A Review of in Situ Chemical Oxidation and Heterogeneity

  • Journal List
  • ACS AuthorChoice
  • PMC4151705

Environ Sci Technol. 2014 Sep 2; 48(17): 10330–10336.

In Situ Chemical Oxidation of Contaminated Groundwater past Persulfate: Decomposition by Fe(Three)- and Mn(IV)-Containing Oxides and Aquifer Materials

Haizhou Liu

Department of Civil and Environmental Engineering science and §Department of Material Science and Engineering, University of California at Berkeley, Berkeley, California 94720, U.s.

Thomas A. Bruton

Department of Civil and Environmental Engineering and §Department of Material Scientific discipline and Applied science, University of California at Berkeley, Berkeley, California 94720, United States

Fiona Thou. Doyle

Department of Civil and Ecology Engineering and §Department of Textile Science and Applied science, University of California at Berkeley, Berkeley, California 94720, United States

David Fifty. Sedlak

Department of Civil and Environmental Technology and §Department of Material Science and Engineering, University of California at Berkeley, Berkeley, California 94720, United States

Received 2014 Apr 26; Revised 2014 Jul thirty; Accepted 2014 Aug 1.

Abstruse

An external file that holds a picture, illustration, etc.  Object name is es-2014-02056d_0006.jpg

Persulfate (S2Oeight 2–) is being used increasingly for in situ chemic oxidation (ISCO) of organic contaminants in groundwater, despite an incomplete understanding of the mechanism through which it is converted into reactive species. In particular, the decomposition of persulfate by naturally occurring mineral surfaces has not been studied in detail. To gain insight into the reaction rates and machinery of persulfate decomposition in the subsurface, and to identify possible approaches for improving its efficacy, the decomposition of persulfate was investigated in the presence of pure metal oxides, clays, and representative aquifer solids collected from field sites in the presence and absence of benzene. Nether weather typical of groundwater, Atomic number 26(III)- and Mn(4)-oxides catalytically converted persulfate into sulfate radical (Then4 •–) and hydroxyl radical (HO) over time scales of several weeks at rates that were 2–twenty times faster than those observed in metal-free systems. Baggy ferrihydrite was the about reactive atomic number 26 mineral with respect to persulfate decomposition, with reaction rates proportional to solid mass and surface expanse. As a result of radical concatenation reactions, the rate of persulfate decomposition increased by every bit much as 100 times when benzene concentrations exceeded 0.1 mM. Due to its relatively wearisome rate of decomposition in the subsurface, it tin can be advantageous to inject persulfate into groundwater, assuasive it to migrate to zones of low hydraulic conductivity where clays, metal oxides, and contaminants volition accelerate its conversion into reactive oxidants.

Introduction

Despite more than three decades of effort, contamination of groundwater with organic pollutants remains a significant threat to drinking h2o supplies.1 Attempts to employ ex situ treatment are frequently expensive and require decades to complete, while bioremediation and physical isolation of contaminants are difficult to use at many complex sites. As a result, engineers are increasingly using in situ chemical oxidation (ISCO) equally an expedient approach for contaminant remediation.2−5

ISCO has been used for the treatment of chlorinated organic solvents and petroleum hydrocarbons for over three decades.vi−eight Many of the early efforts to utilise ISCO used permanganate (MnO4 ) or hydrogen peroxide (H2O2). MnO4 reacts quickly with sure contaminants,9−11 but the germination of manganese oxides (e.g., MnO2(south)) can clog soil pores and inhibit oxidant transport in the subsurface. H2O2 reacts with minerals in the subsurface to produce hydroxyl radicals (HO) via Fenton-like reactions.12 Because HO reacts rapidly with H2O2 to produce O2 and H2O13 and Fenton-like reactions typically exhibit depression yields at circumneutral pH values, the process is relatively inefficient.14−17 In addition, catalase enzyme activity in aquifer materials tin decompose HiiOii through a nonradical pathway.xiii As a result of its high reactivity, HtwoOii decomposes chop-chop in the subsurface, which can make it difficult to remediate contaminants that are located far from the reagent injection bespeak.

While permanganate and hydrogen peroxide are withal used frequently, persulfate (South2Oviii two–) has recently become pop as an ISCO oxidant.5,18−20 The mechanism through which Due southiiO8 2– oxidizes contaminants is similar to the way in which HtwoOtwo is converted to HO by ultraviolet light, with thermolytic cleavage of the peroxide bail slowly producing sulfate radical (SO4 •–) instead of HO:

equation image

ane

The formation of radicals from persulfate is commonly referred to as persulfate activation. However, in many cases it is difficult to measure sulfate radical concentrations and a decrease in persulfate loss is often used as a proxy for radical production. In such cases it may be more appropriate to refer to persulfate decomposition instead of activation, and this is the convention that is followed throughout this manuscript.

