Fourier Transform Ion Cyclotron Resonance Mass Spectrometry

Nov 9, 2014 - complex the structure of the NAs or NAFCs, such as increasing the number .... Instrument control, data acquisition, and data analysis us...
0 downloads 0 Views 838KB Size
Article pubs.acs.org/EF

Fourier Transform Ion Cyclotron Resonance Mass Spectrometry Characterization of Treated Athabasca Oil Sands Processed Waters John V. Headley,*,† Pardeep Kumar,‡ Ajay Dalai,‡ Kerry M. Peru,† Jon Bailey,† Dena W. McMartin,§ Steven M. Rowland,∥ Ryan P. Rodgers,∥,⊥ and Alan G. Marshall∥,⊥ †

Water Science Technology Directorate, Environment Canada, 11 Innovation Boulevard, Saskatoon, Saskatchewan S7N 3H5, Canada Department of Chemical and Biological Engineering, University of Saskatchewan, Saskatoon, Saskatchewan S7N 5A9, Canada § Environmental Systems Engineering, University of Regina, 3737 Wascana Parkway, Regina, Saskatchewan S4S 0A2, Canada ∥ Department of Chemistry and Biochemistry, Florida State University, 95 Chieftain Way, Tallahassee, Florida 32306, United States ⊥ Ion Cyclotron Resonance Program, National High Magnetic Field Laboratory, Florida State University, 1800 East Paul Dirac Drive, Tallahassee, Florida 32310-4005, United States ‡

ABSTRACT: Ultrahigh-resolution negative-ion electrospray ionization Fourier transform ion cyclotron resonance mass spectrometry (FT-ICR MS) was used to characterize Athabasca oil sands processed water (OSPW) treated by (A) coagulation flocculent with lime and bentonite, (B) coagulation flocculent with lime and bentonite followed by activated carbon, and (C) combined ozonation and ultrasonication. Treatment A was ineffective in reducing the level of total naphthenic acid fraction components [NAFCs, defined as the acid-extractable fraction of OSPWs or crude oils (CnH2n + zOwSxNy), where the values n, w, x, and y indicate the number of carbon, oxygen, sulfur, and nitrogen atoms, respectively, and z represents the hydrogen atom deficiency because of the presence of double bonds and ring formation]. Likewise, for treatment A, little or no change was observed for the double bond equivalent (DBE) distributions of the compound classes. Treatments B and C resulted in the reduction of total NAFCs by 26 ± 1.4 and 89 ± 1.1%, respectively. For the latter treatments, there was evidence for selective removal of the S and OxSy heteroatomic species at the molecular level, along with a reduction in the DBE values for all species.



presence of double bonds and ring formation.7 The more complex the structure of the NAs or NAFCs, such as increasing the number of rings, the nature of the alkyl chains, or the number and type of heteroatoms, generally the more persistent the NAFCs to degradation in the environment.4,7 Biodegradation of some NAFCs can be relatively slow, with half-lives of ∼14 years.4,9 Thus, other treatments apart from natural biodegradation are required for reclamation of OSPW within practical time periods.4 In earlier work, Kumar et al.10 reported the treatment of OSPW with combined physical and chemical methods. The earlier studies were limited to monitoring the reduction of chemical oxygen demand (COD) and analyses of total NAs. Changes occurring at the molecular level for the treated OSPW were restricted to the distribution of classical structures of NAs. Because the earlier work provided limited insight into changes occurring at the molecular level for heteroatomic structures of NAFCs, we have conducted follow-up studies to characterize the treated OSPW by ultrahigh-resolution mass spectrometry (MS). The treatments investigated here include two physical methods that were not reported by Kumar et al.,10 namely, (A) coagulation flocculent with lime and bentonite, along with (B) coagulation flocculent with lime and bentonite followed by

