Difference between revisions of "User:Debra Tabron/sandbox"

From Enviro Wiki
Jump to: navigation, search
Line 1: Line 1:
1,4-Dioxane (14D) does not readily biodegrade in most environments. However, a variety of microorganisms have been identified that can biodegrade 14D through either direct growth-related metabolism or cometabolismDuring metabolic biodegradation, microorganisms use 14D as the growth substrate, however growth is slow unless 14D concentrations are very high (>100 mg/L). During cometabolic biodegradation, an additional growth substrate must be supplied to support biomass growth and induce the appropriate 14D-degrading enzymes. Unlike metabolic degradation, cometabolic processes can reduce 14D to very low concentrations.
+
Compound specific isotope analysis (CSIA) usually refers to the measurement of the [[wikipedia: Stable isotope ratio | isotope ratios]] (typically, the stable isotope ratios of carbon, hydrogen, oxygen, nitrogen, sulfur or chlorine) of individual volatile and semi-volatile compounds extracted from complex environmental mixturesThese compounds (or their derivatization products) are generally separable by conventional gas or liquid chromatography and amenable to high-temperature oxidation or pyrolysis to produce simple gases (H<sub>2</sub>, CO<sub>2</sub>, CO, N<sub>2</sub>, O<sub>2</sub>, SO<sub>2</sub>, CH<sub>3</sub>Cl) suitable for analysis by a gas-source isotope-ratio mass spectrometer. The CSIA methodology provides a quantitative means to differentiate reaction pathways for abiotic and biotic degradation, including different pathways of biodegradation, and may serve as a basis for identification of distinct pollutant sources. CSIA can also provide a powerful line of evidence for monitored natural attenuation of sites contaminated with a wide variety of common organic pollutants including petroleum hydrocarbons, fuel oxygenates, BTEX, chlorinated solvents, and nitroaromatics.
 +
<div style="float:right;margin:0 0 2em 2em;">__TOC__</div>
 +
'''Related Article(s)''':
 +
 
 +
*[[Molecular Biological Tools - MBTs]]
 +
 
 +
 
 +
'''CONTRIBUTOR(S):''' [[Dr. Barbara Sherwood Lollar, F.R.S.C.]]
 +
 
 +
 
 +
'''Key Resource(s)''':
  
<div style="float:right;margin:0 0 2em 2em;">__TOC__</div>
+
*[//www.enviro.wiki/images/a/a9/Hunkeler-2008-A_Guide.pdf A Consensus Guide For Assessing Biodegradation and Source Identification Of Organic Contaminants In Groundwater Using Compound Specific Stable Isotope Analysis (CSIA)]<ref name="Hunkeler2008">Hunkeler, D., Meckenstock, R. U., Sherwood Lollar, B., Schmidt, T. C. and Wilson, J. T., 2008. A Guide for Assessing Biodegradation and Source Identification of Organic Groundwater Contaminants Using Compound Specific Isotope Analysis (CSIA). U.S. Environmental Protection Agency, Washington, D.C., EPA/600/R-08/148. [//www.enviro.wiki/images/a/a9/Hunkeler-2008-A_Guide.pdf Report pdf]</ref>
 +
*[https://doi.org/10.1039/c0em00277a Stable Isotope Fractionation to Investigate Natural Transformation Mechanisms of Organic Contaminants: Principles, Prospects and Limitations]<ref name= "Elsner2010.2">Elsner, M., 2010. Stable isotope fractionation to investigate natural transformation mechanisms of organic contaminants: principles, prospects and limitations. Journal of Environmental Monitoring, 12(11), pp.2005-2031. [https://doi.org/10.1039/c0em00277a  doi: 10.1039/C0EM00277A]</ref>
 +
 
 +
==Introduction==
 +
Compound Specific Isotope Analysis (CSIA) refers to the measurement of the [[wikipedia: Stable isotope ratio | isotope ratios]] (typically carbon, hydrogen, oxygen, nitrogen, sulfur or chlorine) of individual organic compounds extracted from complex environmental mixtures. The approach was developed initially during the post-WWII era for application to [[wikipedia: Source rock | source rock]] identification and [[wikipedia: Hydrocarbon exploration | hydrocarbon exploration]], and remains a foundation of the oil and gas industry. In the 1980s, [[wikipedia: John M. Hayes (scientist) | John M. Hayes]] at Indiana University Bloomington and collaborators<ref>Merritt, D.A., Brand, W.A. and Hayes, J.M., 1994. Isotope-ratio-monitoring gas chromatography-mass spectrometry: methods for isotopic calibration. Organic Geochemistry, 21(6-7), 573-583. [https://doi.org/10.1016/0146-6380(94)90003-5 doi: 10.1016/0146-6380(94)90003-5]</ref><ref>Brand, W.A., 1996. High precision isotope ratio monitoring techniques in mass spectrometry. Journal of Mass Spectrometry, 31(3), 225-235. [https://doi.org/10.1002/(sici)1096-9888(199603)31:3 <225::aid-jms319>3.0.co;2-l doi: 10.1002/(SICI)1096-9888(199603)31:3<225::AID-JMS319>3.0.CO;2-L]</ref> introduced the era of continuous flow compound-specific mass spectrometry by interfacing a gas chromatograph via a sample preparatory oxidation system to a stable [[wikipedia:Isotope-ratio mass spectrometry | isotope ratio mass spectrometry]] system, which lowered detection limits by up to 5 orders of magnitude and reduced analytical and sample preparation time. This continuous flow technique allowed CSIA to become widely applied by providing the ability, with instrumentation that became commercially available around 1990, to measure and compare stable isotope ratios for compounds of environmental concern at various spatiotemporal scales in [[wikipedia: Environmental chemistry | environmental chemistry]], [[wikipedia: Biogeochemistry | biogeochemistry]], and contaminant [[wikipedia: Hydrogeology | hydrogeology]].
  
'''Related Article(s):'''
+
==The CSIA Method==
*[[1,4-Dioxane]]
+
The first isotope ratio to which CSIA was widely applied was that of carbon’s stable isotopes.  The element carbon has two stable [[wikipedia:Isotope | isotopes]], <sup>12</sup>C and <sup>13</sup>C. Typically occurring under natural conditions in the ratio of 99:1, the small relative differences in the ratio of <sup>13</sup>C/<sup>12</sup>C for a given compound provide a wealth of information relevant to the investigation and remediation of contaminated sites and the environment<ref name="Hunkeler2008" />. The measured <sup>13</sup>C/<sup>12</sup>C ratio is normalized with respect to international isotopic standard reference materials and expressed in delta notation (e.g. δ<sup>13</sup>C), in units of permil (parts per thousand or [[wikipedia: Per mille | per mille]]). Isotopic standard reference materials for stable isotope ratio measurements of carbon and other light elements (H, N, O, S, Cl and others) have been administered since the 1950s through several sources including the [https://www.iaea.org/ International Atomic Energy Agency], the [https://www.nist.gov/ National Institute of Standards and Technology] and the [https://www.usgs.gov/ U.S. Geological Survey]. All stable isotope laboratories worldwide must report their isotopic data normalized to the appropriate standard reference materials to ensure global consistency and inter-comparability of results<ref name="Hunkeler2008" /><ref>Coplen, T.B., 2011. Guidelines and recommended terms for expression of stable‐isotope‐ratio and gas‐ratio measurement results. Rapid Communications in Mass Spectrometry, 25(17), pp.2538-2560. [https://doi.org/10.1002/rcm.5129 doi: 10.1002/rcm.5129]</ref>.
*[[Biodegradation - Cometabolic]]
 
  
 +
==Applications to Environmental Remediation and Restoration – Forensics==
 +
Both naturally sourced contaminants (e.g., petroleum hydrocarbons) and man-made industrial organic contaminants (e.g., [[wikipedia: Organochloride | organochlorides]], [[wikipedia: Chlorofluorocarbon | chlorofluorocarbons (CFCs)]], [[wikipedia: Nitro compound | nitroaromatics]], and gasoline additives such as [[wikipedia: Methyl tert-butyl ether | methyl tert-butyl ether (MTBE)]]) can have distinct isotopic compositions due to differences in the raw materials and industrial synthesis processes used in their manufacture. This provides the basis for using CSIA as a [[wikipedia: Forensic science | forensic science]] tool<ref>Mancini, S.A., Lacrampe-Couloume, G. and Sherwood Lollar, B., 2008. Source differentiation for benzene and chlorobenzene groundwater contamination: A field application of stable carbon and hydrogen isotope analyses. Environmental Forensics, 9(2-3), 177-186. [http://dx.doi.org/10.1080/15275920802119086 doi: 10.1080/15275920802119086]</ref>. Different contamination sources or different spills from the same source may have distinct isotopic compositions that can be used to apportion responsibility in a mixed contaminant plume, such as in a situation where there is off-site migration and impact. In many cases, knowledge of the isotopic composition of the initial spill material is not available, but the forensic utility of the isotope ratios does not depend on that precondition. Distinguishing between potential source areas at a site can be done by comparing isotopic data from different areas of the site in the context of the geologic and hydrogeologic conceptual models to test source zone apportionment <ref name="Hunkeler2008" />. Although there may be significant overlap in isotopic compositions, successful forensic applications have taken advantage of the additional constraints afforded by coupling carbon isotope ratio measurements with other isotopes (usually [[wikipedia: Isotopes of hydrogen | hydrogen]], [[wikipedia: Isotopes of nitrogen | nitrogen]] and/or [[wikipedia: Isotopes of chlorine | chlorine]] isotope ratios)<ref name="Hunkeler2008" /><ref name="Hunkeler2001">Hunkeler, D., Andersen, N., Aravena, R., Bernasconi, S.M. and Butler, B.J., 2001. Hydrogen and carbon isotope fractionation during aerobic biodegradation of benzene. Environmental Science & Technology, 35(17), 3462-3467. [https://doi.org/10.1021/es0105111 doi: 10.1021/es0105111]</ref><ref>Mancini, S.A., Ulrich, A.C., Lacrampe-Couloume, G., Sleep, B., Edwards, E.A. and Sherwood Lollar, B., 2003. Carbon and hydrogen isotopic fractionation during anaerobic biodegradation of benzene. Applied and Environmental Microbiology, 69(1), 191-198. [https://doi.org/10.1128/aem.69.1.191-198.2003 doi: 10.1128/AEM.69.1.191-198.2003]</ref><ref>Shouakar-Stash, O., Frape, S.K. and Drimmie, R.J., 2003. Stable hydrogen, carbon and chlorine isotope measurements of selected chlorinated organic solvents. Journal of Contaminant Hydrology, 60(3), 211-228. [https://doi.org/10.1016/s0169-7722(02)00085-2 doi: 10.1016/S0169-7722(02)00085-2]</ref><ref name="Kuder2005">Kuder, T., Wilson, J.T., Kaiser, P., Kolhatkar, R., Philp, P. and Allen, J., 2005. Enrichment of stable carbon and hydrogen isotopes during anaerobic biodegradation of MTBE: microcosm and field evidence. Environmental Science & Technology, 39(1), 213-220. [https://doi.org/10.1021/es040420e doi: 10.1021/es040420e]</ref><ref name="Zwank2005">Zwank, L., Berg, M., Elsner, M., Schmidt, T.C., Schwarzenbach, R.P. and Haderlein, S.B., 2005. New evaluation scheme for two-dimensional isotope analysis to decipher biodegradation processes: Application to groundwater contamination by MTBE. Environmental Science & Technology, 39(4), 1018-1029. [https://doi.org/10.1021/es049650j doi: 10.1021/es049650j]</ref><ref>Sessions, A.L., 2006. Isotope‐ratio detection for gas chromatography. Journal of Separation Science, 29(12), 1946-1961. [https://doi.org/10.1002/jssc.200600002 doi: 10.1002/jssc.200600002]</ref><ref>Hartenbach, A., Hofstetter, T. B., Berg, M., Bolotin, J., Schwarzenbach, R. P., 2006. Using nitrogen isotope fractionation to assess abiotic reduction of nitroaromatic compounds. Environmental Science & Technology, 40(24), 7710-7719. [http://pubs.acs.org/doi/abs/10.1021/es061074z doi: 10.1021/es061074z]</ref><ref name="Penning2007">Penning, H. and Elsner, M., 2007. Intramolecular carbon and nitrogen isotope analysis by quantitative dry fragmentation of the phenylurea herbicide isoproturon in a combined injector/capillary reactor prior to GC separation. Analytical Chemistry, 79(21), 8399-8405. [https://doi.org/10.1021/ac071420a doi 10.1021/ac071420a]</ref><ref>Meyer, A.H., Penning, H., Lowag, H. and Elsner, M., 2008. Precise and accurate compound-specific carbon and nitrogen isotope analysis of atrazine: critical role of combustion oven conditions. Environmental Science & Technology, 42(21), 7757-7763. [https://doi.org/10.1021/es800534h doi: 10.1021/es800534h]</ref>.
  