At 25 °C, the half-life of South2O8 two– is approximately 600 days, which ways that groundwater must be heated if this mechanism is to be applied for remediation.21−25 Other approaches for accelerating the rate of StwoO8 ii– decomposition include UV irradiation,26−28 electrochemical systems,29 zerovalent iron,30 ferrous31 or cobalt salts,32 and base of operations activation.33 Base activation is used most unremarkably in field applications.33 Each of these mechanisms has its limitations. For example, use of ultraviolet low-cal in subsurface remediation is impractical, cobalt is a toxic metallic, Fe(II) salts are rapidly oxidized to insoluble Fe(Iii)-oxides in the presence of persulfate, and thermal activation is energy intensive nether the weather condition encountered in soils and aquifers.

Compared to the mechanisms listed to a higher place, persulfate decomposition past reactions at mineral surfaces are not also-known or understood. Results from recent studies propose that aquifer materials and natural organic affair (NOM) can increase persulfate decomposition rates.34−36 The charge per unit of decomposition was positively correlated with the content of amorphous atomic number 26 in aquifer solids.34 Empirical studies besides demonstrated that the decomposition rate was enhanced by approximately an order of magnitude by the presence of aquifer solids.21,34 Although the disappearance of persulfate during ISCO is slow compared to that of H2O2, the depression reactivity of persulfate in aquifer solids can show advantageous because information technology may allow the oxidant to reach source zones. It also may be useful in remediation of dilute plumes. All the same, there is a lack of key understanding of the factors that control the rate of persulfate decomposition by aquifer materials. In add-on to Fe(III) and Mn(Iv) content, other factors that may bear on the rate of persulfate decomposition include pH and the concentration of radical scavengers.

To provide insight into the role of naturally occurring minerals in persulfate decomposition and to develop a means of predicting the rate of persulfate loss during ISCO treatment, persulfate decomposition was studied under well-controlled conditions. Experiments were conducted using suspensions of pure minerals and aquifer solids to decompose persulfate in the presence of a pH buffer. The production of SO4 •– or HO was followed using benzene equally a probe compound. Concentrations of persulfate and phenol (i.due east., a benzene oxidation production) were measured nether conditions expected in groundwater.

Materials and Methods

All chemicals used in this study were reagent grade or college. All solutions were prepared using deionized water (resistivity >18.ii MĪ©, Millipore system). Ferrihydrite (Fe(OH)three(s)) and pyrolusite (Ī²-MnO2(s)) were obtained from Sigma-Aldrich. Crystallized silica (SiO2(s), i.e., pure sand, ACROS Organics, Inc.) was cleaned to remove trace amounts of metals co-ordinate to the procedure described in the Supporting Data (SI) section. Goethite (Ī±-FeOOH(s)) was synthesized by mixing 1 M Atomic number 26(NO3)three solution with 5 One thousand KOH and and so aging the precipitates at 70 °C for 60 h.37 The synthesized goethite particles were rinsed with deionized water and dried using a Labconco Freezone iv.5 freeze-dry system (Labconco Corp., Kansas Urban center, MO). The synthesized goethite was characterized by Ten-ray diffraction (XRD) to confirm its purity (SI Figure S1).

Two clay materials were obtained from the Clay Minerals Society: Australian nontronite (26.0% Iron by weight) and Wyoming montmorillonite (2.59% Atomic number 26 by weight). In addition, aquifer materials were obtained from five sites in California and Arizona. Details of the physicochemical characteristics of pure minerals, clays, and aquifer materials are provided in SI Table S1. Additional information was provided previously.17 Prior to use, each fabric was crushed with a mortar and pestle and sieved to collect particles with sizes between 38 and 150 Ī¼m (sieves with mesh 400 and 100, respectively). The sieved particles were autoclaved at 120 °C and 20 psi for 45 min and stored under sterile conditions.