INTRODUCTION The Athabasca oil sands deposits in Alberta, Canada, are one of the world’s largest reserves of petroleum.1 The ever increasing energy demands and dwindling conventional oil reserves have drawn more attention to alternative energy resources. The oil sands crude production is projected to grow from 2.1 to 3.7 million barrels per day by 2020. Bitumen is extracted from oil sands by use of hot alkaline water or in situ methods, such as steam-assisted gravity drainage.2−4 The extraction of the bitumen from surface mining results in vast amounts of oil sands processed water (OSPW) or produced water. The current practice of storing the OSPW in tailing ponds is not sustainable, and much research is geared at developing reclamation strategies.2−4 Polar organic compounds, such as naphthenic acids, are suspected to be among the principal toxic components, and thus, much attention is given to treatment methods that specifically target the acid fraction of OSPW.3 The naphthenic acid (NA) concentrations [or more correctly the naphthenic acid fraction components (NAFCs)] in OSPW may be as high as 20−120 mg/L.5,6 NAFCs (CnH2n + zOwSxNy) are defined as the acid-extractable fraction of OSPWs, and the term will also be used here for brevity to extend to crude oils.7,8 This fraction was traditionally called NAs (CnH2n + zO2) but is now known to include a wide range of organic acids containing N and S atoms, along with other components with various levels of unsaturation and aromaticity.7,8 Here, n, w, x, and y indicate the number of carbon, oxygen, sulfur, and nitrogen atoms, respectively, and z represents the hydrogen atom deficiency because of the © XXXX American Chemical Society

Special Issue: 15th International Conference on Petroleum Phase Behavior and Fouling Received: September 8, 2014 Revised: November 7, 2014

A

dx.doi.org/10.1021/ef502007b | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

Table 1. Measured Values of NAFCs in OSPW before and after Treatments A, B, and Ca description

NAFC concentration (mg/L)

additional information

treatments A and B initial OSPW treatment A OSPW exposed to flocculent after filtration treatment A initial OSPW treatment A OSPW exposed to flocculent after filtration experiment 1 CF initial experiment 1 AT

32.5 30.1

initial OSPW control treatment A: T1-AT (after treatment A) with 1 mL of alum (50%) and 2 mL (polymer)

28.6 26.3

treatment A: T3-initial treatment A: T3-AT (after treatment A) with 5 mL of lime (50%) and 2 mL (polymer)

36.3 34.9

0 C, 15 min 1 G C, 30 min

23.1 18.2

1 G C, 90 min

16.8

experiment 3 initial before

71.5

experiment 3 1 h

10.7

experiment 4 initial before

87.7

experiment 4 1 h

11.3

treatment A: experiment 1 coagulation floculation initial (samples from an OSPW tailing pond) treatment A: experiment 1 coagulation floculation treated with 250 mg/L bentonite and 2 mL of polymer treatment B: control/initial pretreated sample with 250 mg/L bentonite and 2 mL of polymer treatment B: coagulation floculation treated followed by activated carbon for 30 min with 10 g/L AC treatment B: coagulation floculation treated followed by activated carbon for 90 min with 10 g/L AC treatment C: 100 mg/L OSPW extract diluted from 2800 ppm concentrate, treated with ozone and ultrasonication, [ozone] = 2.3 mg/L treatment C: 100 mg/L OSPW extract diluted from 2800 ppm concentrate, treated with ozone and ultrasonication, [ozone] = 2.3 mg/L treatment C: 100 mg/L OSPW extract diluted from 2800 ppm concentrate, treated with ozone and ultrasonication, [ozone] = 3.4 mg/L treatment C: 100 mg/L OSPW extract diluted from 2800 ppm concentrate, treated with ozone and ultrasonication, [ozone] = 3.4 mg/L

a A denotes treatment by coagulation-flocculent with lime and bentonite; B denotes treatment A followed by activated carbon; C denotes treatment by combined ozonation and ultrasonication. The difference in the relative abundance of the sample before treatment A/B and C corresponds to different batches of OSPW. As treatment A was not effective for removal of total NAFCs nor effective for the selective removal of specific species; this treatment was used as the control for sorption to activated carbon.

