 to 800/sup 0/C is carried out in the presence of about 50 to 500% molar <span class="hlt">concentration</span> of ammonium <span class="hlt">nitrate</span> to total metal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024957','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024957"><span>Surface Water Qualit: Revisiting <span class="hlt">Nitrate</span> <span class="hlt">Concentrations</span> in the Des Moines River: 1945 and 1976-2001</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>McIsaac, G.F.; Libra, R.D.</p> <p>2003-01-01</p> <p>Recent compilations of historical and contemporary riverine <span class="hlt">nitrate</span> (NO3) <span class="hlt">concentrations</span> indicate that <span class="hlt">concentrations</span> in many rivers in the north-central USA increased during the second half of the 20th century. The Des Moines River near Des Moines, Iowa, however, was reported to have had similar NO3 <span class="hlt">concentrations</span> in 1945 and the 1980s, in spite of substantially greater N input to the watershed during the latter period. The objective of this study was to reconsider the comparison of historical and contemporary NO3 <span class="hlt">concentrations</span> in the Des Moines River near Des Moines in light of the following: (i) possible errors in the historical data used, (ii) variations in methods of sample collection, (iii) variations in location of sampling, and (iv) additional data collected since 1990. We discovered that an earlier study had compared the flow-weighted average <span class="hlt">concentration</span> in 1945 to arithmetic annual average <span class="hlt">concentrations</span> in the 1980s. The intertemporal comparison also appeared to be influenced by differences in sample collection methods and locations used at different times. Depending on the model used and the estimated effects of composite sample collection, the 1945 arithmetic average NO3 <span class="hlt">concentration</span> was between 44 and 57% of the expected mean value at a similar water yield during 1976-2001. The flow-weighted average NO3 <span class="hlt">concentration</span> for 1945 was between 54 and 73% of the expected mean value at a similar water yield during 1976-2001. The difference between NO3 <span class="hlt">concentrations</span> in 1945 and the contemporary period are larger than previously reported for the Des Moines River.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..1610876K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..1610876K"><span>Parsimonious Hydrologic and <span class="hlt">Nitrate</span> Response Models For Silver Springs, Florida</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Klammler, Harald; Yaquian-Luna, Jose Antonio; Jawitz, James W.; Annable, Michael D.; Hatfield, Kirk</p> <p>2014-05-01</p> <p>Silver Springs with an approximate discharge of 25 m3/sec is one of Florida's first magnitude springs and among the largest springs worldwide. Its 2500-km2 springshed overlies the mostly unconfined Upper Floridan Aquifer. The aquifer is approximately 100 m thick and predominantly consists of porous, fractured and cavernous limestone, which leads to excellent surface drainage properties (no major stream network other than Silver Springs run) and complex groundwater flow patterns through both rock matrix and fast conduits. Over the past few decades, discharge from Silver Springs has been <span class="hlt">observed</span> to slowly but continuously decline, while <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the spring water have enormously increased from a background level of 0.05 mg/l to over 1 mg/l. In combination with concurrent increases in algae growth and turbidity, for example, and despite an otherwise relatively stable water quality, this has given rise to concerns about the ecological equilibrium in and near the spring run as well as possible impacts on tourism. The purpose of the present work is to elaborate parsimonious lumped parameter models that may be used by resource managers for evaluating the springshed's hydrologic and <span class="hlt">nitrate</span> transport responses. Instead of attempting to explicitly consider the complex hydrogeologic features of the aquifer in a typically numerical and / or stochastic approach, we use a transfer function approach wherein input signals (i.e., time series of groundwater recharge and <span class="hlt">nitrate</span> loading) are transformed into output signals (i.e., time series of spring discharge and spring <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>) by some linear and time-invariant law. The dynamic response types and parameters are inferred from comparing input and output time series in frequency domain (e.g., after Fourier transformation). Results are converted into impulse (or step) response functions, which describe at what time and to what magnitude a unitary change in input manifests at the output. For the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28095125','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28095125"><span><span class="hlt">Nitrate</span> in drinking water and risk of colorectal cancer in Yogyakarta, Indonesia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fathmawati; Fachiroh, Jajah; Gravitiani, Evi; Sarto; Husodo, Adi Heru</p> <p>2017-01-01</p> <p><span class="hlt">Nitrate</span> <span class="hlt">concentration</span> in well water in Yogyakarta, Indonesia, and its surroundings tended to increase rapidly from time to time, and it may be associated with an elevated risk for several types of cancer. The purpose of this study was to examine the association between <span class="hlt">nitrate</span> in drinking water and colorectal cancer (CRC) risk occurrence. A case-control study was conducted in Yogyakarta Special Province. Pathologically confirmed 75 CRC patients and 75 controls were consulted and their individual well water was sampled and examined for <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Logistic regression analysis was conducted to establish the association between <span class="hlt">nitrate</span> and CRC risk development. There was a significant correlation between <span class="hlt">nitrate</span> in drinking water and CRC occurrence, and this value was relatively stable after being adjusted for protein intake, smoking history, age, and family history of cancer. These findings demonstrated that the risk of CRC development was fourfold among those with >10 years of <span class="hlt">nitrate</span> exposure from well water compared with those with ≤10 years of <span class="hlt">nitrate</span> exposure. Consequently, a significant association between <span class="hlt">nitrate</span> in drinking water and occurrence of CRC in Yogyakarta was established.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3237823','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3237823"><span>Inhaled NO therapy increases blood nitrite, <span class="hlt">nitrate</span> and S-nitrosohemoglobin <span class="hlt">concentrations</span> in infants with pulmonary hypertension</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ibrahim, Yomna I.; Ninnis, Janet R.; Hopper, Andrew O.; Deming, Douglas D.; Zhang, Amy X.; Herring, Jason L.; Sowers, Lawrence C.; McMahon, Timothy J.; Power, Gordon G.; Blood, Arlin B.</p> <p>2011-01-01</p> <p>Objective To measure the circulating <span class="hlt">concentrations</span> of nitric oxide (NO) adducts with NO bioactivity following inhaled NO therapy in infants with pulmonary hypertension. Study design In this single center study five sequential blood samples were collected from infants with pulmonary hypertension before, during and after therapy with iNO (n=17). Samples were collected from a control group of hospitalized infants without pulmonary hypertension (n=16) and from healthy adults for comparison (n=12). Results After beginning iNO (20 ppm) whole blood nitrite increased about two-fold within two hours (P<0.01). Whole blood <span class="hlt">nitrate</span> increased to four-fold higher than baseline during treatment with 20ppm iNO (P<0.01). S-nitrosohemoglobin (SNO-Hb) increased measurably after beginning iNO (P<0.01) whereas iron nitrosyl hemoglobin and total Hb-bound NO-species compounds did not change. Conclusion Treatment of pulmonary hypertensive infants with iNO results in increases in nitrite, <span class="hlt">nitrate</span>, and SNO-Hb in circulating blood. We speculate that these compounds may be carriers of NO bioactivity throughout the body and account for peripheral effects of iNO in the brain, heart and other organs. PMID:21907348</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27810489','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27810489"><span>The influence of urea and <span class="hlt">nitrate</span> nutrients on the bioavailability and toxicity of nickel to Prorocentrum donghaiense (Dinophyta) and Skeletonema costatum (Bacillariophyta).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huang, Xu-Guang; Lin, Xie-Chang; Li, Shun-Xing; Xu, Song-Li; Liu, Feng-Jiao</p> <p>2016-12-01</p> <p>Nitrogen nutrients and nickel(Ni) are ubiquitous in aquatic environments, and they are important for primary production of ocean ecosystem. This study examined the interaction of nitrogen nutrients (specifically urea and <span class="hlt">nitrate</span>) and Ni on chlorophyll (Chl a) <span class="hlt">concentration</span> and photosynthesis parameters values of Prorocentrum donghaiense and Skeletonema costatum. The data presented here indicate that low <span class="hlt">concentration</span> of Ni for P. donghaiense and S. costatum can enhance both Chl a <span class="hlt">concentration</span> and photosynthesis parameters values when grown in urea containing environment. Despite this increase there was also an <span class="hlt">observed</span> depression in both species tested when incubated in high <span class="hlt">concentration</span> of Ni for P. donghaiense and S. costatum regardless of incubating in urea or <span class="hlt">nitrate</span>. Additionally, EC 50 values of Chl a and Fv/Fm for Ni at different time intervals were calculated in this study. These <span class="hlt">observations</span> indicated that the Ni tolerance was higher in P. donghaiense as compared to S. costatum. The Ni tolerance of P. donghaiense incubated in urea was higher than that incubating in <span class="hlt">nitrate</span>. The same phenomenon was not <span class="hlt">observed</span> in S. costatum, which indicated that the influence of urea was dependent on the species investigated. Thus, urea input could impact Ni bioavailability and toxicity, and then affect the biodynamics thereafter. Copyright © 2016. Published by Elsevier B.V.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApSS..380..309K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApSS..380..309K"><span>Enhanced nitrogen selectivity for <span class="hlt">nitrate</span> reduction on Cu-nZVI by TiO2 photocatalysts under UV irradiation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krasae, Nalinee; Wantala, Kitirote</p> <p>2016-09-01</p> <p>The aims of this work were to study the effect of Cu-nZVI with and without TiO2 on <span class="hlt">nitrate</span> reduction and to study the pathway of <span class="hlt">nitrate</span> reduction utilizing to nitrogen gas. The chemical and physical properties of Cu-nZVI and Cu-nZVI/TiO2 such as specific surface area, crystalline phase, oxidation state of Cu and Fe and morphology were determined by N2 adsorption-desorption Brunauer-Emmett-Teller (BET) analytical technique, X-ray diffraction (XRD), X-ray Absorption Near Edge Structure (XANES) technique and Transmittance Electron Microscopy (TEM). The full factorial design (FFD) was used in this experiment for the effect of Cu-nZVI with and without TiO2, where the initial solution pH was varied at 4, 5.5, and 7 and initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was varied at 50, 75, and 100 ppm. Finally, the pathway of <span class="hlt">nitrate</span> reduction was examined to calculate the nitrogen gas selectivity. The specific area of Cu-nZVI and Cu-nZVI/TiO2 was found to be about 4 and 36 m2/g, respectively. The XRD pattern of Fe0 in Cu-nZVI was found at 45° (2θ), whereas Cu-nZVI/TiO2 cannot be <span class="hlt">observed</span>. TEM images can confirm the position of the core and the shell of nZVI for Fe0 and ferric oxide. Cu-nZVI/TiO2 proved to have higher activity in nitrogen reduction performance than that without TiO2 and <span class="hlt">nitrate</span> can be completely degraded in both of solution pH of 4 and 7 in 75 ppm of initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. It can be highlighted that the nitrogen gas selectivity of Cu-nZVI/TiO2 greater than 82% was found at an initial solution pH of 4 and 7. The main effects of Cu-nZVI with and without TiO2 and the initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span> on <span class="hlt">nitrate</span> reduction were significant. The interaction between solution pH and initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and the interaction of all effects at a reaction time of 15 min on <span class="hlt">nitrate</span> reduction were also significant.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23340115','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23340115"><span>Drinking-water exposure to a mixture of <span class="hlt">nitrate</span> and low-dose atrazine metabolites and small-for-gestational age (SGA) babies: a historic cohort study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Migeot, V; Albouy-Llaty, M; Carles, C; Limousi, F; Strezlec, S; Dupuis, A; Rabouan, S</p> <p>2013-04-01</p> <p>Groundwater, surface water and drinking water are contaminated by <span class="hlt">nitrates</span> and atrazine, an herbicide. They are present as a mixture in drinking water and with their endocrine-disrupting activity, they may alter fetal growth. To study an association between drinking-water atrazine metabolites/<span class="hlt">nitrate</span> mixture exposure and small-for-gestational-age(SGA). A historic cohort study based on birth records and drinking-water <span class="hlt">nitrate</span> and pesticide measurements in Deux-Sèvres (France) between 2005 and 2009 was carried out. Exposure to drinking-water atrazine metabolites/<span class="hlt">nitrate</span> mixture was divided into 6 classes according to the presence or absence of atrazine metabolites and to terciles of <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in each trimester of pregnancy. Regression analysis of SGA by mixture exposure at second trimester was subsequently conducted. We included 11,446 woman-neonate couples of whom 37.0% were exposed to pesticides, while 99.9% of the women were exposed to <span class="hlt">nitrates</span>. Average <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was from 0 to 63.30 mg/L. In the second trimester of pregnancy, the risk of SGA was different with mixture exposure when drinking-water atrazine metabolites, mainly 2 hydroxyatrazine and desethylatrazine, were present and <span class="hlt">nitrate</span> dose exposure increased: compared to single first tercile of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> exposure, single second tercile exposure OR was 1.74 CI 95% [1.10; 2.75] and atrazine metabolites presence in the third tercile of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> exposure OR was 0.87 CI 95% [0.45;1.67]. It is possible that the association found at the second trimester of exposure with regard to birth weight may likewise be <span class="hlt">observed</span> before birth, with regard to the estimated fetal weight, and that it might change in the event that the atrazine metabolites dose were higher or the <span class="hlt">nitrate</span> dose lower. It would appear necessary to further explore the variability of effects. Crown Copyright © 2013. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title21-vol3/pdf/CFR-2011-title21-vol3-sec181-33.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title21-vol3/pdf/CFR-2011-title21-vol3-sec181-33.pdf"><span>21 CFR 181.33 - Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</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>2011-04-01</p> <p>... 21 Food and Drugs 3 2011-04-01 2011-04-01 false Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. 181.33...-Sanctioned Food Ingredients § 181.33 Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span> are subject to prior sanctions issued by the U.S. Department of Agriculture for use as sources of...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title21-vol3/pdf/CFR-2010-title21-vol3-sec181-33.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title21-vol3/pdf/CFR-2010-title21-vol3-sec181-33.pdf"><span>21 CFR 181.33 - Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-04-01</p> <p>... 21 Food and Drugs 3 2010-04-01 2009-04-01 true Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. 181.33...-Sanctioned Food Ingredients § 181.33 Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span> are subject to prior sanctions issued by the U.S. Department of Agriculture for use as sources of...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title21-vol3/pdf/CFR-2014-title21-vol3-sec181-33.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title21-vol3/pdf/CFR-2014-title21-vol3-sec181-33.pdf"><span>21 CFR 181.33 - Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-04-01</p> <p>... 21 Food and Drugs 3 2014-04-01 2014-04-01 false Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. 181.33... <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span> are subject to prior sanctions issued... potassium nitrite, in the production of cured red meat products and cured poultry products. [48 FR 1705, Jan...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25176303','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25176303"><span><span class="hlt">Nitrate</span> removal with lateral flow sulphur autotrophic denitrification reactor.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lv, Xiaomei; Shao, Mingfei; Li, Ji; Xie, Chuanbo</p> <p>2014-01-01</p> <p>An innovative lateral flow sulphur autotrophic denitrification (LFSAD) reactor was developed in this study; the treatment performance was evaluated and compared with traditional sulphur/limestone autotrophic denitrification (SLAD) reactor. Results showed that nitrite accumulation in the LFSAD reactor was less than 1.0 mg/L during the whole operation. Denitrification rate increased with the increased initial alkalinity and was approaching saturation when initial alkalinity exceeded 2.5 times the theoretical value. Higher influent <span class="hlt">nitrate</span> <span class="hlt">concentration</span> could facilitate <span class="hlt">nitrate</span> removal capacity. In addition, denitrification efficiency could be promoted under an appropriate reflux ratio, and the highest <span class="hlt">nitrate</span> removal percentage was achieved under reflux ratio of 200%, increased by 23.8% than that without reflux. Running resistance was only about 1/9 of that in SLAD reactor with equal amount of <span class="hlt">nitrate</span> removed, which was the prominent excellence of the new reactor. In short, this study indicated that the developed reactor was feasible for <span class="hlt">nitrate</span> removal from waters with lower <span class="hlt">concentrations</span>, including contaminated surface water, groundwater or secondary effluent of municipal wastewater treatment with fairly low running resistance. The innovation in reactor design in this study may bring forth new ideas of reactor development of sulphur autotrophic denitrification for <span class="hlt">nitrate</span>-contaminated water treatment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24020705','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24020705"><span>Nitrites and <span class="hlt">nitrates</span> in exhaled breath condensate in cystic fibrosis: relation to clinical parameters.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fila, L; Chladek, J; Maly, M; Musil, J</p> <p>2013-01-01</p> <p>To evaluate correlation of exhaled breath condensate (EBC) nitrite and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> with disease severity in cystic fibrosis (CF) patients. Nitrites and <span class="hlt">nitrates</span> are products of oxidative metabolism of nitric oxide. Impaired metabolism of nitric oxide plays a role in pathogenesis of CF. EBC was collected from 46 stable CF patients and from 21 healthy controls. EBC <span class="hlt">concentrations</span> of nitrites and <span class="hlt">nitrates</span> were correlated with parameters of lung disease and nutritional status and with systemic inflammatory markers. EBC <span class="hlt">nitrates</span> <span class="hlt">concentrations</span> in CF patients were lower than in healthy subjects (5.8 vs 14.3 μmol/l, p<0.001). They correlated positively with FEV1 (p=0.025) and serum albumin values (p=0.016) and negatively with chest radiograph Northern score (p=0.015) and serum C-reactive protein values (p=0.005). EBC nitrites <span class="hlt">concentrations</span> in CF patients did not differ from those in healthy subjects and were not correlated to any studied parameter. EBC <span class="hlt">nitrates</span> <span class="hlt">concentrations</span> correlate with disease severity in CF patients and are lower than in healthy subjects (Tab. 4, Fig. 1, Ref. 48).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5037521','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5037521"><span>Effects of a Short-Term High-<span class="hlt">Nitrate</span> Diet on Exercise Performance</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Porcelli, Simone; Pugliese, Lorenzo; Rejc, Enrico; Pavei, Gaspare; Bonato, Matteo; Montorsi, Michela; La Torre, Antonio; Rasica, Letizia; Marzorati, Mauro</p> <p>2016-01-01</p> <p>It has been reported that <span class="hlt">nitrate</span> supplementation can improve exercise performance. Most of the studies have used either beetroot juice or sodium <span class="hlt">nitrate</span> as a supplement; there is lack of data on the potential ergogenic benefits of an increased dietary <span class="hlt">nitrate</span> intake from a diet based on fruits and vegetables. Our aim was to assess whether a high-<span class="hlt">nitrate</span> diet increases nitric oxide bioavailability and to evaluate the effects of this nutritional intervention on exercise performance. Seven healthy male subjects participated in a randomized cross-over study. They were tested before and after 6 days of a high (HND) or control (CD) <span class="hlt">nitrate</span> diet (~8.2 mmol∙day−1 or ~2.9 mmol∙day−1, respectively). Plasma <span class="hlt">nitrate</span> and nitrite <span class="hlt">concentrations</span> were significantly higher in HND (127 ± 64 µM and 350 ± 120 nM, respectively) compared to CD (23 ± 10 µM and 240 ± 100 nM, respectively). In HND (vs. CD) were <span class="hlt">observed</span>: (a) a significant reduction of oxygen consumption during moderate-intensity constant work-rate cycling exercise (1.178 ± 0.141 vs. 1.269 ± 0.136 L·min−1); (b) a significantly higher total muscle work during fatiguing, intermittent sub-maximal isometric knee extension (357.3 ± 176.1 vs. 253.6 ± 149.0 Nm·s·kg−1); (c) an improved performance in Repeated Sprint Ability test. These findings suggest that a high-<span class="hlt">nitrate</span> diet could be a feasible and effective strategy to improve exercise performance. PMID:27589795</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.A53D..03M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.A53D..03M"><span>Ammonium <span class="hlt">Nitrate</span> Formation near the Colorado Front Range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Middlebrook, A. M.; Bahreini, R.; Brock, C. A.; Brown, S. S.; Cozic, J.; Frost, G. J.; Langford, A. O.; Lerner, B. M.; Matthew, B.; McKeen, S. A.; Neuman, J.; Nowak, J. B.; Peischl, J. W.; Quinn, P.; Ryerson, T. B.; Schultz, K.; Stark, H.; Trainer, M.; Wagner, N.; Williams, E. J.; Wollny, A. G.</p> <p>2009-12-01</p> <p>A significant air quality issue during wintertime temperature inversions along the Colorado Front Range urban corridor is the infamous “Brown Cloud” which is dominated by ammonium <span class="hlt">nitrate</span> particles. Aerosol composition, size distribution, and gas phase measurements were obtained along with meteorology in Boulder-based ground studies during the winters of 2005 and 2009 and in an airborne survey over the Colorado Front Range urban corridor and northeastern Colorado on April 1, 2008. New in these campaigns was the fast time response data which showed that nitric acid was partitioned mainly into the aerosol phase as ammonium <span class="hlt">nitrate</span>. During the survey flight, ammonium <span class="hlt">nitrate</span> mass <span class="hlt">concentrations</span> were highest on the west side of the urban corridor whereas nitrogen oxide <span class="hlt">concentrations</span> were highest directly west and south of Denver. Nitric acid <span class="hlt">concentrations</span> were highest south of the city. The calculated equilibrium gas phase ammonia was highest close to the ground directly around large feed lots near Brush and west of Greeley. These differences are consistent with what is known about the locations of emission sources, the predominant flow during the experiments, and the chemistry. Indeed, the ammonia emissions in the northern part of the region are sufficiently high to cause ammonium <span class="hlt">nitrate</span> formation to be limited by nitric acid whereas in the southern part of the region ammonium <span class="hlt">nitrate</span> formation was limited by low ammonia emissions. Although NOx (NO + NO2) emissions in the region are much larger than those for ammonia, NOx must be converted into nitric acid in order for ammonium <span class="hlt">nitrate</span> to form. In the survey data, aerosol <span class="hlt">nitrate</span> was correlated with the daytime nitric acid production rate but with higher slopes in the northern parts of the region. In the longer Boulder datasets, the calculated daytime production rate was slow and comparable to nighttime heterogeneous production via N2O5 hydrolysis. During periods of low aerosol surface area, daytime and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24086836','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24086836"><span>Histopathological changes in the liver of rabbits exposed to high <span class="hlt">nitrate</span> ingestion in drinking water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sharma, Manoj Kumar; Sharma, Hemlata; Bapna, Neelam</p> <p>2013-08-01</p> <p>In India, especially in Rajasthan, people drink water which contains high level of <span class="hlt">nitrates</span> and the possibility of finding <span class="hlt">concentrations</span> of up to 500 mg of <span class="hlt">nitrate</span> ions per litre of water is not unusual. Excessive use of <span class="hlt">nitrate</span> fertilisers and herbicides results in accumulation of <span class="hlt">nitrate</span> in plants and methemoglobinaemia in cattle as consequences of <span class="hlt">nitrate</span> poisoning. The ingested <span class="hlt">nitrate</span> is converted to nitrite in the digestive system and it is absorbed in blood, thus causing methemoglobinaemia. Methaemoglobinaemia is not restricted to infants alone, but it is prevalent in higher age groups also. Therefore, an experimental study was conducted on 10 rabbits which were between three and a half months to four months of age, which had weights which ranged from 1.310 kg to 1.720 kg. Five groups A, B, C,D and E were formed, with two rabbits in each group. The control Group A was given water orally, which had 45 mg/litres of <span class="hlt">nitrate</span>. Groups B to E (experimental groups) were administered water orally, which had <span class="hlt">concentrations</span> of 100mg/litre, 200mg/litre, 400mg/litre and 500mg/litre of <span class="hlt">nitrate</span> respectively, for 120 days. During experimental period, the differences in general behaviour of rabbits were noted. After this, rabbits were anaesthetised and sacrificed according to guidelines of ICMR and their livers were removed and processed for making paraffin sections,.Hematoxyllin and eosin staining was done for microscopic <span class="hlt">observations</span>. During experimental period, the animals were found to be lethargic on 75(th) day. Quantity of intake of food and water was not altered in the rabbits which were undergoing experiments in different groups. Rabbits of all groups i.e. A to E showed a continuous increase in heart rate (up to 218/minute in Group E) and respiration rate (up to 84/minute in Group E) respectively. The microscopic study showed mild necrosis of hepatocytes, with infiltration of inflammatory cells in between hepatocytes. In higher groups, the liver showed bridging</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20082016','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20082016"><span>The unintended energy impacts of increased <span class="hlt">nitrate</span> contamination from biofuels production.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Twomey, Kelly M; Stillwell, Ashlynn S; Webber, Michael E</p> <p>2010-01-01</p> <p>Increases in corn cultivation for biofuels production, due to the Energy Independence and Security Act of 2007, are likely to lead to increases in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in both surface and groundwater resources in the United States. These increases might trigger the requirement for additional energy consumption for water treatment to remove the <span class="hlt">nitrates</span>. While these increasing <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> might pose a human health concern, most water resources were found to be within current maximum contaminant level (MCL) limits of 10 mg L(-1) NO(3)-N. When water resources exceed this MCL, energy-intensive drinking water treatment is required to reduce <span class="hlt">nitrate</span> levels below 10 mg L(-1). Based on prior estimates of water supplies currently exceeding the <span class="hlt">nitrate</span> MCL, we calculate that advanced drinking water treatment might require an additional 2360 million kWh annually (for <span class="hlt">nitrate</span> affected areas only)--a 2100% increase in energy requirements for water treatment in those same areas--to mitigate <span class="hlt">nitrate</span> contamination and meet the MCL requirement. We predict that projected increases in <span class="hlt">nitrate</span> contamination in water may impact the energy consumed in the water treatment sector, because of the convergence of several related trends: (1) increasing cornstarch-based ethanol production, (2) increasing nutrient loading in surface water and groundwater resources as a consequence of increased corn-based ethanol production, (3) additional drinking water sources that exceed the MCL for <span class="hlt">nitrate</span>, and (4) potentially more stringent drinking water standards for <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70044000','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70044000"><span><span class="hlt">Nitrate</span> in watersheds: straight from soils to streams?</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Sudduth, Elizabeth B.; Perakis, Steven S.; Bernhardt, Emily S.</p> <p>2013-01-01</p> <p>Human activities are rapidly increasing the global supply of reactive N and substantially altering the structure and hydrologic connectivity of managed ecosystems. There is long-standing recognition that N must be removed along hydrologic flowpaths from uplands to streams, yet it has proven difficult to assess the generality of this removal across ecosystem types, and whether these patterns are influenced by land-use change. To assess how well upland <span class="hlt">nitrate</span> (NO3-) loss is reflected in stream export, we gathered information from >50 watershed biogeochemical studies that reported <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> ([NO3-]) for stream water and for either upslope soil solution or groundwater NO3- to examine whether stream export of NO3- accurately reflects upland NO3- losses. In this dataset, soil solution and streamwater [NO3-] were correlated across 40 undisturbed forest watersheds, with streamwater [NO3-] typically half (median = 50%) soil solution [NO3-]. A similar relationship was seen in 10 disturbed forest watersheds. However, for 12 watersheds with significant agricultural or urban development, the intercept and slope were both significantly higher than the relationship seen in forest watersheds. Differences in <span class="hlt">concentration</span> between soil solution or groundwater and stream water may be attributed to biological uptake, microbial processes including denitrification, and/or preferential flow routing. The results of this synthesis are consistent with the hypotheses that undisturbed watersheds have a significant capacity to remove <span class="hlt">nitrate</span> after it passes below the rooting zone and that land use changes tend to alter the efficiency or the length of watershed flowpaths, leading to reductions in <span class="hlt">nitrate</span> removal and increased stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27130094','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27130094"><span><span class="hlt">Nitrate</span> biosensors and biological methods for <span class="hlt">nitrate</span> determination.