Persulfate decomposition experiments were carried out under chemical conditions typical of groundwater. Prior to initiation of each experiment, suspensions were prepared with solids concentrations between 12.5 and 125 grand/50 using air-saturated deionized water or constructed groundwater (SI Table S2). In some experiments, the suspension was purged with N2 gas to achieve Oii-free atmospheric condition and tests were conducted in a glovebox. In about experiments, the solution was maintained at pH eight.0 with l mM borate buffer prepared by adding 15 mL of a solution of 1 M of boric acid (H2BOthree) and 62.5 mM of sodium tetraborate (NaiiB4O7). The use of loftier buffer concentration was necessary to maintain a constant pH throughout the experiment. For some experiments, boosted NaHCOthree was added along with the borate buffer to reach a bicarbonate concentration between 0.ane and 10 mM.

Several experiments were conducted with benzene at concentrations ranging from 0.ane to 1 mM. Benzene solutions were prepared by diluting from a 1.2 mM stock solution that was freshly prepared by dissolving 10 Ī¼L of anhydrous benzene (purity ≥99.8%, Sigma-Aldrich Inc.) in a 1 L glass volumetric flask filled with deionized water without headspace. No benzene was volatilized during this preparation step.

To beginning a decomposition experiment, persulfate was added to yield initial concentrations ranging from one to 50 mM, using aliquots of a freshly prepared 100 mM ThoutwoSouthward2O8 solution. Later on mixing, the interruption was immediately transferred to multiple sealed glass tubes with no headspace and placed on a rotating mixer (Labquake Tube Rotators, Thermo Scientific Inc.). Each tube was treated as a sacrificial reactor and discarded afterwards sampling. Nearly weather condition were tested in triplicate, forth with a homogeneous control consisting of only persulfate. For experiments with benzene, 2 additional controls consisting of just benzene and benzene with solids were included. All experiments were conducted at room temperature of 23 ± ii °C.

At predetermined sampling intervals, each sealed sacrificial tube reactor was centrifuged at 3000 m for iii min to dissever the solids from the solution. The supernatant was immediately removed and filtered through a 0.22-Ī¼m nylon filter prior to analysis for persulfate, benzene and oxidation products. The solids were resuspended in acetonitrile, agitated vigorously with a vortex mixer, and so placed on a rotating mixer for 3 h to extract benzene and phenol from the solids. The acetonitrile was separated from the solids using the same centrifugation and filtration procedure and analyzed for adsorbed benzene and phenol.

Persulfate was measured using the KI colorimetric method38 with a Lambda-14 UV spectrophotometer (PerkinElmer Inc., Waltham, MA). Benzene and phenol were analyzed on a Waters Brotherhood 2695 HPLC (Waters Corp., Milford, MA) equipped with a diode array detector. A Waters Symmetry-C18 cavalcade was used with a mobile phase of 40% acetonitrile and 60% 10 mM formic acid at a flow charge per unit of 1 mL/min. Dissolved oxygen concentrations were determined using a YSI Model 58 oxygen probe (YSI Inc., Yellow Springs, OH).

Results and Discussion

Decomposition of Persulfate by Minerals

As expected, the rate of persulfate loss in the absence of solids (i.eastward., homogeneous command) at room temperature was very boring (Effigy 1A). The small loss of persulfate in the homogeneous command was consequent with the predicted thermal decomposition of S2O8 two– to Then4 •– (reaction one). Using the activation energy measured previously,21 a loss of only 3% of persulfate was predicted over 30 days, which was similar to the measured loss of 2% in our experiments. The heterogeneous arrangement with silica exhibited less than 3% persulfate loss over 32 days, which was indistinguishable from the homogeneous control. The addition of fe- and manganese-containing oxides accelerated the decomposition charge per unit significantly, with an average of 12%, 65%, and 45% of persulfate loss over 32 days for 100 one thousand/L suspensions of goethite, ferrihydrite, and pyrolusite, respectively (Figure 1A). The decomposition charge per unit was afflicted by the mass of solids in the suspension. For case, reducing the mass of pyrolusite from 100 to 50 k/L decreased persulfate loss from 45% to x% over 32 days, while increasing its mass from 100 to 200 and 500 g/L increased the loss to 60% and xc% over the same menstruation, respectively (Figure 1B). Increasing the mass of silica did non affect the rate of persulfate loss (SI Figure S2).

An external file that holds a picture, illustration, etc.  Object name is es-2014-02056d_0001.jpg

Stability and decomposition of persulfate in different systems. Initial persulfate concentration one mM, pH eight.0, ionic strength l mM. (A) Impact of unlike solids along with a homogeneous control, solid mass loading 100 g/L. (B) affect of pyrolusite mass loading on persulfate decomposition.