suspended matter from the OSPW. In contrast, treatment C combines a chemical oxidation step with ultrasonication and is intended for removal of dissolved contaminants. For all of the experiments, samples were withdrawn from the reactors prior to and immediately after treatment to determine the NAFC concentration and distribution of classes of compounds. Treatment A and B experiments were conducted at room temperature and ambient pH of the OSPW. Designated amounts (summarized in Table 1) of coagulants were added to 1 L of OSPW at the beginning of experiments. In the case of lime and bentonite as coagulants for OSPW, the mixing conditions were initial rapid mixing for 5 min at 180 rpm, followed by 10 min of slow mixing at 50 rpm. A high-molecular-weight cationic polyacrylamide-based polymer (1−5 mg/L) was used as a flocculent. The supernatant from treatment A was filtered by passage through a 120 μm mesh and collected for further treatment by activated carbon. For treatment B, adsorption by granular activated carbon was conducted in 250 mL flasks containing 100 mL of OSPW treated by A. The flasks were agitated at 250 rpm and 22 °C for 24 h in an orbital shaker. One flask containing OSPW treated by A but with no activated carbon served as a control. Experimental conditions for treatment C were adapted from earlier work for degradation of dicyclohexylacetic acid.10 In brief, oxidation of OSPW was conducted in a 150 mL jacketed reactor as described by Kumar et al.10 A volume of 125 mL of OSPW was added to the reactor. Ultrapure oxygen was bubbled through an ozone generator (OZV-8S, Ozone Solutions, Inc., Hull, IA) adjusted to pH 8 with 1 M NaOH to produce ozone. Cold water was circulated through the jacket of the reactor to maintain a constant temperature of 22 °C. The ozone concentrations used for the treatments were 2.3−3.4 mg/L. The ultrasonication conditions were as follows: ultrasonic processor (VCX 130, Sonics & Materials, Inc., Newtown, CT) at 130 W output and operating frequency of 20 kHz and diameter of the ultrasonic probe tip of 6 mm, dipped ∼10 mm into the reaction solution. A pulse mode of 59 and 1 s stop at a maximum output power of 130 W was used for the irradiations. Nitrogen was used to purge residual ozone in the solution, and samples were stored at 4 °C prior to analysis. MS Characterization of NAFCs in Treated OSPW. Total NAFCs were measured by negative-ion ESI MS as described

activated carbon, and re-investigation of the combined physical and chemical method, (C) combined ozonation and ultrasonication. The use of physical methods based on sorption of contaminants to activated carbon, coagulants, lime, and bentonite is well-documented in the literature.11−13 Likewise, physical−chemical oxidation of petroleum contaminants by filtration ozonation,14−17 ozonation with ultrasonication,18−20 and photochemical methods21,22 has been reported by several groups.11−22 What is emerging in recent investigations is a preliminary understanding of the effects of such treatments at the molecular level on the degradation of OSPW.19,10,13,22 The studies reported herein contribute further to the understanding of changes occurring at the molecular level for treated OSPW. We report the first application of ultrahigh-resolution negativeion electrospray ionization (ESI) Fourier transform ion cyclotron resonance mass spectrometry (FT-ICR MS) for the characterization of treated OSPW by the above methods. We demonstrate that physical sorption based on coagulation flocculent with lime and bentonite without further treatment with activated charcoal was ineffective compared to the other treatments investigated. For physical treatments that incorporate sorption to activated carbon and treatments based on combination of ozonation with ultrasonification, there were pronounced changes in the double bond equivalent (DBE) distributions for the compound classes and a reduction of total NAFC by 26 and 89%. For treatments B and C, we present evidence for selective removal of S and OxSy heteroatomic species, along with a reduction in DBE values for all species detected.



EXPERIMENTAL SECTION

Athabasca OSPW was exposed to three sets of treatments (A, B, and C). Treatments A and B are physical methods for removal of primarily B

dx.doi.org/10.1021/ef502007b | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