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sohail, Manzar; Adeloju, Samuel B</p> <p>2016-06-01</p> <p>The inorganic <span class="hlt">nitrate</span> (NO3‾) anion is present under a variety of both natural and artificial environmental conditions. <span class="hlt">Nitrate</span> is ubiquitous within the environment, food, industrial and physiological systems and is mostly present as hydrated anion of a corresponding dissolved salt. Due to the significant environmental and toxicological effects of <span class="hlt">nitrate</span>, its determination and monitoring in environmental and industrial waters are often necessary. A wide range of analytical techniques are available for <span class="hlt">nitrate</span> determination in various sample matrices. This review discusses biosensors available for <span class="hlt">nitrate</span> determination using the enzyme <span class="hlt">nitrate</span> reductase (NaR). We conclude that <span class="hlt">nitrate</span> determination using biosensors is an excellent non-toxic alternative to all other available analytical methods. Over the last fifteen years biosensing technology for <span class="hlt">nitrate</span> analysis has progressed very well, however, there is a need to expedite the development of <span class="hlt">nitrate</span> biosensors as a suitable alternative to non-enzymatic techniques through the use of different polymers, nanostructures, mediators and strategies to overcome oxygen interference. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19221887','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19221887"><span>Fluoride, boron and <span class="hlt">nitrate</span> toxicity in ground water of northwest Rajasthan, India.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chaudhary, Veena; Kumar, Mukesh; Sharma, Mukesh; Yadav, B S</p> <p>2010-02-01</p> <p>The study was carried out to access the fluoride, boron, and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in ground water samples of different villages in Indira Gandhi, Bhakra, and Gang canal catchment area of northwest Rajasthan, India. Rural population, in the study site, is using groundwater for drinking and irrigation purposes, without any quality test of water. All water samples (including canal water) were contaminated with fluoride. Fluoride, boron, and <span class="hlt">nitrate</span> were <span class="hlt">observed</span> in the ranges of 0.50-8.50, 0.0-7.73, and 0.0-278.68 mg/l, respectively. Most of the water samples were in the categories of fluoride 1.50 mg/l, of boron 2.0-4.0 mg/l, and of <span class="hlt">nitrate</span> < 45 mg/l. There was no industrial pollution in the study site; hence, availability of these compounds in groundwater was due to natural reasons and by the use of chemical fertilizers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/14599146','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14599146"><span><span class="hlt">Nitrate</span> removal from drinking water through the use of encapsulated microorganisms in alginate beads.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, S X; Hermanowicz, S W; Peng, M</p> <p>2003-09-01</p> <p>Biological treatment for removal of <span class="hlt">nitrate</span> from drinking water is of great significance, as traditional physical and chemical methods could not effectively remove soluble <span class="hlt">nitrate</span>. In this report immobilized microorganisms with co-immobilized calcium tartrate were used for reducing <span class="hlt">nitrate</span> <span class="hlt">concentration</span> (110 mg l(-1) NO3-N) in a model solution. The carbon source also functions as a stabilizing agent for the immobilization matrix. Experiments of denitrification showed a high <span class="hlt">nitrate</span> removal rate while nitrite residual was at a <span class="hlt">concentration</span> higher than expected. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was reduced to nearly zero (0.2-1.4 mg l(-1)) after 3 days of operation. The calcium tartrate (4%, w/w) co-immobilized alginate beads had better <span class="hlt">nitrate</span> removal performance than tartrate in solution. The nitrite-N residual <span class="hlt">concentration</span> was approximately 1.1-2.9 mg l(-1) at the end of the experiments, showing the desirability of further denitrification. The stability of alginate beads was also tested both to evaluate their behaviors and investigate the efficacy of bead recycling. It was found that the beads could be used for 8-13 days consecutively without any structural deterioration and leaking of microbes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17936327','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17936327"><span>Perchlorate and <span class="hlt">nitrate</span> treatment by ion exchange integrated with biological brine treatment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lehman, S Geno; Badruzzaman, Mohammad; Adham, Samer; Roberts, Deborah J; Clifford, Dennis A</p> <p>2008-02-01</p> <p>Groundwater contaminated with perchlorate and <span class="hlt">nitrate</span> was treated in a pilot plant using a commercially available ion exchange (IX) resin. Regenerant brine <span class="hlt">concentrate</span> from the IX process, containing high perchlorate and <span class="hlt">nitrate</span>, was treated biologically and the treated brine was reused in IX resin regeneration. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of the feed water determined the exhaustion lifetime (i.e., regeneration frequency) of the resin; and the regeneration condition was determined by the perchlorate elution profile from the exhausted resin. The biological brine treatment system, using a salt-tolerant perchlorate- and <span class="hlt">nitrate</span>-reducing culture, was housed in a sequencing batch reactor (SBR). The biological process consistently reduced perchlorate and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the spent brine to below the treatment goals of 500 microg ClO4(-)/L and 0.5mg NO3(-)-N/L determined by equilibrium multicomponent IX modeling. During 20 cycles of regeneration, the system consistently treated the drinking water to below the MCL of <span class="hlt">nitrate</span> (10 mgNO3(-)-N/L) and the California Department of Health Services (CDHS) notification level of perchlorate (i.e., 6 microg/L). A conceptual cost analysis of the IX process estimated that perchlorate and <span class="hlt">nitrate</span> treatment using the IX process with biological brine treatment to be approximately 20% less expensive than using the conventional IX with brine disposal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/wri01-4117/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/wri01-4117/"><span>Effects of Land Use and Travel Time on the Distribution of <span class="hlt">Nitrate</span> in the Kirkwood-Cohansey Aquifer System in Southern New Jersey</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kauffman, Leon J.; Baehr, Arthur L.; Ayers, Mark A.; Stackelberg, Paul E.</p> <p>2001-01-01</p> <p>Residents of the southern New Jersey Coastal Plain are increasingly reliant on the unconfined Kirkwood-Cohansey aquifer system for public water supply as a result of increasing population and restrictions on withdrawals from the deeper, confined aquifers. Elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above background levels have been found in wells in the surficial aquifer system in agricultural and urban parts of this area. A three-dimensional steady-state ground-water-flow model of a 400-square-mile study area near Glassboro, New Jersey, was used in conjunction with particle tracking to examine the effects of land use and travel time on the distribution of <span class="hlt">nitrate</span> in ground and surface water in southern New Jersey. Contributing areas and ground-water ages, or travel times, of water at ground-water discharge points (streams and wells) in the study area were simulated. <span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> were computed by linking land use and age-dependent <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in recharge to the discharge points. Median <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> in water samples collected during 1996 from shallow monitoring wells in different land-use areas were used to represent the <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> in aquifer recharge since 1990. On the basis of upward trends in the use of nitrogen fertilizer, the <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> in aquifer recharge in agricultural and urban areas were assumed to have increased linearly from the background value in 1940 (0.07 mg/L as N) to the 1990 (2.5-14 mg/L as N) <span class="hlt">concentrations</span>. Model performance was evaluated by comparing the simulation results to measured <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and apparent ground-water ages. Apparent ground-water ages at 32 monitoring wells in the study area determined from tritium/helium-3 ratios and sulfur hexafluoride <span class="hlt">concentrations</span> favorably matched simulated travel times to these wells. Simulated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were comparable to <span class="hlt">concentrations</span> measured in 27 water-supply wells in the study area. A time series (1987-98) of <span class="hlt">nitrate</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003CRGeo.335..307L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003CRGeo.335..307L"><span>Pollution par les <span class="hlt">nitrates</span> des eaux souterraines du bassin d'Essaouira (Maroc)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Laftouhi, Nour-Eddine; Vanclooster, Marnik; Jalal, Mohammed; Witam, Omar; Aboufirassi, Mohamed; Bahir, Mohamed; Persoons, Étienne</p> <p>2003-03-01</p> <p>The Essaouira Basin (Morocco) contains a multi-layered aquifer situated in fractured and karstic materials from the Middle and Upper Cretaceous (the Cenomanian, Turonian and Senonian). Water percolates through the limestone and dolomite formations of the Turonian stage either through the marls and calcareous marls of the Cenomanian or through the calcareous marly materials of the Senonian. The aquifer system may be interconnected since the marl layer separating the Turonian, Cenomanian and Senonian aquifers is thin or intensively fractured. In that case, the water is transported through a network of fractures and stratification joints. This paper describes the extent of the <span class="hlt">nitrate</span> pollution in the area and its origin. Most of the wells and drillholes located in the Kourimat perimeter are contaminated by <span class="hlt">nitrates</span> with some <span class="hlt">concentrations</span> over 400 mg l-1. <span class="hlt">Nitrate</span> contamination is also <span class="hlt">observed</span> in the surface water of the Qsob River, which constitutes the natural outlet of the multi-layered complex aquifer system. In this area, agriculture is more developed than in the rest of the Essaouira Basin. Diffuse pollution of the karstic groundwater body by agricultural fertiliser residues may therefore partially explain the <span class="hlt">observed</span> <span class="hlt">nitrate</span> pollution. However, point pollution around the wells, springs and drillholes from human wastewater, livestock faeces and the mineralisation of organic debris close to the Muslim cemeteries cannot be excluded.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28391923','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28391923"><span>Evolutionary processes and sources of high-<span class="hlt">nitrate</span> haze episodes over Beijing, Spring.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Ting; Sun, Yele; Zhang, Wei; Wang, Zifa; Liu, Xingang; Fu, Pingqing; Wang, Xiquan</p> <p>2017-04-01</p> <p>Rare and consecutive high-<span class="hlt">nitrate</span> haze pollution episodes were <span class="hlt">observed</span> in Beijing in spring 2012. We present detailed characterization of the sources and evolutionary mechanisms of this haze pollution, and focus on an episode that occurred between 15 and 26 April. Submicron aerosol species were found to be substantially elevated during haze episodes, and <span class="hlt">nitrates</span> showed the largest increase and occupation (average: 32.2%) in non-refractory submicron particles (NR-PM 1 ), which did not occur in other seasons as previously reported. The haze episode (HE) was divided into three sub-episodes, HEa, HEb, and HEc. During HEa and HEc, a shallow boundary layer, stagnant meteorological conditions, and high humidity favored the formation of high-<span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, which were mainly produced by three different processes - daytime photochemical production, gas-particle partitioning, and nighttime heterogeneous reactions - and the decline in visibility was mainly induced by NR-PM 1. However, unlike HEa and HEc, during HEb, the contribution of high <span class="hlt">nitrates</span> was partly from the transport of haze from the southeast of Beijing - the transport pathway was <span class="hlt">observed</span> at ~800-1000m by aerosol Lidar - and the decline in visibility during HEb was primarily caused by PM 2.5 . Our results provide useful information for air quality improvement strategies in Beijing during Spring. Copyright © 2016. Published by Elsevier B.V.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25439583','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25439583"><span>Effects of <span class="hlt">nitrate</span> addition to a diet on fermentation and microbial populations in the rumen of goats, with special reference to Selenomonas ruminantium having the ability to reduce <span class="hlt">nitrate</span> and nitrite.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Asanuma, Narito; Yokoyama, Shota; Hino, Tsuneo</p> <p>2015-04-01</p> <p>This study investigated the effects of dietary <span class="hlt">nitrate</span> addition on ruminal fermentation characteristics and microbial populations in goats. The involvement of Selenomonas ruminantium in <span class="hlt">nitrate</span> and nitrite reduction in the rumen was also examined. As the result of <span class="hlt">nitrate</span> feeding, the total <span class="hlt">concentration</span> of ruminal volatile fatty acids decreased, whereas the acetate : propionate ratio and the <span class="hlt">concentrations</span> of ammonia and lactate increased. Populations of methanogens, protozoa and fungi, as estimated by real-time PCR, were greatly decreased as a result of <span class="hlt">nitrate</span> inclusion in the diet. There was modest or little impact of <span class="hlt">nitrate</span> on the populations of prevailing species or genus of bacteria in the rumen, whereas Streptococcus bovis and S. ruminantium significantly increased. Both the activities of <span class="hlt">nitrate</span> reductase (NaR) and nitrite reductase (NiR) per total mass of ruminal bacteria were increased by <span class="hlt">nitrate</span> feeding. Quantification of the genes encoding NaR and NiR by real-time PCR with primers specific for S. ruminantium showed that these genes were increased by feeding <span class="hlt">nitrate</span>, suggesting that the growth of <span class="hlt">nitrate</span>- and nitrite-reducing S. ruminantium is stimulated by <span class="hlt">nitrate</span> addition. Thus, S. ruminantium is likely to play a major role in <span class="hlt">nitrate</span> and nitrite reduction in the rumen. © 2014 Japanese Society of Animal Science.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22341469','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22341469"><span>Gravel pit lake ecosystems reduce <span class="hlt">nitrate</span> and phosphate <span class="hlt">concentrations</span> in the outflowing groundwater.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Weilhartner, Andreas; Muellegger, Christian; Kainz, Martin; Mathieu, Francine; Hofmann, Thilo; Battin, Tom J</p> <p>2012-03-15</p> <p>Gravel excavation often bears conflicts with the use of drinking water as under-water-table mining can directly impact groundwater quality downstream of the open gravel pit lake due to exposure of the groundwater aquifer to the atmosphere and to human activities. To assess this potential impact of GPLs on groundwater, we assessed the mass balance for <span class="hlt">nitrate</span> (NO(3)) and phosphate (PO(4)) and whole-ecosystem metabolism of five post-excavation GPLs in Austria. GPLs differed in both age and residence time of lake water. We found that GPLs significantly reduced the <span class="hlt">concentration</span> of NO(3) and PO(4) as groundwater passes through the lake ecosystem, which in most cases acted as a net sink for these nutrients. Groundwater-derived nutrients enhanced both epilithic and pelagic net primary production in the GPLs, which ultimately leads to biomass accrual. Our data also suggest that this biomass accrual may induce, at least in part, clogging of the GPLs and their successive hydrodynamic isolation from the adjacent groundwater. Despite continuous biomass build-up and elevated <span class="hlt">concentrations</span> of dissolved organic carbon (DOC) in the lake water compared to the inflowing groundwater, DOC export into the outflowing groundwater remained low. Our data suggest that GPLs could contribute to groundwater amelioration where agricultural land use increases nutrient <span class="hlt">concentrations</span> in the groundwater given a proper management of these man-made ecosystems. Copyright © 2012 Elsevier B.V. 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_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040088294&hterms=1062&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3D%2526%25231062','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040088294&hterms=1062&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3D%2526%25231062"><span><span class="hlt">Nitrate</span> <span class="hlt">concentration</span> effects on NO3-N uptake and reduction, growth, and fruit yield in strawberry</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Darnell, R. L.; Stutte, G. W.; Sager, J. C. (Principal Investigator)</p> <p>2001-01-01</p> <p>Strawberries (Fragaria xananassa Duch. 'Osogrande') were grown hydroponically with three NO3-N <span class="hlt">concentrations</span> (3.75, 7.5, or 15.0 mM) to determine effects of varying <span class="hlt">concentration</span> on NO3-N uptake and reduction rates, and to relate these processes to growth and fruit yield. Plants were grown for 32 weeks, and NO3-N uptake and <span class="hlt">nitrate</span> reductase (NR) activities in roots and shoots were measured during vegetative and reproductive growth. In general, NO3-N uptake rates increased as NO3-N <span class="hlt">concentration</span> in the hydroponics system increased. Tissue NO3- <span class="hlt">concentration</span> also increased as external NO3-N <span class="hlt">concentration</span> increased, reflecting the differences in uptake rates. There was no effect of external NO3-N <span class="hlt">concentration</span> on NR activities in leaves or roots during either stage of development. Leaf NR activity averaged approximately 360 nmol NO2 formed/g fresh weight (FW)/h over both developmental stages, while NR activity in roots was much lower, averaging approximately 115 nmol NO2 formed/g FW/h. Vegetative organ FW, dry weight (DW), and total fruit yield were unaffected by NO3-N <span class="hlt">concentration</span>. These data suggest that the inability of strawberry to increase growth and fruit yield in response to increasing NO3-N <span class="hlt">concentrations</span> is not due to limitations in NO3-N uptake rates, but rather to limitations in NO3- reduction and/or assimilation in both roots and leaves.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26901734','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26901734"><span>Study of novel mechano-chemical activation process of red mud to optimize <span class="hlt">nitrate</span> removal from water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Alighardashi, A; Gharibi, H R; Raygan, Sh; Akbarzadeh, A</p> <p>2016-01-01</p> <p>Red mud (RM) is the industrial waste of alumina production and causes serious environmental risks. In this paper, a novel activation procedure for RM (mechano-chemical processing) is proposed in order to improve the <span class="hlt">nitrate</span> adsorption from water. High-energy milling and acidification were selected as mechanical and chemical activation methods, respectively. Synthesized samples of adsorbent were produced considering two parameters of activation: acid <span class="hlt">concentrations</span> and acidification time in two selected milling times. Optimization of the activation process was based on <span class="hlt">nitrate</span> removal from a stock solution. Experimental data were analyzed with two-way analysis of variance and Kruskal-Wallis methods to verify and discover the accuracy and probable errors. Best conditions (acceptable removal percentage > 75) were 17.6% w/w for acid <span class="hlt">concentrate</span> and 19.9 minutes for acidification time in 8 hours for milling time. A direct relationship between increase in <span class="hlt">nitrate</span> removal and increasing the acid <span class="hlt">concentration</span> and acidification time was <span class="hlt">observed</span>. The adsorption isotherms were studied and compared with other <span class="hlt">nitrate</span> adsorbents. Characterization tests (X-ray fluorescence, X-ray diffraction, Fourier transform infrared spectrophotometry, dynamic light scattering, surface area analysis and scanning electron microscopy) were conducted for both raw and activated adsorbents. Results showed noticeable superiority in characteristics after activation: higher specific area and porosity, lower particle size and lower agglomeration in structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70185710','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70185710"><span>Variability and prediction of freshwater and <span class="hlt">nitrate</span> fluxes for the Louisiana-Texas shelf: Mississippi and Atchafalaya River source functions</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Bratkovich, A.; Dinnel, S.P.; Goolsby, D.A.</p> <p>1994-01-01</p> <p>Time histories of riverine water discharge, <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, and <span class="hlt">nitrate</span>, flux have been analyzed for the Mississippi and Atchafalaya rivers. Results indicate that water discharge variability is dominated by the annual cycle and shorter-time-scale episodic events presumably associated with snowmelt runoff and spring or summer rains. Interannual variability in water discharge is relatively small compared to the above. In contrast, <span class="hlt">nitrate</span> <span class="hlt">concentration</span> exhibits strongest variability at decadal time scales. The interannual variability is not monotonic but more complicated in structure. Weak covariability between water discharge and <span class="hlt">nitrate</span> <span class="hlt">concentration</span> leads to a relatively “noisy” <span class="hlt">nitrate</span> flux signal. <span class="hlt">Nitrate</span> flux variations exhibit a low-amplitude, long-term modulation of a dominant annual cycle. Predictor-hindcastor analyses indicate that skilled forecasts of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> and <span class="hlt">nitrate</span> flux fields are feasible. Water discharge was the most reliably hindcast (on seasonal to interannual time scales) due to the fundamental strength of the annual hydrologic cycle. However, the forecasting effort for this variable was less successful than the hindcasting effort, mostly due to a phase shift in the annual cycle during our relatively short test period (18 mo). <span class="hlt">Nitrate</span> <span class="hlt">concentration</span> was more skillfully predicted (seasonal to interannual time scales) due to the relative dominance of the decadal-scale portion of the signal. <span class="hlt">Nitrate</span> flux was also skillfully forecast even though historical analyses seemed to indicate that it should be more difficult to predict than either water discharge or <span class="hlt">nitrate</span> <span class="hlt">concentration</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70034263','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70034263"><span>Climate-induced changes in high elevation stream <span class="hlt">nitrate</span> dynamics</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Baron, Jill S.; Schmidt, T.M.; Hartman, M.D.</p> <p>2009-01-01</p> <p>Mountain terrestrial and aquatic ecosystems are responsive to external drivers of change, especially climate change and atmospheric deposition of nitrogen (N). We explored the consequences of a temperature-warming trend on stream <span class="hlt">nitrate</span> in an alpine and subalpine watershed in the Colorado Front Range that has long been the recipient of elevated atmospheric N deposition. Mean annual stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> since 2000 are higher by 50% than an earlier monitoring period of 1991-1999. Mean annual N export increased by 28% from 2.03 kg N ha-1yr-1 before 2000 to 2.84 kg N ha-1yr-1 in Loch Vale watershed since 2000. The substantial increase in N export comes as a surprise, since mean wet atmospheric N deposition from 1991 to 2006 (3.06 kg N ha-1 yr-1) did not increase. There has been a period of below average precipitation from 2000 to 2006 and a steady increase in summer and fall temperatures of 0.12??C yr-1 in both seasons since 1991. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span>, as well as the weathering products calcium and sulfate, were higher for the period 2000-2006 in rock glacier meltwater at the top of the watershed above the influence of alpine and subalpine vegetation and soils. We conclude the <span class="hlt">observed</span> recent N increases in Loch Vale are the result of warmer summer and fall mean temperatures that are melting ice in glaciers and rock glaciers. This, in turn, has exposed sediments from which N produced by nitrification can be flushed. We suggest a water quality threshold may have been crossed around 2000. The phenomenon <span class="hlt">observed</span> in Loch Vale may be indicative of N release from ice features such as rock glaciers worldwide as mountain glaciers retreat. ?? 2009 Blackwell Publishing Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title21-vol3/pdf/CFR-2012-title21-vol3-sec181-33.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title21-vol3/pdf/CFR-2012-title21-vol3-sec181-33.pdf"><span>21 CFR 181.33 - Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</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>2012-04-01</p> <p>... 21 Food and Drugs 3 2012-04-01 2012-04-01 false Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. 181.33...-Sanctioned Food Ingredients § 181.33 Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. Sodium <span class="hlt">nitrate</span> and potassium... nitrite, with or without sodium or potassium nitrite, in the production of cured red meat products and...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title21-vol3/pdf/CFR-2013-title21-vol3-sec181-33.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title21-vol3/pdf/CFR-2013-title21-vol3-sec181-33.pdf"><span>21 CFR 181.33 - Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</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>2013-04-01</p> <p>... 21 Food and Drugs 3 2013-04-01 2013-04-01 false Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. 181.33...-Sanctioned Food Ingredients § 181.33 Sodium <span class="hlt">nitrate</span> and potassium <span class="hlt">nitrate</span>. Sodium <span class="hlt">nitrate</span> and potassium... nitrite, with or without sodium or potassium nitrite, in the production of cured red meat products and...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.5623S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.5623S"><span>Understanding the relationship between DOC and <span class="hlt">nitrate</span> export and dominant rainfall-runoff processes through long-term high frequency measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schwab, Michael; Klaus, Julian; Pfister, Laurent; Weiler, Markus</p> <p>2016-04-01</p> <p>Over the past decades, stream sampling protocols for hydro-geochemical parameters were often limited by logistical and technological constraints. While long-term monitoring protocols were typically based on weekly sampling intervals, high frequency sampling was commonly limited to a few single events. In our study, we combined high frequency and long-term measurements to understand the DOC and <span class="hlt">nitrate</span> behaviour and dynamics for different runoff events and seasons. Our study area is the forested Weierbach catchment (0.47 km2) in Luxembourg. The fractured schist bedrock is covered by cambisol soils. The runoff response of the catchment is characterized by a double peak behaviour. A first discharge peak occurs during or right after a rainfall event (triggered by fast near surface runoff generation processes), while a second delayed peak lasts several days (generated by subsurface flow/ shallow groundwater flow). Peaks in DOC <span class="hlt">concentrations</span> are closely linked to the first discharge peak, whereas <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> follow the second peak. Our <span class="hlt">observations</span> were carried out with the field deployable instrument spectro::lyser (scan Messtechnik GmbH). This instrument relies on the principles of UV-Vis spectrometry and measures DOC and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. The measurements were carried out at a high frequency of 15 minutes in situ in the Weierbach creek for more than two years. In addition, a long-term validation was carried out with data obtained from the analysis of water collected with automatic samplers. The long-term, high-frequency measurements allowed us to calculate a complete and detailed balance of DOC and <span class="hlt">nitrate</span> export over two years. Transport behaviour of the DOC and <span class="hlt">nitrate</span> showed different dynamics between the first and second hydrograph peaks. DOC is mainly exported during first peaks, while <span class="hlt">nitrate</span> is mostly exported during the delayed second peaks. In combination with other measurements in the catchment, the long and detailed <span class="hlt">observations</span> have</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2015/1065/pdf/ofr2015-1065.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2015/1065/pdf/ofr2015-1065.pdf"><span>Results from laboratory and field testing of <span class="hlt">nitrate</span> measuring spectrophotometers</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Snazelle, Teri T.</p> <p>2015-01-01</p> <p>In Phase II, the analyzers were deployed in field conditions at three diferent USGS sites. The measured <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were compared to discrete (reference) samples analyzed by the Direct UV method on a Shimadzu UV1800 bench top spectrophotometer, and by the National Environmental Methods Index (NEMI) method I-2548-11 at the USGS National Water Quality Laboratory. The first deployment at USGS site 0249620 on the East Pearl River in Hancock County, Mississippi, tested the ability of the TriOs ProPs (10-mm path length), Hach NITRATAX (5 mm), Satlantic SUNA (10 mm), and the S::CAN Spectro::lyser (5 mm) to accurately measure low-level (less than 2 mg-N/L) <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> while <span class="hlt">observing</span> the effect turbidity and colored dissolved organic matter (CDOM) would have on the analyzers' measurements. The second deployment at USGS site 01389005 Passaic River below Pompton River at Two Bridges, New Jersey, tested the analyzer's accuracy in mid-level (2-8 mg-N/L) <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. This site provided the means to test the analyzers' performance in two distinct matrices—the Passaic and the Pompton Rivers. In this deployment, three instruments tested in Phase I (TriOS, Hach, and SUNA) were deployed with the S::CAN Spectro::lyser (35 mm) already placed by the New Jersey Water Science Center (WSC). The third deployment at USGS site 05579610 Kickapoo Creek at 2100E Road near Bloomington, Illinois, tested the ability of the analyzers to measure high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> (greater than 8 mg-N/L) in turbid waters. For Kickapoo Creek, the HIF provided the TriOS (10 mm) and S::CAN (5 mm) from Phase I, and a SUNA V2 (5 mm) to be deployed adjacent to the Illinois WSC-owned Hach (2 mm). A total of 40 discrete samples were collected from the three deployment sites and analyzed. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of the samples ranged from 0.3–22.2 mg-N/L. The average absolute difference between the TriOS measurements and discrete samples was 0.46 mg-N/L. For the combined data</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170003444','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170003444"><span>Organic <span class="hlt">Nitrate</span> Chemistry and Its Implications for Nitrogen Budgets in an Isoprene- and Monoterpene-Rich Atmosphere: Constraints From Aircraft (SEAC4RS) and Ground-Based (SOAS) <span class="hlt">Observations</span> in the Southeast US</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fisher, Jenny; Jacob, D. J.; Travis, K. R.; Kim, P. S.; Marais, E. A.; Miller, C. Chan; Yu, K.; Zhu, L.; Yantosca, R. M.; Sulprizio, M. P.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170003444'); toggleEditAbsImage('author_20170003444_show'); toggleEditAbsImage('author_20170003444_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170003444_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170003444_hide"></p> <p>2016-01-01</p> <p>Formation of organic <span class="hlt">nitrates</span> (RONO2) during oxidation of biogenic volatile organic compounds (BVOCs: isoprene, monoterpenes) is a significant loss pathway for atmospheric nitrogen oxide radicals (NOx), but the chemistry of RONO2 formation and degradation remains uncertain. Here we implement a new BVOC oxidation mechanism (including updated isoprene chemistry, new monoterpene chemistry, and particle uptake of RONO2) in the GEOS-Chem global chemical transport model with approximately 25 times 25 km(exp 2) resolution over North America. We evaluate the model using aircraft (SEAC4RS) and ground-based (SOAS) <span class="hlt">observations</span> of NOx, BVOCs, and RONO2 from the Southeast US in summer 2013. The updated simulation successfully reproduces the <span class="hlt">concentrations</span> of individual gas- and particle-phase RONO2 species measured during the campaigns. Gas-phase isoprene <span class="hlt">nitrates</span> account for 2550 of <span class="hlt">observed</span> RONO2 in surface air, and we find that another 10 is contributed by gas-phase monoterpene <span class="hlt">nitrates</span>. <span class="hlt">Observations</span> in the free troposphere show an important contribution from long-lived <span class="hlt">nitrates</span> derived from anthropogenic VOCs. During both campaigns, at least 10 of <span class="hlt">observed</span> boundary layer RONO2 were in the particle phase. We find that aerosol uptake followed by hydrolysis to HNO3 accounts for 60 of simulated gas-phase RONO2 loss in the boundary layer. Other losses are 20 by photolysis to recycle NOx and 15 by dry deposition. RONO2 production accounts for 20 of the net regional NOx sink in the Southeast US in summer, limited by the spatial segregation between BVOC and NOx emissions. This segregation implies that RONO2 production will remain a minor sink for NOx in the Southeast US in the future even as NOx emissions continue to decline. XXXX We have used airborne and ground-based <span class="hlt">observations</span> from two summer 2013 campaigns in the Southeast US (SEAC4RS, SOAS) to better understand the chemistry and impacts of alkyl and multi-functional organic <span class="hlt">nitrates</span> (RONO2). We used the <span class="hlt">observations</span>, along</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28610958','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28610958"><span>Sulfate, <span class="hlt">nitrate</span> and blood pressure - An EPIC interaction between sulfur and nitrogen.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kuhnle, Gunter G; Luben, Robert; Khaw, Kay-Tee; Feelisch, Martin</p> <p>2017-08-01</p> <p><span class="hlt">Nitrate</span> (NO 3 - )-rich foods such as green leafy vegetables are not only part of a healthy diet, but increasingly marketed for primary prevention of cardiovascular disease (CVD) and used as ergogenic aids by competitive athletes. While there is abundant evidence for mild hypotensive effects of <span class="hlt">nitrate</span> on acute application there is limited data on chronic intake in humans, and results from animal studies suggest no long-term benefit. This is important as <span class="hlt">nitrate</span> can also promote the formation of nitrosamines. It is therefore classified as 'probably carcinogenic to humans', although a beneficial effect on CVD risk might compensate for an increased cancer risk. Dietary <span class="hlt">nitrate</span> requires reduction to nitrite (NO 2 - ) by oral commensal bacteria to contribute to the formation of nitric oxide (NO). The extensive crosstalk between NO and hydrogen sulfide (H 2 S) related metabolites may further affect <span class="hlt">nitrate</span>'s bioactivity. Using <span class="hlt">nitrate</span> and nitrite <span class="hlt">concentrations</span> of drinking water - the only dietary source continuously monitored for which detailed data exist - in conjunction with data of >14,000 participants of the EPIC-Norfolk study, we found no inverse associations with blood pressure or CVD risk. Instead, we found a strong interaction with sulfate (SO 4 2- ). At low sulfate <span class="hlt">concentrations</span>, <span class="hlt">nitrate</span> was inversely associated with BP (-4mmHg in top quintile) whereas this was reversed at higher <span class="hlt">concentrations</span> (+3mmHg in top quintile). Our findings have a potentially significant impact for pharmacology, physiology and public health, redirecting our attention from the oral microbiome and mouthwash use to interaction with sulfur-containing dietary constituents. These results also indicate that <span class="hlt">nitrate</span> bioactivation is more complex than hitherto assumed. The modulation of <span class="hlt">nitrate</span> bioactivity by sulfate may render dietary lifestyle interventions aimed at increasing <span class="hlt">nitrate</span> intake ineffective and even reverse potential antihypertensive effects, warranting further investigation</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ACP....1714181U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ACP....1714181U"><span>Seasonal variation of fine- and coarse-mode <span class="hlt">nitrates</span> and related aerosols over East Asia: synergetic <span class="hlt">observations</span> and chemical transport model analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Uno, Itsushi; Osada, Kazuo; Yumimoto, Keiya; Wang, Zhe; Itahashi, Syuichi; Pan, Xiaole; Hara, Yukari; Kanaya, Yugo; Yamamoto, Shigekazu; Fairlie, Thomas Duncan</p> <p>2017-11-01</p> <p>We analyzed long-term fine- and coarse-mode synergetic <span class="hlt">observations</span> of <span class="hlt">nitrate</span> and related aerosols (SO42-, NO3-, NH4+, Na+, Ca2+) at Fukuoka (33.52° N, 130.47° E) from August 2014 to October 2015. A Goddard Earth <span class="hlt">Observing</span> System chemical transport model (GEOS-Chem) including dust and sea salt acid uptake processes was used to assess the <span class="hlt">observed</span> seasonal variation and the impact of long-range transport (LRT) from the Asian continent. For fine aerosols (fSO42-, fNO3-, and fNH4+), numerical results explained the seasonal changes, and a sensitivity analysis excluding Japanese domestic emissions clarified the LRT fraction at Fukuoka (85 % for fSO42-, 47 % for fNO3-, 73 % for fNH4+). <span class="hlt">Observational</span> data confirmed that coarse NO3- (cNO3-) made up the largest proportion (i.e., 40-55 %) of the total <span class="hlt">nitrate</span> (defined as the sum of fNO3-, cNO3-, and HNO3) during the winter, while HNO3 gas constituted approximately 40 % of the total <span class="hlt">nitrate</span> in summer and fNO3- peaked during the winter. Large-scale dust-<span class="hlt">nitrate</span> (mainly cNO3-) outflow from China to Fukuoka was confirmed during all dust events that occurred between January and June. The modeled cNO3- was in good agreement with <span class="hlt">observations</span> between July and November (mainly coming from sea salt NO3-). During the winter, however, the model underestimated cNO3- levels compared to the <span class="hlt">observed</span> levels. The reason for this underestimation was examined statistically using multiple regression analysis (MRA). We used cNa+, nss-cCa2+, and cNH4+ as independent variables to describe the <span class="hlt">observed</span> cNO3- levels; these variables were considered representative of sea salt cNO3-, dust cNO3-, and cNO3- accompanied by cNH4+), respectively. The MRA results explained the <span class="hlt">observed</span> seasonal changes in dust cNO3- and indicated that the dust-acid uptake scheme reproduced the <span class="hlt">observed</span> dust-<span class="hlt">nitrate</span> levels even in winter. The annual average contributions of each component were 43 % (sea salt cNO3-), 19 % (dust cNO3-), and 38 % (cNH4+ term). The MRA dust</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009PhDT.......204P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009PhDT.......204P"><span>The atmospheric chemistry of isoprene- and other multifunctional-<span class="hlt">nitrates</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Perring, Anne Elizabeth</p> <p></p> <p>Formation of alkyl and multifunctional <span class="hlt">nitrates</span> significantly reduces ozone production rates in their source regions, their transport and subsequent chemistry can impact secondary organic aerosol formation and NOy removal rates and they may lead to the re-release of NOx in regions far-removed from the original source. In this dissertation, the atmospheric chemistry of alkyl and multifunctional <span class="hlt">nitrates</span> is investigated through a combination of laboratory and field measurements. In contrast to many previous studies that have focused on <span class="hlt">observations</span> of specific individual <span class="hlt">nitrate</span> compounds, the work described here uses a technique (Thermal Dissociation-Laser Induced Fluorescence or TD-LIF) that allows for measurements of the sum of all alkyl and multifunctional <span class="hlt">nitrates</span>. These <span class="hlt">observations</span> show that alkyl and multifunctional <span class="hlt">nitrates</span> are a significant fraction of NOy in a number of different chemical regimes representing diverse hydrocarbon mixtures. In what follows, I show that their formation impacts both ozone 1 formation and NOy transport in ways that are not accounted for by currently accepted chemical mechanisms. Aircraft measurements are used to constrain <span class="hlt">nitrate</span> yields following isoprene oxidation by OH, the atmospheric lifetimes of these <span class="hlt">nitrates</span>, and the retention rate of the <span class="hlt">nitrate</span> functional group upon oxidation of the initial isoprene <span class="hlt">nitrates</span>. It is found that <span class="hlt">nitrate</span> functionality is maintained upon further oxidation at least 75% of the time indicating that the lifetime of isoprene <span class="hlt">nitrates</span> as a pool of compounds is considerably longer than the lifetime of the individual isoprene <span class="hlt">nitrates</span> with respect to reaction with OH. We examine the products of NO3-initiated oxidation of isoprene in a smog-chamber, propose a detailed reaction scheme, and find that <span class="hlt">nitrates</span> are produced with a yield of 65+/-12%, the majority of which are carbonyl <span class="hlt">nitrates</span>. We investigate the role of alkyl and multifunctional <span class="hlt">nitrates</span> in the Mexico City plume where they are <span class="hlt">observed</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23745394','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23745394"><span>[Removal of <span class="hlt">nitrate</span> from groundwater using permeable reactive barrier].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Xiu-Li; Yang, Jun-Jun; Lu, Xiao-Xia; Zhang, Shu; Hou, Zhen</p> <p>2013-03-01</p> <p>To provide a cost-effective method for the remediation of <span class="hlt">nitrate</span>-polluted groundwater, column experiments were performed to study the removal of <span class="hlt">nitrate</span> by permeable reactive barrier filled with fermented mulch and sand (biowall), and the mechanisms and influence factors were explored. The experimental results showed that the environmental condition in the simulated biowall became highly reduced after three days of operation (oxidation-reduction potential was below - 100 mV), which was favorable for the reduction of <span class="hlt">nitrate</span>. During the 15 days of operation, the removal rate of <span class="hlt">nitrate</span> nitrogen (NO3(-) -N) by the simulated biowall was 80%-90% (NO3(-)-N was reduced from 20 mg x L(-1) in the inlet water to 1.6 mg x L(-1) in the outlet water); the <span class="hlt">concentration</span> of nitrite nitrogen (NO2(-) -N) in the outlet water was below 2.5 mg x L(-1); the <span class="hlt">concentration</span> of ammonium nitrogen (NH4(+) -N) was low in the first two days but increased to about 12 mg x L(-1) since day three. The major mechanisms involved in the removal of <span class="hlt">nitrate</span> nitrogen were adsorption and biodegradation. When increasing the water flow velocity in the simulated biowall, the removal rate of NO3(-) -N was reduced and the <span class="hlt">concentration</span> of NH4(+) -N in the outlet water was significantly reduced. A simulated zeolite wall was set up following the simulated biowall and 98% of the NH4(+) -N could be removed from the water.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19937342','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19937342"><span>Effect of <span class="hlt">nitrate</span> supply and mycorrhizal inoculation on characteristics of tobacco root plasma membrane vesicles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Moche, Martin; Stremlau, Stefanie; Hecht, Lars; Göbel, Cornelia; Feussner, Ivo; Stöhr, Christine</p> <p>2010-01-01</p> <p>Plant plasma membrane (pm) vesicles from mycorrhizal tobacco (Nicotiana tabacum cv. Samsun) roots were isolated with negligible fungal contamination by the aqueous two-phase partitioning technique as proven by fatty acid analysis. Palmitvaccenic acid became apparent as an appropriate indicator for fungal membranes in root pm preparations. The pm vesicles had a low specific activity of the vanadate-sensitive ATPase and probably originated from non-infected root cells. In a phosphate-limited tobacco culture system, root colonisation by the vesicular arbuscular mycorrhizal fungus, Glomus mosseae, is inhibited by external <span class="hlt">nitrate</span> in a dose-dependent way. However, detrimental high <span class="hlt">concentrations</span> of 25 mM <span class="hlt">nitrate</span> lead to the highest colonisation rate <span class="hlt">observed</span>, indicating that the defence system of the plant is impaired. Nitric oxide formation by the pm-bound nitrite:NO reductase increased in parallel with external <span class="hlt">nitrate</span> supply in mycorrhizal roots in comparison to the control plants, but decreased under excess <span class="hlt">nitrate</span>. Mycorrhizal pm vesicles had roughly a twofold higher specific activity as the non-infected control plants when supplied with 10-15 mM <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27481896','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27481896"><span><span class="hlt">Nitrate</span> Protects Cucumber Plants Against Fusarium oxysporum by Regulating Citrate Exudation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Min; Sun, Yuming; Gu, Zechen; Wang, Ruirui; Sun, Guomei; Zhu, Chen; Guo, Shiwei; Shen, Qirong</p> <p>2016-09-01</p> <p>Fusarium wilt causes severe yield losses in cash crops. Nitrogen plays a critical role in the management of plant disease; however, the regulating mechanism is poorly understood. Using biochemical, physiological, bioinformatic and transcriptome approaches, we analyzed how nitrogen forms regulate the interactions between cucumber plants and Fusarium oxysporum f. sp. cucumerinum (FOC). <span class="hlt">Nitrate</span> significantly suppressed Fusarium wilt compared with ammonium in both pot and hydroponic experiments. Fewer FOC colonized the roots and stems under <span class="hlt">nitrate</span> compared with ammonium supply. Cucumber grown with <span class="hlt">nitrate</span> accumulated less fusaric acid (FA) after FOC infection and exhibited increased tolerance to chemical FA by decreasing FA absorption and transportation in shoots. A lower citrate <span class="hlt">concentration</span> was <span class="hlt">observed</span> in <span class="hlt">nitrate</span>-grown cucumbers, which was associated with lower MATE (multidrug and toxin compound extrusion) family gene and citrate synthase (CS) gene expression, as well as lower CS activity. Citrate enhanced FOC spore germination and infection, and increased disease incidence and the FOC population in ammonium-treated plants. Our study provides evidence that <span class="hlt">nitrate</span> protects cucumber plants against F. oxysporum by decreasing root citrate exudation and FOC infection. Citrate exudation is essential for regulating disease development of Fusarium wilt in cucumber plants. © The Author 2016. Published by Oxford University Press on behalf of Japanese Society of Plant Physiologists. All rights reserved. For permissions, please email: journals.permissions@oup.com.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H41F1509H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H41F1509H"><span>Consequences of variation in stream-landscape connections for stream <span class="hlt">nitrate</span> retention and export</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Handler, A. M.; Helton, A. M.; Grimm, N. B.</p> <p>2017-12-01</p> <p>Hydrologic and material connections among streams, the surrounding terrestrial landscape, and groundwater systems fluctuate between extremes in dryland watersheds, yet the consequences of this variation for stream nutrient retention and export remain uncertain. We explored how seasonal variation in hydrologic connection among streams, landscapes, and groundwater affect <span class="hlt">nitrate</span> and ammonium <span class="hlt">concentrations</span> across a dryland stream network and how this variation mediates in-stream <span class="hlt">nitrate</span> uptake and watershed export. We conducted spatial surveys of stream <span class="hlt">nitrate</span> and ammonium <span class="hlt">concentration</span> across the 1200 km2 Oak Creek watershed in central Arizona (USA). In addition, we conducted pulse releases of a solution containing biologically reactive sodium <span class="hlt">nitrate</span>, with sodium chloride as a conservative hydrologic tracer, to estimate <span class="hlt">nitrate</span> uptake rates in the mainstem (Q>1000 L/s) and two tributaries. <span class="hlt">Nitrate</span> and ammonium <span class="hlt">concentrations</span> generally increased from headwaters to mouth in the mainstem. Locally elevated <span class="hlt">concentrations</span> occurred in spring-fed tributaries draining fish hatcheries and larger irrigation ditches, but did not have a substantial effect on the mainstem nitrogen load. Ambient <span class="hlt">nitrate</span> <span class="hlt">concentration</span> (as N) ranged from below the analytical detection limit of 0.005 mg/L to 0.43 mg/L across all uptake experiments. Uptake length—average stream distance traveled for a nutrient atom from the point of release to its uptake—at ambient <span class="hlt">concentration</span> ranged from 250 to 704 m and increased significantly with higher discharge, both across streams and within the same stream on different experiment dates. Vertical uptake velocity and aerial uptake rate ranged from 6.6-10.6 mm min-1 and 0.03 to 1.4 mg N m-2 min-1, respectively. Preliminary analyses indicate potentially elevated nitrogen loading to the lower portion of the watershed during seasonal precipitation events, but overall, the capacity for <span class="hlt">nitrate</span> uptake is high in the mainstem and tributaries. Ongoing work</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19439460','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19439460"><span>Food sources of <span class="hlt">nitrates</span> and nitrites: the physiologic context for potential health benefits.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hord, Norman G; Tang, Yaoping; Bryan, Nathan S</p> <p>2009-07-01</p> <p>The presence of <span class="hlt">nitrates</span> and nitrites in food is associated with an increased risk of gastrointestinal cancer and, in infants, methemoglobinemia. Despite the physiologic roles for <span class="hlt">nitrate</span> and nitrite in vascular and immune function, consideration of food sources of <span class="hlt">nitrates</span> and nitrites as healthful dietary components has received little attention. Approximately 80% of dietary <span class="hlt">nitrates</span> are derived from vegetable consumption; sources of nitrites include vegetables, fruit, and processed meats. Nitrites are produced endogenously through the oxidation of nitric oxide and through a reduction of <span class="hlt">nitrate</span> by commensal bacteria in the mouth and gastrointestinal tract. As such, the dietary provision of <span class="hlt">nitrates</span> and nitrites from vegetables and fruit may contribute to the blood pressure-lowering effects of the Dietary Approaches to Stop Hypertension (DASH) diet. We quantified <span class="hlt">nitrate</span> and nitrite <span class="hlt">concentrations</span> by HPLC in a convenience sample of foods. Incorporating these values into 2 hypothetical dietary patterns that emphasize high-<span class="hlt">nitrate</span> or low-<span class="hlt">nitrate</span> vegetable and fruit choices based on the DASH diet, we found that <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in these 2 patterns vary from 174 to 1222 mg. The hypothetical high-<span class="hlt">nitrate</span> DASH diet pattern exceeds the World Health Organization's Acceptable Daily Intake for <span class="hlt">nitrate</span> by 550% for a 60-kg adult. These data call into question the rationale for recommendations to limit <span class="hlt">nitrate</span> and nitrite consumption from plant foods; a comprehensive reevaluation of the health effects of food sources of <span class="hlt">nitrates</span> and nitrites is appropriate. The strength of the evidence linking the consumption of <span class="hlt">nitrate</span>- and nitrite-containing plant foods to beneficial health effects supports the consideration of these compounds as nutrients.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3084252','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3084252"><span><span class="hlt">Nitrate</span> Paradigm Does Not Hold Up for Sugarcane</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Robinson, Nicole; Brackin, Richard; Vinall, Kerry; Soper, Fiona; Holst, Jirko; Gamage, Harshi; Paungfoo-Lonhienne, Chanyarat; Rennenberg, Heinz; Lakshmanan, Prakash; Schmidt, Susanne</p> <p>2011-01-01</p> <p>Modern agriculture is based on the notion that <span class="hlt">nitrate</span> is the main source of nitrogen (N) for crops, but <span class="hlt">nitrate</span> is also the most mobile form of N and easily lost from soil. Efficient acquisition of <span class="hlt">nitrate</span> by crops is therefore a prerequisite for avoiding off-site N pollution. Sugarcane is considered the most suitable tropical crop for biofuel production, but surprisingly high N fertilizer applications in main producer countries raise doubt about the sustainability of production and are at odds with a carbon-based crop. Examining reasons for the inefficient use of N fertilizer, we hypothesized that sugarcane resembles other giant tropical grasses which inhibit the production of <span class="hlt">nitrate</span> in soil and differ from related grain crops with a confirmed ability to use <span class="hlt">nitrate</span>. The results of our study support the hypothesis that N-replete sugarcane and ancestral species in the Andropogoneae supertribe strongly prefer ammonium over <span class="hlt">nitrate</span>. Sugarcane differs from grain crops, sorghum and maize, which acquired both N sources equally well, while giant grass, Erianthus, displayed an intermediate ability to use <span class="hlt">nitrate</span>. We conclude that discrimination against <span class="hlt">nitrate</span> and a low capacity to store <span class="hlt">nitrate</span> in shoots prevents commercial sugarcane varieties from taking advantage of the high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in fertilized soils in the first three months of the growing season, leaving <span class="hlt">nitrate</span> vulnerable to loss. Our study addresses a major caveat of sugarcane production and affords a strong basis for improvement through breeding cultivars with enhanced capacity to use <span class="hlt">nitrate</span> as well as through agronomic measures that reduce nitrification in soil. PMID:21552564</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27294676','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27294676"><span>Can <span class="hlt">nitrate</span> contaminated groundwater be remediated by optimizing flood irrigation rate with high <span class="hlt">nitrate</span> water in a desert oasis using the WHCNS model?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liang, Hao; Qi, Zhiming; Hu, Kelin; Prasher, Shiv O; Zhang, Yuanpei</p> <p>2016-10-01</p> <p><span class="hlt">Nitrate</span> contamination of groundwater is an environmental concern in intensively cultivated desert oases where this polluted groundwater is in turn used as a major irrigation water resource. However, <span class="hlt">nitrate</span> fluxes from root zone to groundwater are difficult to monitor in this complex system. The objectives of this study were to validate and apply the WHCNS (soil Water Heat Carbon Nitrogen Simulator) model to simulate water drainage and <span class="hlt">nitrate</span> leaching under different irrigation and nitrogen (N) management practices, and to assess the utilization of groundwater <span class="hlt">nitrate</span> as an approach to remediate <span class="hlt">nitrate</span> contaminated groundwater while maintain crop yield. A two-year field experiment was conducted in a corn field irrigated with high <span class="hlt">nitrate</span> groundwater (20 mg N L(-1)) in Alxa, Inner Mongolia, China. The experiment consisted of two irrigation treatments (Istd, standard, 750 mm per season; Icsv, conservation, 570 mm per season) factorially combined with two N fertilization treatments (Nstd, standard, 138 kg ha(-1); Ncsv, conservation, 92 kg ha(-1)). The validated results showed that the WHCNS model simulated values of crop dry matter, yield, soil water content and soil N <span class="hlt">concentration</span> in soil profile all agreed well with the <span class="hlt">observed</span> values. Compared to the standard water management (Istd), the simulated drainage and <span class="hlt">nitrate</span> leaching decreased about 65% and 59%, respectively, under the conservation water management (Icsv). Nearly 55% of input N was lost by leaching under the IstdNstd and IstdNcsv treatments, compared to only 26% under the IcsvNstd and IcsvNcsv treatments. Simulations with more than 240 scenarios combing different levels of irrigation and fertilization indicated that irrigation was the main reason leading to the high risk of <span class="hlt">nitrate</span> leaching, and the <span class="hlt">nitrate</span> in irrigation groundwater can be best utilized without corn yield loss when the total irrigation was reduced from the current 750 mm to 491 mm. This reduced irrigation rate facilitated</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25663374','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25663374"><span>Exposure to vancomycin causes a shift in the microbial community structure without affecting <span class="hlt">nitrate</span> reduction rates in river sediments.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Laverman, Anniet M; Cazier, Thibaut; Yan, Chen; Roose-Amsaleg, Céline; Petit, Fabienne; Garnier, Josette; Berthe, Thierry</p> <p>2015-09-01</p> <p>Antibiotics and antibiotic resistance genes have shown to be omnipresent in the environment. In this study, we investigated the effect of vancomycin (VA) on denitrifying bacteria in river sediments of a Waste Water Treatment Plant, receiving both domestic and hospital waste. We exposed these sediments continuously in flow-through reactors to different VA <span class="hlt">concentrations</span> under denitrifying conditions (<span class="hlt">nitrate</span> addition and anoxia) in order to determine potential <span class="hlt">nitrate</span> reduction rates and changes in sedimentary microbial community structures. The presence of VA had no effect on sedimentary <span class="hlt">nitrate</span> reduction rates at environmental <span class="hlt">concentrations</span>, whereas a change in bacterial (16S rDNA) and denitrifying (nosZ) community structures was <span class="hlt">observed</span> (determined by polymerase chain reaction-denaturing gradient gel electrophoresis). The bacterial and denitrifying community structure within the sediment changed upon VA exposure indicating a selection of a non-susceptible VA population.</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/24645474','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24645474"><span><span class="hlt">Nitrate</span> removal by Fe0/Pd/Cu nano-composite in groundwater.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Hongyuan; Guo, Min; Zhang, Yan</p> <p>2014-01-01</p> <p><span class="hlt">Nitrate</span> pollution in groundwater shows a great threat to the safety of drinking water. Chemical reduction by zero-valent iron is being considered as a promising technique for <span class="hlt">nitrate</span> removal from contaminated groundwater. In this paper, Fe0/Pd/Cu nano-composites were prepared by the liquid-phase reduction method, and batch experiments of <span class="hlt">nitrate</span> reduction by the prepared Fe0/Pd/Cu nano-composites under various operating conditions were carried out. It has been found that nano-Fe0/Pd/Cu composites processed dual functions: catalytic reduction and chemical reduction. The introduction of Pd and Cu not only improved <span class="hlt">nitrate</span> removal rate, but also reduced the generation of ammonia. <span class="hlt">Nitrate</span> removal rate was affected by the amount of Fe0/Pd/Cu, initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, solution pH, dissolved oxygen (DO), reaction temperature, the presence of anions, and organic pollutant. Moreover, <span class="hlt">nitrate</span> reduction by Fe0/Pd/Cu composites followed the pseudo-first-order reaction kinetics. The removal rate of <span class="hlt">nitrate</span> and total nitrogen were about 85% and 40.8%, respectively, under the reaction condition of Fe-6.0%Pd-3.0%Cu amount of 0.25 g/L, pH value of 7.1, DO of 0.42 mg/L, and initial <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of 100 mg/L. Compared with the previous studies with Fe0 alone or Fe-Cu, nano-Fe-6%Pd-3%Cu composites showed a better selectivity to N2.