To examine the effect of suspended mineral mass on persulfate decomposition, information were converted to pseudo-first-order kinetics with the following expression:

equation image

In this expression, thou obs represents the observed pseudo-outset-order rate constant (due south–i), k SA is the surface area normalized pseudo-first-order rate constant (Fifty·1000–2·s–1), and C S is the BET surface area of the suspended solids (grand2·L–one).

The surface area normalized decomposition charge per unit constants for different solids varied by approximately 6 orders of magnitude (Figure 2). The relative reactivity of the minerals followed the guild pyrolusite > ferrihydrite > goethite > silica. Nontronite, an iron-rich dirt and montmorillonite, a clay with a relatively high manganese content, were both capable of persulfate decomposition, with montmorillonite decomposing persulfate at a rate that was approximately an guild of magnitude higher than that of nontronite (Figure ii). The aquifer materials exhibited decomposition rates that were similar to those observed for the iron oxides (Figure 2). These aquifer materials contained between 0.7 and ii.5% Iron and between 0.01 and 0.12% Mn. Iron- and manganese-containing minerals likewise decomposed persulfate at initial persulfate concentrations ranging from 5 to 50 mM with 50 g/L of each mineral (SI Effigy S3).

An external file that holds a picture, illustration, etc.  Object name is es-2014-02056d_0002.jpg

Comparison of rate constants of persulfate decomposition past different solids at varying concentrations. Rate constants are normalized by BET surface areas and corrected for thermal decomposition at room temperature. Initial persulfate i mM, pH 8.0. Error bar represents ane standard divergence.

Reaction Mechanism for Heterogeneous Persulfate Decomposition

Past illustration to the Fenton system, nosotros propose that the rate-determining pace for persulfate decomposition by an iron-containing mineral can be described past the following reaction:

equation image

2

where ≡Iron(III) represents the redox-active iron surface site. Ane-electron reduction of ≡Fe(3) would effect in the formation of persulfate radical (StwoO8 •–), which could then initiate radical chain reactions. The existence of Southward2O8 •– is supported by previous studies of the reaction between S2O8 2– and Sofour •–, in which SiiO8 ii– was oxidized to SiiO8 •– via a one-electron transfer reaction.39,40

Similar to the Fenton reaction, the overall reaction rate of persulfate decomposition is likely afflicted past the number of reactive surface sites and electron transfer rates.41 Every bit in the Fenton system, the surface reactivity of each blazon of mineral is likely affected by crystallinity and metallic coordination. For example, in the Fenton organisation, the HtwoO2 decomposition charge per unit was much higher in the presence of amorphous ferrihydrite than crystalline goethite.14 A similar trend was too observed in our experiments, with ferrihydrite exhibiting college charge per unit constants than goethite. Similarly, our observation of higher reactivity for manganese oxides was consistent with observations from the Fenton organisation.17

Previous research has demonstrated persulfate decomposition by dissolved transition metals, including Fetwo+, Mn2+, Ag+, Coii+, and Ni2+, under acidic conditions (i.e., pH < three).31,42 Acidic weather condition are potentially problematic in groundwater because they tin can release toxic metals and alter minerals. Under circumneutral pH weather, transition metals associated with surfaces can undergo similar reactions. Later on surface metals are reduced via reaction 2, ≡Fe(II) should be oxidized quickly past persulfate to regenerate ≡Fe(III) with the production of Then4 •– and sulfate:

equation image

3

Therefore, the metallic catalyzed decomposition of persulfate proceeds as follows:

equation image

4

The fact that persulfate tin be both oxidized and reduced is like to the machinery by which H2O2 acts as both an oxidant and reductant in the Fenton arrangement. In addition, S2Oeight •– can be produced through a radical chain propagation reaction:40,43,44

equation image

v

The fate of SouthiiO8 •– is unclear. Previous studies on the fate of the radical did not examine its decomposition machinery.45,46 1 possible fate of S2Oeight •– is oxidation of water, which would occur through a multistep procedure with the following stoichiometry:

equation image

half dozen

OH could besides serve as a source of HO46 and And so4 •– can also oxidize water,45 and nether element of group i weather condition

equation image

7

equation image

8

equation image

nine

In the absence of organic solutes, Reactions half-dozen and nine serve as concatenation termination steps.

The overall reaction of persulfate decomposition in water is

equation image

10

Results from selected experiments in which sulfate production was measured along with persulfate loss showed skilful agreement with the overall stoichiometry of reaction 10 (SI Effigy S4A). For Mn(4)-containing materials, a similar reaction pathway tin can take identify through a redox cycle involving Mn(IV)/Mn(3) surface species to generate sulfate radical and initiate chain reactions.

