Fourier Transform Ion Cyclotron Resonance Mass Spectrometry for the Analysis of Small Ubiquitin-like...

6
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 occulent with lime and bentonite, (B) coagulation occulent with lime and bentonite followed by activated carbon, and (C) combined ozonation and ultrasonication. Treatment A was ineective in reducing the level of total naphthenic acid fraction components [NAFCs, dened as the acid-extractable fraction of OSPWs or crude oils (C n H 2n + z O w S x N y ), 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 deciency 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 O x S y heteroatomic species at the molecular level, along with a reduction in the DBE values for all species. INTRODUCTION The Athabasca oil sands deposits in Alberta, Canada, are one of the worlds 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. 24 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. 24 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 specically 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 20120 mg/L. 5,6 NAFCs (C n H 2n + z O w S x N y ) are dened 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 (C n H 2n + z O 2 ) 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 deciency because of the 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 Biodegrada- tion 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 occulent with lime and bentonite, along with (B) coagulation occulent with lime and bentonite followed by Special Issue: 15th International Conference on Petroleum Phase Behavior and Fouling Received: September 8, 2014 Revised: November 7, 2014 Published: November 9, 2014 Article pubs.acs.org/EF © 2014 American Chemical Society 2768 dx.doi.org/10.1021/ef502007b | Energy Fuels 2015, 29, 27682773

Transcript of Fourier Transform Ion Cyclotron Resonance Mass Spectrometry for the Analysis of Small Ubiquitin-like...

Fourier Transform Ion Cyclotron Resonance Mass SpectrometryCharacterization of Treated Athabasca Oil Sands Processed WatersJohn 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 massspectrometry (FT-ICR MS) was used to characterize Athabasca oil sands processed water (OSPW) treated by (A) coagulationflocculent 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 fractioncomponents [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 atomdeficiency because of the presence of double bonds and ring formation]. Likewise, for treatment A, little or no change wasobserved for the double bond equivalent (DBE) distributions of the compound classes. Treatments B and C resulted in thereduction of total NAFCs by 26 ± 1.4 and 89 ± 1.1%, respectively. For the latter treatments, there was evidence for selectiveremoval of the S and OxSy heteroatomic species at the molecular level, along with a reduction in the DBE values for all species.

■ INTRODUCTION

The Athabasca oil sands deposits in Alberta, Canada, are one ofthe world’s largest reserves of petroleum.1 The ever increasingenergy demands and dwindling conventional oil reserves havedrawn more attention to alternative energy resources. The oilsands crude production is projected to grow from 2.1 to 3.7million barrels per day by 2020. Bitumen is extracted from oilsands by use of hot alkaline water or in situ methods, such assteam-assisted gravity drainage.2−4 The extraction of thebitumen from surface mining results in vast amounts of oilsands processed water (OSPW) or produced water. Thecurrent practice of storing the OSPW in tailing ponds is notsustainable, and much research is geared at developingreclamation strategies.2−4 Polar organic compounds, such asnaphthenic acids, are suspected to be among the principal toxiccomponents, and thus, much attention is given to treatmentmethods that specifically target the acid fraction of OSPW.3

The naphthenic acid (NA) concentrations [or more correctlythe naphthenic acid fraction components (NAFCs)] in OSPWmay be as high as 20−120 mg/L.5,6

NAFCs (CnH2n + zOwSxNy) are defined as the acid-extractablefraction of OSPWs, and the term will also be used here forbrevity to extend to crude oils.7,8 This fraction was traditionallycalled NAs (CnH2n + zO2) but is now known to include a widerange of organic acids containing N and S atoms, along withother components with various levels of unsaturation andaromaticity.7,8 Here, n, w, x, and y indicate the number ofcarbon, oxygen, sulfur, and nitrogen atoms, respectively, and zrepresents the hydrogen atom deficiency because of the

presence of double bonds and ring formation.7 The morecomplex the structure of the NAs or NAFCs, such as increasingthe number of rings, the nature of the alkyl chains, or thenumber and type of heteroatoms, generally the more persistentthe NAFCs to degradation in the environment.4,7 Biodegrada-tion of some NAFCs can be relatively slow, with half-lives of∼14 years.4,9 Thus, other treatments apart from naturalbiodegradation are required for reclamation of OSPW withinpractical time periods.4