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4546371','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4546371"><span><span class="hlt">Nitrate</span> Accumulation and Leaching in Surface and Ground Water Based on Simulated Rainfall Experiments</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, Hong; Gao, Jian-en; Li, Xing-hua; Zhang, Shao-long; Wang, Hong-jie</p> <p>2015-01-01</p> <p>To evaluate the process of <span class="hlt">nitrate</span> accumulation and leaching in surface and ground water, we conducted simulated rainfall experiments. The experiments were performed in areas of 5.3 m2 with bare slopes of 3° that were treated with two nitrogen fertilizer inputs, high (22.5 g/m2 NH4NO3) and control (no fertilizer), and subjected to 2 hours of rainfall, with. From the 1st to the 7th experiments, the same content of fertilizer mixed with soil was uniformly applied to the soil surface at 10 minutes before rainfall, and no fertilizer was applied for the 8th through 12th experiments. Initially, the time-series <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the surface flow quickly increased, and then it rapidly decreased and gradually stabilized at a low level during the fertilizer experiments. The nitrogen loss in the surface flow primarily occurred during the first 18.6 minutes of rainfall. For the continuous fertilizer experiments, the mean <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the groundwater flow remained at less than 10 mg/L before the 5th experiment, and after the 7th experiment, these <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were greater than 10 mg/L throughout the process. The time-series process of the changing <span class="hlt">concentration</span> in the groundwater flow exhibited the same parabolic trend for each fertilizer experiment. However, the time at which the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> began to change lagged behind the start time of groundwater flow by approximately 0.94 hours on average. The experiments were also performed with no fertilizer. In these experiments, the mean <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of groundwater initially increased continuously, and then, the process exhibited the same parabolic trend as the results of the fertilization experiments. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> decreased in the subsequent experiments. Eight days after the 12 rainfall experiments, 50.53% of the total <span class="hlt">nitrate</span> applied remained in the experimental soil. <span class="hlt">Nitrate</span> residues mainly existed at the surface and in the bottom soil layers, which represents a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26291616','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26291616"><span><span class="hlt">Nitrate</span> Accumulation and Leaching in Surface and Ground Water Based on Simulated Rainfall Experiments.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Hong; Gao, Jian-en; Li, Xing-hua; Zhang, Shao-long; Wang, Hong-jie</p> <p>2015-01-01</p> <p>To evaluate the process of <span class="hlt">nitrate</span> accumulation and leaching in surface and ground water, we conducted simulated rainfall experiments. The experiments were performed in areas of 5.3 m2 with bare slopes of 3° that were treated with two nitrogen fertilizer inputs, high (22.5 g/m2 NH4NO3) and control (no fertilizer), and subjected to 2 hours of rainfall, with. From the 1st to the 7th experiments, the same content of fertilizer mixed with soil was uniformly applied to the soil surface at 10 minutes before rainfall, and no fertilizer was applied for the 8th through 12th experiments. Initially, the time-series <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the surface flow quickly increased, and then it rapidly decreased and gradually stabilized at a low level during the fertilizer experiments. The nitrogen loss in the surface flow primarily occurred during the first 18.6 minutes of rainfall. For the continuous fertilizer experiments, the mean <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in the groundwater flow remained at less than 10 mg/L before the 5th experiment, and after the 7th experiment, these <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were greater than 10 mg/L throughout the process. The time-series process of the changing <span class="hlt">concentration</span> in the groundwater flow exhibited the same parabolic trend for each fertilizer experiment. However, the time at which the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> began to change lagged behind the start time of groundwater flow by approximately 0.94 hours on average. The experiments were also performed with no fertilizer. In these experiments, the mean <span class="hlt">nitrate</span> <span class="hlt">concentration</span> of groundwater initially increased continuously, and then, the process exhibited the same parabolic trend as the results of the fertilization experiments. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> decreased in the subsequent experiments. Eight days after the 12 rainfall experiments, 50.53% of the total <span class="hlt">nitrate</span> applied remained in the experimental soil. <span class="hlt">Nitrate</span> residues mainly existed at the surface and in the bottom soil layers, which represents a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA121040','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA121040"><span>Reactivity of Metal <span class="hlt">Nitrates</span>.</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1982-07-20</p> <p>02NOCuOH Any mechanism suggested for the <span class="hlt">nitration</span> of aromatic systems by titanium(IV) <span class="hlt">nitrate</span> must take into account the <span class="hlt">observed</span> similarity, in...occurs. -26- References 1. For recent reviews see (a) R. B. Moodie and K. Schofield, Accounts Chem. Res., 1976, 9, 287; (b) G. A. Olah and S. J. Kuhn...Ithaca, N.Y., 1969, Chapter VI; L. M. Stock, Prog. Phys. Org. Chem., 1976, 12, 21; J. G. Hoggett , R. B. Moodie, J. R. Penton, and K. Schofield</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21907348','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21907348"><span>Inhaled nitric oxide therapy increases blood nitrite, <span class="hlt">nitrate</span>, and s-nitrosohemoglobin <span class="hlt">concentrations</span> in infants with pulmonary hypertension.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ibrahim, Yomna I; Ninnis, Janet R; Hopper, Andrew O; Deming, Douglas D; Zhang, Amy X; Herring, Jason L; Sowers, Lawrence C; McMahon, Timothy J; Power, Gordon G; Blood, Arlin B</p> <p>2012-02-01</p> <p>To measure the circulating <span class="hlt">concentrations</span> of nitric oxide (NO) adducts with NO bioactivity after inhaled NO (iNO) therapy in infants with pulmonary hypertension. In this single center study, 5 sequential blood samples were collected from infants with pulmonary hypertension before, during, and after therapy with iNO (n = 17). Samples were collected from a control group of hospitalized infants without pulmonary hypertension (n = 16) and from healthy adults for comparison (n = 12). After beginning iNO (20 ppm) whole blood nitrite levels increased approximately two-fold within 2 hours (P<.01). Whole blood <span class="hlt">nitrate</span> levels increased to 4-fold higher than baseline during treatment with 20 ppm iNO (P<.01). S-nitrosohemoglobin increased measurably after beginning iNO (P<.01), whereas iron nitrosyl hemoglobin and total hemoglobin-bound NO-species compounds did not change. Treatment of pulmonary hypertensive infants with iNO results in increases in levels of nitrite, <span class="hlt">nitrate</span>, and S-nitrosohemoglobin in circulating blood. We speculate that these compounds may be carriers of NO bioactivity throughout the body and account for peripheral effects of iNO in the brain, heart, and other organs. Copyright © 2012 Mosby, Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1997/4139/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1997/4139/report.pdf"><span><span class="hlt">Nitrate</span> and selected pesticides in ground water of the Mid-Atlantic region</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ator, Scott W.; Ferrari, Matthew J.</p> <p>1997-01-01</p> <p>Data from more than 850 sites were compiled and analyzed to document the occurrence of <span class="hlt">nitrate</span> and pesticides in ground water of the Mid-Atlantic region as part of the Mid-Atlantic Integrated Assessment program of the U.S. Environmental Protection Agency. Only those data collected by the U.S. Geological Survey as part of regional networks between October 1985 and September 1996 (inclusive) were used in the analyses, and the data were examined to ensure analytical results are not biased toward sites at the same location or sites sampled multiple times during this period. Regional data are available for most of the Mid-Atlantic region but large spatial gaps in available data do exist. <span class="hlt">Nitrate</span> was detected in nearly three-quarters of the samples for which it was analyzed, commonly at levels that suggest anthropogenic sources. Ten percent of samples contained <span class="hlt">nitrate</span> at <span class="hlt">concentrations</span> exceeding the Federal Maximum Contaminant Level (MCL) of 10 milligrams per liter as nitrogen. Pesticide compounds (including atrazine, metolachlor, prometon, simazine, and desethylatrazine, an atrazine degradate) were detected in about half of the samples for which they were analyzed, but rarely at <span class="hlt">concentrations</span> exceeding established MCL?s. The most commonly detected pesticide compounds were desethylatrazine and atrazine. The occurrence of <span class="hlt">nitrate</span> and pesticides in ground water of the Mid-Atlantic region is related to land cover and rock type. Likely sources of <span class="hlt">nitrate</span> and pesticides to ground water include agricultural and urban land-use practices; rock type affects the movement of these compounds into and through the ground-water system. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in the compiled data set are significantly higher in ground water in agricultural areas than in urban or forested areas, but <span class="hlt">concentrations</span> in areas of row crops are statistically indistinguishable from those in areas of pastures. Detection frequencies of atrazine, desethylatrazine, and simazine are indistinguishable among urban</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24964574','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24964574"><span><span class="hlt">Nitrates</span> and nitrites in selected vegetables purchased at supermarkets in Siedlce, Poland.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Raczuk, Jolanta; Wadas, Wanda; Głozak, Katarzyna</p> <p>2014-01-01</p> <p>Vegetables constitute a vital part of the human diet, being the main source of minerals, vitamins, dietary fibre and phytochemicals. They however, also contain <span class="hlt">nitrates</span> and nitrites, which adversely affect human health. To determine <span class="hlt">nitrate</span> and nitrite content in selected vegetables purchased at supermarket chains in Siedlce and to assess their impact on consumer health. Vegetable samples were purchased from local supermarkets in Siedlce, town situated in the Mazovian province (Voivodeship) of Poland. These consisted of 116 samples of nine vegetables types including butterhead and iceberg lettuce, beetroot, white cabbage, carrot, cucumber, radish, tomato and potato collected between April and September 2011. <span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> and nitrite were determined by standard colorimetric methods used in Poland, with results expressed as mg per kg fresh weight of vegetables. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> varied between 10 mg x kg(-1) to 4800 mg x kg(-1). The highest mean <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were found in radishes (2132 mgkg(-1)), butterhead lettuce (1725 mg x kg(-1)), beetroots (1306 mg x kg(-1)) and iceberg lettuce (890 mg x kg(-1)), whereas the lowest were found in cucumber (32 mg x kg(-1)) and tomato (35 mg x kg(-1)). Nitrite levels were also variable; the highest <span class="hlt">concentrations</span> measured were in beetroot (mean 9.19 mg x kg(-1)) whilst much smaller amounts were present in carrot, cucumbers, iceberg lettuce, white cabbage, tomatoes and potatoes. The daily adult consumption of 100 g amounts of the studied vegetables were found not exceed the ADI for both <span class="hlt">nitrates</span> and nitrites. Findings indicated the need for monitoring <span class="hlt">nitrate</span> and nitrite content in radishes, butterhead lettuce and beetroot due to consumer health concerns.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4672050','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4672050"><span>Control of Microbial Sulfide Production with Biocides and <span class="hlt">Nitrate</span> in Oil Reservoir Simulating Bioreactors</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Xue, Yuan; Voordouw, Gerrit</p> <p>2015-01-01</p> <p>Oil reservoir souring by the microbial reduction of sulfate to sulfide is unwanted, because it enhances corrosion of metal infrastructure used for oil production and processing. Reservoir souring can be prevented or remediated by the injection of <span class="hlt">nitrate</span> or biocides, although injection of biocides into reservoirs is not commonly done. Whether combined application of these agents may give synergistic reservoir souring control is unknown. In order to address this we have used up-flow sand-packed bioreactors injected with 2 mM sulfate and volatile fatty acids (VFA, 3 mM each of acetate, propionate and butyrate) at a flow rate of 3 or 6 pore volumes (PV) per day. Pulsed injection of the biocides glutaraldehyde (Glut), benzalkonium chloride (BAC) and cocodiamine was used to control souring. Souring control was determined as the recovery time (RT) needed to re-establish an aqueous sulfide <span class="hlt">concentration</span> of 0.8–1 mM (of the 1.7–2 mM before the pulse). Pulses were either for a long time (120 h) at low <span class="hlt">concentration</span> (long-low) or for a short time (1 h) at high <span class="hlt">concentration</span> (short-high). The short-high strategy gave better souring control with Glut, whereas the long-low strategy was better with cocodiamine. Continuous injection of 2 mM <span class="hlt">nitrate</span> alone was not effective, because 3 mM VFA can fully reduce both 2 mM <span class="hlt">nitrate</span> to nitrite and N2 and, subsequently, 2 mM sulfate to sulfide. No synergy was <span class="hlt">observed</span> for short-high pulsed biocides and continuously injected <span class="hlt">nitrate</span>. However, use of continuous <span class="hlt">nitrate</span> and long-low pulsed biocide gave synergistic souring control with BAC and Glut, as indicated by increased RTs in the presence, as compared to the absence of <span class="hlt">nitrate</span>. Increased production of nitrite, which increases the effectiveness of souring control by biocides, is the most likely cause for this synergy. PMID:26696994</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002AGUFM.H72C0865K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002AGUFM.H72C0865K"><span>Sources and Chronology of <span class="hlt">Nitrate</span> Contamination of Spring Waters: Integrating Science and Policy Decisions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Katz, B. G.; Stevenson, J. A.</p> <p>2002-12-01</p> <p>Human health and ecological concerns have arisen regarding spring waters in Florida as a steady increase in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> has been <span class="hlt">observed</span> during the past 30 years. The extensive aesthetic, cultural, and recreational value of these springs, which also supply water for human consumption and support critical ecological habitats, could be threatened by the presence of <span class="hlt">nitrate</span>. As part of the response to these concerns by the State of Florida, several research studies have used various chemical and isotopic tracers to determine sources of <span class="hlt">nitrate</span> contamination and age of ground water discharging from springs. Since 1997, 60 water samples have been collected from 44 springs and analyzed for isotopic (15N, 3H/3He, 18O, 2H, 13C) and other chemical tracers (CFCs, major ions, dissolved gases, SF6). Delta 15N values of <span class="hlt">nitrate</span> ranged from 2.6 to 12.9 per mil (median = 5.8 per mil) and indicated that <span class="hlt">nitrate</span> in most spring waters originated from synthetic fertilizers. CFCs, 3H/3He, and SF6, used to estimate the residence time of ground water discharging from springs, indicated that spring-water ages ranged from 5 to 39 years. <span class="hlt">Concentrations</span> of these multiple transient tracers are consistent with a two-component hydrologic model with mixtures of varying proportions of young water (less than 8 years) from the shallow part of the aquifer system and older water (20-50 years) from the deeper part of the flow system. Given residence times of 20-40 years for ground water discharging from most springs, it could take decades for <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> to decrease to near background levels, even with immediate reductions in nitrogen inputs to the land surface. These research results are being used by the State of Florida to inform elected officials, water-resource mangers, and planners that decisions about land use today will affect the quality of ground water in springs for decades.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27733522','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27733522"><span>Inorganic <span class="hlt">Nitrate</span> Supplementation in Young and Old Obese Adults Does Not Affect Acute Glucose and Insulin Responses but Lowers Oxidative Stress.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ashor, Ammar W; Chowdhury, Shakir; Oggioni, Clio; Qadir, Othman; Brandt, Kirsten; Ishaq, Abbas; Mathers, John C; Saretzki, Gabriele; Siervo, Mario</p> <p>2016-11-01</p> <p>Aging and obesity are associated with raised oxidative stress and a reduction of nitric oxide (NO) bioavailability, with subsequent decline in insulin sensitivity and endothelial function. Inorganic <span class="hlt">nitrate</span> is converted into NO via a 2-step reduction process and may be an effective nutritional intervention to modify vascular and metabolic functions. This study tested whether inorganic <span class="hlt">nitrate</span> supplementation improved glucose disposal and attenuated the acute effects of hyperglycemia on oxidative stress, inflammation, and vascular function in young and old obese participants. Ten young (aged 18-44 y) and 10 old (aged 55-70 y) obese participants consumed 75 g glucose followed by either potassium <span class="hlt">nitrate</span> (7 mg/kg body weight) or potassium chloride (placebo) in a randomized, double-blind crossover design. Resting blood pressure (BP), endothelial function, and blood biomarkers were measured for 3 h postintervention. Biomarkers included plasma <span class="hlt">nitrate</span>/nitrite (NOx), glucose, insulin, cyclic GMP, interleukin 6, 3-nitrotyrosine, E- and P-selectins, intercellular adhesion molecule 3 (ICAM-3), and thrombomodulin, as well as superoxide in freshly isolated peripheral blood mononuclear cells (PBMCs). Inorganic <span class="hlt">nitrate</span> supplementation did not affect plasma glucose (P = 0.18) or insulin (P = 0.26) responses. The increase in plasma NOx <span class="hlt">concentrations</span> 3 h after the administration of inorganic <span class="hlt">nitrate</span> was significantly higher in young than in old participants (234% increase compared with 149% increase, respectively, P < 0.001). Plasma 3-nitrotyrosine <span class="hlt">concentrations</span> declined significantly after inorganic <span class="hlt">nitrate</span> supplementation compared with placebo (3 h postdose, 46% decrease compared with 27% increase, respectively, P = 0.04), and a similar nonsignificant trend was <span class="hlt">observed</span> for superoxide <span class="hlt">concentrations</span> (3 h postdose, 16% decrease compared with 23% increase, respectively, P = 0.06). Plasma cyclic GMP, ICAM-3, and thrombomodulin <span class="hlt">concentrations</span> differed between young and old</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12145942','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12145942"><span>[<span class="hlt">Nitrate</span> accumulation in vegetables and its residual in vegetable fields].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Zhaohui; Zong, Zhiqiang; Li, Shengxiu; Chen, Baoming</p> <p>2002-05-01</p> <p>Determinations of 11 kinds, 48 varieties of vegetables were carried out at different seasons. The results showed that <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> in 20 vegetables reached Pollution Level 4 (NO3(-)-N > 325 mg.kg-1), which accounted for 41.7% of the total number of the sampled vegetables and included all of the leafy, and most of the melon, root, onion and garlic vegetables. Among them, 5 leafy vegetables even exceeded Level 4 (NO3(-)-N > 700 mg.kg-1). Although leafy vegetables were usually apt to heavily accumulate <span class="hlt">nitrate</span>, most of them were with <span class="hlt">nitrate</span>-N <span class="hlt">concentrations</span> lower than Level 3 (NO3(-)-N < 325 mg.kg-1) in leave blades. Further investigation showed that vegetable soils accumulated more <span class="hlt">nitrates</span> in each layer from 0 cm to 200 cm than did cereal crop soil. The total amount of residual <span class="hlt">nitrate</span>-N was 1358.8 kg.hm-2 in the 200 cm soil profile of usual vegetable fields, and 1411.8 kg.hm-2 and 1520.9 kg.hm-2 in the 2-yaers and the 5-years long plastic greenhouse fields respectively, however that in the cereal crop fields was only 245.4 kg.hm-2. <span class="hlt">Nitrate</span> residual in vegetable soils formed serious threats to underground water in vegetable growing areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1712574T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1712574T"><span>Assessment of Water and <span class="hlt">Nitrate</span>-N deep percolation fluxes in soil as affected by irrigation and nutrient management practices</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsehaye, Habte; Ceglie, Francesco; Mimiola, Giancarlo; dragonetti, giovanna; Lamaddalena, Nicola; Coppola, Antonio</p> <p>2015-04-01</p> <p> conventional agriculture and is a widely adopted organic production system, especially in greenhouse. So called because substituting the conventional agrochemicals with the organic allowed products; ii. AGROMAN was characterized by a cover crop mixture and green manure, which are flattened on the ground; iii. AGROCOM made of the mixed cover crop species and are incorporated into soil together with on-farm composting. The SUBST was characterized by significantly lower water losses than the other two systems. In the first stage, very high <span class="hlt">nitrate</span> fluxes were <span class="hlt">observed</span> in all the three management systems. After, <span class="hlt">nitrate</span> fluxes were practically null for the SUBST system, but in the second stage where some <span class="hlt">nitrate</span> losses comes from the combination of low water fluxes and higher <span class="hlt">concentrations</span>. Similar losses were <span class="hlt">observed</span> for the AGROMAN system, but coming from a combination of higher fluxes and lower <span class="hlt">concentrations</span>. Significant losses were <span class="hlt">observed</span> in the AGROCOM system in the middle stage, coming from the combination of high fluxes and high <span class="hlt">concentrations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70046619','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70046619"><span>Vulnerability of shallow groundwater and drinking-water wells to <span class="hlt">nitrate</span> in the United States</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Nolan, Bernard T.; Hitt, Kerie J.</p> <p>2006-01-01</p> <p>Two nonlinear models were developed at the national scale to (1) predict contamination of shallow ground water (typically < 5 m deep) by <span class="hlt">nitrate</span> from nonpoint sources and (2) to predict ambient <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in deeper supplies used for drinking. The new models have several advantages over previous national-scale approaches. First, they predict <span class="hlt">nitrate</span> <span class="hlt">concentration</span> (rather than probability of occurrence), which can be directly compared with water-quality criteria. Second, the models share a mechanistic structure that segregates nitrogen (N) sources and physical factors that enhance or restrict <span class="hlt">nitrate</span> transport and accumulation in ground water. Finally, data were spatially averaged to minimize small-scale variability so that the large-scale influences of N loading, climate, and aquifer characteristics could more readily be identified. Results indicate that areas with high N application, high water input, well-drained soils, fractured rocks or those with high effective porosity, and lack of attenuation processes have the highest predicted <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. The shallow groundwater model (mean square error or MSE = 2.96) yielded a coefficient of determination (R2) of 0.801, indicating that much of the variation in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> is explained by the model. Moderate to severe <span class="hlt">nitrate</span> contamination is predicted to occur in the High Plains, northern Midwest, and selected other areas. The drinking-water model performed comparably (MSE = 2.00, R2 = 0.767) and predicts that the number of users on private wells and residing in moderately contaminated areas (>5 to ≤10 mg/L <span class="hlt">nitrate</span>) decreases by 12% when simulation depth increases from 10 to 50 m.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17256535','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17256535"><span>Vulnerability of shallow groundwater and drinking-water wells to <span class="hlt">nitrate</span> 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>Nolan, Bernard T; Hitt, Kerie J</p> <p>2006-12-15</p> <p>Two nonlinear models were developed at the national scale to (1) predict contamination of shallow ground water (typically < 5 m deep) by <span class="hlt">nitrate</span> from nonpoint sources and (2) to predict ambient <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in deeper supplies used for drinking. The new models have several advantages over previous national-scale approaches. First, they predict <span class="hlt">nitrate</span> <span class="hlt">concentration</span> (rather than probability of occurrence), which can be directly compared with water-quality criteria. Second, the models share a mechanistic structure that segregates nitrogen (N) sources and physical factors that enhance or restrict <span class="hlt">nitrate</span> transport and accumulation in ground water. Finally, data were spatially averaged to minimize small-scale variability so that the large-scale influences of N loading, climate, and aquifer characteristics could more readily be identified. Results indicate that areas with high N application, high water input, well-drained soils, fractured rocks or those with high effective porosity, and lack of attenuation processes have the highest predicted <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. The shallow groundwater model (mean square error or MSE = 2.96) yielded a coefficient of determination (R(2)) of 0.801, indicating that much of the variation in <span class="hlt">nitrate</span> <span class="hlt">concentration</span> is explained by the model. Moderate to severe <span class="hlt">nitrate</span> contamination is predicted to occur in the High Plains, northern Midwest, and selected other areas. The drinking-water model performed comparably (MSE = 2.00, R(2) = 0.767) and predicts that the number of users on private wells and residing in moderately contaminated areas (>5 to < or =10 mg/L <span class="hlt">nitrate</span>) decreases by 12% when simulation depth increases from 10 to 50 m.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27075614','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27075614"><span>Long-term effect of linseed plus <span class="hlt">nitrate</span> fed to dairy cows on enteric methane emission and <span class="hlt">nitrate</span> and nitrite residuals in milk.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Guyader, J; Doreau, M; Morgavi, D P; Gérard, C; Loncke, C; Martin, C</p> <p>2016-07-01</p> <p>A previous study showed the additive methane (CH4)-mitigating effect of <span class="hlt">nitrate</span> and linseed fed to non-lactating cows. Before practical application, the use of this new strategy in dairy cows requires further investigation in terms of persistency of methanogenesis reduction and absence of residuals in milk products. The objective of this experiment was to study the long-term effect of linseed plus <span class="hlt">nitrate</span> on enteric CH4 emission and performance in dairy cows. We also assessed the effect of this feeding strategy on the presence of <span class="hlt">nitrate</span> residuals in milk products, total tract digestibility, nitrogen (N) balance and rumen fermentation. A total of 16 lactating Holstein cows were allocated to two groups in a randomised design conducted in parallel for 17 weeks. Diets were on a dry matter (DM) basis: (1) control (54% maize silage, 6% hay and 40% <span class="hlt">concentrate</span>; CON) or (2) control plus 3.5% added fat from linseed and 1.8% <span class="hlt">nitrate</span> (LIN+NIT). Diets were equivalent in terms of CP (16%), starch (28%) and NDF (33%), and were offered twice daily. Cows were fed ad libitum, except during weeks 5, 16 and 17 in which feed was restricted to 95% of dry matter intake (DMI) to ensure complete consumption of meals during measurement periods. Milk production and DMI were measured weekly. <span class="hlt">Nitrate</span> and nitrite <span class="hlt">concentrations</span> in milk and milk products were determined monthly. Daily CH4 emission was quantified in open circuit respiration chambers (weeks 5 and 16). Total tract apparent digestibility, N balance and rumen fermentation parameters were determined in week 17. Daily DMI tended to be lower with LIN+NIT from week 4 to 16 (-5.1 kg/day on average). The LIN+NIT diet decreased milk production during 6 non-consecutive weeks (-2.5 kg/day on average). <span class="hlt">Nitrate</span> or nitrite residuals were not detected in milk and associated products. The LIN+NIT diet reduced CH4 emission to a similar extent at the beginning and end of the trial (-47%, g/day; -30%, g/kg DMI; -33%, g/kg fat- and protein</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26078154','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26078154"><span>Modified graphene oxide sensors for ultra-sensitive detection of <span class="hlt">nitrate</span> ions in water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ren, Wen; Mura, Stefania; Irudayaraj, Joseph M K</p> <p>2015-10-01</p> <p><span class="hlt">Nitrate</span> ions is a very common contaminant in drinking water and has a significant impact on the environment, necessitating routine monitoring. Due to its chemical and physical properties, it is hard to directly detect <span class="hlt">nitrate</span> ions with high sensitivity in a simple and inexpensive manner. Herein with amino group modified graphene oxide (GO) as a sensing element, we show a direct and ultra-sensitive method to detect <span class="hlt">nitrate</span> ions, at a lowest detected <span class="hlt">concentration</span> of 5 nM in river water samples, much lower than the reported methods based on absorption spectroscopy. Furthermore, unlike the reported strategies based on absorption spectroscopy wherein the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> is determined by monitoring an increase in aggregation of gold nanoparticles (GNPs), our method evaluates the <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> ions based on reduction in aggregation of GNPs for monitoring in real samples. To improve sensitivity, several optimizations were performed, including the assessment of the amount of modified GO required, <span class="hlt">concentration</span> of GNPs and incubation time. The detection methodology was characterized by zeta potential, TEM and SEM. Our results indicate that an enrichment of modified GO with <span class="hlt">nitrate</span> ions contributed to excellent sensitivity and the entire detection procedure could be completed within 75 min with only 20 μl of sample. This simple and rapid methodology was applied to monitor <span class="hlt">nitrate</span> ions in real samples with excellent sensitivity and minimum pretreatment. The proposed approach paves the way for a novel means to detect anions in real samples and highlights the potential of GO based detection strategy for water quality monitoring. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1998/4040a/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1998/4040a/report.pdf"><span><span class="hlt">Nitrate</span> and pesticides in ground water in the eastern San Joaquin Valley, California : occurrence and trends</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Burow, Karen R.; Stork, Sylvia V.; Dubrovsky, N.M.