'''CONTRIBUTOR(S):''' [[Dr. Shaily Mahendra]] and [[Dr. Michael Hyman]]
+
==Quantifying and Monitoring Remediation Processes==
 +
Abiotic and biotic degradation reactions that transform contaminant compounds in the environment generally affect their isotopic compositions in systematic ways. This overall process is called  [[wikipedia: Isotope fractionation | isotope fractionation]] and is the key to CSIA providing insight and information on transformation pathways<ref name="Faure2004">Faure, G. and Mensing, T.M., 2005. Isotopes: principles and applications. John Wiley & Sons Inc.</ref><ref name= "Elsner2010.2"/>. While both degradative (e.g. chemical or biological transformation of contaminant to degradation products) and non-degradative processes (e.g. phase changes such as volatilization, sorption, diffusion) have the potential to result in carbon isotope fractionation, to date the largest fractionation signals are related to degradative processes in which bonds are broken<ref name="Hunkeler2008" />. This results from the [[wikipedia: Kinetic isotope effect | kinetic isotope effect]] and the fact that bonds involving a heavy stable isotope (e.g. <sup>13</sup>C-<sup>12</sup>C bond) have a lower [[wikipedia: Zero-point energy | zero-point energy]] and a larger [[wikipedia: Activation energy | activation energy]] than bonds containing exclusively light isotopes (e.g. <sup>12</sup>C-<sup>12</sup>C bond). Effectively this means that the rate of transformation of molecules containing exclusively light isotopes at the reactive site is faster than the rate of transformation of compounds containing a heavy isotope at the reactive site <ref name="Faure2004" /><ref name= "Elsner2010.2"/>.
  
 +
==CSIA Signals of Transformation and Remediation==
 +
The net outcome of isotope fractionation is that a contaminant that has been undergoing degradation can provide a significant signal of transformation, as its isotopic composition generally (but not always) increases to heavier values as a function of reaction progress<ref>Meckenstock, R.U., Morasch, B., Warthmann, R., Schink, B., Annweiler, E., Michaelis, W. and Richnow, H.H., 1999. 13C/12C isotope fractionation of aromatic hydrocarbons during microbial degradation. Environmental Microbiology, 1(5), 409-414. [https://doi.org/10.1046/j.1462-2920.1999.00050.x doi: 10.1046/j.1462-2920.1999.00050.x]</ref><ref>Hunkeler, D., Aravena, R. and Butler, B.J., 1999. Monitoring microbial dechlorination of tetrachloroethene (PCE) in groundwater using compound-specific stable carbon isotope ratios: microcosm and field studies. Environmental Science & Technology, 33(16), 2733-2738. [https://doi.org/10.1021/es981282u doi: 10.1021/es981282u]</ref><ref>Sherwood Lollar, B., Slater, G.F., Ahad, J., Sleep, B., Spivack, J., Brennan, M. and MacKenzie, P., 1999. Contrasting carbon isotope fractionation during biodegradation of trichloroethylene and toluene: Implications for intrinsic bioremediation. Organic Geochemistry, 30(8), 813-820. [https://doi.org/10.1016/s0146-6380(99)00064-9 doi: 10.1016/S0146-6380(99)00064-9]</ref><ref name= "Elsner2010.2"/>. The obvious corollary is that the products of degradation will generally be preferentially enriched in the lighter isotopes relative to the instantaneous isotopic composition of the parent compound from which they are derived. This principle holds for both chemical transformation and biologically mediated transformation reactions, and the principles described above apply to other elements such as hydrogen, nitrogen, oxygen, sulfur, and chlorine as well.  The largest isotope effects have always been found at the reactive sites where bonds are broken or formed.
  
'''Key Resource(s)''':
+
Laboratory experiments have shown that not only does fractionation during transformation provide a strong signal of degradation, but also that the signal is highly reproducible<ref name="Hunkeler2008" />. For many organic contaminants of interest, the relationship between the change in isotopic composition and the degree of degradation is governed by a quantitative relationship – the Rayleigh equation<ref>Mariotti, A., Germon, J.C., Hubert, P., Kaiser, P., Letolle, R., Tardieux, A. and Tardieux, P., 1981. Experimental determination of nitrogen kinetic isotope fractionation: some principles; illustration for the denitrification and nitrification processes. Plant and Soil, 62(3), 413-430. [https://doi.org/10.1007/bf02374138 doi: 10.1007/BF02374138]</ref>. Specifically, for a given compound and degradation pathway or mechanism, the measured difference in isotopic composition can be quantitatively related to the extent of transformation (e.g. fraction or percentage of contaminant remaining) by the equation:
*[https://doi.org/10.1016/j.jenvman.2017.05.033 Advances in bioremediation of 1,4-dioxane-contaminated waters]<ref>Zhang, S., Gedalanga, P.B. and Mahendra, S., 2017. Advances in bioremediation of 1, 4-dioxane-contaminated waters. Journal of environmental management, 204, pp.765-774. [https://doi.org/10.1016/j.jenvman.2017.05.033 doi: 10.1016/j.jenvman.2017.05.033]</ref>
 
  
==1,4-Dioxane Biodegradation==
+
::'''Equation 1:'''&nbsp;&nbsp;&nbsp;&nbsp;&nbsp;&nbsp;<big>''R<sub>t</sub> = R<sub>0</sub> f<sup> (α -1)</sup>''</big>
[[File:Mahendra1w2 Fig1.png|thumb|right|Figure 1.  Growth Rates of Two 14D Metabolizers Versus 14D Concentration]]
 
  
1,4-Dioxane (14D) does not readily biodegrade in most anaerobic environments<ref name= "Shen2008">Shen, W., Chen, H. and Pan, S., 2008. Anaerobic biodegradation of 1, 4-dioxane by sludge enriched with iron-reducing microorganisms. Bioresource technology, 99(7), pp.2483-2487. [https://doi.org/10.1016/j.biortech.2007.04.054 doi: 10.1016/j.biortech.2007.04.054]</ref>. However, multiple studies have demonstrated that 14D can be biodegraded by a variety of microorganisms under aerobic conditions<ref name= "Mahendra2006">Mahendra, S. and Alvarez-Cohen, L., 2006. Kinetics of 1, 4-dioxane biodegradation by monooxygenase-expressing bacteria. Environmental Science & Technology, 40(17), pp.5435-5442. [https://doi.org/10.1021/es060714v doi:10.1021/es060714v]</ref><ref>Mahendra, S., Petzold, C.J., Baidoo, E.E., Keasling, J.D. and Alvarez-Cohen, L., 2007. Identification of the intermediates of in vivo oxidation of 1, 4-dioxane by monooxygenase-containing bacteria. Environmental Science & Technology, 41(21), pp.7330-7336. [https://doi.org/10.1021/es0705745 doi: 10.1021/es0705745]</ref><ref name= "Skinner1536">Skinner, K., Cuiffetti, L. and Hyman, M., 2009. Metabolism and cometabolism of cyclic ethers by a filamentous fungus, a Graphium sp. Appl. Environ. Microbiol., 75(17), pp.5514-5522. [https://doi.org/10.1128/aem.00078-09 doi:10.1128/AEM.00078-09]</ref><ref name = "Sun2011">Sun, B., Ko, K. and Ramsay, J.A., 2011. Biodegradation of 1, 4-dioxane by a Flavobacterium. Biodegradation, 22(3), pp.651-659. [https://doi.org/10.1007/s10532-010-9438-9 doi: 10.1007/s10532-010-9438-9]</ref>. A summary of microorganisms that are reported to aerobically biodegrade 14D is presented in Table 1.
+
where ''R<sub>t</sub>'' is the stable isotope ratio (<sup>13</sup>C/<sup>12</sup>C) of the compound at time ''t'', ''R<sub>0</sub>'' is the initial isotope ratio of the compound and ''f'' is the fraction of contaminant remaining where ''f'' = 1 at ''t'' = 0 and decreases to ''f'' = 0 when the reactant compound is fully transformed to products. The stable isotope fractionation factor (''α'') is defined as:  
  
Under aerobic conditions 14D can be biodegraded through two physiologically distinct processes: (a) metabolism; and (b) cometabolism.  Metabolism is a process in which microorganisms use the organic contaminants as a carbon and energy source to support their growth. [https://enviro.wiki/index.php?title=Biodegradation_-_Cometabolic Cometabolism] occurs when microorganisms degrade contaminants using non-specific enzymes but do not gain carbon or energy to support growth from the degradation process.
+
::'''Equation 2:'''&nbsp;&nbsp;&nbsp;&nbsp;&nbsp;&nbsp;<big>''α'' = (1000 + ''δ''<sup>13</sup>C<sub>''a''</sub>)/(1000 + ''δ''<sup>13</sup>C<sub>''b''</sub>)</big>
 