In earlier work, Kumar et al.10 reported the treatment ofOSPW with combined physical and chemical methods. Theearlier studies were limited to monitoring the reduction ofchemical oxygen demand (COD) and analyses of total NAs.Changes occurring at the molecular level for the treated OSPWwere restricted to the distribution of classical structures of NAs.Because the earlier work provided limited insight into changesoccurring at the molecular level for heteroatomic structures ofNAFCs, we have conducted follow-up studies to characterizethe treated OSPW by ultrahigh-resolution mass spectrometry(MS). The treatments investigated here include two physicalmethods 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

Special Issue: 15th International Conference on Petroleum PhaseBehavior and Fouling

Received: September 8, 2014Revised: November 7, 2014Published: November 9, 2014

Article

pubs.acs.org/EF

© 2014 American Chemical Society 2768 dx.doi.org/10.1021/ef502007b | Energy Fuels 2015, 29, 2768−2773

activated carbon, and re-investigation of the combined physicaland chemical method, (C) combined ozonation and ultra-sonication.The use of physical methods based on sorption of

contaminants to activated carbon, coagulants, lime, andbentonite is well-documented in the literature.11−13 Likewise,physical−chemical oxidation of petroleum contaminants byfiltration ozonation,14−17 ozonation with ultrasonication,18−20

and photochemical methods21,22 has been reported by severalgroups.11−22 What is emerging in recent investigations is apreliminary understanding of the effects of such treatments atthe molecular level on the degradation of OSPW.19,10,13,22 Thestudies reported herein contribute further to the understandingof changes occurring at the molecular level for treated OSPW.We report the first application of ultrahigh-resolution negative-ion electrospray ionization (ESI) Fourier transform ioncyclotron resonance mass spectrometry (FT-ICR MS) for thecharacterization of treated OSPW by the above methods. Wedemonstrate that physical sorption based on coagulationflocculent with lime and bentonite without further treatmentwith activated charcoal was ineffective compared to the othertreatments investigated. For physical treatments that incorpo-rate sorption to activated carbon and treatments based oncombination of ozonation with ultrasonification, there werepronounced changes in the double bond equivalent (DBE)distributions for the compound classes and a reduction of totalNAFC by 26 and 89%. For treatments B and C, we presentevidence for selective removal of S and OxSy heteroatomicspecies, along with a reduction in DBE values for all speciesdetected.

■ EXPERIMENTAL SECTIONAthabasca OSPW was exposed to three sets of treatments (A, B, andC). Treatments A and B are physical methods for removal of primarily

suspended matter from the OSPW. In contrast, treatment C combinesa chemical oxidation step with ultrasonication and is intended forremoval of dissolved contaminants. For all of the experiments, sampleswere withdrawn from the reactors prior to and immediately aftertreatment to determine the NAFC concentration and distribution ofclasses of compounds.

Treatment A and B experiments were conducted at roomtemperature and ambient pH of the OSPW. Designated amounts(summarized in Table 1) of coagulants were added to 1 L of OSPW atthe beginning of experiments. In the case of lime and bentonite ascoagulants for OSPW, the mixing conditions were initial rapid mixingfor 5 min at 180 rpm, followed by 10 min of slow mixing at 50 rpm. Ahigh-molecular-weight cationic polyacrylamide-based polymer (1−5mg/L) was used as a flocculent. The supernatant from treatment Awas filtered by passage through a 120 μm mesh and collected forfurther treatment by activated carbon. For treatment B, adsorption bygranular activated carbon was conducted in 250 mL flasks containing100 mL of OSPW treated by A. The flasks were agitated at 250 rpmand 22 °C for 24 h in an orbital shaker. One flask containing OSPWtreated by A but with no activated carbon served as a control.