To assess the possible office of dissolved O2 in persulfate decomposition, experiments were conducted in the absence of oxygen with a pyrolusite suspension. Results showed that the presence of dissolved Otwo had no effect on the rate of persulfate loss (SI Effigy S4B). In add-on, dissolved Otwo was produced equally S2Oviii ii– decomposed. The observed stoichiometric ratio of O2 produced per StwoO8 ii– lost was approximately 0.v (SI Figure S5), which was consistent with reaction 10.

Impact of Organic Compounds on Persulfate Decomposition

In the presence of organic contaminants, the radical chain decomposition of persulfate changes. Using benzene as a representative organic contaminant, the stability of persulfate was examined (Figures 3 and 4). The rate of persulfate decomposition increased equally benzene concentrations increased from 0 to thou Ī¼M. For example, in the presence of 50 g/Fifty ferrihydrite, approximately 15% of persulfate was decomposed over thirty days in the absenteeism of benzene. When 100 Ī¼M of benzene was added, persulfate decomposition accelerated significantly, with 45% loss over 30 days. At a benzene concentration of 1000 Ī¼M, approximately 75% of persulfate was decomposed over the same catamenia (Effigy 3). These results indicated that the half-life of persulfate would decrease from 170 days in the absence of benzene to 45 and 20 days in the presence of 100 and thou Ī¼M of benzene, respectively. A like trend was observed for other pure minerals, with the exception of silica, which was not capable of catalyzing persulfate decomposition. For the aquifer materials, benzene appeared to have an effect, but it was harder to notice due to higher variability amidst replicates due to heterogeneity of the aquifer materials.

An external file that holds a picture, illustration, etc.  Object name is es-2014-02056d_0003.jpg

Bear upon of benzene concentration on the decomposition of persulfate in the presence of 50 g/Fifty ferrihydrite, pH viii.0. No mineral was nowadays in the homogeneous control.

An external file that holds a picture, illustration, etc.  Object name is es-2014-02056d_0004.jpg

Bear on of benzene concentration on the decomposition of persulfate in the presence of different solids. Solids mass loading 50 chiliad/L, pH 8.0.

The acceleration of persulfate decomposition in the presence of benzene may be owing to radical concatenation reactions initiated by oxidation of benzene. Previous enquiry using HO has shown that the intermediate products of benzene oxidation include organic radicals that can propagate the chain.28,48 For case, the reaction of HO with benzene results in formation of a cationic radical that is subsequently hydrolyzed to hydroxyl-cyclohexadienyl (HCHD) radical.48 HCHD radical farther reacts with oxygen to produce an organic peroxy radical that generates HOtwo and phenol. Under circumneutral pH weather condition, HOii will deprotonate to superoxide (Oii •–), which can undergo iron-catalyzed dismutation:

equation image

11

equation image

12

During this process, the faster rates of reduction of Fe(III) relative to oxidation of Atomic number 26(2) favor the presence of reduced Fe(II),49 thereby contributing to oxidant decomposition. H2O2 produced past superoxide dismutation reacts quickly with iron and manganese through Fenton-like reactions on mineral surfaces.13,14,17

Presumably a similar radical concatenation machinery occurs during the reaction of benzene with SO4 •–. A second-order rate constant of 3 × 10nine Grand–ane·s–one has been reported for the initial step of this reaction:50

equation image

13

Because the branching ratio of reaction thirteen is much bigger than those of reactions 5 and 7 under experimental weather condition used in this study (i.due east., yard 13[benzene] ≫ 1000 v[S2O8 ii–]; thou 13[benzene] ≫ thou 7[H2O], reaction rate constants provided in SI Table S4), essentially all SO4 •– reacted with benzene instead of persulfate or water when benzene was nowadays.

The acceleration of persulfate decomposition by benzene was besides observed in experiments using synthetic groundwater that contained chloride, bromide, carbonate and natural organic matter (SI Table S2). In these experiments, the rates of persulfate decomposition by iron and manganese-containing minerals were consistent with those measured in benzene solutions prepared with borate-buffered water (SI Table S3).

Under typical groundwater atmospheric condition, boosted reactions between SO4 •– and chloride or bromide tin can take place:44,51

equation image

14

equation image

15

Cl and Br radicals can further react with water to produce other brusque-lived radicals.52−54 However, in the presence of 1 mM benzene and under typical groundwater weather condition, chloride and bromide are pocket-sized sinks for sulfate radical (i.due east., yard 13[benzene] > k 14[Cl]; k thirteen[benzene] > one thousand 15[Br]). Therefore, groundwater constituents only had a small impact on the scavenging of Theniv •– radicals in the presence of benzene.