{| class="wikitable floatright"
 
|+ Table 1.  List of 1,4-Dioxane Degrading Microorganisms and Biodegradation Rates
 
|-
 
! Strain
 
! Induced Enzyme
 
! Biodegradation Rate
 
! Reference
 
|-
 
| colspan="4" | '''Metabolism'''
 
|-
 
| ''Pseudonocardia dioxanivorans'' CB1190 || THFMO || 0.19 ± 0.007 mg/hr/mg-protein || Mahendra and Alvarez-Cohen (2005, 2006)<ref name= "Mahendra2005">Mahendra, S. and Alvarez-Cohen, L., 2005. Pseudonocardia dioxanivorans sp. nov., a novel actinomycete that grows on 1, 4-dioxane. International Journal of Systematic and Evolutionary Microbiology, 55(2), pp.593-598. [https://doi.org/10.1099/ijs.0.63085-0 doi: 10.1099/ijs.0.63085-0]</ref><ref name= "Mahendra2006"/>
 
|-
 
| ''Actinomycete'' CB1190* || N/A || 0.33 mg/min/mg-protein || Parales et al. (1994)<ref name= "Parales1994">Parales, R.E., Adamus, J.E., White, N. and May, H.D., 1994. Degradation of 1, 4-dioxane by an actinomycete in pure culture. Applied and Environmental Microbiology, 60(12), pp.4527-4530. [[media:1994-Parales-Degradation_of_1%2C4-Dioxane_by_an_Actinomycete_in_Pure_Culture.pdf| Report.pdf]]</ref>
 
|-
 
| ''Amycolata'' sp. CB1190* || N/A || 0.038 ± 0.012 mg/hr/mg-protein || Kelley et al. (2001)
 
|-
 
| ''Pseudonocardia benzenivorans'' B5 || THFMO || 0.01± 0.003 mg/hr/mg-protein || Mahendra and Alvarez-Cohen (2006)<ref name= "Mahendra2006"/>
 
|-
 
| ''Pseudonocardia carboxydivorans'' RM-31 || ? || 31.6 mg/L/hr || Matsui et al. (2016)<ref>Matsui, R., Takagi, K., Sakakibara, F., Abe, T. and Shiiba, K., 2016. Identification and characterization of 1, 4-dioxane-degrading microbe separated from surface seawater by the seawater-charcoal perfusion apparatus. Biodegradation, 27(2-3), pp.155-163. [https://doi.org/10.1007/s10532-016-9763-8 doi: 10.1007/s10532-016-9763-8</ref>
 
|-
 
| ''Afipia'' sp. D1 || ? || 0.052 to 0.263 mg/hr/mg-protein || Sei et al. (2013)<ref name =  "Sei2013">Sei, K., Miyagaki, K., Kakinoki, T., Fukugasako, K., Inoue, D. and Ike, M., 2013. Isolation and characterization of bacterial strains that have high ability to degrade 1, 4-dioxane as a sole carbon and energy source. Biodegradation, 24(5), pp.665-674. [https://doi.org/10.1007/s10532-012-9614-1 doi: 10.1007/s10532-012-9614-1]</ref>
 
|-
 
| ''Mycobacterium'' sp. PH-06 || PrMO || 2.5 mg/L/hr || Kim et al. (2009)<ref name = "Kim2009">Kim, Y.M., Jeon, J.R., Murugesan, K., Kim, E.J. and Chang, Y.S., 2009. Biodegradation of 1, 4-dioxane and transformation of related cyclic compounds by a newly isolated Mycobacterium sp. PH-06. Biodegradation, 20(4), p.511. [https://doi.org/10.1007/s10532-008-9240-0 doi: 10.1007/s10532-008-9240-0]</ref>
 
|-
 
| ''Acinetobacter baumannii'' DD1 || ? || 2.38 mg/L/hr || Huang et al. (2014)<ref name = "Huang2014">Huang, H., Shen, D., Li, N., Shan, D., Shentu, J. and Zhou, Y., 2014. Biodegradation of 1, 4-dioxane by a novel strain and its biodegradation pathway. Water, Air, & Soil Pollution, 225(9), p.2135. [https://doi.org/10.1007/s11270-014-2135-2 doi: 10.1007/s11270-014-2135-2]</ref>
 
|-
 
| ''Xanthobacter flavus'' DT8 || ? || Similar to CB1190 || Chen et al. (2016)<ref>Chen, D.Z., Jin, X.J., Chen, J., Ye, J.X., Jiang, N.X. and Chen, J.M., 2016. Intermediates and substrate interaction of 1, 4-dioxane degradation by the effective metabolizer Xanthobacter flavus DT8. International Biodeterioration & Biodegradation, 106, pp.133-140. [https://doi.org/10.1016/j.ibiod.2015.09.018 doi: 10.1016/j.ibiod.2015.09.018]</ref>
 
|-
 
| ''Cordyceps sinensis'' (fungus) || ? || 0.011 mol/day || Nakamiya et al. (2005)<ref>Nakamiya, K., Hashimoto, S., Ito, H., Edmonds, J.S. and Morita, M., 2005. Degradation of 1, 4-dioxane and cyclic ethers by an isolated fungus. Appl. Environ. Microbiol., 71(3), pp.1254-1258. [https://doi.org/10.1128/aem.71.3.1254-1258.2005 doi: 10.1128/AEM.71.3.1254-1258.2005]</ref> 
 
|-
 
| colspan="4" | '''Cometabolism'''
 
|-
 
| ''Mycobacterium austroafricanum'' JOB5 || ? || 0.40 ± 0.06 mg/hr/mg-protein || House and Hyman (2010)<ref>House, A.J. and Hyman, M.R., 2010. Effects of gasoline components on MTBE and TBA cometabolism by Mycobacterium austroafricanum JOB5. Biodegradation, 21(4), pp.525-541. [https://doi.org/10.1007/s10532-009-9321-8 doi: 10.1007/s10532-009-9321-8]</ref> Lan et al. (2013)<ref>Lan, R.S., Smith, C.A. and Hyman, M.R., 2013. Oxidation of cyclic ethers by alkane‐grown Mycobacterium vaccae JOB5. Remediation Journal, 23(4), pp.23-42. [https://doi.org/10.1002/rem.21364 doi: 10.1002/rem.21364}</ref><ref name= "Mahendra2006"/>
 
|-
 
| ''Rhodococcus ruber'' ENV425 || ? || 10 mg/hr/g TSS || Lippincott et al. (2015)<ref name= "Lippincott2015">Lippincott, D., Streger, S.H., Schaefer, C.E., Hinkle, J., Stormo, J. and Steffan, R.J., 2015. Bioaugmentation and propane biosparging for in situ biodegradation of 1, 4‐dioxane. Groundwater Monitoring & Remediation, 35(2), pp.81-92. [https://doi.org/10.1111/gwmr.12093 doi: 10.1111/gwmr.12093]</ref>; Vainberg et al. (2006)<ref name= "Vainberg2006">Vainberg, S., McClay, K., Masuda, H., Root, D., Condee, C., Zylstra, G.J. and Steffan, R.J., 2006. Biodegradation of ether pollutants by Pseudonocardia sp. strain ENV478. Appl. Environ. Microbiol., 72(8), pp.5218-5224. [https://doi.org/10.1128/aem.00160-06 doi: 10.1128/AEM.00160-06]</ref>
 
|-
 
| ''Pseudonocardia'' sp. ENV478 || THFMO || 21 mg/hr/g TSS || Masuda et al. (2012)<ref>Masuda, H., McClay, K., Steffan, R.J. and Zylstra, G.J., 2012. Biodegradation of tetrahydrofuran and 1, 4-dioxane by soluble diiron monooxygenase in Pseudonocardia sp. strain ENV478. Journal of Molecular Microbiology and Biotechnology, 22(5), pp.312-316. [https://doi.org/10.1159/000343817  doi: 10.1159/000343817]</ref> Vainberg et al. (2006)<ref name= "Vainberg2006"/>
 
|-
 
| ''Rhodococcus'' RR1 || ? || 0.38 ± 0.03 mg/hr/mg-protein || Mahendra and Alvarez-Cohen (2006)<ref name= "Mahendra2006"/>
 
|-
 
| ''Rhodococcus jostii'' RHA1 || PrMO || N/A || Hand et al. (2015)<ref name= "Hand2015">Hand, S., Wang, B. and Chu, K.H., 2015. Biodegradation of 1, 4-dioxane: effects of enzyme inducers and trichloroethylene. Science of the Total Environment, 520, pp.154-159. [https://doi.org/10.1016/j.scitotenv.2015.03.031 doi: 10.1016/j.scitotenv.2015.03.031]</ref>; Li et al. (2013)<ref name= "Li2013">Li, M., Mathieu, J., Yang, Y., Fiorenza, S., Deng, Y., He, Z., Zhou, J. and Alvarez, P.J., 2013. Widespread distribution of soluble di-iron monooxygenase (SDIMO) genes in arctic groundwater impacted by 1, 4-dioxane. Environmental science & technology, 47(17), pp.9950-9958. [https://doi.org/10.1021/es402228x doi: 10.1021/es402228x]</ref>
 
|-
 
| ''Flavobacterium'' || ? || N/A || Sun et al. (2011)<ref name = "Sun2011"/>
 
|-
 
| ''Pseudonocardia'' K1 || THFMO || 0.26 ± 0.013 mg/hr/mg-protein || Mahendra and Alvarez-Cohen (2006)<ref name= "Mahendra2006"/>
 
|-
 
| ''Burkholderia cepacia'' G4 || T2MO || 0.1± 0.006 mg/hr/mg-protein || Mahendra and Alvarez-Cohen (2006)<ref name= "Mahendra2006"/>
 
|-
 
| ''Ralstonia pickettii'' PKO1 || T3MO || 0.31± 0.007 mg/hr/mg-protein || Mahendra and Alvarez-Cohen (2006)<ref name= "Mahendra2006"/>
 
|-
 
| ''Pseudomonas mendocina'' KR1 || T4MO || 0.37± 0.04 mg/hr/mg-protein || Mahendra and Alvarez-Cohen (2006)<ref name= "Mahendra2006"/>
 
|-
 
| ''Aureobasidium pullmans'' NRRL 21064 || ? || 6-8 mg/L within a day || Patt and Abebe (1995)<ref name= "Patt1995">Patt, T.E. and Abebe, H.M., Upjohn Co, 1995. Microbial degradation of chemical pollutants. U.S. Patent 5,399,495. [[media:1995-Patt-Microbial_degradation_of_chemical_pollutants.pdf| Report.pdf]]</ref>
 
|-
 
| ''Graphium'' sp. ATCC 58400 (fungus) || CYP || 4 ± 1 nmol/min/mg dry weight (with Propane) <br/>9 ± 5 nmol/min/mg dry weight (with THF) || Skinner et al. (2009)<ref name= "Skinner2009">Skinner, K., Cuiffetti, L. and Hyman, M., 2009. Metabolism and cometabolism of cyclic ethers by a filamentous fungus, a Graphium sp. Appl. Environ. Microbiol., 75(17), pp.5514-5522. [https://doi.org/10.1128/aem.00078-09 doi:10.1128/AEM.00078-09]</ref>
 
|}
 
==Aerobic Metabolism==
 
While several microorganisms have been isolated that can metabolize 14D<ref name= "Mahendra2006"/><ref name = "Kim2009"/><ref>Sei, K., Kakinoki, T., Inoue, D., Soda, S., Fujita, M. and Ike, M., 2010. Evaluation of the biodegradation potential of 1, 4-dioxane in river, soil and activated sludge samples. Biodegradation, 21(4), pp.585-591. [https://doi.org/10.1007/s10532-010-9326-3 doi: 10.1007/s10532-010-9326-3]</ref><ref name =  "Sei2013"/><ref name = "Huang2014"/>, these organisms are not common.  The best characterized 14D-metabolizing strain is ''Pseudonocardia dioxanivorans'' CB1190<ref name= "Mahendra2005"/>. This bacterium was originally enriched from industrial activated sludge, fed with tetrahydrofuran (THF) and then subsequently fed with 14D<ref name= "Parales1994"/>.  The doubling time of CB1190 was about 30 hours when it was grown in ammonium mineral salts medium at 30 °C amended with 5.5 mM (484 mg/L) 14D<ref name= "Parales1994"/>.
 