Experimental conditions for treatment C were adapted from earlierwork for degradation of dicyclohexylacetic acid.10 In brief, oxidation ofOSPW was conducted in a 150 mL jacketed reactor as described byKumar et al.10 A volume of 125 mL of OSPW was added to thereactor. Ultrapure oxygen was bubbled through an ozone generator(OZV-8S, Ozone Solutions, Inc., Hull, IA) adjusted to pH 8 with 1 MNaOH to produce ozone. Cold water was circulated through the jacketof the reactor to maintain a constant temperature of 22 °C. The ozoneconcentrations used for the treatments were 2.3−3.4 mg/L. Theultrasonication conditions were as follows: ultrasonic processor (VCX130, Sonics & Materials, Inc., Newtown, CT) at 130 W output andoperating frequency of 20 kHz and diameter of the ultrasonic probe tipof 6 mm, dipped ∼10 mm into the reaction solution. A pulse mode of59 and 1 s stop at a maximum output power of 130 W was used for theirradiations. 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. TotalNAFCs were measured by negative-ion ESI MS as described

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

description

NAFCconcentration

(mg/L) additional information

treatments A and B initial OSPW 32.5 initial OSPW controltreatment A OSPW exposed toflocculent after filtration

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

treatment A initial OSPW 28.6 treatment A: T3-initialtreatment A OSPW exposed toflocculent after filtration

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

experiment 1 CF initial 36.3 treatment A: experiment 1 coagulation floculation initial (samples from an OSPW tailing pond)experiment 1 AT 34.9 treatment A: experiment 1 coagulation floculation treated with 250 mg/L bentonite and 2 mL of

polymer0 C, 15 min 23.1 treatment B: control/initial pretreated sample with 250 mg/L bentonite and 2 mL of polymer1 G C, 30 min 18.2 treatment B: coagulation floculation treated followed by activated carbon for 30 min with

10 g/L AC1 G C, 90 min 16.8 treatment B: coagulation floculation treated followed by activated carbon for 90 min with

10 g/L ACexperiment 3 initial before 71.5 treatment C: 100 mg/L OSPW extract diluted from 2800 ppm concentrate, treated with ozone and

ultrasonication, [ozone] = 2.3 mg/Lexperiment 3 1 h 10.7 treatment C: 100 mg/L OSPW extract diluted from 2800 ppm concentrate, treated with ozone and

ultrasonication, [ozone] = 2.3 mg/Lexperiment 4 initial before 87.7 treatment C: 100 mg/L OSPW extract diluted from 2800 ppm concentrate, treated with ozone and

ultrasonication, [ozone] = 3.4 mg/Lexperiment 4 1 h 11.3 treatment C: 100 mg/L OSPW extract diluted from 2800 ppm concentrate, treated with ozone and

ultrasonication, [ozone] = 3.4 mg/LaA denotes treatment by coagulation-flocculent with lime and bentonite; B denotes treatment A followed by activated carbon; C denotes treatmentby combined ozonation and ultrasonication. The difference in the relative abundance of the sample before treatment A/B and C corresponds todifferent 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.