</p> <p>1998-01-01</p> <p>The occurrence of <span class="hlt">nitrate</span> and pesticides in ground water in California's eastern San Joaquin Valley may be greatly influenced by the long history of intensive farming and irrigation and the generally permeable sediments. This study, which is part of the U.S. Geological Survey National Water-Quality Assessment Program, was done to assess the quality of the ground water and to do a preliminary evaluation of the temporal trends in <span class="hlt">nitrate</span> and pesticides in the alluvial fans of the eastern San Joaquin Valley. Ground-water samples were collected from 30 domestic wells in 1995 (each well was sampled once during 1995). The results of the analyses of these samples were related to various physical and chemical factors in an attempt to understand the processes that control the occurrence and the <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> and pesticides. A preliminary evaluation of the temporal trends in the occurrence and the <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> and pesticides was done by comparing the results of the analyses of the 1995 ground-water samples with the results of the analyses of the samples collected in 1986-87 as part of the U.S. Geological Survey Regional Aquifer-System Analysis Program. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> (dissolved <span class="hlt">nitrate</span> plus nitrite, as nitrogen) in ground water sampled in 1995 ranged from less than 0.05 to 34 milligrams per liter, with a median <span class="hlt">concentration</span> of 4.6 milligrams per liter. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> exceeded the maximum contaminant level of 10 milligrams per liter (as nitrogen) in 5 of the 30 ground-water samples (17 percent), whereas 12 of the 30 samples (40 percent) had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> less than 3.0 milligrams per liter. The high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were associated with recently recharged, well-oxygenated ground water that has been affected by agriculture (indicated by the positive correlations between <span class="hlt">nitrate</span>, dissolved-oxygen, tritium, and specific conductance). Twelve pesticides were detected in 21 of the 30 ground-water samples (70 percent) in 1995</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.H31E0830L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.H31E0830L"><span>Differentiation among Multiple Sources of Anthropogenic <span class="hlt">Nitrate</span> in a Complex Groundwater System using Dual Isotope Systematics: A case study from Mortandad Canyon, New Mexico</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Larson, T. E.; Perkins, G.; Longmire, P.; Heikoop, J. M.; Fessenden, J. E.; Rearick, M.; Fabyrka-Martin, J.; Chrystal, A. E.; Dale, M.; Simmons, A. M.</p> <p>2009-12-01</p> <p>The groundwater system beneath Los Alamos National Laboratory has been affected by multiple sources of anthropogenic <span class="hlt">nitrate</span> contamination. Average NO3-N <span class="hlt">concentrations</span> of up to 18.2±1.7 mg/L have been found in wells in the perched intermediate aquifer beneath one of the more affected sites within Mortandad Canyon. Sources of <span class="hlt">nitrate</span> potentially reaching the alluvial and intermediate aquifers include: (1) sewage effluent, (2) neutralized nitric acid, (3) neutralized 15N-depleted nitric acid (treated waste from an experiment enriching nitric acid in 15N), and (4) natural background <span class="hlt">nitrate</span>. Each of these sources is unique in δ18O and δ15N space. Using <span class="hlt">nitrate</span> stable isotope ratios, a mixing model for the three anthropogenic sources of <span class="hlt">nitrate</span> was established, after applying a linear subtraction of the background component. The spatial and temporal variability in <span class="hlt">nitrate</span> contaminant sources through Mortandad Canyon is clearly shown in ternary plots. While microbial denitrification has been shown to change groundwater <span class="hlt">nitrate</span> stable isotope ratios in other settings, the redox potential, relatively high dissolved oxygen content, increasing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> over time, and lack of <span class="hlt">observed</span> NO2 in these wells suggest minimal changes to the stable isotope ratios have occurred. Temporal trends indicate that the earliest form of anthropogenic <span class="hlt">nitrate</span> in this watershed was neutralized nitric acid. Alluvial wells preserve a trend of decreasing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and mixing models show decreasing contributions of 15N-depleted nitric acid. Nearby intermediate wells show increasing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and mixing models indicate a larger component derived from 15N-depleted nitric acid. These data indicate that the pulse of neutralized 15N-depleted nitric acid that was released into Mortandad Canyon between 1986 and 1989 has infiltrated through the alluvial aquifer and is currently affecting two intermediate wells. This hypothesis is consistent with previous</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70045272','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70045272"><span>Vulnerability of streams to legacy <span class="hlt">nitrate</span> sources</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Tesoriero, Anthony J.; Duff, John H.; Saad, David A.; Spahr, Norman E.; Wolock, David M.</p> <p>2013-01-01</p> <p>The influence of hydrogeologic setting on the susceptibility of streams to legacy <span class="hlt">nitrate</span> was examined at seven study sites having a wide range of base flow index (BFI) values. BFI is the ratio of base flow to total streamflow volume. The portion of annual stream <span class="hlt">nitrate</span> loads from base flow was strongly correlated with BFI. Furthermore, dissolved oxygen <span class="hlt">concentrations</span> in streambed pore water were significantly higher in high BFI watersheds than in low BFI watersheds suggesting that geochemical conditions favor <span class="hlt">nitrate</span> transport through the bed when BFI is high. Results from a groundwater-surface water interaction study at a high BFI watershed indicate that decades old <span class="hlt">nitrate</span>-laden water is discharging to this stream. These findings indicate that high <span class="hlt">nitrate</span> levels in this stream may be sustained for decades to come regardless of current practices. It is hypothesized that a first approximation of stream vulnerability to legacy nutrients may be made by geospatial analysis of watersheds with high nitrogen inputs and a strong connection to groundwater (e.g., high BFI).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70185377','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70185377"><span><span class="hlt">Nitrate</span> transport and transformation processes in unsaturated porous media</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Tindall, James A.; Petrusak, Robin L.; McMahon, Peter B.</p> <p>1995-01-01</p> <p>A series of experiments was conducted on two contrasting agricultural soils to <span class="hlt">observe</span> the influence of soil texture, preferential flow, and plants on <span class="hlt">nitrate</span> transport and denitrification under unsaturated conditions. Calcium <span class="hlt">nitrate</span> fertilizer was applied to the surface of four large undisturbed soil cores (30 cm diameter by 40 cm height). Two of the cores were a structured clay obtained from central Missouri and two were an unstructured fine sand obtained from central Florida. The cores were irrigated daily and maintained at a matric potential of -20 kPa, representative of soil tension in the rooting zone of irrigated agricultural fields. Volumetric water content (θ), <span class="hlt">concentration</span> of <span class="hlt">nitrate</span>-N in the soil solution, and nitrous oxide flux at the surface, 10, 20, and 30 cm were monitored daily. Leaching loss of surface-applied N03− -N was significant in both the sand and the clay. In unplanted sand cores, almost all of the applied <span class="hlt">nitrate</span> was leached below 30 cm within 10 days. Gaseous N loss owing to denitrification was no greater than 2% of the <span class="hlt">nitrate</span>-N applied to the unplanted sand cores and, in general, was less than 1 %. Although leaching was somewhat retarded in the clay cores, about 60% of the applied <span class="hlt">nitrate</span>-N was leached from the unplanted clay soil in 5–6 weeks. Under unsaturated conditions, the clay had little to no tendency to denitrify despite the greater moisture content of the clay and retarded leaching of <span class="hlt">nitrate</span> in the clay. The planted sand cores had surprisingly large gaseous N loss owing to denitrification, as much as 17% of the <span class="hlt">nitrate</span>-N. Results from both the clay and sand experiments show that the dynamics of <span class="hlt">nitrate</span> transport and transformation in unsaturated soils are affected by small, localized variations in the soil moisture content profile, the gaseous diffusion coefficient of the soil, the rate at which the <span class="hlt">nitrate</span> pulse passes through the soil, the solubility of N2O and N2 and the diffusion of the gasses through the soil</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://www.ncbi.nlm.nih.gov/pubmed/24758896','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24758896"><span>Effects of nitrogen fertilizers on the growth and <span class="hlt">nitrate</span> content of lettuce (Lactuca sativa L.).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Cheng-Wei; Sung, Yu; Chen, Bo-Ching; Lai, Hung-Yu</p> <p>2014-04-22</p> <p>Nitrogen is an essential element for plant growth and development; however, due to environmental pollution, high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> accumulate in the edible parts of these leafy vegetables, particularly if excessive nitrogen fertilizer has been applied. Consuming these crops can harm human health; thus, developing a suitable strategy for the agricultural application of nitrogen fertilizer is important. Organic, inorganic, and liquid fertilizers were utilized in this study to investigate their effect on <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and lettuce growth. The results of this pot experiment show that the total nitrogen <span class="hlt">concentration</span> in soil and the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in lettuce increased as the amount of nitrogen fertilizer increased. If the recommended amount of inorganic fertilizer (200 kg·N·ha⁻¹) is used as a standard of comparison, lettuce augmented with organic fertilizers (200 kg·N·ha⁻¹) have significantly longer and wider leaves, higher shoot, and lower <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4025000','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4025000"><span>Effects of Nitrogen Fertilizers on the Growth and <span class="hlt">Nitrate</span> Content of Lettuce (Lactuca sativa L.)</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Liu, Cheng-Wei; Sung, Yu; Chen, Bo-Ching; Lai, Hung-Yu</p> <p>2014-01-01</p> <p>Nitrogen is an essential element for plant growth and development; however, due to environmental pollution, high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> accumulate in the edible parts of these leafy vegetables, particularly if excessive nitrogen fertilizer has been applied. Consuming these crops can harm human health; thus, developing a suitable strategy for the agricultural application of nitrogen fertilizer is important. Organic, inorganic, and liquid fertilizers were utilized in this study to investigate their effect on <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and lettuce growth. The results of this pot experiment show that the total nitrogen <span class="hlt">concentration</span> in soil and the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in lettuce increased as the amount of nitrogen fertilizer increased. If the recommended amount of inorganic fertilizer (200 kg·N·ha−1) is used as a standard of comparison, lettuce augmented with organic fertilizers (200 kg·N·ha−1) have significantly longer and wider leaves, higher shoot, and lower <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>. PMID:24758896</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2012/5235/pdf/sir2012-5235_508.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2012/5235/pdf/sir2012-5235_508.pdf"><span>Residence time, chemical and isotopic analysis of <span class="hlt">nitrate</span> in the groundwater and surface water of a small agricultural watershed in the Coastal Plain, Bucks Branch, Sussex County, Delaware</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Clune, John W.; Denver, Judith M.</p> <p>2012-01-01</p> <p><span class="hlt">Nitrate</span> is a common contaminant in groundwater and surface water throughout the Nation, and water-resource managers need more detailed small-scale watershed research to guide conservation efforts aimed at improving water quality. <span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> in Bucks Branch are among the highest in the state of Delaware and a scientific investigation was performed to provide water-quality information to assist with the management of agriculture and water resources. A combination of major-ion chemistry, nitrogen isotopic composition and age-dating techniques was used to estimate the residence time and provide a chemical and isotopic analysis of <span class="hlt">nitrate</span> in the groundwater in the surficial aquifer of the Bucks Branch watershed in Sussex County, Delaware. The land use was more than 90 percent agricultural and most nitrogen inputs were from manure and fertilizer. The apparent median age of sampled groundwater is 18 years and the estimated residence time of groundwater contributing to the streamflow for the entire Bucks Branch watershed at the outlet is approximately 19 years. <span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> exceeded the U.S. Environmental Protection Agency drinking-water standard of 10 milligrams per liter (as nitrogen) in 60 percent of groundwater samples and 42 percent of surface-water samples. The overall geochemistry in the Bucks Branch watershed indicates that agriculture is the predominant source of <span class="hlt">nitrate</span> contamination and the <span class="hlt">observed</span> patterns in major-ion chemistry are similar to those <span class="hlt">observed</span> in other studies on the Mid-Atlantic Coastal Plain. The pattern of enrichment in nitrogen and oxygen isotopes (δ15N and δ18O) of <span class="hlt">nitrate</span> in groundwater and surface water indicates there is some loss of <span class="hlt">nitrate</span> through denitrification, but this process is not sufficient to remove all of the <span class="hlt">nitrate</span> from groundwater discharging to streams, and <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> in streams remain elevated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17982383','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17982383"><span>[Can <span class="hlt">nitrates</span> lead to indirect toxicity?].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hamon, M</p> <p>2007-09-01</p> <p>For many years, <span class="hlt">nitrates</span> have been used, at low dosages, as an additive in salted food. New laws have been promulgated to limit their <span class="hlt">concentration</span> in water due to increased levels found in soils, rivers and even the aquifer. Although <span class="hlt">nitrate</span> ions themselves have not toxic properties, bacterial reduction into nitrite ions (occurring even in aqueous medium) can lead to nitrous anhydride, which in turn generates nitrosonium ions. Nitrosium ions react with secondary amine to give nitrosamines, many of which are cancer-inducing agents at very low doses. Opinions on this toxicity are clear-cut and difficult to reconcile. In fact, increased levels are due, in a large part, to the use of <span class="hlt">nitrates</span> as fertiliéers but also to bacterial transformation of human and animal nitrogenous wastes such as urea.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70023992','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70023992"><span>Simulation of stream discharge and transport of <span class="hlt">nitrate</span> and selected herbicides in the Mississippi River Basin</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Broshears, R.E.; Clark, G.M.; Jobson, H.E.</p> <p>2001-01-01</p> <p>Stream discharge and the transport of <span class="hlt">nitrate</span>, atrazine, and metolachlor in the Mississippi River Basin were simulated using the DAFLOW/BLTM hydrologic model. The simulated domain for stream discharge included river reaches downstream from the following stations in the National Stream Quality Accounting Network: Mississippi River at Clinton, IA; Missouri River at Hermann, MO: Ohio River at Grand Chain, IL: And Arkansas River at Little Rock, AR. Coefficients of hydraulic geometry were calibrated using data from water year 1996; the model was validated by favourable simulation of <span class="hlt">observed</span> discharges in water years 1992-1994. The transport of <span class="hlt">nitrate</span>, atrazine, and metolachlor was simulated downstream from the Mississippi River at Thebes, IL, and the Ohio River at Grand Chain. Simulated <span class="hlt">concentrations</span> compared favourably with <span class="hlt">observed</span> <span class="hlt">concentrations</span> at Baton Rouge, LA. Development of this model is a preliminary step in gaining a more quantitative understanding of the sources and fate of nutrients and pesticides delivered from the Mississippi River Basin to the Gulf of Mexico.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15..252N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15..252N"><span><span class="hlt">Nitrate</span> Remediation of Soil and Groundwater Using Phytoremediation: Transfer of Nitrogen Containing Compounds from the Subsurface to Surface Vegetation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nelson, Sheldon</p> <p>2013-04-01</p> <p><span class="hlt">Nitrate</span> Remediation of Soil and Groundwater Using Phytoremediation: Transfer of Nitrogen Containing Compounds from the Subsurface to Surface Vegetation Sheldon Nelson Chevron Energy Technology Company 6001 Bollinger Canyon Road San Ramon, California 94583 snne@chevron.com The basic concept of using a plant-based remedial approach (phytoremediation) for nitrogen containing compounds is the incorporation and transformation of the inorganic nitrogen from the soil and/or groundwater (<span class="hlt">nitrate</span>, ammonium) into plant biomass, thereby removing the constituent from the subsurface. There is a general preference in many plants for the ammonium nitrogen form during the early growth stage, with the uptake and accumulation of <span class="hlt">nitrate</span> often increasing as the plant matures. The synthesis process refers to the variety of biochemical mechanisms that use ammonium or <span class="hlt">nitrate</span> compounds to primarily form plant proteins, and to a lesser extent other nitrogen containing organic compounds. The shallow soil at the former warehouse facility test site is impacted primarily by elevated <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>, with a minimal presence of ammonium. Dissolved <span class="hlt">nitrate</span> (NO3-) is the primary dissolved nitrogen compound in on-site groundwater, historically reaching <span class="hlt">concentrations</span> of 1000 mg/L. The initial phases of the project consisted of the installation of approximately 1750 trees, planted in 10-foot centers in the areas impacted by <span class="hlt">nitrate</span> and ammonia in the shallow soil and groundwater. As of the most recent groundwater analytical data, dissolved <span class="hlt">nitrate</span> reductions of 40% to 96% have been <span class="hlt">observed</span> in monitor wells located both within, and immediately downgradient of the planted area. In summary, an evaluation of time series groundwater analytical data from the initial planted groves suggests that the trees are an effective means of transfering nitrogen compounds from the subsurface to overlying vegetation. The mechanism of <span class="hlt">concentration</span> reduction may be the uptake of residual <span class="hlt">nitrate</span> from the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1150423','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1150423"><span>Light-Dark Changes in Cytosolic <span class="hlt">Nitrate</span> Pools Depend on <span class="hlt">Nitrate</span> Reductase Activity in Arabidopsis Leaf Cells1[w</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Cookson, Sarah J.; Williams, Lorraine E.; Miller, Anthony J.</p> <p>2005-01-01</p> <p>Several different cellular processes determine the size of the metabolically available <span class="hlt">nitrate</span> pool in the cytoplasm. These processes include not only ion fluxes across the plasma membrane and tonoplast but also assimilation by the activity of <span class="hlt">nitrate</span> reductase (NR). In roots, the maintenance of cytosolic <span class="hlt">nitrate</span> activity during periods of <span class="hlt">nitrate</span> starvation and resupply (M. van der Leij, S.J. Smith, A.J. Miller [1998] Planta 205: 64–72; R.-G. Zhen, H.-W. Koyro, R.A. Leigh, A.D. Tomos, A.J. Miller [1991] Planta 185: 356–361) suggests that this pool is regulated. Under <span class="hlt">nitrate</span>-replete conditions vacuolar <span class="hlt">nitrate</span> is a membrane-bound store that can release <span class="hlt">nitrate</span> to the cytoplasm; after depletion of cytosolic <span class="hlt">nitrate</span>, tonoplast transporters would serve to restore this pool. To study the role of assimilation, specifically the activity of NR in regulating the size of the cytosolic <span class="hlt">nitrate</span> pool, we have compared wild-type and mutant plants. In leaf mesophyll cells, light-to-dark transitions increase cytosolic <span class="hlt">nitrate</span> activity (1.5–2.8 mm), and these changes were reversed by dark-to-light transitions. Such changes were not <span class="hlt">observed</span> in nia1nia2 NR-deficient plants indicating that this change in cytosolic <span class="hlt">nitrate</span> activity was dependent on the presence of functional NR. Furthermore, in the dark, the steady-state cytosolic <span class="hlt">nitrate</span> activities were not statistically different between the two types of plant, indicating that NR has little role in determining resting levels of <span class="hlt">nitrate</span>. Epidermal cells of both wild type and NR mutants had cytosolic <span class="hlt">nitrate</span> activities that were not significantly different from mesophyll cells in the dark and were unaltered by dark-to-light transitions. We propose that the NR-dependent changes in cytosolic <span class="hlt">nitrate</span> provide a cellular mechanism for the diurnal changes in vacuolar <span class="hlt">nitrate</span> storage, and the results are discussed in terms of the possible signaling role of cytosolic <span class="hlt">nitrate</span>. PMID:15908593</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=93047','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=93047"><span>Presence of <span class="hlt">Nitrate</span>-Accumulating Sulfur Bacteria and Their Influence on Nitrogen Cycling in a Shallow Coastal Marine Sediment</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Sayama, Mikio</p> <p>2001-01-01</p> <p><span class="hlt">Nitrate</span> flux between sediment and water, <span class="hlt">nitrate</span> <span class="hlt">concentration</span> profile at the sediment-water interface, and in situ sediment denitrification activity were measured seasonally at the innermost part of Tokyo Bay, Japan. For the determination of sediment <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, undisturbed sediment cores were sectioned into 5-mm depth intervals and each segment was stored frozen at −30°C. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> was determined for the supernatants after centrifuging the frozen and thawed sediments. <span class="hlt">Nitrate</span> in the uppermost sediment showed a remarkable seasonal change, and its seasonal maximum of up to 400 μM was found in October. The directions of the diffusive <span class="hlt">nitrate</span> fluxes predicted from the interfacial <span class="hlt">concentration</span> gradients were out of the sediment throughout the year. In contrast, the directions of the total <span class="hlt">nitrate</span> fluxes measured by the whole-core incubation were into the sediment at all seasons. This contradiction between directions indicates that a large part of the <span class="hlt">nitrate</span> pool extracted from the frozen surface sediments is not a pore water constituent, and preliminary examinations demonstrated that the <span class="hlt">nitrate</span> was contained in the intracellular vacuoles of filamentous sulfur bacteria dwelling on or in the surface sediment. Based on the comparison between in situ sediment denitrification activity and total <span class="hlt">nitrate</span> flux, it is suggested that intracellular <span class="hlt">nitrate</span> cannot be directly utilized by sediment denitrification, and the probable fate of the intracellular <span class="hlt">nitrate</span> is hypothesized to be dissimilatory reduction to ammonium. The presence of <span class="hlt">nitrate</span>-accumulating sulfur bacteria therefore may lower nature's self-purification capacity (denitrification) and exacerbate eutrophication in shallow coastal marine environments. PMID:11472923</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4635503','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4635503"><span>High-affinity <span class="hlt">nitrate</span>/nitrite transporters NrtA and NrtB of Aspergillus nidulans exhibit high specificity and different inhibitor sensitivity</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Akhtar, Naureen; Karabika, Eugenia; Kinghorn, James R.; Glass, Anthony D.M.; Unkles, Shiela E.</p> <p>2015-01-01</p> <p>The NrtA and NrtB <span class="hlt">nitrate</span> transporters are paralogous members of the major facilitator superfamily in Aspergillus nidulans. The availability of loss-of-function mutations allowed individual investigation of the specificity and inhibitor sensitivity of both NrtA and NrtB. In this study, growth response tests were carried out at a growth-limiting <span class="hlt">concentration</span> of <span class="hlt">nitrate</span> (1 mM) as the sole nitrogen source, in the presence of a number of potential <span class="hlt">nitrate</span> analogues at various <span class="hlt">concentrations</span>, to evaluate their effect on <span class="hlt">nitrate</span> transport. Both chlorate and chlorite inhibited fungal growth, with chlorite exerting the greater inhibition. The main transporter of <span class="hlt">nitrate</span>, NrtA, proved to be more sensitive to chlorate than the minor transporter, NrtB. Similarly, the cation caesium was shown to exert differential effects, strongly inhibiting the activity of NrtB, but not NrtA. In contrast, no inhibition of <span class="hlt">nitrate</span> uptake by NrtA or NrtB transporters was <span class="hlt">observed</span> in either growth tests or uptake assays in the presence of bicarbonate, formate, malonate or oxalate (sulphite could not be tested in uptake assays owing to its reaction with <span class="hlt">nitrate</span>), indicating significant specificity of <span class="hlt">nitrate</span> transport. Kinetic analyses of <span class="hlt">nitrate</span> uptake revealed that both chlorate and chlorite inhibited NrtA competitively, while these same inhibitors inhibited NrtB in a non-competitive fashion. The caesium ion appeared to inhibit NrtA in a non-competitive fashion, while NrtB was inhibited uncompetitively. The results provide further evidence of the distinctly different characteristics as well as the high specificity of <span class="hlt">nitrate</span> uptake by these two transporters. PMID:25855763</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26654079','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26654079"><span>On-site semi-quantitative analysis for ammonium <span class="hlt">nitrate</span> detection using digital image colourimetry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Choodum, Aree; Boonsamran, Pichapat; NicDaeid, Niamh; Wongniramaikul, Worawit</p> <p>2015-12-01</p> <p> the various imaging devices and spectrophotometer used to determine ammonium <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>. Five post-blast soil samples were also analysed and pink-violet product were <span class="hlt">observed</span> using Griess's test without zinc reduction indicating the absence of ammonium <span class="hlt">nitrate</span>. This demonstrates significant potential for practical and accurate on-site semi-quantitative determinations of ammonium <span class="hlt">nitrate</span> <span class="hlt">concentration</span>. Copyright © 2015 The Chartered Society of Forensic Sciences. Published by Elsevier Ireland Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title46-vol5/pdf/CFR-2011-title46-vol5-sec148-205.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title46-vol5/pdf/CFR-2011-title46-vol5-sec148-205.pdf"><span>46 CFR 148.205 - Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers.</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>2011-10-01</p> <p>... 46 Shipping 5 2011-10-01 2011-10-01 false Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers. 148... Materials § 148.205 Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers. (a) This section applies to the stowage and transportation in bulk of ammonium <span class="hlt">nitrate</span> and the following fertilizers composed of uniform...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title46-vol5/pdf/CFR-2014-title46-vol5-sec148-205.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title46-vol5/pdf/CFR-2014-title46-vol5-sec148-205.pdf"><span>46 CFR 148.205 - Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-10-01</p> <p>... 46 Shipping 5 2014-10-01 2014-10-01 false Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers. 148... Materials § 148.205 Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers. (a) This section applies to the stowage and transportation in bulk of ammonium <span class="hlt">nitrate</span> and the following fertilizers composed of uniform...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title46-vol5/pdf/CFR-2012-title46-vol5-sec148-205.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title46-vol5/pdf/CFR-2012-title46-vol5-sec148-205.pdf"><span>46 CFR 148.205 - Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers.</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>2012-10-01</p> <p>... 46 Shipping 5 2012-10-01 2012-10-01 false Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers. 148... Materials § 148.205 Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers. (a) This section applies to the stowage and transportation in bulk of ammonium <span class="hlt">nitrate</span> and the following fertilizers composed of uniform...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title46-vol5/pdf/CFR-2013-title46-vol5-sec148-205.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title46-vol5/pdf/CFR-2013-title46-vol5-sec148-205.pdf"><span>46 CFR 148.205 - Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers.</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>2013-10-01</p> <p>... 46 Shipping 5 2013-10-01 2013-10-01 false Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers. 148... Materials § 148.205 Ammonium <span class="hlt">nitrate</span> and ammonium <span class="hlt">nitrate</span> fertilizers. (a) This section applies to the stowage and transportation in bulk of ammonium <span class="hlt">nitrate</span> and the following fertilizers composed of uniform...