In addition, under groundwater atmospheric condition, And then4 •– generated through reaction 3 can react with bicarbonate:47

equation image

16

COiii •– can then react with persulfate to cause farther decomposition of persulfate:

equation image

17

All the same, for the range of concentrations studied in these experiments ([HCOthree ] = ane–10 mM and [Southward2O8 2–] = 1 mM) and in the presence of 1 mM benzene, bicarbonate is non a major sink for sulfate radical (i.e., yard 16[HCO3 ] < k 13[benzene]) and reaction 17 will not be important to persulfate decomposition. This is consistent with observations from experiments in which persulfate decomposition was measured in the presence of benzene at varying concentrations of bicarbonate (SI Figure S6).

The presence of And then4 •– or HO during persulfate decomposition in heterogeneous systems was confirmed by measurements of phenol production. In heterogeneous systems with Fe(III)- and Mn(IV)-containing minerals, an initial persulfate concentration of ane mM and an initial benzene concentration of 100 Ī¼M, phenol was formed in the presence of different solids (Figure 5). For example, approximately 25 Ī¼M of phenol was generated in the presence of 50 g/50 goethite, nontronite, and montmorillonite afterward 30 days. Fifteen micromolar of phenol was produced in pyrolusite, and 5 Ī¼M, in ferrihydrite. Benzene loss ranged from twenty to 40 Ī¼M in these experiments.

An external file that holds a picture, illustration, etc.  Object name is es-2014-02056d_0005.jpg

Formation of phenol as oxidation product of benzene by sulfate radical in the presence of fifty g/L dissimilar Fe- or Mn-containing oxides and aquifer materials. Initial benzene concentration 100 Ī¼M, initial persulfate concentration 1 mM, pH viii.0, ionic strength 55 mM.

Environmental Implications

The findings from this study accept significant implications for the employ of persulfate for ISCO applications. First, the observation that fe- and manganese-containing minerals decompose persulfate under conditions encountered in the subsurface allow for prediction of the rate at which SOiv •– or HO is produced in the absence of added activators, such as heat or base. In aquifers dominated by pure sand, persulfate is very stable with one-half-lives of more ii years. If Fe(III)- or Mn(IV)-containing oxides comprise greater than approximately 2% of aquifer solids, the rate of persulfate decomposition will increase, with half-lives decreasing to less than a twelvemonth. When persulfate encounters a zone that is rich in iron- or manganese-oxides or dirt, its half-life tin can decrease essentially.

Second, one time persulfate is injected into the subsurface, its treatment efficacy may exist affected past the distance that it travels in the surface. In many situations, information technology may be preferable for persulfate to persist in the aquifer for many months to allow it to reach contaminants afar from injection wells. Results from this study prove that persulfate compares favorably to H2Oii-based ISCO systems in this respect. Persulfate decomposition was much slower than hydrogen peroxide decomposition in the Fenton organisation, with a half-life that was more than than 2 orders of magnitude longer (SI Effigy S7). Hydrogen peroxide was decomposed over several hours (SI Figure S8A), whereas persulfate decomposition took several weeks (SI Figure S8B). Considering different hydraulic conductivities in the subsurface and uncertainties associated with the rate of oxidant transport in the aquifer, the high stability of persulfate in uncontaminated aquifer zones indicates more effective ship and delivery of oxidants.

Finally, the rate of persulfate decomposition accelerates when the injected oxidant reaches a contaminant feather. The half-life of persulfate in the aquifer can be more one twelvemonth in the absence of an organic contaminant. Once persulfate encounters a contaminant such as benzene, radical chain reactions increase the persulfate decomposition rate and its half-life can decrease to as piffling equally a few days. Its half-life tin can likewise decrease in the presence of iron- or manganese-containing oxides or clays, which are frequently associated with low hydraulic conductivity regions where contaminants frequently accrue.

Acknowledgments

This written report was supported by the U.Due south. National Constitute for Environmental Health Sciences (NIEHS) Superfund Inquiry Program (Grant P42 ES004705) and the Superfund Research Center at Academy of California, Berkeley. Nosotros thank our grouping member Anh Pham for stimulating discussions and Dr. Urs Jans for his aid in conducting initial experiments that helped us to develop a means of studying these slow reactions.

Funding Statement

National Institutes of Health, The states

Supporting Information Available

Additional description of fe and manganese mineral preparation and viii figures on the rate of persulfate decomposition and product formation in different atmospheric condition. This fabric is available costless of charge via the Internet at http://pubs.acs.org.