Energy & Fuels Article

dx.doi.org/10.1021/ef502007b | Energy Fuels 2015, 29, 2768−27732769

previously by Headley et al.7,23 The calibration is based on actualAthabsca oil sands acid extract generated in-house as oppose to use ofa commercial standard, such as Fluka or Acros NA mixtures. Theresponse factors of the multitude of NAFCs present in the mixturesare not known, and clearly ionization energies, particularly those fordifferent heteroatom classes, are not expected to play equal roles. Themeasurements are thus considered at best semi-quantitative. Thus,emphasis is placed throughout the paper on the relative differences inNAFCs obtained under the same experimental conditions. A ThermoScientific Linear Trap Quadrupole (LTQ) Orbitrap Velos massspectrometer was used for determination of NAFC concentrations andpreliminary screening of the treated OSPW samples prior to fullcharacterization by FT-ICR MS. The mass resolution of the OrbitrapVelos mass spectrometer was set to 100 000 at m/z 250, and full-scanmass spectra were acquired over a m/z range of 100−600. The ESIinterface was set to negative ionization mode. Ionization and full-scanMS conditions were optimized by use of the automated tune functionfor transmission of ions of m/z 212.075 07. The ion at m/z 212.075 07is derived from n-butyl benzenesulfonamide, which is present in thebackground mass spectra and was not derived from a compound addedto the Orbitrap mass spectrometer. Parameters for the heated ESIinterface were as follows: source heater temperature, 53 °C; sprayvoltage, 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 Inbrief, the FT-ICR mass spectrometer was custom-built and equippedwith a 9.4 T horizontal 220 mm bore diameter superconductingsolenoid magnet.25 For a given sample of 1 mL, a volume of 10 μL ofNH4OH (2% in methanol) was added to facilitate deprotonation ofthe NAFCs.26−28

A syringe pump was used to infuse samples at 400 nL/min to amicroelectrospray ion source27,28 (50 μm inner diameter fused silicaemitter). The mass resolving power of the FT-ICR MS was ∼950 000at m/z 400, and full-scan mass spectra was acquired over a m/z rangeof 185−1000. The ESI interface was set to negative ionization mode.Negative ions were accumulated in an external linear octopole iontrap29 for 0.1−5 s and transferred by radio frequency (rf)-onlyoctopoles to a 10 cm diameter, 30 cm long open cylindrical Penningion trap. Octopoles were operated at 2.0 MHz and 120 Vp−p rfamplitude. Broadband frequency sweep (chirp) dipolar excitation(70−641 kHz at 50 Hz/μs sweep rate and 0.67 Vp−p amplitude) wasfollowed by direct-mode image current detection to yield 8 Mwordtime-domain data. Multiple time-domain acquisitions (≥100) wereconditionally co-added for each sample Hanning-apodized and zero-filled once before fast Fourier transform and magnitude calculation.The spectra were then internally calibrated with a “walking” calibrationbased on the most abundant homologous alkylation series.30,31

Instrument control, data acquisition, and data analysis used a modularICR data station (PREDATOR).32 Formation of dimers wasminimized as described earlier27,28 by irradiation with a high-intensitySynrad (Mukilteo, WA) CW CO2 laser (λ = 10.6 μm) operating at 40W with irradiation of ions for 200 ms. FT-ICR mass spectra wereinternally calibrated with respect to a high-abundance homologousalkylation series of ions that each contained two oxygen atoms aspreviously described.27,28 Molecular formulas were assigned to peaksof lowest m/z value for a Kendrick mass defect (KMD) series.Calculations were limited to formulas containing less than 100 12C, 213C, 200 1H, 5 14N, 10 16O, 3 32S, and 1 34S. All formula assignmentswere <1 ppm mass error.

■ RESULTS AND DISCUSSION

As shown in Table 1, treatment A was not effective for removalof total NAFCs.7

Likewise, this treatment was not effective for the selectiveremoval of specific species. The total NAFC concentration wasunchanged within experimental error for duplicate analyses(32.5 and 30.1 mg/L and 28.6 and 26.3 mg/L). Additional

treatment with activated carbon, however, resulted in areduction of NAFCs by 26%, as discussed below.The distribution of NOx species for the present OSPW

treatments is a topic of ongoing investigations and is deferredto a future paper. Discussion will thus focus on NAFCs with thegeneral formula CnH2n + zOxSy. As illustrated in Figure 1,