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.4074L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.4074L"><span>Modelling <span class="hlt">nitrate</span> from land-surface to wells-perforations under Mediterranean agricultural land: success, failure, and future scenarios</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Levy, Yehuda; Chefetz, Benny; Shapira, Roi; Kurtzman, Daniel</p> <p>2017-04-01</p> <p>Contamination of groundwater resources by <span class="hlt">nitrate</span> due to leaching under agricultural land is probably the most troublesome agriculture-related water contamination, worldwide. Deep soil sampling (10 m) were used for calibrating vertical flow and nitrogen-transport numerical models of the unsaturated zone, under different agricultural land uses. Vegetables fields (potato and strawberries) and deciduous (persimmon) orchards in the Sharon area overlaying the coastal aquifer of Israel, were examined. Average <span class="hlt">nitrate</span>-nitrogen fluxes below vegetables fields were 210-290 kg ha-1 a-1 and under deciduous orchards were 110-140 kg ha-1 a-1. The output water and <span class="hlt">nitrate</span>-nitrogen fluxes of the unsaturated zone models were used as input for a three dimensional flow and <span class="hlt">nitrate</span>-transport model in the aquifer under an area of 13.3 square kilometers of agricultural land. The area was subdivided to 4 agricultural land-uses: vegetables, deciduous, citrus orchards and non-cultivated. Fluxes of water and <span class="hlt">nitrate</span>-nitrogen below citrus orchards were taken from a previous study in this area (Kurtzman et al., 2013, j. Contam. Hydrol.). The groundwater flow model was calibrated to well heads only by changing the hydraulic conductivity while transient recharge fluxes were constraint to the bottom-fluxes of the unsaturated zone flow models. The <span class="hlt">nitrate</span>-transport model in the aquifer, which was fed at the top by the <span class="hlt">nitrate</span> fluxes of the unsaturated zone models, succeeded in reconstructing the average <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the wells. On the other hand, this transport model failed in calculating the high <span class="hlt">concentrations</span> in the most contaminated wells and the large spatial variability of <span class="hlt">nitrate-concentrations</span> in the aquifer. In order to reconstruct the spatial variability and enable predictions <span class="hlt">nitrate</span>-fluxes from the unsaturated zone were multiplied by local multipliers. This action was rationalized by the fact that the high <span class="hlt">concentrations</span> in some wells cannot be explained by regular</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.B23A0558R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.B23A0558R"><span>Persistent Influences of the 2002 Hayman Fire on Stream <span class="hlt">Nitrate</span> and Dissolved Organic Carbon</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rhoades, C.; Pierson, D. N.; Fegel, T. S., II; Chow, A. T.; Covino, T. P.</p> <p>2016-12-01</p> <p>Large, high severity wildfires alter the physical and biological conditions that determine how watersheds retain and release nutrients and regulate stream water quality. For five years after the 2002 Hayman Fire burned in Colorado conifer forests, stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and export increased steadily in watersheds with extensive high-severity burning. Stream temperature and turbidity also increased in relation to the extent of high-severity burning and remained elevated above background levels throughout the initial five year post-fire period. Our recent sampling documents that 14 years after the Hayman Fire stream <span class="hlt">nitrate</span> remains an order of magnitude higher in extensively-burned (35-90%) compared to unburned watersheds (0.2 vs 2.8 mg L-1). <span class="hlt">Nitrate</span> represents 83% of the total dissolved N in extensively-burned watersheds compared to 29% in unburned watersheds. In contrast, dissolved organic carbon (DOC), was highest in watersheds that burned to a moderate extent (10-20%) and lowest in those with extensive burning. Catchments with a moderate extent burned had DOC <span class="hlt">concentrations</span> 2.5 and 1.7 times more than those with extensive burning and unburned catchments, respectively. Peak <span class="hlt">concentrations</span> of DOC and <span class="hlt">nitrate</span> track the rising limb of the streamflow hydrograph and reach a maximum in May, but patterns among burn extent categories were seasonally consistent. Current riparian conditions are linked to stream <span class="hlt">nitrate</span> in burned watersheds. For example, stream <span class="hlt">nitrate</span> increases proportionally to the extent of riparian zones with low shrub cover (R2 = 0.76). We found signs of watershed recovery compared to the initial post-fire period; stream temperature and turbidity remained elevated in extensively burned catchments, but increases were only significant during the spring season. The persistent stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> as well as the relation between riparian cover and post-fire stream <span class="hlt">nitrate</span> may help prioritize restoration planting efforts and mitigate chronic</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AGUFM.H42B0360K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AGUFM.H42B0360K"><span>Algal <span class="hlt">Nitrate</span> Assimilation and Productivity in an Urban, Concrete-Lined Stream Dominated by Tertiary Treated Municipal Waste-Water</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kent, R. H.; Burton, C. A.</p> <p>2001-12-01</p> <p>This study examined the extent and variabiltity of <span class="hlt">nitrate</span> loss in a 2.85 km reach of Cucamonga Creek, which is concrete-lined and dominated by treated municipal waste-water. Primary production was measured to determine if the loss could be attributed to algal assimilation. Samples for nitrite plus <span class="hlt">nitrate</span> analysis were collected at the top and bottom of the study reach every hour throughout the 24-hour sampling period; samples for analyses of other parameters were collected less frequently. Water temperature, dissolved oxygen (DO), pH and specific conductance were monitored continuously throughout the sampling period using in-stream probes. During the two weeks prior to the study, periphyton samples were collected periodically at four stations along the reach for standing crop measurements and a growth rate time-series using Chlorophyll A and ash-free-dry mass. Water samples from the upstream station were compared to those taken an hour later (the approximate travel time) at the downstream station. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> were lower at the downstream station in 21 of 25 of the paired samples, indicating nearly continuous loss in the reach. The total loss of NO3 for the day was about 0.71 g as N/m2. Most of the loss occurred during daylight hours, with the peak occurring at midday. During the night, CO2 <span class="hlt">concentrations</span> were relatively constant at about 25 mg/L. <span class="hlt">Concentrations</span> began to decline at sunrise, and declined to 0 mg/L at the lower site after midday. Peak <span class="hlt">nitrate</span> loss occurred at about the same time as the CO2 <span class="hlt">concentration</span> was at its minimum. DO declined slightly during the night, began to rise at sunrise, reached a peak during midday, and declined in late afternoon through evening; pH followed a similar pattern. Net primary productivity, as measured by the differences in DO between the two sites was 13 g O2/m2 for the day. Using the Redfield ratio, the predicted <span class="hlt">nitrate</span> assimilation is about 0.66 g NO3 as N/m2. The continuous loss of <span class="hlt">nitrate</span> between the two</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16934380','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16934380"><span>Assessment of <span class="hlt">nitrate</span> pollution in the Grand Morin aquifers (France): combined use of geostatistics and physically based modeling.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Flipo, Nicolas; Jeannée, Nicolas; Poulin, Michel; Even, Stéphanie; Ledoux, Emmanuel</p> <p>2007-03-01</p> <p>The objective of this work is to combine several approaches to better understand <span class="hlt">nitrate</span> fate in the Grand Morin aquifers (2700 km(2)), part of the Seine basin. cawaqs results from the coupling of the hydrogeological model newsam with the hydrodynamic and biogeochemical model of river ProSe. cawaqs is coupled with the agronomic model Stics in order to simulate <span class="hlt">nitrate</span> migration in basins. First, kriging provides a satisfactory representation of aquifer <span class="hlt">nitrate</span> contamination from local <span class="hlt">observations</span>, to set initial conditions for the physically based model. Then associated confidence intervals, derived from data using geostatistics, are used to validate cawaqs results. Results and evaluation obtained from the combination of these approaches are given (period 1977-1988). Then cawaqs is used to simulate <span class="hlt">nitrate</span> fate for a 20-year period (1977-1996). The mean <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> increase in aquifers is 0.09 mgN L(-1)yr(-1), resulting from an average infiltration flux of 3500 kgN.km(-2)yr(-1).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.B23C1103B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.B23C1103B"><span>Oxygen Isotopic Composition of <span class="hlt">Nitrate</span> and Sulfate in Fog and River water in Podocarpus National Forest, Ecuador</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brothers, L. A.; Fabian, P.; Thiemens, M. H.</p> <p>2006-12-01</p> <p>The eastern slopes of the Andean rainforests of Ecuador possess some of the highest plant biodiversity found on the planet; however, these ecosystems are in jeopardy because region is experiences one of the highest deforestation rates in South America. This rainforest characterized by high acidity and low nutrient soils and experiences natural process which are both destabilizing and stabilizing to biodiversity rendering this a unique, though sensitive environment. There is increased concern that anthropogenic activities are affecting rainforests and could lead to higher extinction rates, changes in the biodiversity and far reaching effects on the global troposphere. Measurements of <span class="hlt">nitrate</span> and sulfate in rain and fog water have shown periods of elevated <span class="hlt">concentrations</span> in the Podocarpus National Park near Loja, Ecuador. These high episodes contribute to annual deposition rates that are comparable to polluted central Europe. Significant anthropogenic sources near this region are lacking and it is believed that the majority of the <span class="hlt">nitrate</span> and sulfate pollution can be attributed to biomass burning in the Amazon basin. <span class="hlt">Concentration</span> measurements do not elucidate the source of high <span class="hlt">nitrate</span> and sulfate pollution; however, by measuring all three stable isotopes of oxygen in <span class="hlt">nitrate</span> and sulfate from fog and river water provides a new way to examine the impacts of biomass burning on the region. By using stable isotope techniques atmospheric <span class="hlt">nitrate</span> and sulfate can be resolved from terrestrial sources. This provides an unique way to trace the contributions from the biomass burning and farming sources. Current research at the field station monitors sulfate and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in rain and fog water by standard methods to investigate water and nutrient pathways along with data from satellite and ground based remote sensing, in-situ <span class="hlt">observations</span> and numerical models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://water.usgs.gov/GIS/metadata/usgswrd/XML/sir095082_no3.xml','USGSPUBS'); return false;" href="http://water.usgs.gov/GIS/metadata/usgswrd/XML/sir095082_no3.xml"><span>Probability of Elevated <span class="hlt">Nitrate</span> <span class="hlt">Concentrations</span> in Groundwater in the Eagle River Watershed Valley-Fill Aquifer, Eagle County, North-Central Colorado, 2006-2007</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Rupert, Michael G.; Plummer, Niel</p> <p>2009-01-01</p> <p>This raster data set delineates the predicted probability of elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater in the Eagle River watershed valley-fill aquifer, Eagle County, North-Central Colorado, 2006-2007. This data set was developed by a cooperative project between the U.S. Geological Survey, Eagle County, the Eagle River Water and Sanitation District, the Town of Eagle, the Town of Gypsum, and the Upper Eagle Regional Water Authority. This project was designed to evaluate potential land-development effects on groundwater and surface-water resources so that informed land-use and water management decisions can be made. This groundwater probability map and its associated probability maps was developed as follows: (1) A point data set of wells with groundwater quality and groundwater age data was overlaid with thematic layers of anthropogenic (related to human activities) and hydrogeologic data by using a geographic information system to assign each well values for depth to groundwater, distance to major streams and canals, distance to gypsum beds, precipitation, soils, and well depth. These data then were downloaded to a statistical software package for analysis by logistic regression. (2) Statistical models predicting the probability of elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, the probability of unmixed young water (using chlorofluorocarbon-11 <span class="hlt">concentrations</span> and tritium activities), and the probability of elevated volatile organic compound <span class="hlt">concentrations</span> were developed using logistic regression techniques. (3) The statistical models were entered into a GIS and the probability map was constructed.</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/19130411','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19130411"><span>Bioactivation of organic <span class="hlt">nitrates</span> and the mechanism of <span class="hlt">nitrate</span> tolerance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Klemenska, Emila; Beresewicz, Andrzej</p> <p>2009-01-01</p> <p>Organic <span class="hlt">nitrates</span>, such as nitroglycerin, are commonly used in the therapy of cardiovascular disease. Long-term therapy with these drugs, however, results in the rapid development of <span class="hlt">nitrate</span> tolerance, limiting their hemodynamic and anti-ischemic efficacy. In addition, <span class="hlt">nitrate</span> tolerance is associated with the expression of potentially deleterious modifications such as increased oxidative stress, endothelial dysfunction, and sympathetic activation. In this review we discuss current concepts regarding the mechanisms of organic <span class="hlt">nitrate</span> bioactivation, <span class="hlt">nitrate</span> tolerance, and <span class="hlt">nitrate</span>-mediated oxidative stress and endothelial dysfunction. We also examine how hydralazine may prevent <span class="hlt">nitrate</span> tolerance and related endothelial dysfunction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25957718','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25957718"><span>Immobilization of <span class="hlt">nitrate</span> reductase onto epoxy affixed silver nanoparticles for determination of soil <span class="hlt">nitrates</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sachdeva, Veena; Hooda, Vinita</p> <p>2015-08-01</p> <p>Epoxy glued silver nanoparticles were used as immobilization support for <span class="hlt">nitrate</span> reductase (NR). The resulting epoxy/AgNPs/NR conjugates were characterized at successive stages of fabrication by scanning electron microscopy and fourier transform infrared spectroscopy. The immobilized enzyme system exhibited reasonably high conjugation yield (37.6±0.01 μg/cm(2)), with 93.54±0.88% retention of specific activity. Most favorable working conditions of pH, temperature and substrate <span class="hlt">concentration</span> were ascertained to optimize the performance of epoxy/AgNPs/NR conjugates for soil <span class="hlt">nitrate</span> quantification. The analytical results for soil <span class="hlt">nitrate</span> determination were consistent, reliable and reproducible. Minimum detection limit of the method was 0.05 mM with linearity from 0.1 to 11.0 mM. The % recoveries of added <span class="hlt">nitrates</span> (0.1 and 0.2 mM) were<95.0% and within-day and between-day coefficients of variations were 0.556% and 1.63% respectively. The method showed good correlation (R(2)=0.998) with the popular Griess reaction method. Epoxy/AgNPs bound NR had a half-life of 18 days at 4 °C and retained 50% activity after 15 reuses. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/49058','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/49058"><span>Triple <span class="hlt">nitrate</span> isotopes indicate differing <span class="hlt">nitrate</span> source contributions to streams across a nitrogen saturation gradient</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Lucy A. Rose; Emily M. Elliott; Mary Beth. Adams</p> <p>2015-01-01</p> <p>Nitrogen (N) deposition affects forest biogeochemical cycles worldwide, often contributing to N saturation. Using long-term (>30-year) records of stream <span class="hlt">nitrate</span> (NO3-) <span class="hlt">concentrations</span> at Fernow Experimental Forest (West Virginia, USA), we classified four watersheds into N saturation stages ranging from Stage 0 (N-...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A51J..07S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A51J..07S"><span>Aircraft-based <span class="hlt">Observations</span> and Modeling of Wintertime Submicron Aerosol Composition over the Northeastern U.S.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shah, V.; Jaegle, L.; Schroder, J. C.; Campuzano-Jost, P.; Jimenez, J. L.; Guo, H.; Sullivan, A.; Weber, R. J.; Green, J. R.; Fiddler, M.; Bililign, S.; Lopez-Hilfiker, F.; Lee, B. H.; Thornton, J. A.</p> <p>2017-12-01</p> <p>Submicron aerosol particles (PM1) remain a major air pollution concern in the urban areas of northeastern U.S. While SO2 and NOx emission controls have been effective at reducing summertime PM1 <span class="hlt">concentrations</span>, this has not been the case for wintertime sulfate and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, suggesting a nonlinear response during winter. During winter, organic aerosol (OA) is also an important contributor to PM1 mass despite low biogenic emissions, suggesting the presence of important urban sources. We use aircraft-based <span class="hlt">observations</span> collected during the Wintertime INvestigation of Transport, Emissions and Reactivity (WINTER) campaign (Feb-March 2015), together with the GEOS-Chem chemical transport model, to investigate the sources and chemical processes governing wintertime PM1 over the northeastern U.S. The mean <span class="hlt">observed</span> <span class="hlt">concentration</span> of PM1 between the surface and 1 km was 4 μg m-3, about 30% of which was composed of sulfate, 20% <span class="hlt">nitrate</span>, 10% ammonium, and 40% OA. The model reproduces the <span class="hlt">observed</span> sulfate, <span class="hlt">nitrate</span> and ammonium <span class="hlt">concentrations</span> after updates to HNO3 production and loss, SO2 oxidation, and NH3 emissions. We find that 65% of the sulfate formation occurs in the aqueous phase, and 55% of <span class="hlt">nitrate</span> formation through N2O5 hydrolysis, highlighting the importance of multiphase and heterogeneous processes during winter. Aqueous-phase sulfate production and the gas-particle partitioning of <span class="hlt">nitrate</span> and ammonium are affected by atmospheric acidity, which in turn depends on the <span class="hlt">concentration</span> of these species. We examine these couplings with GEOS-Chem, and assess the response of wintertime PM1 <span class="hlt">concentrations</span> to further emission reductions based on the U.S. EPA projections for the year 2023. For OA, we find that the standard GEOS-Chem simulation underestimates the <span class="hlt">observed</span> <span class="hlt">concentrations</span>, but a simple parameterization developed from previous summer field campaigns is able to reproduce the <span class="hlt">observations</span> and the contribution of primary and secondary OA. We find that</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/6799750','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/6799750"><span>Antimicrobial action of silver <span class="hlt">nitrate</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Richards, R M</p> <p>1981-01-01</p> <p>Silver <span class="hlt">nitrate</span> 3 mug/ml prevented the separation into two daughter cells of sensitive dividing cells of Pseudomonas aeruginosa growing in nutrient broth plus the chemical. Cell size of sensitive cells was increased and the cytoplasmic contents, cytoplasmic membrane and external cell envelope structures were all abnormal. P. aeruginosa cells grown in the presence of silver <span class="hlt">nitrate</span> 9 mug/ml showed all these changes to a marked degree except inhibition of cell division was not <span class="hlt">observed</span>. Silver <span class="hlt">nitrate</span> (1.5 mug/ml) in distilled water inactivated bacteriophage T2 particles as determined by their infectivity to Escherichia coli B cultures. Lysozyme (50 mug/ml) reduced, and sodium chloride (0.9%) blocked this activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..GECMW1070A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..GECMW1070A"><span>Comparison of plasma generated nitrogen fertilizer to conventional fertilizers ammonium <span class="hlt">nitrate</span> and sodium <span class="hlt">nitrate</span> for pre-emergent and seedling growth</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Andhavarapu, A.; King, W.; Lindsay, A.; Byrns, B.; Knappe, D.; Fonteno, W.; Shannon, S.</p> <p>2014-10-01</p> <p>Plasma source generated nitrogen fertilizer is compared to conventional nitrogen fertilizers in water for plant growth. Root, shoot sizes, and weights are used to examine differences between plant treatment groups. With a simple coaxial structure creating a large-volume atmospheric glow discharge, a 162 MHz generator drives the air plasma. The VHF plasma source emits a steady state glow; the high drive frequency is believed to inhibit the glow-to-arc transition for non-thermal discharge generation. To create the plasma activated water (PAW) solutions used for plant treatment, the discharge is held over distilled water until a 100 ppm <span class="hlt">nitrate</span> aqueous <span class="hlt">concentration</span> is achieved. The discharge is used to incorporate nitrogen species into aqueous solution, which is used to fertilize radishes, marigolds, and tomatoes. In a four week experiment, these plants are watered with four different solutions: tap water, dissolved ammonium <span class="hlt">nitrate</span> DI water, dissolved sodium <span class="hlt">nitrate</span> DI water, and PAW. Ammonium <span class="hlt">nitrate</span> solution has the same amount of total nitrogen as PAW; sodium <span class="hlt">nitrate</span> solution has the same amount of <span class="hlt">nitrate</span> as PAW. T-tests are used to determine statistical significance in plant group growth differences. PAW fertilization chemical mechanisms are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18499297','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18499297"><span>Determination of the origin of groundwater <span class="hlt">nitrate</span> at an air weapons range using the dual isotope approach.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bordeleau, Geneviève; Savard, Martine M; Martel, Richard; Ampleman, Guy; Thiboutot, Sonia</p> <p>2008-06-06</p> <p><span class="hlt">Nitrate</span> is one of the most common contaminants in shallow groundwater, and many sources may contribute to the <span class="hlt">nitrate</span> load within an aquifer. Groundwater <span class="hlt">nitrate</span> plumes have been detected at several ammunition production sites. However, the presence of multiple potential sources and the lack of existing isotopic data concerning explosive degradation-induced <span class="hlt">nitrate</span> constitute a limitation when it comes to linking both types of contaminants. On military training ranges, high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater were reported for the first time as part of the hydrogeological characterization of the Cold Lake Air Weapons Range (CLAWR), Alberta, Canada. Explosives degradation is thought to be the main source of <span class="hlt">nitrate</span> contamination at CLAWR, as no other major source is present. Isotopic analyses of N and O in <span class="hlt">nitrate</span> were performed on groundwater samples from the unconfined and confined aquifers; the dual isotopic analysis approach was used in order to increase the chances of identifying the source of <span class="hlt">nitrate</span>. The isotopic ratios for the groundwater samples with low <span class="hlt">nitrate</span> <span class="hlt">concentration</span> suggested a natural origin with a strong contribution of anthropogenic atmospheric NOx. For the samples with <span class="hlt">nitrate</span> <span class="hlt">concentration</span> above the expected background level the isotopic ratios did not correspond to any source documented in the literature. Dissolved RDX samples were degraded in the laboratory and results showed that all reproduced degradation processes released <span class="hlt">nitrate</span> with a strong fractionation. Laboratory isotopic values for RDX-derived NO(3)(-) produced a trend of high delta(18)O-low delta(15)N to low delta(18)O-high delta(15)N, and groundwater samples with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above the expected background level appeared along this trend. Our results thus point toward a characteristic field of isotopic ratios for <span class="hlt">nitrate</span> being derived from the degradation of RDX.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25762424','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25762424"><span>Seasonal <span class="hlt">nitrate</span> algorithms for <span class="hlt">nitrate</span> retrieval using OCEANSAT-2 and MODIS-AQUA satellite data.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Durairaj, Poornima; Sarangi, Ranjit Kumar; Ramalingam, Shanthi; Thirunavukarassu, Thangaradjou; Chauhan, Prakash</p> <p>2015-04-01</p> <p>In situ datasets of <span class="hlt">nitrate</span>, sea surface temperature (SST), and chlorophyll a (chl a) collected during the monthly coastal samplings and organized cruises along the Tamilnadu and Andhra Pradesh coast between 2009 and 2013 were used to develop seasonal <span class="hlt">nitrate</span> algorithms. The <span class="hlt">nitrate</span> algorithms have been built up based on the three-dimensional regressions between SST, chl a, and <span class="hlt">nitrate</span> in situ data using linear, Gaussian, Lorentzian, and paraboloid function fittings. Among these four functions, paraboloid was found to be better with the highest co-efficient of determination (postmonsoon: R2=0.711, n=357; summer: R2=0.635, n=302; premonsoon: R2=0.829, n=249; and monsoon: R2=0.692, n=272) for all seasons. Based on these fittings, seasonal <span class="hlt">nitrate</span> images were generated using the concurrent satellite data of SST from Moderate Resolution Imaging Spectroradiometer (MODIS) and chlorophyll (chl) from Ocean Color Monitor (OCM-2) and MODIS. The best retrieval of modeled <span class="hlt">nitrate</span> (R2=0.527, root mean square error (RMSE)=3.72, and mean normalized bias (MNB)=0.821) was <span class="hlt">observed</span> for the postmonsoon season due to the better retrieval of both SST MODIS (28 February 2012, R2=0.651, RMSE=2.037, and MNB=0.068) and chl OCM-2 (R2=0.534, RMSE=0.317, and MNB=0.27). Present results confirm that the chl OCM-2 and SST MODIS retrieve <span class="hlt">nitrate</span> well than the MODIS-derived chl and SST largely due to the better retrieval of chl by OCM-2 than MODIS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70023782','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70023782"><span>In situ stimulation of groundwater denitrification with formate to remediate <span class="hlt">nitrate</span> contamination</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, R.L.; Miller, D.N.; Brooks, M.H.; Widdowson, M.A.; Killingstad, M.W.</p> <p>2001-01-01</p> <p>In situ stimulation of denitrification has been proposed as a mechanism to remediate groundwater <span class="hlt">nitrate</span> contamination. In this study, sodium formate was added to a sand and gravel aquifer on Cape Cod, MA, to test whether formate could serve as a potential electron donor for subsurface denitrification. During 16- and 10-day trials, groundwater from an anoxic <span class="hlt">nitrate</span>-containing zone (0.5-1.5 mM) was continuously withdrawn, amended with formate and bromide, and pumped back into the aquifer. <span class="hlt">Concentrations</span> of groundwater constituents were monitored in multilevel samplers after up to 15 m of transport by natural gradient flow. <span class="hlt">Nitrate</span> and formate <span class="hlt">concentrations</span> were decreased 80-100% and 60-70%, respectively, with time and subsequent travel distance, while nitrite <span class="hlt">concentrations</span> inversely increased. The field experiment breakthrough curves were simulated with a two-dimensional site-specific model that included transport, denitrification, and microbial growth. Initial values for model parameters were obtained from laboratory incubations with aquifer core material and then refined to fit field breakthrough curves. The model and the lab results indicated that formate-enhanced nitrite reduction was nearly 4-fold slower than <span class="hlt">nitrate</span> reduction, but in the lab, nitrite was completely consumed with sufficient exposure time. Results of this study suggest that a long-term injection of formate is necessary to test the remediation potential of this approach for <span class="hlt">nitrate</span> contamination and that adaptation to nitrite accumulation will be a key determinative factor.In situ stimulation of denitrification has been proposed as a mechanism to remediate groundwater <span class="hlt">nitrate</span> contamination. In this study, sodium formate was added to a sand and gravel aquifer on Cape Cod, MA, to test whether formate could serve as a potential electron donor for subsurface denitrification. During 16- and 10-day trials, groundwater from an anoxic <span class="hlt">nitrate</span>-containing zone (0.5-1.5 mM) was continuously withdrawn</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70043343','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70043343"><span>Fractionation of stable isotopes in perchlorate and <span class="hlt">nitrate</span> during in situ biodegradation in a sandy aquifer</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hatzinger, P.B.; Bohlke, John Karl; Sturchio, N.C.; Gu, B.; Heraty, L.J.; Borden, R.C.</p> <p>2009-01-01</p> <p>An in situ experiment was performed in a shallow alluvial aquifer in Maryland to quantify the fractionation of stable isotopes in perchlorate (Cl and O) and <span class="hlt">nitrate</span> (N and O) during biodegradation. An emulsified soybean oil substrate that was previously injected into this aquifer provided the electron donor necessary for biological perchlorate reduction and denitrification. During the field experiment, groundwater extracted from an upgradient well was pumped into an injection well located within the in situ oil barrier, and then groundwater samples were withdrawn for the next 30 h. After correction for dilution (using Br– as a conservative tracer of the injectate), perchlorate <span class="hlt">concentrations</span> decreased by 78% and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> decreased by 82% during the initial 8.6 h after the injection. The <span class="hlt">observed</span> ratio of fractionation effects of O and Cl isotopes in perchlorate (e18O/e37Cl) was 2.6, which is similar to that <span class="hlt">observed</span> in the laboratory using pure cultures (2.5). Denitrification by indigenous bacteria fractionated O and N isotopes in <span class="hlt">nitrate</span> at a ratio of ~0.8 (e18O/e15N), which is within the range of values reported previously for denitrification. However, the magnitudes of the individual apparent in situ isotope fractionation effects for perchlorate and <span class="hlt">nitrate</span> were appreciably smaller than those reported in homogeneous closed systems (0.2 to 0.6 times), even after adjustment for dilution. These results indicate that (1) isotope fractionation factor ratios (e18O/e37Cl, e18O/e15N) derived from homogeneous laboratory systems (e.g. pure culture studies) can be used qualitatively to confirm the occurrence of in situ biodegradation of both perchlorate and <span class="hlt">nitrate</span>, but (2) the magnitudes of the individual apparent e values cannot be used quantitatively to estimate the in situ extent of biodegradation of either anion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/964323-fractionation-stable-isotopes-perchlorate-nitrate-during-situ-biodegradation-sandy-aquifer','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/964323-fractionation-stable-isotopes-perchlorate-nitrate-during-situ-biodegradation-sandy-aquifer"><span>Fractionation of stable isotopes in perchlorate and <span class="hlt">nitrate</span> during in situ biodegradation in a sandy aquifer</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>Hatzinger, Paul B.; Bohlke, J. K.; Sturchio, N. C.