previously by Headley et al.7,23 The calibration is based on actual Athabsca oil sands acid extract generated in-house as oppose to use of a commercial standard, such as Fluka or Acros NA mixtures. The response factors of the multitude of NAFCs present in the mixtures are not known, and clearly ionization energies, particularly those for different heteroatom classes, are not expected to play equal roles. The measurements are thus considered at best semi-quantitative. Thus, emphasis is placed throughout the paper on the relative differences in NAFCs obtained under the same experimental conditions. A Thermo Scientific Linear Trap Quadrupole (LTQ) Orbitrap Velos mass spectrometer was used for determination of NAFC concentrations and preliminary screening of the treated OSPW samples prior to full characterization by FT-ICR MS. The mass resolution of the Orbitrap Velos mass spectrometer was set to 100 000 at m/z 250, and full-scan mass spectra were acquired over a m/z range of 100−600. The ESI interface was set to negative ionization mode. Ionization and full-scan MS conditions were optimized by use of the automated tune function for transmission of ions of m/z 212.075 07. The ion at m/z 212.075 07 is derived from n-butyl benzenesulfonamide, which is present in the background mass spectra and was not derived from a compound added to the Orbitrap mass spectrometer. Parameters for the heated ESI interface were as follows: source heater temperature, 53 °C; spray voltage, 2.86 kV; capillary temperature, 275 °C; sheath gas flow rate, 25 L h−1; auxiliary gas flow rate, 5 L h−1. Ultrahigh-Resolution FT-ICR MS. Samples were characterized by negative-ion FT-ICR MS as previously described by Headley et al.24 In brief, the FT-ICR mass spectrometer was custom-built and equipped with a 9.4 T horizontal 220 mm bore diameter superconducting solenoid magnet.25 For a given sample of 1 mL, a volume of 10 μL of NH4OH (2% in methanol) was added to facilitate deprotonation of the NAFCs.26−28 A syringe pump was used to infuse samples at 400 nL/min to a microelectrospray ion source27,28 (50 μm inner diameter fused silica emitter). The mass resolving power of the FT-ICR MS was ∼950 000 at m/z 400, and full-scan mass spectra was acquired over a m/z range of 185−1000. The ESI interface was set to negative ionization mode. Negative ions were accumulated in an external linear octopole ion trap29 for 0.1−5 s and transferred by radio frequency (rf)-only octopoles to a 10 cm diameter, 30 cm long open cylindrical Penning ion trap. Octopoles were operated at 2.0 MHz and 120 Vp−p rf amplitude. Broadband frequency sweep (chirp) dipolar excitation (70−641 kHz at 50 Hz/μs sweep rate and 0.67 Vp−p amplitude) was followed by direct-mode image current detection to yield 8 Mword time-domain data. Multiple time-domain acquisitions (≥100) were conditionally co-added for each sample Hanning-apodized and zerofilled once before fast Fourier transform and magnitude calculation. The spectra were then internally calibrated with a “walking” calibration based on the most abundant homologous alkylation series.30,31 Instrument control, data acquisition, and data analysis used a modular ICR data station (PREDATOR).32 Formation of dimers was minimized as described earlier27,28 by irradiation with a high-intensity Synrad (Mukilteo, WA) CW CO2 laser (λ = 10.6 μm) operating at 40 W with irradiation of ions for 200 ms. FT-ICR mass spectra were internally calibrated with respect to a high-abundance homologous alkylation series of ions that each contained two oxygen atoms as previously described.27,28 Molecular formulas were assigned to peaks of lowest m/z value for a Kendrick mass defect (KMD) series. Calculations were limited to formulas containing less than 100 12C, 2 13 C, 200 1H, 5 14N, 10 16O, 3 32S, and 1 34S. All formula assignments were 10 for Z between −8 and −12; suggesting the preferential oxidation of higher molecular weight NAs, in agreement with Kumar et al.10 Characterization of treatment C by FT-ICR MS is illustrated in Figure 2. No significant difference in the FT-ICR MS data

Figure 2. FT-ICR MS characterization of the distribution of Ox (x = 1−4), S1, and OxS1 (x = 1−3) classes of OSPW treated by combined ozonation and ultrasonication.