coagulation flocculent with lime and bentonite followed byactivated carbon was effective in the partial removal of specificNAFCs. All heteroatom classes showed a decrease in relativeabundance at both the 30 and 90 min time points, with acomplete removal of the O2S1 and O4S1 classes at 90 min.Likewise, there was a time-dependent removal of the most

prevalent Ox species, in which O2, O3, and O4 species werereduced by 54, 44, and 27% at 30 min and then by 80, 76, and61% after 90 min of treatment. Note that the samples analyzedhere were not subjected to extraction procedures (i.e., liquid−liquid or solid-phase extractions) prior to mass spectral analysis.The low summed relative abundance for the three samplesshown in Figure 1 (∼2−10% total relative abundance forassigned compounds) was due to the presence of salt adducts inthe mass spectrum.Kumar et al.10 reported a reduction in COD values of 65 ±

1.2 and 78 ± 1.4%, for oxidation of OSPW by ozonation aloneand combined ozonation and ultrasonication. The combinationof ozonation and ultrasonication was thus anticipated to beeffective for possible treatment of OSPW.18−20 Ozone mayoxidize the NAFCs directly or through the generation ofhydroxyl radicals.10 Ultrasonication is thought to generatelocalized regions in which thermolytic dissociation of watermolecules occurs by acoustic cavitation to produce hydroxylradicals.10 Degradation of NAFCs will thus be a function of theozone concentration and mass transfer rate of ozone insolution, along with the ultrasonic irradiation intensity andfrequency.10,19

The concentration of ozone was essentially the same for thetwo sets of conditions investigated. The Orbitrap MSprescreening (data not shown) confirmed that the profiles forthe treated and non-treated OSPWs were similar, with carbon

Figure 1. FT-ICR MS characterization of the distribution of Ox (x =1−5), S1, and OxS1 (x = 1−4) classes of OSPW treated by coagulationflocculent with lime and bentonite followed by activated carbon. Onlyclasses with a relative abundance of ≥0.5% prior to treatment wereplotted. Relative abundance of classes O1, O1S1, and S1 was lower than0.5%.

Energy & Fuels Article

dx.doi.org/10.1021/ef502007b | Energy Fuels 2015, 29, 2768−27732770

number (n) ranging from 5 to 20 and Z from −2 to −12. Theprescreening revealed removal of NAs with n > 10 for Zbetween −8 and −12; suggesting the preferential oxidation ofhigher molecular weight NAs, in agreement with Kumar et al.10

Characterization of treatment C by FT-ICR MS is illustratedin Figure 2. No significant difference in the FT-ICR MS data

was observed for duplicate conditions. The FT-ICR MScharacterization revealed the distribution of Ox (x = 1−5), S1,and OxS1 (x = 1−4) classes. The O2 species, presumablyclassical NAs were reduced by 89 ± 1.1% by combinedozonation and ultrasonication. This reduction in NAs comparesfavorably with earlier combined approaches based on pretreat-ment with petroleum coke followed by ozonation, for whichdegradation of more than 76% of NAFCs was reported.13

Note that specific heteroatomic sulfur species werecompletely removed, namely, O2S1, S1, and O3S1, along withO1 species. The combined ozonation and ultrasonicationtreatment was thus particularly effective for removal of sulfurheteroatoms. However, some NAFC species are recalcitrantand persist after the treatment. The latter include residual levelsof O4, and O3 species, along with some O2 species. Clearly,there is a diverse range of structures9 for a given class ofcomponents, some of which are less susceptible to treatment bycombined ozonation and ultrasonication. Samples shown inFigure 2 were also analyzed with no prior extraction. Theunassigned 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 combinedozonation and ultrasonication. Degradation thus results in lossof unsaturation and/or loss of ring structures. For example, theO3 class exhibits a bimodal distribution, with DBE valuescentered at 4 and 8. Following treatment, only traces of thelower DBE values (2−4) are observed, indicating removal ofthe 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 andother Ox and OxS1 species before and after treatment incontrast to class O3. The lower DBE O3 components appear tobe resistant to combined ozonation and ultrasonication. For therecalcitrant O2 class (11%) observed under negative-ion ESI,

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 combinedozonation and ultrasonication.