  
Degradation of 14D by strain CB1190 may be inhibited by elevated concentrations of chlorinated solvents and their degradation products, and by some metals. Inhibition of 14D degradation was strongest for 1,1-dichloroethene (1,1-DCE) followed by ''cis''-1,2-diochloroethene (cDCE) > trichloroethene (TCE) > 1,1,1-trichloroethane (TCA). 14D biodegradation was completely inhibited by 5 mg/L 1,1-DCE<ref>Mahendra, S., Grostern, A. and Alvarez-Cohen, L., 2013. The impact of chlorinated solvent co-contaminants on the biodegradation kinetics of 1, 4-dioxane. Chemosphere, 91(1), pp.88-92. [https://doi.org/10.1016/j.chemosphere.2012.10.104 doi: 10.1016/j.chemosphere.2012.10.104]</ref><ref name= "Hand2015"/><ref name= "Zhang2016">Zhang, S., Gedalanga, P.B. and Mahendra, S., 2016. Biodegradation kinetics of 1, 4-dioxane in chlorinated solvent mixtures. Environmental Science & Technology, 50(17), pp.9599-9607. [https://doi.org/10.1021/acs.est.6b02797 doi: 10.1021/acs.est.6b02797]</ref>. Cu(II) was the strongest metal inhibitor of 14D degradation by CB1190, causing an increase in the lag period at 1 mg/L and an order of magnitude reduction in 14D degradation rates at 10 and 20 mg/L Cu(II).  14D degradation was less sensitive to Cd(II) and Ni(II), while Zn(II) had no impact on 14D biodegradation at the maximum concentration tested (20 mg/L Zn)<ref>Pornwongthong, P., Mulchandani, A., Gedalanga, P.B. and Mahendra, S., 2014. Transition metals and organic ligands influence biodegradation of 1, 4-dioxane. Applied biochemistry and biotechnology, 173(1), pp.291-306. [https://doi.org/10.1007/s12010-014-0841-2 doi: 10.1007/s12010-014-0841-2]</ref>.
+
where subscripts ''a'' and ''b'' may represent a compound at time zero (''t''<sub>0</sub>) and at a later point (''t'') in a reaction; or a compound in a source zone, versus the compound in a downgradient well for instance.
  
An important challenge for in situ bioremediation of 14D is the slow growth of 14D metabolizers at typical groundwater concentrations.  Figure 1 shows estimated growth rates for two 14D metabolizing organisms (''P. dioxanivorans'' CB1190 and ''P. benzenivorans'' B5) computed using published kinetic parameters<ref name= "Mahendra2006"/>.  At a 14D concentrations of 1000 µg/L, CB1190 and B2 have growth rates of 0.015 and 0.002 per day.  At typical groundwater concentrations (<1000 µg/L), growth rates are expected to be less than the endogenous decay rate, resulting in a steady decline in the number of 14D metabolizers.
+
Equation 1 can be rearranged to produce Equation 3<ref name="Hunkeler2008" />:
  
==Aerobic Cometabolism==
+
::'''Equation 3:'''&nbsp;&nbsp;&nbsp;&nbsp;&nbsp;&nbsp;<big>''f'' = ''e''</big><sup>(''δ''<sup>13</sup>C<sub>groundwater</sub> - ''δ''<sup>13</sup>C<sub>source</sub>)<big> '''/''' ''ε''</big></sup>
Unlike 14D-metabolizing microorganisms, 14D-cometabolizing organisms do not grow on 14D but can degrade this compound after growth on a primary, growth-supporting substrate. As cometabolically active microorganisms do not use co-substrates as either major carbon or energy sources, they can often degrade co-substrates at concentrations well below those that can be achieved by organisms that metabolize and grow on these co-substrates. A wide variety of bacteria<ref name= "Mahendra2006"/> and some fungi<ref name= "Patt1995"/><ref name= "Skinner2009"/> can cometabolically degrade 14D (see Table 1).  These include model organisms that grow on primary substrates such as toluene or methane<ref name= "Mahendra2006"/> and involve well-characterized enzymes such as soluble methane monooxygenase (sMMO) or toluene monooxygenases (TxMO), respectively.  Other 14D-cometabolizing strains that grow on THF<ref name= "Vainberg2006"/>, ethane<ref>Hatzinger, P.B., Banerjee, R., Rezes, R., Streger, S.H., McClay, K. and Schaefer, C.E., 2017. Potential for cometabolic biodegradation of 1, 4-dioxane in aquifers with methane or ethane as primary substrates. Biodegradation, 28(5-6), pp.453-468. [https://doi.org/10.1007/s10532-017-9808-7 doi: 0.1007/s10532-017-9808-7 ]</ref>, and isobutane<ref name= "Bennett2017">Bennett, P., Hyman, M., Smith, C., El Mugammar, H., Chu, M.Y., Nickelsen, M. and Aravena, R., 2018. Enrichment with carbon-13 and deuterium during monooxygenase-mediated biodegradation of 1, 4-dioxane. Environmental Science & Technology Letters, 5(3), pp.148-153. [https://doi.org/10.1021/acs.estlett.7b00565 doi: 10.1021/acs.estlett.7b00565]</ref> have also been described.
 
  
Much of the research into 14D cometabolism has focused on high concentrations of 14D (≥100 mg/L), and less is currently known about the activity of specific microorganisms and their monooxygenases at lower, more environmentally relevant 14D concentrations (<100 µg/L). Despite this, many bacterial monooxygenases can concurrently oxidize multiple co-substrates, and recent field studies have demonstrated that cometabolic approaches involving either propane biostimulation and bioaugmentation<ref name= "Lippincott2015"/> or propane biostimulation alone<ref>Chu, M.Y.J., Bennett, P.J., Dolan, M.E., Hyman, M.R., Peacock, A.D., Bodour, A., Anderson, R.H., Mackay, D.M. and Goltz, M.N., 2018. Concurrent Treatment of 1, 4‐Dioxane and Chlorinated Aliphatics in a Groundwater Recirculation System Via Aerobic Cometabolism. Groundwater Monitoring & Remediation, 38(3), pp.53-64. [https://doi.org/10.1111/gwmr.12293 doi: 10.1111/gwmr.12293]</ref> can be highly effective treatment strategies for degrading contaminant mixtures that contain parts-per-billion concentrations of both 14D and associated chlorinated co-contaminants.
+
where ''δ''<sup>13</sup>C<sub>groundwater</sub> is the measure of the isotope ratio in the organic contaminant in the sample of groundwater, ''δ''<sup>13</sup>C<sub>source</sub> is the isotopic ratio in the un-fractionated organic contaminant before biodegradation has occurred, and epsilon (''ε'') is the stable isotope enrichment factor, defined as:  
  
==Anaerobic Biodegradation==
+
::'''Equation 4:'''&nbsp;&nbsp;&nbsp;&nbsp;&nbsp;&nbsp;<big>''ε'' = (''α'' -1) * 1000</big>
To date, there is little evidence of metabolic or cometabolic anaerobic 14D biodegradation.  In a microcosm study using samples of aquifer material from several different 14D impacted sites, there was no evidence of 14D biodegradation<ref>Zenker, M.J., Borden, R.C. and Barlaz, M.A., 1999. Investigation of the intrinsic biodegradation of alkyl and cyclic ethers. The Fifth International In Situ and On-Site Bioremediation Symposium</ref>.  However, there is one study that reported anaerobic growth of an iron-reducing bacterium on 14D<ref name= "Shen2008"/>.
 