</p> <p></p> <p>An in situ experiment was performed in a shallow alluvial aquifer in Maryland to quantify the fractionation of stable isotopes in perchlorate (Cl and O) and <span class="hlt">nitrate</span> (N and O) during biodegradation. An emulsified soybean oil substrate that was previously injected into this aquifer provided the electron donor necessary for biological perchlorate reduction and denitrification. During the field experiment, groundwater extracted from an upgradient well was pumped into an injection well located within the in situ oil barrier, and then groundwater samples were withdrawn for the next 30 h. After correction for dilution (using Br-as a conservative tracer of themore » injectate), perchlorate <span class="hlt">concentrations</span> decreased by 78 % and <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> decreased by 87 %, during the initial 8.6 h after the injection. The <span class="hlt">observed</span> ratio of fractionation effects of O and Cl isotopes in perchlorate (ε18O/ε37Cl) was 2.6, which is similar to that <span class="hlt">observed</span> in the laboratory using pure cultures (2.5). Denitrification by indigenous bacteria fractionated O and N isotopes in <span class="hlt">nitrate</span> at a ratio of approximately 0.8 (ε18O/ε15N), which is within the range of values reported previously for denitrification. However, the magnitudes of the individual apparent in situ isotope fractionation effects for perchlorate and <span class="hlt">nitrate</span> were appreciably smaller than those reported in homogeneous closed systems (0.2 to 0.6 times), even after adjustment for dilution. These results indicate that (1) isotope fractionation factor ratios (ε18O/ε37Cl, ε18O/ε15N) derived from homogeneous laboratory systems (e.g., pure culture studies) can be used qualitatively to confirm the occurrence of in situ biodegradation of both perchlorate and <span class="hlt">nitrate</span>, but (2) the magnitudes of the individual apparent  values cannot be used quantitatively to estimate the in situ extent of biodegradation of either anion.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4189604','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4189604"><span>Effect of nano silver and silver <span class="hlt">nitrate</span> on seed yield of (Ocimum basilicum L.)</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 The aim of this study was to evaluate the effect of nano silver and silver <span class="hlt">nitrate</span> on yield of seed in basil plant. The study was carried out in a randomized block design with three replications. Results Four levels of either silver <span class="hlt">nitrate</span> (0, 100, 200 and 300 ppm) or nano silver (0, 20, 40, and 60 ppm) were sprayed on basil plant at seed growth stage. The results showed that there was no significant difference between 100 ppm of silver <span class="hlt">nitrate</span> and 60 ppm <span class="hlt">concentration</span> of nano silver on the shoot silver <span class="hlt">concentration</span>. However, increasing the <span class="hlt">concentration</span> of silver <span class="hlt">nitrate</span> from 100 to 300 ppm caused a decrease in seed yield. In contrast, a raise in the <span class="hlt">concentration</span> of nano silver from 20 to 60 ppm has led to an improvement in the seed yield. Additionally, the lowest amount of seed yield was found with control plants. Conclusions Finally, with increasing level of silver <span class="hlt">nitrate</span>, the polyphenol compound content was raised but the enhancing level of nano silver resulting in the reduction of these components. In conclusion, nano silver can be used instead of other compounds of silver. PMID:25383311</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70032167','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70032167"><span>Stream denitrification across biomes and its response to anthropogenic <span class="hlt">nitrate</span> loading</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Mulholland, P.J.; Helton, A.M.; Poole, G.C.; Hall, R.O.; Hamilton, S.K.; Peterson, B.J.; Tank, J.L.; Ashkenas, L.R.; Cooper, L.W.; Dahm, Clifford N.; Dodds, W.K.; Findlay, S.E.G.; Gregory, S.V.; Grimm, N. B.; Johnson, S.L.; McDowell, W.H.; Meyer, J.L.; Valett, H.M.; Webster, J.R.; Arango, C.P.; Beaulieu, J.J.; Bernot, M.J.; Burgin, A.J.; Crenshaw, C.L.; Johnson, L.T.; Niederlehner, B.R.; O'Brien, J. M.; Potter, J.D.; Sheibley, R.W.; Sobota, D.J.; Thomas, S.M.</p> <p>2008-01-01</p> <p>Anthropogenic addition of bioavailable nitrogen to the biosphere is increasing and terrestrial ecosystems are becoming increasingly nitrogen-saturated, causing more bioavailable nitrogen to enter groundwater and surface waters. Large-scale nitrogen budgets show that an average of about 20-25 per cent of the nitrogen added to the biosphere is exported from rivers to the ocean or inland basins, indicating that substantial sinks for nitrogen must exist in the landscape. Streams and rivers may themselves be important sinks for bioavailable nitrogen owing to their hydrological connections with terrestrial systems, high rates of biological activity, and streambed sediment environments that favour microbial denitrification. Here we present data from nitrogen stable isotope tracer experiments across 72 streams and 8 regions representing several biomes. We show that total biotic uptake and denitrification of <span class="hlt">nitrate</span> increase with stream <span class="hlt">nitrate</span> <span class="hlt">concentration</span>, but that the efficiency of biotic uptake and denitrification declines as <span class="hlt">concentration</span> increases, reducing the proportion of in-stream <span class="hlt">nitrate</span> that is removed from transport. Our data suggest that the total uptake of <span class="hlt">nitrate</span> is related to ecosystem photosynthesis and that denitrification is related to ecosystem respiration. In addition, we use a stream network model to demonstrate that excess <span class="hlt">nitrate</span> in streams elicits a disproportionate increase in the fraction of <span class="hlt">nitrate</span> that is exported to receiving waters and reduces the relative role of small versus large streams as <span class="hlt">nitrate</span> sinks. ??2008 Nature Publishing Group.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.H33G1115A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.H33G1115A"><span>Anomalous electrical signals associated with microbial activity: Results from Iron and <span class="hlt">Nitrate</span>-Reducing Columns</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aaron, R. B.; Zheng, Q.; Flynn, P.; Singha, K.; Brantley, S.</p> <p>2008-12-01</p> <p>Three flow-through columns outfitted with Ag/AgCl electrodes were constructed to test the effects of different microbial processes on the geophysical measurements of self potential (SP), bulk electrical conductivity (σ b), and induced polarization (IP). The columns were filled with sieved, Fe-bearing subsurface sediment from the Delmarva Peninsula near Oyster, VA, inoculated (9:1 ratio) with a freshly-collected, shallow subsurface sediment from a wetland floodplain (Dorn Creek) near Madison, WI. Each of the columns was fed anoxic and sterile PIPES buffered artificial groundwater (PBAGW) containing different <span class="hlt">concentrations</span> of acetate and <span class="hlt">nitrate</span>. The medium fed to Column 1 (<span class="hlt">nitrate</span>-reducing) was amended with 100 μM acetate and 2 mM <span class="hlt">nitrate</span>. Column 2 (iron-reducing) was run with PBAGW containing 1.0 mM acetate and 0 mM <span class="hlt">nitrate</span>. Column 3 (alternating redox state) was operated under conditions designed to alternately stimulate <span class="hlt">nitrate</span>-reducing and iron-reducing populations to provide conditions, i.e., the presence of both <span class="hlt">nitrate</span> and microbially-produced Fe(II), that would allow growth of <span class="hlt">nitrate</span>-dependent Fe(II)-oxidizing populations. We operated Column 3 with a cycling strategy of 14-18 days of high C medium (1 mM acetate and 100 μ M <span class="hlt">nitrate</span>) followed by 14-18 days of low C medium (100 μ M acetate and 2 mM <span class="hlt">nitrate</span>). Effluent chemistry (NO3-, NO2-, NH4+, acetate, and Fe2+) was sampled daily for four months so as to be concurrent with the electrical measurements. We <span class="hlt">observed</span> chemical evidence of iron reduction (dissolved [Fe(II)] = 0.2mM) in the effluent from the iron reduction and alternating redox columns. Chemical depletion of NO3- ([NO3-] ranged from 1 to 0.02mM), the production of NO2-, and possible production of NH4+ (0.2 mM) was <span class="hlt">observed</span> in the <span class="hlt">nitrate</span> reducing column as well as the alternating redox column. All three columns displayed loss of acetate as microbial activity progressed. σ b remained constant in the alternating redox column (~0.15 S</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22486675','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22486675"><span>Biofiltration of air polluted with methane at <span class="hlt">concentration</span> levels similar to swine slurry emissions: influence of ammonium <span class="hlt">concentration</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Veillette, Marc; Avalos Ramirez, Antonio; Heitz, Michèle</p> <p>2012-01-01</p> <p>An evaluation of the effect of ammonium on the performance of two up-flow inorganic packed bed biofilters treating methane was conducted. The air flow rate was set to 3.0 L min(-1) for an empty bed residence time of 6.0 min. The biofilter was fed with a methane <span class="hlt">concentration</span> of 0.30% (v/v). The ammonium <span class="hlt">concentration</span> in the nutrient solution was increased by small increments (from 0.01 to 0.025 gN-NH(4) (+) L(-1)) for one biofilter and by large increments of 0.05 gN-NH(4) (+) L(-1) in the other biofilter. The total <span class="hlt">concentration</span> of nitrogen was kept constant at 0.5 gN-NH(4) (+) L(-1) throughout the experiment by balancing ammonium with <span class="hlt">nitrate</span>. For both biofilters, the methane elimination capacity, carbon dioxide production, nitrogen bed retention and biomass content decreased with the ammonium <span class="hlt">concentration</span> in the nutrient solution. The biofilter with smaller ammonium increments featured a higher elimination capacity and carbon dioxide production rate, which varied from 4.9 to 14.3 g m(-3) h(-1) and from 11.5 to 30 g m(-3) h(-1), respectively. Denitrification was <span class="hlt">observed</span> as some values of the <span class="hlt">nitrate</span> production rate were negative for ammonium <span class="hlt">concentrations</span> below 0.2 gN-NH(4) (+) L(-1). A Michalelis-Menten-type model fitted the ammonium elimination rate and the <span class="hlt">nitrate</span> production rate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.tandfonline.com/doi/abs/10.1080/10402381.2013.876132','USGSPUBS'); return false;" href="http://www.tandfonline.com/doi/abs/10.1080/10402381.2013.876132"><span>Experimental additions of aluminum sulfate and ammonium <span class="hlt">nitrate</span> to in situ mesocosms to reduce cyanobacterial biovolume and microcystin <span class="hlt">concentration</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>Harris, Ted D.; Wilhelm, Frank M.; Graham, Jennifer L.; Loftin, Keith A.</p> <p>2014-01-01</p> <p>Recent studies suggest that nitrogen additions to increase the total nitrogen:total phosphorus (TN:TP) ratio may reduce cyanobacterial biovolume and microcystin <span class="hlt">concentration</span> in reservoirs. In systems where TP is >100 μg/L, however, nitrogen additions to increase the TN:TP ratio could cause ammonia, <span class="hlt">nitrate</span>, or nitrite toxicity to terrestrial and aquatic organisms. Reducing phosphorus via aluminum sulfate (alum) may be needed prior to nitrogen additions aimed at increasing the TN:TP ratio. We experimentally tested this sequential management approach in large in situ mesocosms (70.7 m3) to examine effects on cyanobacteria and microcystin <span class="hlt">concentration</span>. Because alum removes nutrients and most seston from the water column, alum treatment reduced both TN and TP, leaving post-treatment TN:TP ratios similar to pre-treatment ratios. Cyanobacterial biovolume was reduced after alum addition, but the percent composition (i.e., relative) cyanobacterial abundance remained unchanged. A single ammonium <span class="hlt">nitrate</span> (nitrogen) addition increased the TN:TP ratio 7-fold. After the TN:TP ratio was >50 (by weight), cyanobacterial biovolume and abundance were reduced, and chrysophyte and cryptophyte biovolume and abundance increased compared to the alum treatment. Microcystin was not detectable until the TN:TP ratio was <50. Although both treatments reduced cyanobacteria, only the nitrogen treatment seemed to stimulate energy flow from primary producers to zooplankton, which suggests that combining alum and nitrogen treatments may be a viable in-lake management strategy to reduce cyanobacteria and possibly microcystin <span class="hlt">concentrations</span> in high-phosphorus systems. Additional studies are needed to define best management practices before combined alum and nitrogen additions are implemented as a reservoir management strategy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H53H1569R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H53H1569R"><span>Nutrient Losses from Non-Point Sources or from Unidentified Point Sources? Application Examples of the Smartphone Based <span class="hlt">Nitrate</span> App.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rozemeijer, J.; Ekkelenkamp, R.; van der Zaan, B.</p> <p>2017-12-01</p> <p>In 2016 Deltares launched the free to use <span class="hlt">Nitrate</span> App which accurately reads and interprets <span class="hlt">nitrate</span> test strips. The app directly displays the measured <span class="hlt">concentration</span> and gives the option to share the result. Shared results are visualised in map functionality within the app and online. Since its introduction we've been seeing an increasing number of <span class="hlt">nitrate</span> app applications. In this presentation we show some unanticipated types of application. The <span class="hlt">Nitrate</span> App was originally intended to enable farmers to measure <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> on their own farms. This may encourage farmers to talk to specialists about the right nutrient best management practices (BMP's) for their farm. Several groups of farmers have recently started to apply the <span class="hlt">Nitrate</span> App and to discuss their results with each other and with the authorities. <span class="hlt">Nitrate</span> <span class="hlt">concentration</span> routings in catchments have proven to be another useful application. Within a day a person can generate a catchment scale <span class="hlt">nitrate</span> <span class="hlt">concentration</span> map identifying <span class="hlt">nitrate</span> loss hotspots. In several routings in agricultural catchments clear point sources were found, for example at small scale manure processing plants. These routings proved that the <span class="hlt">Nitrate</span> App can help water managers to target conservation practices more accurately to areas with the highest <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and loads. Other current applications are the screening of domestic water wells in California, the collection of extra measurements (also pH and NH4) in the National Monitoring Network for the Evaluation of the Manure Policy in the Netherlands, and several educational initiatives in cooperation with schools and universities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70010014','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70010014"><span>Trends of pesticides and <span class="hlt">nitrate</span> in ground water of the Central Columbia Plateau, Washington, 1993-2003</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Frans, L.</p> <p>2008-01-01</p> <p>Pesticide and <span class="hlt">nitrate</span> data for ground water sampled in the Central Columbia Plateau, Washington, between 1993 and 2003 by the U.S. Geological Survey National Water-Quality Assessment Program were evaluated for trends in <span class="hlt">concentration</span>. A total of 72 wells were sampled in 1993-1995 and again in 2002-2003 in three well networks that targeted row crop and orchard land use settings as well as the regional basalt aquifer. The Regional Kendall trend test indicated that only deethylatrazine (DEA) <span class="hlt">concentrations</span> showed a significant trend. Deethylatrazine <span class="hlt">concentrations</span> were found to increase beneath the row crop land use well network, the regional aquifer well network, and for the dataset as a whole. No other pesticides showed a significant trend (nor did <span class="hlt">nitrate</span>) in the 72-well dataset. Despite the lack of a trend in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> within the National Water-Quality Assessment dataset, previous work has found a statistically significant decrease in <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> from 1998-2002 for wells with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> above 10 mg L-1 within the Columbia Basin ground water management area, which is located within the National Water-Quality Assessment study unit boundary. The increasing trend in DEA <span class="hlt">concentrations</span> was found to negatively correlate with soil hydrologic group using logistic regression and with soil hydrologic group and drainage class using Spearman's correlation. The decreasing trend in high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> was found to positively correlate with the depth to which the well was cased using logistic regression, to positively correlate with <span class="hlt">nitrate</span> application rates and sand content of the soil, and to negatively correlate with soil hydrologic group using Spearman's correlation. Copyright ?? 2008 by the American Society of Agronomy, Crop Science Society of America, and Soil Science Society of America. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AtmEn..39.5325K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AtmEn..39.5325K"><span>Control strategies for the reduction of airborne particulate <span class="hlt">nitrate</span> in California's San Joaquin Valley</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kleeman, Michael J.; Ying, Qi; Kaduwela, Ajith</p> <p></p> <p>The effect of NO x, volatile organic compound (VOC), and NH 3 emissions control programs on the formation of particulate ammonium <span class="hlt">nitrate</span> in the San Joaquin Valley (SJV) was examined under the typical winter conditions that existed on 4-6 January, 1996. The UCD/CIT photochemical transport model was used for this study so that the source origin of primary particulate matter and secondary particulate matter could be identified. When averaged across the entire SJV, the model results predict that 13-18% of the reactive nitrogen (NO y=NO x+reaction products of NO x) emitted from local sources within the SJV was converted to <span class="hlt">nitrate</span> at the ground level. Each gram of NO x emitted locally within the SJV (expressed as NO 2) produced 0.23-0.31 g of particulate ammonium <span class="hlt">nitrate</span> (NH 4NO 3), which is much smaller than the maximum theoretical yield of 1.7 g of NH 4NO 3 per gram of NO 2. The fraction of reactive nitrogen converted to <span class="hlt">nitrate</span> varied strongly as a function of location. Urban regions with large amounts of fresh NO emissions converted little reactive nitrogen to <span class="hlt">nitrate</span>, while remote areas had up to 70% conversion (equivalent to approximately 1.2 g of NH 4NO 3 per gram of NO 2). The use of a single spatially averaged ratio of NH 4NO 3/NO x as a predictor of how changes to NO x emissions would affect particulate <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> would not be accurate at all locations in the SJV under the conditions studied. The largest local sources of particulate <span class="hlt">nitrate</span> in the SJV were predicted to be diesel engines and catalyst equipped gasoline engines under the conditions experienced on 6 January, 1996. Together, these sources accounted for less than half of the ground-level <span class="hlt">nitrate</span> aerosol in the SJV. The remaining fraction of the aerosol <span class="hlt">nitrate</span> originated from reactive nitrogen originally released upwind of the SJV. The majority of this upwind reactive nitrogen was already transformed to <span class="hlt">nitrate</span> by the time it entered the SJV. The effect of local emissions controls on</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1994/4162/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1994/4162/report.pdf"><span><span class="hlt">Nitrate</span> in ground water and spring water near four dairy farms in North Florida, 1990-93</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Andrews, W.J.</p> <p>1994-01-01</p> <p><span class="hlt">Concentrations</span> of <span class="hlt">nitrate</span> and other selected water- quality characteristics were analyzed periodically for two years in water from 51 monitoring wells installed at four farms and in water discharging from three nearby springs along the Suwannee River in Lafayette and Suwannee Counties to examine the quality of ground water at these farms and the transport of nutrients in ground water to the nearby spring-fed Suwannee River: Ground water from shallow wells, which were completed in the top ten feet of the saturated zone in a surficial sandy aquifer and in the karstic Upper Floridan aquifer generally had the highest <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>, ranging from <.02 to 130 mg/L as nitrogen. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> commonly exceeded the primary drinking water standard of 10 mg/L for <span class="hlt">nitrate</span> as nitrogen in water from shallow wells, which tapped the top ten feet of the uppermost aquifers near waste-disposal areas such as wastewater lagoons and defoliated, intensive-use areas near milking barns. Upgradient from waste-disposal areas, <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> in ground water were commonly less than 1 mg/L as nitrogen. Water samples from deep wells (screened 20 feet deeper than shallow wells in these aquifers) generally had lower <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span> (ranging from <0.02 to 84 mg/L) than water from shallow wells. Water samples from the three monitored springs (Blue, Telford, and Convict Springs) had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> ranging from 1.5 to 6.5 mg/L as nitrogen, which were higher than those typically occurring in water from upgradient wells at the monitored dairy farms or from back- ground wells sampled in the region. Analyses of nitrogen isotope ratios in <span class="hlt">nitrate</span> indicated that leachate from animal wastes was the principal source of <span class="hlt">nitrate</span> in ground water adjacent to waste-disposal areas at the monitored and unmonitored dairy farms. Leachate from a combi- nation of fertilizers, soils, and animal wastes appeared to be the source of <span class="hlt">nitrate</span> in ground- water downgradient from</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24122771','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24122771"><span>Monitoring of nitrites and <span class="hlt">nitrates</span> levels in leafy vegetables (spinach and lettuce): a contribution to risk assessment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Iammarino, Marco; Di Taranto, Aurelia; Cristino, Marianna</p> <p>2014-03-15</p> <p>Nitrites and <span class="hlt">nitrates</span> are compounds considered harmful to humans and the major part of the daily intake of <span class="hlt">nitrates</span> in foodstuffs is related to vegetable consumption. In this work, 150 leafy vegetables samples (75 spinach and 75 lettuce) were analysed in order to assess the levels of nitrites and <span class="hlt">nitrates</span>. The analyses were carried out by a validated ion chromatography method and the samples with <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> higher than legal limits and/or with quantifiable <span class="hlt">concentrations</span> of nitrites were confirmed by an alternative ion chromatography method. <span class="hlt">Nitrate</span> levels higher than legal limits were detected both in spinach (four samples) and in lettuce (five samples). Nitrite residues were registered both at low <span class="hlt">concentrations</span>--lower than 28.5 mg kg⁻¹ (12 spinach samples)--and at high <span class="hlt">concentrations</span>, up to 197.5 mg kg⁻¹ (three spinach and one lettuce sample). Considering the non-negligible percentage of 'not-compliant' samples for <span class="hlt">nitrates</span> (6.0%), control is needed. Moreover, it is possible to suggest the introduction in the Communities Regulations of a 'maximum admissible level' for nitrites in leafy vegetables. © 2013 Society of Chemical Industry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016WRR....52.9046G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016WRR....52.9046G"><span>Quantifying an aquifer <span class="hlt">nitrate</span> budget and future <span class="hlt">nitrate</span> discharge using field data from streambeds and well nests</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gilmore, Troy E.; Genereux, David P.; Solomon, D. Kip; Farrell, Kathleen M.; Mitasova, Helena</p> <p>2016-11-01</p> <p>Novel groundwater sampling (age, flux, and <span class="hlt">nitrate</span>) carried out beneath a streambed and in wells was used to estimate (1) the current rate of change of <span class="hlt">nitrate</span> storage, dSNO3/dt, in a contaminated unconfined aquifer, and (2) future [NO3-]FWM (the flow-weighted mean <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in groundwater discharge) and fNO3 (the <span class="hlt">nitrate</span> flux from aquifer to stream). Estimates of dSNO3/dt suggested that at the time of sampling (2013) the <span class="hlt">nitrate</span> storage in the aquifer was decreasing at an annual rate (mean = -9 mmol/m2yr) equal to about one-tenth the rate of <span class="hlt">nitrate</span> input by recharge. This is consistent with data showing a slow decrease in the [NO3-] of groundwater recharge in recent years. Regarding future [NO3-]FWM and fNO3, predictions based on well data show an immediate decrease that becomes more rapid after ˜5 years before leveling out in the early 2040s. Predictions based on streambed data generally show an increase in future [NO3-]FWM and fNO3 until the late 2020s, followed by a decrease before leveling out in the 2040s. Differences show the potential value of using information directly from the groundwater—surface water interface to quantify the future impact of groundwater <span class="hlt">nitrate</span> on surface water quality. The choice of denitrification kinetics was similarly important; compared to zero-order kinetics, a first-order rate law levels out estimates of future [NO3-]FWM and fNO3 (lower peak, higher minimum) as legacy <span class="hlt">nitrate</span> is flushed from the aquifer. Major fundamental questions about nonpoint-source aquifer contamination can be answered without a complex numerical model or long-term monitoring program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3072465','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3072465"><span>Inhibition of amyloidogenesis by non-steroidal anti-inflammatory drugs and their hybrid <span class="hlt">nitrates</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>Schiefer, Isaac T.; Abdul-Hay, Samer; Wang, Huali; Vanni, Michael; Qin, Zhihui; Thatcher, Gregory R. J.</p> <p>2011-01-01</p> <p>Poor blood-brain barrier penetration of non-steroidal anti-inflammatory drugs (NSAIDs) has been blamed for the failure of the selective amyloid lowering agent (SALA) R-flurbiprofen in phase 3 clinical trials for Alzheimer’s disease (AD). NO-donor NSAIDs (NO-NSAIDs) provide an alternative, gastric-sparing approach to NSAID SALAs, which may improve bioavailability. NSAID analogs were studied for anti-inflammatory activity and for SALA activity in N2a neuronal cells transfected with human amyloid precursor protein (APP). Flurbiprofen (1) analogs were obtained with enhanced anti-inflammatory and anti-amyloidogenic properties compared to 1, however, esterification led to elevated Aβ1–42 levels. Hybrid <span class="hlt">nitrate</span> prodrugs possessed superior anti-inflammatory activity and reduced toxicity relative to the parent NSAIDs, including clinical candidate, CHF5074. Although hybrid <span class="hlt">nitrates</span> elevated Aβ1–42 at higher <span class="hlt">concentration</span>, SALA activity was <span class="hlt">observed</span> at low <span class="hlt">concentrations</span> (≤ 1 µM): both Aβ1–42 and the ratio of Aβ1–42/Aβ1–40 were lowered. This biphasic SALA activity was attributed to the intact <span class="hlt">nitrate</span> drug. For several compounds the selective modulation of amyloidogenesis was tested using an immunoprecipitation MALDI-TOF approach. These data support the development of NO-NSAIDs as an alternative approach towards a clinically useful SALA. PMID:21405086</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70170925','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70170925"><span><span class="hlt">Nitrate</span> in groundwater and water sources used by riparian trees in an agricultural watershed: A chemical and isotopic investigation in southern Minnesota</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Komor, Stephen C.; Magner, Joseph A.</p> <p>1996-01-01</p> <p>This study evaluates processes that affect <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater beneath riparian zones in an agricultural watershed. <span class="hlt">Nitrate</span> pathways in the upper 2 m of groundwater were investigated beneath wooded and grass-shrub riparian zones next to cultivated fields. Because trees can be important components of the overall <span class="hlt">nitrate</span> pathway in wooded riparian zones, water sources used by riparian trees and possible effects of trees on <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater were also investigated. Average <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in shallow groundwater beneath the cultivated fields were 5.5 mg/L upgradient of the wooded riparian zone and 3.5 mg/L upgradient of the grass-shrub zone. Shallow groundwater beneath the fields passed through the riparian zones and discharged into streams that had average <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> of 8.5 mg/L (as N). Lateral variations of δD values in groundwater showed that mixing among different water sources occurred beneath the riparian zones. In the wooded riparian zone, <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in shallow groundwater were diluted by upwelling, <span class="hlt">nitrate</span>-poor, deep groundwater. Upwelling deep groundwater contained ammonium with a δ15N of 5‰ that upon nitrification and mixing with <span class="hlt">nitrate</span> in shallow groundwater caused <span class="hlt">nitrate</span> δ15N values in shallow groundwater to decrease by as much as 19.5‰. Stream water penetrated laterally beneath the wooded riparian zone as far as 19 m from the stream's edge and beneath the grass-shrub zone as far as 27 m from the stream's edge. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> in shallow groundwater immediately upgradient of where it mixed with stream water averaged 0.4 mg/L in the wooded riparian zone and 0.8 mg/L near the grass-shrub riparian zone. <span class="hlt">Nitrate</span> <span class="hlt">concentrations</span> increased toward the streams because of mixing with <span class="hlt">nitrate</span>-rich stream water. Because <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were larger in stream water than shallow groundwater, <span class="hlt">concentrated</span> <span class="hlt">nitrate</span> in the streams cannot have come from shallow groundwater at these</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFM.H54D..05M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFM.H54D..05M"><span>Winter Cover Crop Effects on <span class="hlt">Nitrate</span> Leaching in Subsurface Drainage as Simulated by RZWQM-DSSAT</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Malone, R. W.; Chu, X.; Ma, L.; Li, L.; Kaspar, T.; Jaynes, D.; Saseendran, S. A.; Thorp, K.; Yu, Q.