Figure 3. Negative-ion ESI 9.4 T FT-ICR MS DBE versus carbonnumber for the Ox (x = 2−4) and OxS1 classes for OSPW treated withcombined ozonation and ultrasonication: (a) Ox species beforetreatment, (b) OxS1 species before treatment, and (c) species aftertreatment.

Energy & Fuels Article

dx.doi.org/10.1021/ef502007b | Energy Fuels 2015, 29, 2768−27732771

the structures therefore cover the same range of DBE values asthe non-treated OSPW. There are, however, only a few O2components left with high O2 abundance in the DBE versus Cplot after treatment C, namely, C14O28O2 (DBE = 1) andC18H34O2 (DBE = 2). The high abundance of these specificcomponents may indicate preferential formation of oxidationproducts, such as tretradecanoic and octadecenenoic acids.Alternatively, these components may be artifacts of thetreatment process. In general, the structures of the Ox, andOxS1 components are not known. However, with respect to theO3S1 compounds, we did not observe these compounds insolvent blanks. The O3S1 compounds observed are likelypetroleum sulfonates used in the bitumen recovery process asemulsifiers. The sulfonate versus carboxylic acid functionalitiesin O3S1 and O2S1 classes may be the reason for the recalcitrantnature of the O3S1 compounds.A remaining question for ongoing studies is the degree of

toxicity reduction following the respective treatments. Treat-ment with ozone is known to be effective for toxicityremoval.13−17,33 For example, Pereira et al.33 applied reversed-phase liquid chromatography with linear ion trap orbitrap massspectrometry for both positive and negative electrosprayionization of OSPW treated with ozonation alone. Results fora range of heteratomic species were similar to those observedhere for ESI FT-ICR MS characterization of treated OSPWwith ozonation and ultrasonication. However, O2 speciesobserved by (+)ESI were generally more recalcitrant than theO2 species observed by (−)ESI. The O2+ species are part of theOx class. Pereira et al.

33 suggested that the recalcitrant nature ofO2+ and other persistent species observed may help explain theresidual toxicity of ozonated OSPW. Because samples were notextracted prior to MS analysis in the present study, the positive-ion spectra were not informative and yielded only salt clusters.Activated carbon likewise is expected to be effective for toxicityreduction of treated OSPW. In related work, petroleum cokeadsorption was found to be effective in reducing NAFCs andclassical NAs by 91 and 84%, respectively.13 The toxicity of theOSPW following treatment with the petroleum coke wasreduced from 4.3 to 1.1 toxicity units.13 Sequential treatment ofOSPW with petroleum coke followed by ozonation was non-toxic toward Vibrio fischeri.13 Future work is thus warranted toestablish if similar reductions in toxicity occur for treatmentsbased on combined ozonation and ultrasonication.

■ CONCLUSIONFT-ICR MS reveals important changes at the molecular levelfor class distributions and DBE values for various OSPWtreatments. The results show that ultrasonication combinedwith ozonation is an effective approach for 89% reduction ofNAFCs from OSPWs.

■ AUTHOR INFORMATIONCorresponding Author*Telephone: +1-850-644-2398. Fax: +1-306-975-5143. E-mail:[email protected] authors declare no competing financial interest.