  
An indirect method of anaerobic 14D biodegradation exists via a microbially driven Fenton reaction. ''Shewanella oneidensis'', an Fe(III)-reducing facultative anaerobe, is able to generate hydroxyl radicals that can then break down 14D<ref>Sekar, R. and DiChristina, T.J., 2014. Microbially driven Fenton reaction for degradation of the widespread environmental contaminant 1, 4-dioxane. Environmental science & technology, 48(21), pp.12858-12867. [https://doi.org/10.1021/es503454a doi: 10.1021/es503454a]</ref>. Under anaerobic conditions, ''S. oneidensis'' produces Fe(II) that can then interact chemically with H<sub>2</sub>O<sub>2</sub> to yield HO• radicals which can oxidatively degrade 1,4-dioxane.
+
==Implications for Remediation==
 +
The quantitative relationships controlling isotope fractionation during transformation enable CSIA to provide a diagnostic signal that transformation is taking place, as well as a quantitative measure of the extent of transformation independent of conventional metrics based on changes in concentration. In some cases, due to signal sensitivity, changes in stable carbon isotope fractionation can be identified in advance of definitive reduction in contaminant concentrations, or before appearance of daughter products, providing an “early warning system” for confirmation of remediation<ref name="Morrill2005">Morrill, P.L., Lacrampe-Couloume, G., Slater, G.F., Sleep, B.E., Edwards, E.A., McMaster, M.L., Major, D.W. and Sherwood Lollar, B., 2005. Quantifying chlorinated ethene degradation during reductive dechlorination at Kelly AFB using stable carbon isotopes. Journal of Contaminant Hydrology, 76(3), pp.279-293. [https://doi.org/10.1016/j.jconhyd.2004.11.002 doi: 10.1016/j.jconhyd.2004.11.002]</ref>. This is particularly advantageous for field studies since changes in contaminant concentration result not only from transformation processes, but from physical transport and dispersal. For this reason, decreasing concentrations of contaminants alone are insufficient evidence that a site is undergoing transformation towards clean-up goals<ref>Wiedemeier, T.H., Wilson, J.T., Kampbell, D.H., Miller, R.N. and Hansen, J.E., 1995. Technical Protocol for Implementing Intrinsic Remediation with Long-Term Monitoring for Natural Attenuation of Fuel Contamination Dissolved in Groundwater. U.S. Air Force Center for Environmental Excellence, Technology Transfer Division, Brooks Air Force Base, San Antonio, Texas.</ref><ref>Wiedemeier, T.H.,  Swanson, M.A., Moutoux, D.E., Gordon, E.K., Wilson, J.T., Wilson, B.H., Kampbell, D.H., Haas, P.E., Hansen, J.E., and Chapelle, F.H. 1998. Technical protocol for evaluating natural attenuation of chlorinated solvents in groundwater.  EPA-600-R-98-128. [//www.enviro.wiki/images/2/27/Wiedemeier-1998-Technical_Protocol_for_Evaluating_Natuaral_Attenuation.pdf Report pdf]</ref>. In contrast, as physical transport and dispersal processes without concomitant degradation produce negligible changes in isotopic composition, any significant isotope effects measured in the contaminants of concern provide a direct line of evidence that transformation is occurring, and also yields independent quantification of the rate of the transformation<ref>Sherwood Lollar, B., Slater, G.F., Sleep, B., Witt, M., Klecka, G.M., Harkness, M. and Spivack, J., 2001. Stable carbon isotope evidence for intrinsic bioremediation of tetrachloroethene and trichloroethene at area 6, Dover Air Force Base. Environmental Science & Technology, 35(2), 261-269. [https://doi.org/10.1021/es001227x doi: 10.1021/es001227x]</ref><ref name="Morrill2005" /><ref>McKelvie, J.R., Mackay, D.M., de Sieyes, N.R., Lacrampe-Couloume, G. and Sherwood Lollar, B., 2007. Quantifying MTBE biodegradation in the Vandenberg Air Force Base ethanol release study using stable carbon isotopes. Journal of Contaminant Hydrology, 94(3), 157-165. [https://doi.org/10.1016/j.jconhyd.2007.05.008 doi: 10.1016/j.jconhyd.2007.05.008]</ref>.
  
==Molecular Biological Tools==
+
CSIA provides additional value to environmental investigation and remediation in that the degree of fractionation is reaction specific, giving researchers the ability to pinpoint which of a variety of possible degradation mechanisms may be dominating at a contaminated site. A specific example of this is 1,2-dichloroethane, an industrial chemical used in PVC production, production of furniture, upholstery and automobile parts and a common environmental contaminant of concern. Microbial biodegradation of this compound in the environment is common, but different organisms degrade the compound via different pathways (e.g. involving a C-Cl bond cleavage, or a C-H bond cleavage). As a result, CSIA can be used to positively identify which of the biodegradation pathways is operative at a site – information that can be critical to optimizing a remediation strategy<ref>Hunkeler, D. and Aravena, R., 2000. Evidence of Substantial Carbon Isotope Fractionation among Substrate, Inorganic Carbon, and Biomass during Aerobic Mineralization of 1, 2-Dichloroethane by Xanthobacter autotrophicus. Applied and Environmental Microbiology, 66(11), 4870-4876. [https://doi.org/10.1128/aem.66.11.4870-4876.2000 doi: 10.1128/AEM.66.11.4870-4876.2000]</ref><ref>Hirschorn, S.K., Grostern, A., Lacrampe-Couloume, G., Edwards, E.A., MacKinnon, L., Repta, C., Major, D.W. and Sherwood Lollar, B., 2007. Quantification of biotransformation of chlorinated hydrocarbons in a biostimulation study: Added value via stable carbon isotope analysis. Journal of Contaminant Hydrology, 94(3), 249-260. [https://doi.org/10.1016/j.jconhyd.2007.07.001 doi: 10.1016/j.jconhyd.2007.07.001]</ref>. In other examples, CSIA has been a critical tool in deciphering the biodegradation potential and remediation mechanisms for benzene<ref name="Hunkeler2001" /><ref>Mancini, S.A., Lacrampe-Couloume, G., Jonker, H., van Breukelen, B.M., Groen, J., Volkering, F., Sherwood Lollar, B., 2002. Hydrogen isotope enrichment: An indicator of biodegradation at a petroleum hydrocarbon contaminated field site. Environmental Science & Technology, 36(11), 2464-2470. [http://pubs.acs.org/doi/abs/10.1021/es011253a doi: 10.1021/es011253a]</ref><ref>Mancini, S.A., Devine, C.E., Elsner, M., Nandi, M.E., Ulrich, A.C., Edwards, E.A. and Sherwood Lollar, B., 2008. Isotopic evidence suggests different initial reaction mechanisms for anaerobic benzene biodegradation. Environmental Science & Technology, 42(22), 8290-8296. [https://doi.org/10.1021/es801107g doi: 10.1021/es801107g]</ref><ref>Fischer, A., Gehre, M., Breitfeld, J., Richnow, H.-H., 2009. Carbon and hydrogen isotope fractionation of benzene during biodegradation under sulphate-reducing conditions: A laboratory to field site approach. Rapid Communications in Mass Spectrometry, 236, 2439-2447. [https://doi.org/10.1002/rcm.4049 doi:10.1002/rcm.4049]</ref>, methyl tert-butyl ether (MTBE)<ref name="Zwank2005" /><ref>McKelvie, J.R., Hyman, M.R., Elsner, M., Smith, C., Aslett, D.M., Lacrampe-Couloume, G. and Sherwood Lollar, B., 2009. Isotopic fractionation of methyl tert-butyl ether suggests different initial reaction mechanisms during aerobic biodegradation. Environmental Science & Technology, 43(8), 2793-2799. [https://doi.org/10.1021/es803307y doi: 10.1021/es803307y]</ref><ref>Elsner, M., McKelvie, J., Lacrampe Couloume, G. and Sherwood Lollar, B., 2007. Insight into methyl tert-butyl ether (MTBE) stable isotope fractionation from abiotic reference experiments. Environmental Science & Technology, 41(16), 5693-5700. [https://doi.org/10.1021/es070531o doi: 10.1021/es070531o]</ref><ref name="Kuder2005" /> and other priority pollutants. In related applications, where abiotic and biotic transformations of a compound occur via different pathways and mechanisms, CSIA can differentiate between the relative contributions of chemical versus biological transformations<ref>Elsner, M., Chartrand, M., VanStone, N., Lacrampe Couloume, G. and Sherwood Lollar, B., 2008. Identifying abiotic chlorinated ethene degradation: characteristic isotope patterns in reaction products with nanoscale zero-valent iron. Environmental Science & Technology, 42(16), 5963-5970. [https://doi.org/10.1021/es8001986 doi: 10.1021/es8001986]</ref><ref>Elsner, M., Lacrampe Couloume, G., Mancini, S., Burns, L. and Sherwood Lollar, B., 2010. Carbon isotope analysis to evaluate nanoscale Fe (O) treatment at a chlorohydrocarbon contaminated site. Groundwater Monitoring & Remediation, 30(3), 79-95. [https://doi.org/10.1111/j.1745-6592.2010.01294.x doi: 10.1111/j.1745-6592.2010.01294.x]</ref>.  
[https://enviro.wiki/index.php?title=Quantitative_Polymerase_Chain_Reaction_(qPCR) Quantitative polymerase chain reaction (qPCR)] quantifies the abundance of specific microorganisms and functional genes capable of degrading a particular contaminant. qPCR analyses have been employed to quantify the abundance of microorganisms that can degrade 14D through the activity of tetrahydrofuran monooxygenase (THFMO) in samples from industrial activated sludge and groundwater<ref>Chiang, S.Y.D., Mora, R., Diguiseppi, W.H., Davis, G., Sublette, K., Gedalanga, P. and Mahendra, S., 2012. Characterizing the intrinsic bioremediation potential of 1, 4-dioxane and trichloroethene using innovative environmental diagnostic tools. Journal of Environmental Monitoring, 14(9), pp.2317-2326. [https://doi.org/10.1039/c2em30358b doi: 10.1039/C2EM30358B]</ref><ref>Gedalanga, P.B., Pornwongthong, P., Mora, R., Chiang, S.Y.D., Baldwin, B., Ogles, D. and Mahendra, S., 2014. Identification of biomarker genes to predict biodegradation of 1, 4-dioxane. Appl. Environ. Microbiol., 80(10), pp.3209-3218. [https://doi.org/10.1128/aem.04162-13 doi: 10.1128/AEM.04162-13]</ref><ref>Li, M., Mathieu, J., Liu, Y., Van Orden, E.T., Yang, Y., Fiorenza, S. and Alvarez, P.J., 2013. The abundance of tetrahydrofuran/dioxane monooxygenase genes (thmA/dxmA) and 1, 4-dioxane degradation activity are significantly correlated at various impacted aquifers. Environmental Science & Technology Letters, 1(1), pp.122-127. [https://doi.org/10.1021/ez400176h doi: 10.1021/ez400176h]</ref><ref name= "Li2013"/><ref name= "Zhang2016"/>. However, care is warranted in the interpretation of qPCR results in other cases (e.g. propane-stimulated cometabolic 14D degradation) where the role of enzymes such as propane monooxygenase (PrMO) in 14D degradation is less clearly established.  Studies also suggest that these biomarkers can be used to detect the abundance of 14D-degraders ''in situ'' and how they compete within the larger microbial community<ref>Miao, Y., Johnson, N.W., Gedalanga, P.B., Adamson, D., Newell, C. and Mahendra, S., 2019. Response and recovery of microbial communities subjected to oxidative and biological treatments of 1, 4-dioxane and co-contaminants. Water research, 149, pp.74-85. [https://doi.org/10.1016/j.watres.2018.10.070 doi:10.1016/j.watres.2018.10.070]</ref>. Microbial community analyses can be an asset towards guiding treatment strategies and predicting treatment synergies.
 