was observed for duplicate conditions. The FT-ICR MS characterization revealed the distribution of Ox (x = 1−5), S1, and OxS1 (x = 1−4) classes. The O2 species, presumably classical NAs were reduced by 89 ± 1.1% by combined ozonation and ultrasonication. This reduction in NAs compares favorably with earlier combined approaches based on pretreatment with petroleum coke followed by ozonation, for which degradation of more than 76% of NAFCs was reported.13 Note that specific heteroatomic sulfur species were completely removed, namely, O2S1, S1, and O3S1, along with O1 species. The combined ozonation and ultrasonication treatment was thus particularly effective for removal of sulfur heteroatoms. However, some NAFC species are recalcitrant and persist after the treatment. The latter include residual levels of O4, and O3 species, along with some O2 species. Clearly, there is a diverse range of structures9 for a given class of components, some of which are less susceptible to treatment by combined ozonation and ultrasonication. Samples shown in Figure 2 were also analyzed with no prior extraction. The unassigned compounds, such as the samples from Figure 1, correspond to salt clusters. Panels a−c of Figure 3 reveal a reduction in DBE values for all Ox and OxS1 species following treatment by combined ozonation and ultrasonication. Degradation thus results in loss of unsaturation and/or loss of ring structures. For example, the O3 class exhibits a bimodal distribution, with DBE values centered at 4 and 8. Following treatment, only traces of the lower DBE values (2−4) are observed, indicating removal of the more unsaturated or aromatic structures. Collectively, the FT-ICR MS results revealed a relatively small difference in the DBE versus C plot of the class O2 and other Ox and OxS1 species before and after treatment in contrast to class O3. The lower DBE O3 components appear to be resistant to combined ozonation and ultrasonication. For the recalcitrant O2 class (11%) observed under negative-ion ESI,

Figure 3. Negative-ion ESI 9.4 T FT-ICR MS DBE versus carbon number for the Ox (x = 2−4) and OxS1 classes for OSPW treated with combined ozonation and ultrasonication: (a) Ox species before treatment, (b) OxS1 species before treatment, and (c) species after treatment. D

dx.doi.org/10.1021/ef502007b | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

Development (PERD), Mitacs, and EnviroWay R&D, Inc. Saskatoon, Saskatchewan, Canada, the National Science Foundation (NSF) Division of Materials Research through DMR-11-57490, and the State of Florida.

the structures therefore cover the same range of DBE values as the non-treated OSPW. There are, however, only a few O2 components left with high O2 abundance in the DBE versus C plot after treatment C, namely, C14O28O2 (DBE = 1) and C18H34O2 (DBE = 2). The high abundance of these specific components may indicate preferential formation of oxidation products, such as tretradecanoic and octadecenenoic acids. Alternatively, these components may be artifacts of the treatment process. In general, the structures of the Ox, and OxS1 components are not known. However, with respect to the O3S1 compounds, we did not observe these compounds in solvent blanks. The O3S1 compounds observed are likely petroleum sulfonates used in the bitumen recovery process as emulsifiers. The sulfonate versus carboxylic acid functionalities in O3S1 and O2S1 classes may be the reason for the recalcitrant nature of the O3S1 compounds. A remaining question for ongoing studies is the degree of toxicity reduction following the respective treatments. Treatment with ozone is known to be effective for toxicity removal.13−17,33 For example, Pereira et al.33 applied reversedphase liquid chromatography with linear ion trap orbitrap mass spectrometry for both positive and negative electrospray ionization of OSPW treated with ozonation alone. Results for a range of heteratomic species were similar to those observed here for ESI FT-ICR MS characterization of treated OSPW with ozonation and ultrasonication. However, O2 species observed by (+)ESI were generally more recalcitrant than the O2 species observed by (−)ESI. The O2+ species are part of the Ox class. Pereira et al.33 suggested that the recalcitrant nature of O2+ and other persistent species observed may help explain the residual toxicity of ozonated OSPW. Because samples were not extracted prior to MS analysis in the present study, the positiveion spectra were not informative and yielded only salt clusters. Activated carbon likewise is expected to be effective for toxicity reduction of treated OSPW. In related work, petroleum coke adsorption was found to be effective in reducing NAFCs and classical NAs by 91 and 84%, respectively.13 The toxicity of the OSPW following treatment with the petroleum coke was reduced from 4.3 to 1.1 toxicity units.13 Sequential treatment of OSPW with petroleum coke followed by ozonation was nontoxic toward Vibrio fischeri.13 Future work is thus warranted to establish if similar reductions in toxicity occur for treatments based on combined ozonation and ultrasonication.