Author Present Address

Department of Chemical and Environmental Applied science, Academy of California at Riverside, Riverside, CA 92521, The states.

Notes

The authors declare no competing financial interest.

Supplementary Fabric

References

  • Alternatives for Managing the Nation's Complex Contaminated Groundwater Sites; The National Academies Press, 2013. [Google Scholar]
  • Siegrist R. L.; Crimi G.; Simpkin T. J.. In situ chemic oxidation: technology, description and condition. In situ chemical oxidation for groundwater remediation; Springer Media, LLC: New York Urban center, 2011; Affiliate 1. [Google Scholar]
  • Interstate Engineering science and Regulatory Council (ITRC). Technical and Regulatory Guidance for In Situ Chemical Oxidation of Contaminated Soil and Groundwater, 2nd ed.; 2005; http://www.itrcweb.org.
  • Krembs F. J.; Siegrist R. L.; Crimi M. L.; Furrer R. F.; Petri B. G. ISCO for groundwater remediation: analysis of field applications and performance. Ground H2o Monit. Rem. 2010, 30 442–53. [Google Scholar]
  • Tsitonaki A.; Petri B.; Crimi Yard.; Mosbaek H.; Siegrist R. L.; Bjerg P. Fifty. In Situ Chemical Oxidation of Contaminated Soil and Groundwater Using Persulfate: A Review. Crit. Rev. in Environ. Sci. and Technol. 2010, 40 155–91. [Google Scholar]
  • Miller C. M.; Valentine R. L. Hydrogen peroxide decomposition and quinoline degradation in the presence of aquifer material. Water Res. 1995, 29 ten2353–2359. [Google Scholar]
  • Gates D. D.; Siegrist R. Fifty. In-situ chemic oxidation of trichloroethylene using hydrogen-peroxide. J. Environ. Eng. 1995, 121 9639–644. [Google Scholar]
  • Yan Y. E.; Schwartz F. W. Oxidative degradation and kinetics of chlorinated ethylenes by potassium permanganate. J. Contam. Hydrol. 1999, 37 3–four343–365. [Google Scholar]
  • Waldemer R. H.; Tratnyek P. M. Kinetics of contaminant degradation by permanganate. Environ. Sci. Technol. 2006, 40 31055–1061. [PubMed] [Google Scholar]
  • Poulson S. R.; Naraoka H. Carbon isotope fractionation during permanganate oxidation of chlorinated ethylenes (cDCE, TCE, PCE). Environ. Sci. Technol. 2002, 36 153270–3274. [PubMed] [Google Scholar]
  • Huang K. C.; Hoag K. Eastward.; Chheda P.; Woody B. A.; Dobbs One thousand. 1000. Oxidation of chlorinated ethenes by potassium permanganate: a kinetics study. J. Hazard. Mater. 2001, 87 i155–169. [PubMed] [Google Scholar]
  • Pignatello J. J.; Oliveros East.; MacKay A. Advanced oxidation processes for organic contaminant destruction based on the Fenton reaction and related chemistry. Crit. Rev. in Environ. Sci. Technol. 2006, 36 1ane–84. [Google Scholar]
  • Petigara B. R.; Blough Northward. V.; Mignerey A. C. Mechanisms of hydrogen peroxide decomposition in soils. Environ. Sci. Technol. 2002, 36 4639–645. [PubMed] [Google Scholar]
  • Kwan W.; Voelker B. M. Rates of hydroxyl radical generation and organic chemical compound oxidation in mineral-catalyzed Fenton-like systems. Environ. Sci. Technol. 2003, 37 63270–3274. [PubMed] [Google Scholar]
  • Xu X.; Thomson North. R. Hydrogen peroxide persistence in the presence of aquifer materials. Soil Sediment Contam. 2010, 19 5602–616. [Google Scholar]
  • Watts R. J.; Finn D. D.; Cutler L. M.; Schmidt J. T.; Teel A. L. Enhanced stability of hydrogen peroxide in the presence of subsurface solids. J. Contam. Hydrol. 2007, 91 3312–326. [PubMed] [Google Scholar]
  • Pham A. L.; Doyle F. M.; Sedlak D. L. Kinetics and efficiency of HtwoO2 activation by iron-containing minerals and aquifer materials. H2o Res. 2012, 46 196454–6462. [PMC costless commodity] [PubMed] [Google Scholar]