  
[https://enviro.wiki/index.php?title=Compound_Specific_Isotope_Analysis_(CSIA) Compound specific isotope analysis (CSIA)] refers to measurement of the isotopic signatures of individual chemical compounds and can be used to differentiate contaminant sources, delineate reaction pathways, and provide evidence of ''in situ'' contaminant degradation. CSIA methods are now available to measure isotopic fractionation of carbon and hydrogen<ref name= "Bennett2017"/>. For example, the aerobic biodegradation of 14D by a THF-grown 14D-degrading ''Pseudonocardia'' strain exhibited an isotopic fractionation factor (<big>''ε'' </big>) for carbon (<big>''ε''</big><sub>''c''</sub>)  of −4.73 ± 0.9‰ and hydrogen (<big>ε</big><sub>''H''</sub>) of -147  ± 22‰, respectively. Smaller <big>''ε''</big><sub>''c''</sub>and <big>ε</big><sub>''H''</sub> values, (-2.7 ± 0.3‰ and -21 ± 2‰, respectively) were determined for aerobic 14D degradation by a propane-grown ''Rhodococcus'' strain. As many ''Pseudonocardia'' strains use THFMO to initiate 14D degradation, CSIA, in conjunction with qPCR analyses, may be able to discriminate the roles of metabolism and cometabolism in 14D biodegradation at field sites<ref>Gedalanga, P., Madison, A., Miao, Y., Richards, T., Hatton, J., DiGuiseppi, W.H., Wilson, J. and Mahendra, S., 2016. A Multiple Lines of Evidence Framework to Evaluate Intrinsic Biodegradation of 1, 4‐Dioxane. Remediation Journal, 27(1), pp.93-114. [https://doi.org/10.1002/rem.21499 doi: 10.1002/rem.21499]</ref>.  
+
Not all transformation processes necessarily result in measurable isotopic fractionation. Fractionation factors can be small simply due to isotope dilution, e.g., where a C isotope effect occurs at a single reactive site in a molecule that has many carbon atoms. In such cases there is need for the development of models to disambiguate intrinsic versus apparent kinetic isotope effects <ref>Elsner, M., Zwank, L., Hunkeler, D. and Schwarzenbach, R.P., 2005. A new concept linking observable stable isotope fractionation to transformation pathways of organic pollutants. Environmental Science & Technology, 39(18), 6896-6916. [https://doi.org/10.1021/es0504587 doi: 10.1021/es0504587]</ref>, for use of multi-isotope analysis<ref name="Penning2007" />, and for the use of novel techniques such as <sup>13</sup>C NMR that can allow position-specific isotope analysis<ref>McKelvie, J.R., Elsner, M., Simpson, A.J., Sherwood Lollar, B., Simpson, M.J., 2010. Quantitative site-specific 2H NMR investigation of MTBE: Potential for investigating contaminant sources and fate. Environmental Science & Technology, 44(3), 1062-1068. [http://pubs.acs.org/doi/abs/10.1021/es9030276 doi: 10.1021/es9030276]</ref><ref>Julien, M., Parinet, J., Nun, P., Bayle, K., Höhener, P., Robins, R.J. and Remaud, G.S., 2015. Fractionation in position-specific isotope composition during vaporization of environmental pollutants measured with isotope ratio monitoring by <sup>13</sup>C nuclear magnetic resonance spectrometry. Environmental Pollution, 205, 299-306. [https://doi.org/10.1016/j.envpol.2015.05.047 doi: 10.1016/j.envpol.2015.05.047]</ref><ref>Gilbert, A., Yamada, K., Suda, K., Ueno, Y. and Yoshida, N., 2016. Measurement of position-specific <sup>13</sup>C isotopic composition of propane at the nanomole level. Geochimica et Cosmochimica Acta, 177, 205-216. [http://dx.doi.org/10.1016/j.gca.2016.01.017 doi: 10.1016/j.gca.2016.01.017]</ref>. The presence of additional rate-limiting steps in the transformation reaction can complicate the measurement of isotope fractionation in ways that may obscure the extent of transformation, yet may also yield other important information about transport effects<ref>Nijenhuis, I., Andert, J., Beck, K., Kastner, M., Diekert, G., Richnow, H-H., 2005. Stable isotope fractionation of tetrachloroethene during reductive dechlorination by sulfurospirillum multivorans and desulfitobacterium sp. Strain PCE-S and abiotic reactions with cyanocobalamin. Applied and Environmental Microbiology, 71(7), 3413-3419. [https://www.ncbi.nlm.nih.gov/pmc/articles/PMC1169044/ doi: 10.1128/AEM.71.7.3413-3419.2005]</ref> or the efficiency of the enzymes involved in biodegradation<ref>Mancini, S.A., Hirschorn, S.K., Elsner, M., Lacrampe-Couloume, G., Sleep, B.E., Edwards, E.A. and Sherwood Lollar, B., 2006. Effects of trace element concentration on enzyme controlled stable isotope fractionation during aerobic biodegradation of toluene. Environmental Science & Technology, 40(24), 7675-7681. [https://doi.org/10.1021/es061363n doi: 10.1021/es061363n]</ref><ref>Sherwood Lollar, B., Hirschorn, S., Mundle, S.O., Grostern, A., Edwards, E.A. and Lacrampe-Couloume, G., 2010. Insights into enzyme kinetics of chloroethane biodegradation using compound-specific stable isotopes. Environmental Science & Technology, 44(19), 7498-7503. [http://dx.doi.org/10.1021/es101330r doi: 10.1021/es101330r]</ref>.
  
 
==Summary==
 
==Summary==
14D is not readily biodegraded under ambient conditions in most environments.  However, a variety of microorganisms have been identified that can biodegrade 14D through either growth-related metabolism or fortuitous cometabolism. Metabolic growth on the target pollutant is the more traditional approach for both above ground and ''in situ'' bioremediation approaches. However, approaches based on metabolic growth may not be feasible for treatment of 14D at typical concentrations. Cometabolic treatment approaches can reduce 14D to very low levels and have been demonstrated in the field. However, implementation of cometabolic treatment is more complex, and there is currently less engineering experience in the design and operation of these systems.  
+
Applications of CSIA are dependent on background information about the isotope fractionation factors associated with specific chemical reactions or biodegradation pathways. While fractionation can be calculated ''ab initio'' (from the beginning) by using molecular modeling methods, fractionation factors are typically empirically derived from laboratory experiments and other approaches. Recent guidance documents and review papers provide an essential resource with database compilations of this knowledge to date<ref>Meckenstock, R.U., Morasch, B., Griebler, C. and Richnow, H.H., 2004. Stable isotope fractionation analysis as a tool to monitor biodegradation in contaminated acquifers. Journal of Contaminant Hydrology, 75(3), 215-255. [https://doi.org/10.1016/j.jconhyd.2004.06.003 doi: 10.1016/j.jconhyd.2004.06.003]</ref><ref name="Hunkeler2008" /><ref>Elsner, M. and Imfeld, G., 2016. Compound-specific isotope analysis (CSIA) of micropollutants in the environment—current developments and future challenges. Current opinion in biotechnology, 41, pp.60-72. [https://doi.org/10.1016/j.copbio.2016.04.014 doi: 10.1016/j.copbio.2016.04.014]</ref>.
  
 
==References==
 
==References==
 +
 
<references />
 
<references />
  
 
==See Also==
 
==See Also==
 +
 +
*[https://www.serdp-estcp.org/Program-Areas/Environmental-Restoration/Contaminated-Groundwater/Emerging-Issues/ER-200509 Validation of Chlorine and Oxygen Isotope Ratio Analysis to Differentiate Perchlorate Sources and to Document Perchlorate Biodegradation]
 +
*[https://www.serdp-estcp.org/Program-Areas/Environmental-Restoration/Contaminated-Groundwater/ER-201029 Integrated Stable Isotope-Reactive Transport Model Approach for Assessment of Chlorinated Solvent Degradation]
 +
*[https://www.coursera.org/learn/natural-attenuation-of-groundwater-contaminants/lecture/C0EN7/types-of-isotope-analysis-and-sampling-techniques-lab-analysis Online Lecture Course - Stable Isotopes]

Revision as of 13:30, 23 October 2019

Compound specific isotope analysis (CSIA) usually refers to the measurement of the isotope ratios (typically, the stable isotope ratios of carbon, hydrogen, oxygen, nitrogen, sulfur or chlorine) of individual volatile and semi-volatile compounds extracted from complex environmental mixtures. These compounds (or their derivatization products) are generally separable by conventional gas or liquid chromatography and amenable to high-temperature oxidation or pyrolysis to produce simple gases (H2, CO2, CO, N2, O2, SO2, CH3Cl) suitable for analysis by a gas-source isotope-ratio mass spectrometer. The CSIA methodology provides a quantitative means to differentiate reaction pathways for abiotic and biotic degradation, including different pathways of biodegradation, and may serve as a basis for identification of distinct pollutant sources. CSIA can also provide a powerful line of evidence for monitored natural attenuation of sites contaminated with a wide variety of common organic pollutants including petroleum hydrocarbons, fuel oxygenates, BTEX, chlorinated solvents, and nitroaromatics.

Related Article(s):


CONTRIBUTOR(S): Dr. Barbara Sherwood Lollar, F.R.S.C.


Key Resource(s):

Introduction

Compound Specific Isotope Analysis (CSIA) refers to the measurement of the isotope ratios (typically carbon, hydrogen, oxygen, nitrogen, sulfur or chlorine) of individual organic compounds extracted from complex environmental mixtures. The approach was developed initially during the post-WWII era for application to source rock identification and hydrocarbon exploration, and remains a foundation of the oil and gas industry. In the 1980s, John M. Hayes at Indiana University Bloomington and collaborators[3][4] introduced the era of continuous flow compound-specific mass spectrometry by interfacing a gas chromatograph via a sample preparatory oxidation system to a stable isotope ratio mass spectrometry system, which lowered detection limits by up to 5 orders of magnitude and reduced analytical and sample preparation time. This continuous flow technique allowed CSIA to become widely applied by providing the ability, with instrumentation that became commercially available around 1990, to measure and compare stable isotope ratios for compounds of environmental concern at various spatiotemporal scales in environmental chemistry, biogeochemistry, and contaminant hydrogeology.

The CSIA Method

The first isotope ratio to which CSIA was widely applied was that of carbon’s stable isotopes. The element carbon has two stable isotopes, 12C and 13C. Typically occurring under natural conditions in the ratio of 99:1, the small relative differences in the ratio of 13C/12C for a given compound provide a wealth of information relevant to the investigation and remediation of contaminated sites and the environment[1]. The measured 13C/12C ratio is normalized with respect to international isotopic standard reference materials and expressed in delta notation (e.g. δ13C), in units of permil (parts per thousand or per mille). Isotopic standard reference materials for stable isotope ratio measurements of carbon and other light elements (H, N, O, S, Cl and others) have been administered since the 1950s through several sources including the International Atomic Energy Agency, the National Institute of Standards and Technology and the U.S. Geological Survey. All stable isotope laboratories worldwide must report their isotopic data normalized to the appropriate standard reference materials to ensure global consistency and inter-comparability of results[1][5].

Applications to Environmental Remediation and Restoration – Forensics

Both naturally sourced contaminants (e.g., petroleum hydrocarbons) and man-made industrial organic contaminants (e.g., organochlorides, chlorofluorocarbons (CFCs), nitroaromatics, and gasoline additives such as methyl tert-butyl ether (MTBE)) can have distinct isotopic compositions due to differences in the raw materials and industrial synthesis processes used in their manufacture. This provides the basis for using CSIA as a forensic science tool[6]. Different contamination sources or different spills from the same source may have distinct isotopic compositions that can be used to apportion responsibility in a mixed contaminant plume, such as in a situation where there is off-site migration and impact. In many cases, knowledge of the isotopic composition of the initial spill material is not available, but the forensic utility of the isotope ratios does not depend on that precondition. Distinguishing between potential source areas at a site can be done by comparing isotopic data from different areas of the site in the context of the geologic and hydrogeologic conceptual models to test source zone apportionment [1]. Although there may be significant overlap in isotopic compositions, successful forensic applications have taken advantage of the additional constraints afforded by coupling carbon isotope ratio measurements with other isotopes (usually hydrogen, nitrogen and/or chlorine isotope ratios)[1][7][8][9][10][11][12][13][14][15].