■ ACKNOWLEDGMENTSThe Natural Science and Engineering Research Council ofCanada is gratefully acknowledged for financial support of thisresearch, along with the Program of Energy and Research

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

■ 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 2008and Supply/Demand Outlook 2009−2018; Energy Resources Con-servation Board, Government of Alberta: Calgary, Alberta, Canada,2009; p 220.(2) Schramm, L. L.; Stasiuk, E. N.; MacKinnon, M. Surfactants inAthabasca 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 oilsands tailings. Toxicol. Sci. 2002, 66 (2), 347−355.(4) Han, X.; MacKinnon, M. D.; Martin, J. W. Estimating the in situbiodegradation of naphthenic acids in oil sands process waters byHPLC/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. Chemosphere2005, 60 (5), 585−600.(6) Headley, J. V.; McMartin, D. W. A review of the occurrence andfate 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.; DosSantos 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 acidsand oil sands process watersA review of analytical methods forenvironmental 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 Fouriertransform ion cyclotron resonance mass spectrometry characterizationof tunable carbohydrate-based materials for sorption of oil sandsnaphthenic 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 naphthenicacids. 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 usingultrasound, 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. Alaboratory evaluation of the sorption of oil sands naphthenic acids onorganic 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.; Perez-Estrada, L. A.; Martin, J. W.; Anderson, J.; Giesy, J. P. The impact ofmetallic coagulants on the removal of organic compounds from oilsands process-affected water. Environ. Sci. Technol. 2011, 45 (19),8452−8459.(13) Gamal El-Din, M.; Fu, H.; Wang, N.; Chelme-Ayala, P.; Perez-Estrada, L.; Drzewicz, P.; Martin, J. W.; Smith, D. W. Naphthenic acidsspeciation and removal during petroleum-coke adsorption and

Energy & Fuels Article

dx.doi.org/10.1021/ef502007b | Energy Fuels 2015, 29, 2768−27732772

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 thesteroidogenic disruptive effects of sediment free oil sands processwater in the H295 cell line. Chemosphere 2010, 80 (5), 578−584.(15) Anderson, J. C.; Wiseman, S. B.; Wang, N.; Moustafa, A.; Perez-Estrada, L.; Gamal El-Din, M.; Martin, J. W.; Liber, K.; Giesy, J. P.Effectiveness of ozonation treatment in eliminating toxicity of oil sandsprocess-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 sandsprocess-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 sandsprocess-affected water biodegradability and biofilm formationcharacteristics in bioreactors. Bioresour. Technol. 2013, 130, 269−277.(18) Cui, M.; Jang, M.; Cho, S.-H.; Elena, D. K. J. Enhancement ofmineralization of a number of natural refractory organic compoundsby 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 andelectrochemical destruction of 1,3-dinitrobenzene and 2,4-dinitroto-luene. Ultrason. Sonochem. 2006, 13 (4), 303−307.(20) He, Z.; Zhu, R.; Xu, X.; Song, S.; Chen, J. Ozonation combinedwith sonolysis for degradation and detoxification of m-nitrotoluene inaqueous 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 tosunlight and characterization using atmospheric pressure photo-ionization Fourier transform ion cyclotron resonance mass spectrom-etry. 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 photodegradationof 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 β-cyclodextrin-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 oftunable carbohydrate-based materials for sorption of oil sandsnaphthenic 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 signaturesdetermined 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 ultravioletand microwave radiation by negative ion electrospray Fouriertransform 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 fromtreated 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 enhancedelectrospray ionization Fourier transform ion cyclotron resonancemass 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-per-billion Fourier transform ion cyclotron resonance mass measurementaccuracy 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 mostcommon frequency-to-mass calibration functions for Fourier transformion cyclotron resonance mass spectrometry. Int. J. Mass Spectrom.2000, 195−196, 591−598.(32) Blakney, G. T.; Hendrickson, C. L.; Marshall, A. G. Predatordata station: A fast data acquisition system for advanced FT-ICR MSexperiments. 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 oilsands process-affected water; An application of high-performanceliquid chromatography/orbitrap mass spectrometry. Rapid Commun.Mass Spectrom. 2013, 27 (21), 2317−2326.

Energy & Fuels Article

dx.doi.org/10.1021/ef502007b | Energy Fuels 2015, 29, 2768−27732773