</p> <p>2007-12-01</p> <p>Planting winter cover crops such as winter rye (Secale cereale L.) after corn and soybean harvest is one of the more promising practices to reduce <span class="hlt">nitrate</span> loss to streams from tile drainage systems without negatively affecting production. Because availability of replicated tile-drained field data is limited and because use of cover crops to reduce <span class="hlt">nitrate</span> loss has only been tested over a few years with limited environmental and management conditions, estimating the impacts of cover crops under the range of expected conditions is difficult. If properly tested against <span class="hlt">observed</span> data, models can objectively estimate the relative effects of different weather conditions and agronomic practices (e.g., various N fertilizer application rates in conjunction with winter cover crops). In this study, an optimized winter wheat cover crop growth component was integrated into the calibrated RZWQM-DSSAT hybrid model and then we compare the <span class="hlt">observed</span> and simulated effects of a winter cover crop on <span class="hlt">nitrate</span> leaching losses in subsurface drainage water for a corn-soybean rotation with N fertilizer application rates over 225 kg N ha-1 in corn years. Annual <span class="hlt">observed</span> and simulated flow-weighted average <span class="hlt">nitrate</span> <span class="hlt">concentration</span> (FWANC) in drainage from 2002 to 2005 for the cover crop treatments (CC) were 8.7 and 9.3 mg L-1 compared to 21.3 and 18.2 mg L-1 for no cover crop (CON). The resulting <span class="hlt">observed</span> and simulated FWANC reductions due to CC were 59% and 49%. Simulations with the optimized model at various N fertilizer rates resulted in average annual drainage N loss differences between CC and CON to increase exponentially from 12 to 34 kg N ha-1 for rates of 11 to 261 kg N ha-1. The results suggest that RZWQM-DSSAT is a promising tool to estimate the relative effects of a winter crop under different conditions on <span class="hlt">nitrate</span> loss in tile drains and that a winter cover crop can effectively reduce <span class="hlt">nitrate</span> losses over a range of N fertilizer levels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003AGUFM.H41F1050S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003AGUFM.H41F1050S"><span>Tracing <span class="hlt">Nitrate</span> Contributions to Streams During Varying Flow Regimes at the Sleepers River Research Watershed, Vermont, USA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sebestyen, S. D.; Shanley, J. B.; Boyer, E. W.; Ohte, N.; Doctor, D. H.; Kendall, C.</p> <p>2003-12-01</p> <p>Quantifying sources and transformations of <span class="hlt">nitrate</span> in headwater catchments is fundamental to understanding the movement of nitrogen to streams. At the Sleepers River Research Watershed in northeastern Vermont (USA), we are using multiple chemical tracer and mixing model approaches to quantify sources and transport of <span class="hlt">nitrate</span> to streams under varying flow regimes. We sampled streams, lysimeters, and wells at nested locations from the headwaters to the outlet of the 41 ha W-9 watershed under the entire range of flow regimes <span class="hlt">observed</span> throughout 2002-2003, including baseflow and multiple events (stormflow and snowmelt). Our results suggest that nitrogen sources, and consequently stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, are rapidly regenerated during several weeks of baseflow and nitrogen is flushed from the watershed by stormflow events that follow baseflow periods. Both basic chemistry data (anions, cations, & dissolved organic carbon) and isotopic data (<span class="hlt">nitrate</span>, dissolved organic carbon, and dissolved inorganic carbon) indicate that nitrogen source contributions vary depending upon the extent of saturation in the watershed, the initiation of shallow subsurface water inputs, and other hydrological processes. Stream <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> typically peak with discharge and are higher on the falling than the rising limb of the hydrograph. Our data also indicate the importance of terrestrial and aquatic biogeochemical processes, in addition to hydrological connectivity in controlling how <span class="hlt">nitrate</span> moves from the terrestrial landscape to streams. Our detailed sampling data from multiple flow regimes are helping to identify and quantify the "hot spots" and "hot moments" of biogeochemical and hydrological processes that control nitrogen fluxes in streams.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.H11F1413N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.H11F1413N"><span>Evaluating Chemical Tracers in Suburban Groundwater as Indicators of <span class="hlt">Nitrate</span>-Nitrogen Sources</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nitka, A.; DeVita, W. M.; McGinley, P.</p> <p>2015-12-01</p> <p>The CDC reports that over 15 million US households use private wells. These wells are vulnerable to contamination. One of the most common contaminants in private wells is <span class="hlt">nitrate</span>. <span class="hlt">Nitrate</span> has a health standard of 10 mg/L. This standard is set to prevent methemaglobinemia, or "blue baby" syndrome, in infants. In extreme cases it can affect breathing and heart function, and even lead to death. Elevated <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> have also been associated with increased risk of thyroid disease, diabetes, and certain types of cancer. Unlike municipal wells, there is no mandatory testing of private wells. It is the responsibility of users to have their well water tested. The objective of this research was to identify the most useful chemical tracers for determining sources of <span class="hlt">nitrate</span> in private water supplies. Chemical characteristics, such as mobility in groundwater and water solubility, as well as frequency of use, were considered when choosing source indicators. Fourteen pharmaceuticals and personal care products unique to human use were chosen to identify wells impacted by septic waste. A bovine antibiotic and five pesticide metabolites were used to identify contamination from agricultural sources. Eighteen private wells were selected in a suburban area with septic systems and adjacent agricultural land. The wells were sampled five times and analyzed to provide a temporal profile of <span class="hlt">nitrate</span> and the tracers. The artificial sweetener sucralose was found in >70% of private wells. Wells with sucralose detected had <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> between 5-15 mg/L. The herbicide metabolite metolachlor ESA was detected in 50% of the wells. These wells typically had the highest <span class="hlt">nitrate</span> <span class="hlt">concentrations</span>, often >10 mg/L. The common use and frequent detection of these two compounds made them the most reliable indicators of <span class="hlt">nitrate</span> sources evaluated in this study. This information will help well owners determine appropriate treatment and remediation options and could direct future</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ApWS....4..397S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ApWS....4..397S"><span>Chemometric evaluation of <span class="hlt">nitrate</span> contamination in the groundwater of a hard rock area in Dharapuram, south India</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sajil Kumar, P. J.; Jegathambal, P.; James, E. J.</p> <p>2014-12-01</p> <p>This paper presents the results of investigations on groundwater <span class="hlt">nitrate</span> contamination in the Dharapuram area of Tamil Nadu in south India as a primary step to initiate denitrification. Groundwater samples were collected from 26 selected locations during the pre-monsoon season in July 2010 and analysed for <span class="hlt">nitrate</span> and other water quality parameters. Two important water types were identified, viz. Ca-Na-HCO3 and mixed Ca-Mg-Cl. It is found that the majority of samples possess high <span class="hlt">nitrate</span> <span class="hlt">concentration</span>; 57 % of samples exceeded the permissible limit of Indian (45 mg/L) and WHO (50 mg/L) drinking water standard. Spatial distribution map of NO3 suggested that major contamination was <span class="hlt">observed</span> in the SW and NW parts of the study area. This result was in agreement with the corresponding land-use pattern in this study area. Denitrification process at greater depths was evident from the negative correlation between NO3 and well depth. The sources and controlling factors of high <span class="hlt">nitrate</span> were investigated using cross plots of NO3 with other selected hydrochemical parameters. Positive correlation for NO3 was <span class="hlt">observed</span> with EC, K, Cl and SO4. This analysis was capable of differentiating the various sources of <span class="hlt">nitrate</span> in groundwater. The major sources of <span class="hlt">nitrate</span> contamination are identified as areas of high fertilizer application, sewages and animal waste dumping yards. Regulation of these pollutant sources with appropriate and cost-effective denitrification process can restore the water quality in this area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/19305','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/19305"><span>Relationship of stand age to streamwater <span class="hlt">nitrate</span> in New Hampshire</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>William B. Leak; C. Wayne Martin</p> <p>1975-01-01</p> <p>Streamwater <span class="hlt">nitrate</span> content of six watersheds during spring and summer was apparently related to stand age or age since disturbance. <span class="hlt">Nitrate</span> <span class="hlt">concentration</span> averaged 10.3 ppm right after cutting, dropped to a trace in medium-aged stands, and then rose again to a maximum of 4.8 ppm as stands became overmature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AGUSM...U21B03K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AGUSM...U21B03K"><span>A Multi-Tracer Approach to Characterize Sources and Transport of <span class="hlt">Nitrate</span> in Groundwater in Mantled Karst, Northern Florida</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Katz, B. G.; Bohlke, J.; Hornsby, D.</p> <p>2001-05-01</p> <p><span class="hlt">Nitrate</span> is readily transported from agricultural activities at the surface to the Upper Floridan aquifer in northern Florida due to karst features mantled by highly permeable sands and a high recharge rate (50 cm/yr). In Suwannee and Lafayette Counties, <span class="hlt">nitrate</span> contamination of groundwater is widespread due to the 10-30 kg/ha nitrogen (N) applied annually for the past few decades as synthetic fertilizers (the dominant source of N). Water samples were collected from 12 springs during baseflow conditions (1997-99) and monthly from 14 wells (1998-99). Springwaters were analyzed for various chemical (N species, dissolved gases, CFCs) and isotopic tracers (15N, 3H/3He, 18O, D, 13C). Water from wells was analyzed monthly for N species, and during low-flow and high-flow conditions for 15N, 18O, D, and 13C. As a result of oxic conditions in the aquifer, <span class="hlt">nitrate</span> was the dominant N species in water samples. Large monthly fluctuations of groundwater <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were <span class="hlt">observed</span> at most wells. Relatively high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater from 7 wells likely resulted from seasonal agricultural practices including fertilizer applications and manure spreading on cropland. Relatively low <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> in groundwater from two wells during high-flow conditions were related to mixing with river water. Groundwater samples had N-isotope values (3.8-11.7 per mil) that indicated varying mixtures of inorganic and organic N sources, which corresponded in part to varying proportions of synthetic fertilizers and manure applied to fields. In springwaters from Suwannee County, <span class="hlt">nitrate</span> trends and N-isotope data (2.7-6.2 per mil) were consistent with a peak in fertilizer N input in the late 1970's and a relatively high overall ratio of artificial fertilizer/manure. In contrast, springwater <span class="hlt">nitrate</span> trends and N-isotope data (4.5-9.1 per mil) in Lafayette County were consistent with a more monotonic increase in fertilizer N input and relatively low overall ratio of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2016/1014/ofr20161014.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2016/1014/ofr20161014.pdf"><span>The effect of suspended sediment and color on ultraviolet spectrophotometric <span class="hlt">nitrate</span> sensors</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Snazelle, Teri T.</p> <p>2016-03-08</p> <p>Four commercially available ultraviolet <span class="hlt">nitrate</span> spectrophotometric sensors were evaluated by the U.S. Geological Survey Hydrologic Instrumentation Facility (HIF) to determine the effects of suspended sediment <span class="hlt">concentration</span> (SSC) and colored dissolved organic matter (CDOM) on sensor accuracy. The evaluated sensors were: the Hach NITRATAX plus sc (5-millimeters (mm) path length), Hach NITRATAX plus sc (2 mm), S::CAN Spectro::lyser (5 mm), and the Satlantic SUNA V2 (5 mm). A National Institute of Standards and Technology-traceable <span class="hlt">nitrate</span>-free sediment standard was purchased and used to create the turbid environment, and an easily made filtered tea solution was used for the CDOM test. All four sensors performed well in the test that evaluated the effect of suspended sediment on accuracy. The Hach 5 mm, Hach 2 mm, and the SUNA V2 met their respective manufacturer accuracy specifications up to <span class="hlt">concentrations</span> of 4,500 milligrams per liter (mg/L) SSC. The S::CAN failed to meet its accuracy specifications when the SSC <span class="hlt">concentrations</span> exceeded 4,000 mg/L. Test results from the effect of CDOM on accuracy indicated a significant skewing of data from all four sensors and showed an artificial elevation of measured <span class="hlt">nitrate</span> to varying amounts. Of the four sensors tested, the Satlantic SUNA V2’s accuracy was affected the least in the CDOM test. The <span class="hlt">nitrate</span> <span class="hlt">concentration</span> measured by the SUNA V2 was approximately 24 percent higher than the actual <span class="hlt">concentration</span> when estimated total organic carbon values exceeded 44 mg/L. Measured <span class="hlt">nitrate</span> <span class="hlt">concentration</span> falsely increased 49 percent when measured by the Hach 5 mm, and 75 percent when measured by the Hach 2 mm. The S::CAN’s reported <span class="hlt">nitrate</span> <span class="hlt">concentration</span> increased 96 percent. Path length plays an important role in the sensor’s ability to compensate measurements for matrix interferences, but does not solely determine how well a sensor can handle all interferences. The sensor’s proprietary algorithms also play a key role in matrix</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024462','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024462"><span>Forensic applications of nitrogen and oxygen isotopes in tracing <span class="hlt">nitrate</span> sources in urban environments</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Silva, S.R.; Ging, P.B.; Lee, R.W.; Ebbert, J.C.; Tesoriero, A.J.; Inkpen, E.L.</p> <p>2002-01-01</p> <p>Ground and surface waters in urban areas are susceptible to <span class="hlt">nitrate</span> contamination from septic systems, leaking sewer lines, and fertilizer applications. Source identification is a primary step toward a successful remediation plan in affected areas. In this respect, nitrogen and oxygen isotope ratios of <span class="hlt">nitrate</span>, in conjunction with hydrologic data and water chemistry, have proven valuable in urban studies from Austin, Texas, and Tacoma, Washington. In Austin, stream water was sampled during stremflow and baseflow conditions to assess surface and subsurface sources of <span class="hlt">nitrate</span>, respectively. In Tacoma, well waters were sampled in adjacent sewered and un-sewered areas to determine if locally high <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> were caused by septic systems in the un-sewered areas. In both studies, sewage was identified as a <span class="hlt">nitrate</span> source and mixing between sewage and other sources of <span class="hlt">nitrate</span> was apparent. In addition to source identification, combined nitrogen and oxygen isotopes were important in determining the significance of denitrification, which can complicate source assessment by reducing <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and increasing ??15N values. The two studies illustrate the value of nitrogen and oxygen isotopes of <span class="hlt">nitrate</span> for forensic applications in urban areas. ?? Published by Elsevier Science Ltd. on behalf of AEHS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26495763','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26495763"><span>Interaction of organic carbon, reduced sulphur and <span class="hlt">nitrate</span> in anaerobic baffled reactor for fresh leachate treatment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yin, Zhixuan; Xie, Li; Khanal, Samir Kumar; Zhou, Qi</p> <p>2016-01-01</p> <p>Interaction of organic carbon, reduced sulphur and <span class="hlt">nitrate</span> was examined using anaerobic baffled reactor for fresh leachate treatment by supplementing <span class="hlt">nitrate</span> and/or sulphide to compartment 3. <span class="hlt">Nitrate</span> was removed completely throughout the study mostly via denitrification (>80%) without nitrite accumulation. Besides carbon source, various reduced sulphur (e.g. sulphide, elemental sulphur and organic sulphur) could be involved in the <span class="hlt">nitrate</span> reduction process via sulphur-based autotrophic denitrification when dissolved organic carbon/<span class="hlt">nitrate</span> ratio decreased below 1.6. High sulphide <span class="hlt">concentration</span> not only stimulated autotrophic denitrification, but it also inhibited heterotrophic denitrification, resulting in a shift (11-20%) from heterotrophic denitrification to dissimilatory <span class="hlt">nitrate</span> reduction to ammonia. High-throughput 16S rRNA gene sequencing analysis further confirmed that sulphur-oxidizing <span class="hlt">nitrate</span>-reducing bacteria were stimulated with increase in the proportion of bacterial population from 18.6% to 27.2% by high sulphide <span class="hlt">concentration</span>, meanwhile, heterotrophic <span class="hlt">nitrate</span>-reducing bacteria and fermentative bacteria were inhibited with 25.5% and 66.6% decrease in the bacterial population.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24999115','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24999115"><span>Simultaneous removal of cadmium and <span class="hlt">nitrate</span> in aqueous media by nanoscale zerovalent iron (nZVI) and Au doped nZVI particles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Su, Yiming; Adeleye, Adeyemi S; Huang, Yuxiong; Sun, Xiaoya; Dai, Chaomeng; Zhou, Xuefei; Zhang, Yalei; Keller, Arturo A</p> <p>2014-10-15</p> <p>Nanoscale zerovalent iron (nZVI) has demonstrated high efficacy for treating <span class="hlt">nitrate</span> or cadmium (Cd) contamination, but its efficiency for simultaneous removal of <span class="hlt">nitrate</span> and Cd has not been investigated. This study evaluated the reactivity of nZVI to the co-contaminants and by-product formation, employed different catalysts to reduce nitrite yield from <span class="hlt">nitrate</span>, and examined the transformation of nZVI after reaction. <span class="hlt">Nitrate</span> reduction resulted in high solution pH, negatively charged surface of nZVI, formation of Fe3O4 (a stable transformation of nZVI), and no release of ionic iron. Increased pH and negative charge contributed to significant increase in Cd(II) removal capacity (from 40 mg/g to 188 mg/g) with <span class="hlt">nitrate</span> present. In addition, <span class="hlt">nitrate</span> reduction by nZVI could be catalyzed by Cd(II): while 30% of <span class="hlt">nitrate</span> was reduced by nZVI within 2 h in the absence of Cd(II), complete <span class="hlt">nitrate</span> reduction was <span class="hlt">observed</span> in the presence of 40 mg-Cd/L due to the formation of Cd islands (Cd(0) and CdO) on the nZVI particles. While <span class="hlt">nitrate</span> was reduced mostly to ammonium when Cd(II) was not present or at Cd(II) <span class="hlt">concentrations</span> ≥ 40 mg/L, up to 20% of the initial <span class="hlt">nitrate</span> was reduced to nitrite at Cd(II) <span class="hlt">concentrations</span> < 40 mg/L. Among nZVI particles doped with 1 wt. % Cu, Ag, or Au, nZVI deposited with 1 wt. % Au reduced nitrite yield to less than 3% of the initial <span class="hlt">nitrate</span>, while maintaining a high Cd(II) removal capacity. Copyright © 2014 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017E%26ES...82a2059L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017E%26ES...82a2059L"><span><span class="hlt">Nitrate</span>-nitrogen contamination in groundwater: Spatiotemporal variation and driving factors under cropland in Shandong Province, China</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, J.; Jiang, L. H.; Zhang, C. J.; Li, P.; Zhao, T. K.</p> <p>2017-08-01</p> <p>High groundwater <span class="hlt">nitrate</span>-N is a serious problem especially in highly active agricultural areas. In study, the <span class="hlt">concentration</span> and spatialtemporal distribution of groundwater <span class="hlt">nitrate</span>-N under cropland in Shandong province were assessed by statistical and geostatistical techniques. <span class="hlt">Nitrate</span>-N <span class="hlt">concentration</span> reached a maximum of 184.60 mg L-1 and 29.5% of samples had levels in excess of safety threshold <span class="hlt">concentration</span> (20 mg L-1). The median <span class="hlt">nitrate</span>-N contents after rainy season were significantly higher than those before rainy season, and decreased with increasing groundwater depth. <span class="hlt">Nitrate</span>-N under vegetable and orchard area are significantly higher than ones under grain. The kriging map shows that groundwater <span class="hlt">nitrate</span>-N has a strong spatial variability. Many districts, such as Weifang, Linyi in Shandong province are heavily contaminated with <span class="hlt">nitrate</span>-N. However, there are no significant trends of NO3 --N for most cities. Stepwise regression analysis showed influencing factors are different for the groundwater in different depth. But overall, vegetable yield per unit area, percentages of orchard area, per capita agricultural production, unit-area nitrogen fertilizer, livestock per unit area, percentages of irrigation areas, population per unit area and annual mean temperature are significant variables for groundwater <span class="hlt">nitrate</span>-N variation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/872875','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/872875"><span><span class="hlt">Nitrate</span> reduction</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Dziewinski, Jacek J.; Marczak, Stanislaw</p> <p>2000-01-01</p> <p><span class="hlt">Nitrates</span> are reduced to nitrogen gas by contacting the <span class="hlt">nitrates</span> with a metal to reduce the <span class="hlt">nitrates</span> to nitrites which are then contacted with an amide to produce nitrogen and carbon dioxide or acid anions which can be released to the atmosphere. Minor amounts of metal catalysts can be useful in the reduction of the <span class="hlt">nitrates</span> to nitrites. Metal salts which are formed can be treated electrochemically to recover the metals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017WRR....53.7316T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017WRR....53.7316T"><span>Predicting redox-sensitive contaminant <span class="hlt">concentrations</span> in groundwater using random forest classification</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tesoriero, Anthony J.; Gronberg, Jo Ann; Juckem, Paul F.; Miller, Matthew P.; Austin, Brian P.</p> <p>2017-08-01</p> <p>Machine learning techniques were applied to a large (n > 10,000) compliance monitoring database to predict the occurrence of several redox-active constituents in groundwater across a large watershed. Specifically, random forest classification was used to determine the probabilities of detecting elevated <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>, iron, and arsenic in the Fox, Wolf, Peshtigo, and surrounding watersheds in northeastern Wisconsin. Random forest classification is well suited to describe the nonlinear relationships <span class="hlt">observed</span> among several explanatory variables and the predicted probabilities of elevated <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>, iron, and arsenic. Maps of the probability of elevated <span class="hlt">nitrate</span>, iron, and arsenic can be used to assess groundwater vulnerability and the vulnerability of streams to contaminants derived from groundwater. Processes responsible for elevated <span class="hlt">concentrations</span> are elucidated using partial dependence plots. For example, an increase in the probability of elevated iron and arsenic occurred when well depths coincided with the glacial/bedrock interface, suggesting a bedrock source for these constituents. Furthermore, groundwater in contact with Ordovician bedrock has a higher likelihood of elevated iron <span class="hlt">concentrations</span>, which supports the hypothesis that groundwater liberates iron from a sulfide-bearing secondary cement horizon of Ordovician age. Application of machine learning techniques to existing compliance monitoring data offers an opportunity to broadly assess aquifer and stream vulnerability at regional and national scales and to better understand geochemical processes responsible for <span class="hlt">observed</span> conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70190391','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70190391"><span>Predicting redox-sensitive contaminant <span class="hlt">concentrations</span> in groundwater using random forest classification</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Tesoriero, Anthony J.; Gronberg, Jo Ann M.; Juckem, Paul F.; Miller, Matthew P.; Austin, Brian P.</p> <p>2017-01-01</p> <p>Machine learning techniques were applied to a large (n > 10,000) compliance monitoring database to predict the occurrence of several redox-active constituents in groundwater across a large watershed. Specifically, random forest classification was used to determine the probabilities of detecting elevated <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>, iron, and arsenic in the Fox, Wolf, Peshtigo, and surrounding watersheds in northeastern Wisconsin. Random forest classification is well suited to describe the nonlinear relationships <span class="hlt">observed</span> among several explanatory variables and the predicted probabilities of elevated <span class="hlt">concentrations</span> of <span class="hlt">nitrate</span>, iron, and arsenic. Maps of the probability of elevated <span class="hlt">nitrate</span>, iron, and arsenic can be used to assess groundwater vulnerability and the vulnerability of streams to contaminants derived from groundwater. Processes responsible for elevated <span class="hlt">concentrations</span> are elucidated using partial dependence plots. For example, an increase in the probability of elevated iron and arsenic occurred when well depths coincided with the glacial/bedrock interface, suggesting a bedrock source for these constituents. Furthermore, groundwater in contact with Ordovician bedrock has a higher likelihood of elevated iron <span class="hlt">concentrations</span>, which supports the hypothesis that groundwater liberates iron from a sulfide-bearing secondary cement horizon of Ordovician age. Application of machine learning techniques to existing compliance monitoring data offers an opportunity to broadly assess aquifer and stream vulnerability at regional and national scales and to better understand geochemical processes responsible for <span class="hlt">observed</span> conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70019208','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70019208"><span>Detecting changes in the spatial distribution of <span class="hlt">nitrate</span> contamination in ground water</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Liu, Z.-J.; Hallberg, G.R.; Zimmerman, D.L.; Libra, R.D.</p> <p>1997-01-01</p> <p>Many studies of ground water pollution in general and <span class="hlt">nitrate</span> contamination in particular have often relied on a one-time investigation, tracking of individual wells, or aggregate summaries. Studies of changes in spatial distribution of contaminants over time are lacking. This paper presents a method to compare spatial distributions for possible changes over time. The large-scale spatial distribution at a given time can be considered as a surface over the area (a trend surface). The changes in spatial distribution from period to period can be revealed by the differences in the shape and/or height of surfaces. If such a surface is described by a polynomial function, changes in surfaces can be detected by testing statistically for differences in their corresponding polynomial functions. This method was applied to <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in a population of wells in an agricultural drainage basin in Iowa, sampled in three different years. For the period of 1981-1992, the large-scale spatial distribution of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> did not show significant change in the shape of spatial surfaces; while the magnitude of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> in the basin, or height of the computed surfaces showed significant fluctuations. The change in magnitude of <span class="hlt">nitrate</span> <span class="hlt">concentration</span> is closely related to climatic variations, especially in precipitation. The lack of change in the shape of spatial surfaces means that either the influence of land use/nitrogen management was overshadowed by climatic influence, or the changes in land use/management occurred in a random fashion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999JHyd..214...74W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999JHyd..214...74W"><span>A univariate model of river water <span class="hlt">nitrate</span> time series</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Worrall, F.; Burt, T. P.</p> <p>1999-01-01</p> <p>Four time series were taken from three catchments in the North and South of England. The sites chosen included two in predominantly agricultural catchments, one at the tidal limit and one downstream of a sewage treatment works. A time series model was constructed for each of these series as a means of decomposing the elements controlling river water <span class="hlt">nitrate</span> <span class="hlt">concentrations</span> and to assess whether this approach could provide a simple management tool for protecting water abstractions. Autoregressive (AR) modelling of the detrended and deseasoned time series showed a "memory effect". This memory effect expressed itself as an increase in the winter-summer difference in <span class="hlt">nitrate</span> levels that was dependent upon the <span class="hlt">nitrate</span> <span class="hlt">concentration</span> 12 or 6 months previously. Autoregressive moving average (ARMA) modelling showed that one of the series contained seasonal, non-stationary elements that appeared as an increasing trend in the winter-summer difference. The ARMA model was used to predict <span class="hlt">nitrate</span> levels and predictions were tested against data held back from the model construction process - predictions gave average percentage errors of less than 10%. Empirical modelling can therefore provide a simple, efficient method for constructing management models for downstream water abstraction.</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>