CONCLUSION FT-ICR MS reveals important changes at the molecular level for class distributions and DBE values for various OSPW treatments. The results show that ultrasonication combined with ozonation is an effective approach for 89% reduction of NAFCs from OSPWs.



REFERENCES

(1) Burrowes, A.; Marsh, R.; Evans, C.; Teare, M.; Ramos, S.; Rahnama, F.; Kirsch, M.-A.; Philp, L.; Stenson, J.; Yemane, M.; Horne, J. V.; Fong, J.; Sankey, G.; Harrison, P. Alberta’s Energy Reserves 2008 and Supply/Demand Outlook 2009−2018; Energy Resources Conservation Board, Government of Alberta: Calgary, Alberta, Canada, 2009; p 220. (2) Schramm, L. L.; Stasiuk, E. N.; MacKinnon, M. Surfactants in Athabasca oil sands extraction and tailing process. In Surfactants: Fundamentals and Applications in the Petroleum Industry; Schramm, L. L., Ed.; Cambridge University Press: Cambridge, U.K., 2000; pp 365− 430. (3) Rogers, V. V.; Wickstrom, M.; Liber, K.; MacKinnon, M. D. Acute and subchronic mammalian toxicity of naphthenic acids from oil sands tailings. Toxicol. Sci. 2002, 66 (2), 347−355. (4) Han, X.; MacKinnon, M. D.; Martin, J. W. Estimating the in situ biodegradation of naphthenic acids in oil sands process waters by HPLC/HRMS. Chemosphere 2009, 76 (1), 63−70. (5) Clemente, J. S.; Fedorak, P. M. A review of the occurrence, analyses, toxicity and biodegradation of naphthenic acids. Chemosphere 2005, 60 (5), 585−600. (6) Headley, J. V.; McMartin, D. W. A review of the occurrence and fate of naphthenic acids in aquatic environments. J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2004, 39 (8), 1989−2010. (7) Headley, J. V.; Peru, K. M.; Mohamed, M. H.; Frank, R. A.; Martin, J. W.; Hazewinkel, R. R.O.; Humphries, D.; Gurprasad, N. P.; Hewitt, L. M.; Muir, D. C. G.; Lindeman, D.; Strub, R.; Young, R. F.; Grewer, D. M.; Whittal, R. M.; Fedorak, P. M.; Birkholz, D. A.; Hindle, R.; Reisdorph, R.; Wang, X.; Kasperski, K. L.; Hamilton, C.; Woudneh, M.; Wang, G.; Loescher, B.; Farwell, A.; Dixon, D. G.; Ross, M.; Dos Santos Pereira, A.; King, E.; Barrow, M. P.; Fahlman, B.; Bailey, J.; McMartin, D. W.; Borchers, C. H.; Ryan, C. H.; Toor, N. S.; Gillis, H. M.; Zuin, L.; Bickerton, G.; McMaster, M.; Sverko, E.; Shang, D.; Wilson, L. D.; Wrona, F. J. Chemical fingerprinting of naphthenic acids and oil sands process watersA review of analytical methods for environmental samples. J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2013, 48 (10), 1145−1163. (8) Headley, J. V.; Peru, K. M.; Mohamed, M. H.; Wilson, L.; McMartin, D. W.; Mapolelo, M. M.; Lobodin, V. V.; Rodgers, R. P.; Marshall, A. G. Atmospheric pressure photoionization Fourier transform ion cyclotron resonance mass spectrometry characterization of tunable carbohydrate-based materials for sorption of oil sands naphthenic acids. Energy Fuels 2014, 28 (3), 1611−1616. (9) Han, X.; Scott, A. C.; Fedorak, P. M.; Bataineh, M.; Martin, J. W. Influence of molecular structure on the biodegradability of naphthenic acids. Environ. Sci. Technol. 2008, 42 (4), 1290−1295. (10) Kumar, P.; Headley, J. V.; Peru, K. M.; Bailey, J.; Dalai, A. Removal of dicyclohexyl acetic acid from aqueous solution using ultrasound, ozone and their combination. J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2014, 49 (13), 1512−1519. (11) Janfada, A.; Headley, J. V.; Peru, K. M.; Barbour, S. L. A laboratory evaluation of the sorption of oil sands naphthenic acids on organic rich soils. J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2006, 41 (6), 985−997. (12) Pourrezaei, P.; Drzewicz, P.; Wang, Y.; Gamal El-Din, M.; PerezEstrada, L. A.; Martin, J. W.; Anderson, J.; Giesy, J. P. The impact of metallic coagulants on the removal of organic compounds from oil sands process-affected water. Environ. Sci. Technol. 2011, 45 (19), 8452−8459. (13) Gamal El-Din, M.; Fu, H.; Wang, N.; Chelme-Ayala, P.; PérezEstrada, L.; Drzewicz, P.; Martin, J. W.; Smith, D. W. Naphthenic acids speciation and removal during petroleum-coke adsorption and