Quantifying and Monitoring Remediation Processes

Abiotic and biotic degradation reactions that transform contaminant compounds in the environment generally affect their isotopic compositions in systematic ways. This overall process is called isotope fractionation and is the key to CSIA providing insight and information on transformation pathways[16][2]. While both degradative (e.g. chemical or biological transformation of contaminant to degradation products) and non-degradative processes (e.g. phase changes such as volatilization, sorption, diffusion) have the potential to result in carbon isotope fractionation, to date the largest fractionation signals are related to degradative processes in which bonds are broken[1]. This results from the kinetic isotope effect and the fact that bonds involving a heavy stable isotope (e.g. 13C-12C bond) have a lower zero-point energy and a larger activation energy than bonds containing exclusively light isotopes (e.g. 12C-12C bond). Effectively this means that the rate of transformation of molecules containing exclusively light isotopes at the reactive site is faster than the rate of transformation of compounds containing a heavy isotope at the reactive site [16][2].

CSIA Signals of Transformation and Remediation

The net outcome of isotope fractionation is that a contaminant that has been undergoing degradation can provide a significant signal of transformation, as its isotopic composition generally (but not always) increases to heavier values as a function of reaction progress[17][18][19][2]. The obvious corollary is that the products of degradation will generally be preferentially enriched in the lighter isotopes relative to the instantaneous isotopic composition of the parent compound from which they are derived. This principle holds for both chemical transformation and biologically mediated transformation reactions, and the principles described above apply to other elements such as hydrogen, nitrogen, oxygen, sulfur, and chlorine as well. The largest isotope effects have always been found at the reactive sites where bonds are broken or formed.

Laboratory experiments have shown that not only does fractionation during transformation provide a strong signal of degradation, but also that the signal is highly reproducible[1]. For many organic contaminants of interest, the relationship between the change in isotopic composition and the degree of degradation is governed by a quantitative relationship – the Rayleigh equation[20]. Specifically, for a given compound and degradation pathway or mechanism, the measured difference in isotopic composition can be quantitatively related to the extent of transformation (e.g. fraction or percentage of contaminant remaining) by the equation:

Equation 1:      Rt = R0 f (α -1)

where Rt is the stable isotope ratio (13C/12C) of the compound at time t, R0 is the initial isotope ratio of the compound and f is the fraction of contaminant remaining where f = 1 at t = 0 and decreases to f = 0 when the reactant compound is fully transformed to products. The stable isotope fractionation factor (α) is defined as:

Equation 2:      α = (1000 + δ13Ca)/(1000 + δ13Cb)

where subscripts a and b may represent a compound at time zero (t0) and at a later point (t) in a reaction; or a compound in a source zone, versus the compound in a downgradient well for instance.

Equation 1 can be rearranged to produce Equation 3[1]:

Equation 3:      f = e(δ13Cgroundwater - δ13Csource) / ε

where δ13Cgroundwater is the measure of the isotope ratio in the organic contaminant in the sample of groundwater, δ13Csource is the isotopic ratio in the un-fractionated organic contaminant before biodegradation has occurred, and epsilon (ε) is the stable isotope enrichment factor, defined as:

Equation 4:      ε = (α -1) * 1000

Implications for Remediation

The quantitative relationships controlling isotope fractionation during transformation enable CSIA to provide a diagnostic signal that transformation is taking place, as well as a quantitative measure of the extent of transformation independent of conventional metrics based on changes in concentration. In some cases, due to signal sensitivity, changes in stable carbon isotope fractionation can be identified in advance of definitive reduction in contaminant concentrations, or before appearance of daughter products, providing an “early warning system” for confirmation of remediation[21]. This is particularly advantageous for field studies since changes in contaminant concentration result not only from transformation processes, but from physical transport and dispersal. For this reason, decreasing concentrations of contaminants alone are insufficient evidence that a site is undergoing transformation towards clean-up goals[22][23]. In contrast, as physical transport and dispersal processes without concomitant degradation produce negligible changes in isotopic composition, any significant isotope effects measured in the contaminants of concern provide a direct line of evidence that transformation is occurring, and also yields independent quantification of the rate of the transformation[24][21][25].

CSIA provides additional value to environmental investigation and remediation in that the degree of fractionation is reaction specific, giving researchers the ability to pinpoint which of a variety of possible degradation mechanisms may be dominating at a contaminated site. A specific example of this is 1,2-dichloroethane, an industrial chemical used in PVC production, production of furniture, upholstery and automobile parts and a common environmental contaminant of concern. Microbial biodegradation of this compound in the environment is common, but different organisms degrade the compound via different pathways (e.g. involving a C-Cl bond cleavage, or a C-H bond cleavage). As a result, CSIA can be used to positively identify which of the biodegradation pathways is operative at a site – information that can be critical to optimizing a remediation strategy[26][27]. In other examples, CSIA has been a critical tool in deciphering the biodegradation potential and remediation mechanisms for benzene[7][28][29][30], methyl tert-butyl ether (MTBE)[11][31][32][10] and other priority pollutants. In related applications, where abiotic and biotic transformations of a compound occur via different pathways and mechanisms, CSIA can differentiate between the relative contributions of chemical versus biological transformations[33][34].

Not all transformation processes necessarily result in measurable isotopic fractionation. Fractionation factors can be small simply due to isotope dilution, e.g., where a C isotope effect occurs at a single reactive site in a molecule that has many carbon atoms. In such cases there is need for the development of models to disambiguate intrinsic versus apparent kinetic isotope effects [35], for use of multi-isotope analysis[14], and for the use of novel techniques such as 13C NMR that can allow position-specific isotope analysis[36][37][38]. The presence of additional rate-limiting steps in the transformation reaction can complicate the measurement of isotope fractionation in ways that may obscure the extent of transformation, yet may also yield other important information about transport effects[39] or the efficiency of the enzymes involved in biodegradation[40][41].

Summary

Applications of CSIA are dependent on background information about the isotope fractionation factors associated with specific chemical reactions or biodegradation pathways. While fractionation can be calculated ab initio (from the beginning) by using molecular modeling methods, fractionation factors are typically empirically derived from laboratory experiments and other approaches. Recent guidance documents and review papers provide an essential resource with database compilations of this knowledge to date[42][1][43].