AUTHOR INFORMATION

Corresponding Author

*Telephone: +1-850-644-2398. Fax: +1-306-975-5143. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The Natural Science and Engineering Research Council of Canada is gratefully acknowledged for financial support of this research, along with the Program of Energy and Research E

dx.doi.org/10.1021/ef502007b | Energy Fuels XXXX, XXX, XXX−XXX

Energy & Fuels

Article

ozonation of oil sands process-affected water. Sci. Total Environ. 2011, 409 (23), 5119−5125. (14) He, Y.; Wiseman, S. B.; Zhang, X. W.; Hecker, M.; Jones, P. D.; El-Din, M. G.; Martin, J. W.; Giesy, J. P. Ozonation attenuates the steroidogenic disruptive effects of sediment free oil sands process water in the H295 cell line. Chemosphere 2010, 80 (5), 578−584. (15) Anderson, J. C.; Wiseman, S. B.; Wang, N.; Moustafa, A.; PerezEstrada, L.; Gamal El-Din, M.; Martin, J. W.; Liber, K.; Giesy, J. P. Effectiveness of ozonation treatment in eliminating toxicity of oil sands process-affected water to Chironomus dilutus. Environ. Sci. Technol. 2012, 46 (1), 486−493. (16) Martin, J. W.; Barri, T.; Han, X.; Fedorak, P. M.; Gamal El-Din, M.; Perez-Estrada, L.; Scott, A.; Jiang, J. T. Ozonation of oil sands process-affected water accelerates microbial bioremediation. Environ. Sci. Technol. 2010, 44 (21), 8350−8356. (17) Hwang, G.; Dong, T.; Islam, M. S.; Sheng, Z.; Perez-Estrada, L. A.; Liu, Y.; Gamal El-Din, M. The impacts of ozonation on oil sands process-affected water biodegradability and biofilm formation characteristics in bioreactors. Bioresour. Technol. 2013, 130, 269−277. (18) Cui, M.; Jang, M.; Cho, S.-H.; Elena, D. K. J. Enhancement of mineralization of a number of natural refractory organic compounds by the combined process of sonolysis and ozonolysis (US/O3). Ultrason. Sonochem. 2011, 18 (3), 773−780. (19) Abramov, V. O.; Abramov, O. V.; Gekhman, A. E.; Kuznetsov, V. M.; Price, G. J. Ultrasonic intensification of ozone and electrochemical destruction of 1,3-dinitrobenzene and 2,4-dinitrotoluene. Ultrason. Sonochem. 2006, 13 (4), 303−307. (20) He, Z.; Zhu, R.; Xu, X.; Song, S.; Chen, J. Ozonation combined with sonolysis for degradation and detoxification of m-nitrotoluene in aqueous solution. Ind. Eng. Chem. Res. 2009, 48 (12), 5578−5583. (21) Griffiths, M. T.; Da Campo, R.; O’Connor, P. B.; Barrow, M. P. Throwing light on petroleum: Simulated exposure of crude oil to sunlight and characterization using atmospheric pressure photoionization Fourier transform ion cyclotron resonance mass spectrometry. Anal. Chem. 2014, 86 (1), 527−534. (22) Headley, J. V.; Du, J.; Peru, K. M.; McMartin, D. W. Electrospray ionization mass spectrometry of the photodegradation of naphthenic acids mixtures irradiated with titanium dioxide. J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2009, 44 (6), 591−597. (23) Mohamed, M. H.; Wilson, L. D.; Headley, J. V.; Peru, K. M. Sequestration of naphthenic acids from aqueous solution using βcyclodextrin-based polyurethanes. Phys. Chem. Chem. Phys. 2011, 13 (3), 1112−1122. (24) Headley, J. V.; Peru, K. M.; Mohamed, M. H.; Wilson, L.; McMartin, D. W.; Mapolelo, M. M.; Lobodin, V. V.; Rodgers, R. P.; Marshall, A. G. Electrospray ionization FT-ICR MS characterization of tunable carbohydrate-based materials for sorption of oil sands naphthenic acids. Energy Fuels 2013, 27 (4), 1772−1778. (25) Kaiser, N.; Quinn, J.; Blakney, G.; Hendrickson, C.; Marshall, A. A novel 9.4 T FTICR mass spectrometer with improved sensitivity, mass resolution, and mass range. J. Am. Soc. Mass Spectrom. 2011, 22 (8), 1343−1351. (26) Headley, J. V.; Peru, K. M.; Armstrong, S. A.; Han, X.; Martin, J. W.; Mapolelo, M. M.; Smith, D. F.; Rogers, R. P.; Marshall, A. G. Aquatic plant-derived changes in oil sands naphthenic acid signatures determined by low-, high- and ultrahigh-resolution mass spectrometry. Rapid Commun. Mass Spectrom. 2009, 23 (4), 515−522. (27) Headley, J. V.; Peru, K. M.; Mishra, S.; Meda, V.; Dalai, A. K.; McMartin, D. W.; Mapolelo, M. M.; Rodgers, R. P.; Marshall, A. G. Characterization of oil sands naphthenic acids treated with ultraviolet and microwave radiation by negative ion electrospray Fourier transform ion cyclotron resonance mass spectrometry. Rapid Commun. Mass Spectrom. 2010, 24 (21), 3121−3126. (28) Headley, J. V.; Armstrong, S. A.; Peru, K. M.; Mikula, R. J.; Germida, J. J.; Mapolelo, M. M.; Rodgers, R. P.; Marshall, A. G. Ultrahigh-resolution mass spectrometry of simulated runoff from treated oil sands mature fine tailings. Rapid Commun. Mass Spectrom. 2010, 24 (16), 2400−2406.