References

  1. ^ 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 Hunkeler, D., Meckenstock, R. U., Sherwood Lollar, B., Schmidt, T. C. and Wilson, J. T., 2008. A Guide for Assessing Biodegradation and Source Identification of Organic Groundwater Contaminants Using Compound Specific Isotope Analysis (CSIA). U.S. Environmental Protection Agency, Washington, D.C., EPA/600/R-08/148. Report pdf
  2. ^ 2.0 2.1 2.2 2.3 Elsner, M., 2010. Stable isotope fractionation to investigate natural transformation mechanisms of organic contaminants: principles, prospects and limitations. Journal of Environmental Monitoring, 12(11), pp.2005-2031. doi: 10.1039/C0EM00277A
  3. ^ Merritt, D.A., Brand, W.A. and Hayes, J.M., 1994. Isotope-ratio-monitoring gas chromatography-mass spectrometry: methods for isotopic calibration. Organic Geochemistry, 21(6-7), 573-583. doi: 10.1016/0146-6380(94)90003-5
  4. ^ Brand, W.A., 1996. High precision isotope ratio monitoring techniques in mass spectrometry. Journal of Mass Spectrometry, 31(3), 225-235. <225::aid-jms319>3.0.co;2-l doi: 10.1002/(SICI)1096-9888(199603)31:3<225::AID-JMS319>3.0.CO;2-L
  5. ^ Coplen, T.B., 2011. Guidelines and recommended terms for expression of stable‐isotope‐ratio and gas‐ratio measurement results. Rapid Communications in Mass Spectrometry, 25(17), pp.2538-2560. doi: 10.1002/rcm.5129
  6. ^ Mancini, S.A., Lacrampe-Couloume, G. and Sherwood Lollar, B., 2008. Source differentiation for benzene and chlorobenzene groundwater contamination: A field application of stable carbon and hydrogen isotope analyses. Environmental Forensics, 9(2-3), 177-186. doi: 10.1080/15275920802119086
  7. ^ 7.0 7.1 Hunkeler, D., Andersen, N., Aravena, R., Bernasconi, S.M. and Butler, B.J., 2001. Hydrogen and carbon isotope fractionation during aerobic biodegradation of benzene. Environmental Science & Technology, 35(17), 3462-3467. doi: 10.1021/es0105111
  8. ^ Mancini, S.A., Ulrich, A.C., Lacrampe-Couloume, G., Sleep, B., Edwards, E.A. and Sherwood Lollar, B., 2003. Carbon and hydrogen isotopic fractionation during anaerobic biodegradation of benzene. Applied and Environmental Microbiology, 69(1), 191-198. doi: 10.1128/AEM.69.1.191-198.2003
  9. ^ Shouakar-Stash, O., Frape, S.K. and Drimmie, R.J., 2003. Stable hydrogen, carbon and chlorine isotope measurements of selected chlorinated organic solvents. Journal of Contaminant Hydrology, 60(3), 211-228. doi: 10.1016/S0169-7722(02)00085-2
  10. ^ 10.0 10.1 Kuder, T., Wilson, J.T., Kaiser, P., Kolhatkar, R., Philp, P. and Allen, J., 2005. Enrichment of stable carbon and hydrogen isotopes during anaerobic biodegradation of MTBE: microcosm and field evidence. Environmental Science & Technology, 39(1), 213-220. doi: 10.1021/es040420e
  11. ^ 11.0 11.1 Zwank, L., Berg, M., Elsner, M., Schmidt, T.C., Schwarzenbach, R.P. and Haderlein, S.B., 2005. New evaluation scheme for two-dimensional isotope analysis to decipher biodegradation processes: Application to groundwater contamination by MTBE. Environmental Science & Technology, 39(4), 1018-1029. doi: 10.1021/es049650j
  12. ^ Sessions, A.L., 2006. Isotope‐ratio detection for gas chromatography. Journal of Separation Science, 29(12), 1946-1961. doi: 10.1002/jssc.200600002
  13. ^ Hartenbach, A., Hofstetter, T. B., Berg, M., Bolotin, J., Schwarzenbach, R. P., 2006. Using nitrogen isotope fractionation to assess abiotic reduction of nitroaromatic compounds. Environmental Science & Technology, 40(24), 7710-7719. doi: 10.1021/es061074z
  14. ^ 14.0 14.1 Penning, H. and Elsner, M., 2007. Intramolecular carbon and nitrogen isotope analysis by quantitative dry fragmentation of the phenylurea herbicide isoproturon in a combined injector/capillary reactor prior to GC separation. Analytical Chemistry, 79(21), 8399-8405. doi 10.1021/ac071420a
  15. ^ Meyer, A.H., Penning, H., Lowag, H. and Elsner, M., 2008. Precise and accurate compound-specific carbon and nitrogen isotope analysis of atrazine: critical role of combustion oven conditions. Environmental Science & Technology, 42(21), 7757-7763. doi: 10.1021/es800534h
  16. ^ 16.0 16.1 Faure, G. and Mensing, T.M., 2005. Isotopes: principles and applications. John Wiley & Sons Inc.
  17. ^ Meckenstock, R.U., Morasch, B., Warthmann, R., Schink, B., Annweiler, E., Michaelis, W. and Richnow, H.H., 1999. 13C/12C isotope fractionation of aromatic hydrocarbons during microbial degradation. Environmental Microbiology, 1(5), 409-414. doi: 10.1046/j.1462-2920.1999.00050.x
  18. ^ Hunkeler, D., Aravena, R. and Butler, B.J., 1999. Monitoring microbial dechlorination of tetrachloroethene (PCE) in groundwater using compound-specific stable carbon isotope ratios: microcosm and field studies. Environmental Science & Technology, 33(16), 2733-2738. doi: 10.1021/es981282u
  19. ^ Sherwood Lollar, B., Slater, G.F., Ahad, J., Sleep, B., Spivack, J., Brennan, M. and MacKenzie, P., 1999. Contrasting carbon isotope fractionation during biodegradation of trichloroethylene and toluene: Implications for intrinsic bioremediation. Organic Geochemistry, 30(8), 813-820. doi: 10.1016/S0146-6380(99)00064-9
  20. ^ Mariotti, A., Germon, J.C., Hubert, P., Kaiser, P., Letolle, R., Tardieux, A. and Tardieux, P., 1981. Experimental determination of nitrogen kinetic isotope fractionation: some principles; illustration for the denitrification and nitrification processes. Plant and Soil, 62(3), 413-430. doi: 10.1007/BF02374138
  21. ^ 21.0 21.1 Morrill, P.L., Lacrampe-Couloume, G., Slater, G.F., Sleep, B.E., Edwards, E.A., McMaster, M.L., Major, D.W. and Sherwood Lollar, B., 2005. Quantifying chlorinated ethene degradation during reductive dechlorination at Kelly AFB using stable carbon isotopes. Journal of Contaminant Hydrology, 76(3), pp.279-293. doi: 10.1016/j.jconhyd.2004.11.002
  22. ^ Wiedemeier, T.H., Wilson, J.T., Kampbell, D.H., Miller, R.N. and Hansen, J.E., 1995. Technical Protocol for Implementing Intrinsic Remediation with Long-Term Monitoring for Natural Attenuation of Fuel Contamination Dissolved in Groundwater. U.S. Air Force Center for Environmental Excellence, Technology Transfer Division, Brooks Air Force Base, San Antonio, Texas.
  23. ^ Wiedemeier, T.H., Swanson, M.A., Moutoux, D.E., Gordon, E.K., Wilson, J.T., Wilson, B.H., Kampbell, D.H., Haas, P.E., Hansen, J.E., and Chapelle, F.H. 1998. Technical protocol for evaluating natural attenuation of chlorinated solvents in groundwater. EPA-600-R-98-128. Report pdf
  24. ^ Sherwood Lollar, B., Slater, G.F., Sleep, B., Witt, M., Klecka, G.M., Harkness, M. and Spivack, J., 2001. Stable carbon isotope evidence for intrinsic bioremediation of tetrachloroethene and trichloroethene at area 6, Dover Air Force Base. Environmental Science & Technology, 35(2), 261-269. doi: 10.1021/es001227x
  25. ^ McKelvie, J.R., Mackay, D.M., de Sieyes, N.R., Lacrampe-Couloume, G. and Sherwood Lollar, B., 2007. Quantifying MTBE biodegradation in the Vandenberg Air Force Base ethanol release study using stable carbon isotopes. Journal of Contaminant Hydrology, 94(3), 157-165. doi: 10.1016/j.jconhyd.2007.05.008
  26. ^ Hunkeler, D. and Aravena, R., 2000. Evidence of Substantial Carbon Isotope Fractionation among Substrate, Inorganic Carbon, and Biomass during Aerobic Mineralization of 1, 2-Dichloroethane by Xanthobacter autotrophicus. Applied and Environmental Microbiology, 66(11), 4870-4876. doi: 10.1128/AEM.66.11.4870-4876.2000
  27. ^ Hirschorn, S.K., Grostern, A., Lacrampe-Couloume, G., Edwards, E.A., MacKinnon, L., Repta, C., Major, D.W. and Sherwood Lollar, B., 2007. Quantification of biotransformation of chlorinated hydrocarbons in a biostimulation study: Added value via stable carbon isotope analysis. Journal of Contaminant Hydrology, 94(3), 249-260. doi: 10.1016/j.jconhyd.2007.07.001
  28. ^ Mancini, S.A., Lacrampe-Couloume, G., Jonker, H., van Breukelen, B.M., Groen, J., Volkering, F., Sherwood Lollar, B., 2002. Hydrogen isotope enrichment: An indicator of biodegradation at a petroleum hydrocarbon contaminated field site. Environmental Science & Technology, 36(11), 2464-2470. doi: 10.1021/es011253a
  29. ^ Mancini, S.A., Devine, C.E., Elsner, M., Nandi, M.E., Ulrich, A.C., Edwards, E.A. and Sherwood Lollar, B., 2008. Isotopic evidence suggests different initial reaction mechanisms for anaerobic benzene biodegradation. Environmental Science & Technology, 42(22), 8290-8296. doi: 10.1021/es801107g
  30. ^ Fischer, A., Gehre, M., Breitfeld, J., Richnow, H.-H., 2009. Carbon and hydrogen isotope fractionation of benzene during biodegradation under sulphate-reducing conditions: A laboratory to field site approach. Rapid Communications in Mass Spectrometry, 236, 2439-2447. doi:10.1002/rcm.4049
  31. ^ McKelvie, J.R., Hyman, M.R., Elsner, M., Smith, C., Aslett, D.M., Lacrampe-Couloume, G. and Sherwood Lollar, B., 2009. Isotopic fractionation of methyl tert-butyl ether suggests different initial reaction mechanisms during aerobic biodegradation. Environmental Science & Technology, 43(8), 2793-2799. doi: 10.1021/es803307y
  32. ^ Elsner, M., McKelvie, J., Lacrampe Couloume, G. and Sherwood Lollar, B., 2007. Insight into methyl tert-butyl ether (MTBE) stable isotope fractionation from abiotic reference experiments. Environmental Science & Technology, 41(16), 5693-5700. doi: 10.1021/es070531o
  33. ^ Elsner, M., Chartrand, M., VanStone, N., Lacrampe Couloume, G. and Sherwood Lollar, B., 2008. Identifying abiotic chlorinated ethene degradation: characteristic isotope patterns in reaction products with nanoscale zero-valent iron. Environmental Science & Technology, 42(16), 5963-5970. doi: 10.1021/es8001986
  34. ^ Elsner, M., Lacrampe Couloume, G., Mancini, S., Burns, L. and Sherwood Lollar, B., 2010. Carbon isotope analysis to evaluate nanoscale Fe (O) treatment at a chlorohydrocarbon contaminated site. Groundwater Monitoring & Remediation, 30(3), 79-95. doi: 10.1111/j.1745-6592.2010.01294.x
  35. ^ Elsner, M., Zwank, L., Hunkeler, D. and Schwarzenbach, R.P., 2005. A new concept linking observable stable isotope fractionation to transformation pathways of organic pollutants. Environmental Science & Technology, 39(18), 6896-6916. doi: 10.1021/es0504587
  36. ^ McKelvie, J.R., Elsner, M., Simpson, A.J., Sherwood Lollar, B., Simpson, M.J., 2010. Quantitative site-specific 2H NMR investigation of MTBE: Potential for investigating contaminant sources and fate. Environmental Science & Technology, 44(3), 1062-1068. doi: 10.1021/es9030276
  37. ^ Julien, M., Parinet, J., Nun, P., Bayle, K., Höhener, P., Robins, R.J. and Remaud, G.S., 2015. Fractionation in position-specific isotope composition during vaporization of environmental pollutants measured with isotope ratio monitoring by 13C nuclear magnetic resonance spectrometry. Environmental Pollution, 205, 299-306. doi: 10.1016/j.envpol.2015.05.047
  38. ^ Gilbert, A., Yamada, K., Suda, K., Ueno, Y. and Yoshida, N., 2016. Measurement of position-specific 13C isotopic composition of propane at the nanomole level. Geochimica et Cosmochimica Acta, 177, 205-216. doi: 10.1016/j.gca.2016.01.017
  39. ^ Nijenhuis, I., Andert, J., Beck, K., Kastner, M., Diekert, G., Richnow, H-H., 2005. Stable isotope fractionation of tetrachloroethene during reductive dechlorination by sulfurospirillum multivorans and desulfitobacterium sp. Strain PCE-S and abiotic reactions with cyanocobalamin. Applied and Environmental Microbiology, 71(7), 3413-3419. doi: 10.1128/AEM.71.7.3413-3419.2005
  40. ^ Mancini, S.A., Hirschorn, S.K., Elsner, M., Lacrampe-Couloume, G., Sleep, B.E., Edwards, E.A. and Sherwood Lollar, B., 2006. Effects of trace element concentration on enzyme controlled stable isotope fractionation during aerobic biodegradation of toluene. Environmental Science & Technology, 40(24), 7675-7681. doi: 10.1021/es061363n
  41. ^ Sherwood Lollar, B., Hirschorn, S., Mundle, S.O., Grostern, A., Edwards, E.A. and Lacrampe-Couloume, G., 2010. Insights into enzyme kinetics of chloroethane biodegradation using compound-specific stable isotopes. Environmental Science & Technology, 44(19), 7498-7503. doi: 10.1021/es101330r
  42. ^ Meckenstock, R.U., Morasch, B., Griebler, C. and Richnow, H.H., 2004. Stable isotope fractionation analysis as a tool to monitor biodegradation in contaminated acquifers. Journal of Contaminant Hydrology, 75(3), 215-255. doi: 10.1016/j.jconhyd.2004.06.003
  43. ^ Elsner, M. and Imfeld, G., 2016. Compound-specific isotope analysis (CSIA) of micropollutants in the environment—current developments and future challenges. Current opinion in biotechnology, 41, pp.60-72. doi: 10.1016/j.copbio.2016.04.014

See Also