(29) Senko, M. W.; Hendrickson, C. L.; Emmett, M. R.; Shi, S. D. H.; Marshall, A. G. External accumulation of ions for enhanced electrospray ionization Fourier transform ion cyclotron resonance mass spectrometry. J. Am. Soc. Mass Spectrom. 1997, 8 (9), 970−976. (30) Savory, J. J.; Kaiser, N. K.; McKenna, A. M.; Xian, F.; Blakney, G. T.; Rodgers, R. P.; Hendrickson, C. L.; Marshall, A. G. Parts-perbillion Fourier transform ion cyclotron resonance mass measurement accuracy with a “walking” calibration equation. Anal. Chem. 2011, 83 (5), 1732−1736. (31) Shi, S. D. H.; Drader, J. J.; Freitas, M. A.; Hendrickson, C. L.; Marshall, A. G. Comparison and interconversion of the two most common frequency-to-mass calibration functions for Fourier transform ion cyclotron resonance mass spectrometry. Int. J. Mass Spectrom. 2000, 195−196, 591−598. (32) Blakney, G. T.; Hendrickson, C. L.; Marshall, A. G. Predator data station: A fast data acquisition system for advanced FT-ICR MS experiments. Int. J. Mass Spectrom. 2011, 306 (2−3), 246−252. (33) Pereira, A. S.; Islam, M. S.; Gamal El-Din, M.; Martin, J. W. Ozonation degrades all detectable organic compound classes in oil sands process-affected water; An application of high-performance liquid chromatography/orbitrap mass spectrometry. Rapid Commun. Mass Spectrom. 2013, 27 (21), 2317−2326.

F

dx.doi.org/10.1021/ef502007b | Energy Fuels XXXX, XXX, XXX−XXX