  • Liang C.; Bruell C. J.; Marley M. C.; Sperry K. L. Thermally activated persulfate oxidation of trichloroethylene (TCE) and 1,1,1-trichloroethane (TCA) in aqueous systems and soil slurries. Soil Sediment Contam. 2003, 12 two207–228. [Google Scholar]
  • Waldemer R. H.; Tratnyek P. One thousand.; Johnson R. L.; Nurmi J. T. Oxidation of chlorinated ethenes past heat-activated persulfate: kinetics and products. Environ. Sci. Technol. 2007, 41 31010–1015. [PubMed] [Google Scholar]
  • Costanza J.; OtaƱo G.; Callaghan J.; Pennell Thousand. D. PCE oxidation past sodium persulfate in the presence of solids. Environ. Sci. Technol. 2010, 44 249445–9450. [PubMed] [Google Scholar]
  • Johnson R. L.; Tratnyek P. Thou.; Johnson R. O. Persulfate persistence under thermal activation weather condition. Environ. Sci. Technol. 2008, 42 249350–9356. [PubMed] [Google Scholar]
  • Hori H.; Nagaoka Y.; Murayama M.; Kutsuna S. Efficient decomposition of perfluorocarboxylic acids and alternative fluorochemical surfactants in hot water. Environ. Sci. Technol. 2008, 42 197438–7443. [PubMed] [Google Scholar]
  • Liu C. S.; Higgins C. P.; Wang F.; Shih K. Effect of temperature on oxidative transformation of perfluorooctanoic acid (PFOA) past persulfate activation in water. Sep. Purif. Technol. 2012, 91, 46–51. [Google Scholar]
  • Lee Y.; Lo Due south.; Kuo J.; Hsieh C. Decomposition of perfluorooctanoic acid past microwave-activated persulfate: Effects of temperature, pH, and chloride ions. Front end. Environ. Sci. Engr. Communist china. 2012, 6 117–25. [Google Scholar]
  • Deng Y.; Ezyske C. M. Sulfate radical-advanced oxidation process (SR-AOP) for simultaneous removal of refractory organic contaminants and ammonia in landfill leachate. H2o. Res. 2011, 45 186189–6194. [PubMed] [Google Scholar]
  • Hori H.; Yamamoto A.; Hayakawa E.; Taniyasu Due south.; Yamashita Due north.; Kutsuna South. Efficient decomposition of environmentally persistent perfluorocarboxylic acids past use of persulfate as a photochemical oxidant. Eniron. Sci. Technol. 2005, 39 72383–2388. [PubMed] [Google Scholar]
  • Lau T. Chiliad.; Chu W.; Graham North. J. D. The aqueous degradation of butylated hydroxyanisole by UV/StwoO8 2-: study of reaction mechanisms via dimerization and mineralization. Environ. Sci. Technol. 2007, 41 2613–619. [PubMed] [Google Scholar]
  • Antoniou Yard. G.; De La Cruz A. A.; Dionysiou D. D. Intermediates and reaction pathways from the degradation of microcystin-LR with sulfate radicals. Environ. Sci. Technol. 2010, 37 197238–7244. [PubMed] [Google Scholar]
  • Yuan South.; Liao P.; Alshawabkeh A. N. Electrolytic manipulation of persulfate reactivity by atomic number 26 electrodes for trichloroethylene degradation in groundwater. Environ. Sci. Technol. 2014, 48 1656–663. [PMC free article] [PubMed] [Google Scholar]
  • Ahn Due south.; Peterson T. D.; Righter J.; Miles D. K.; Tratnyek P. M. Disinfection of ballast water with fe activated persulfate. Environ. Sci. Technol. 2013, 47 xx11717–11725. [PubMed] [Google Scholar]
  • Liang C.; Bruell C. J.; Marley M. C.; Sperry K. L. Persulfate oxidation for in situ remediation of TCE. Two. Activated past chelated ferrous ion. Chemosphere 2004, 55 ix1225–1233. [PubMed] [Google Scholar]
  • Anipsitakis G. P.; Dionysiou D. D.; Gonzalez K. A. Cobalt-mediated activation of peroxymonosulfate and sulfate radical set on on phenolic compounds. Implications of chloride ions. Environ. Sci. Technol. 2006, 40 3chiliad–1007. [PubMed] [Google Scholar]
  • Furman O. S.; Teel A. L.; Watts R. J. Mechanism of base activation of persulfate. Environ. Sci. Technol. 2010, 44 166423–6428. [PubMed] [Google Scholar]
  • Sra K. South.; Thomson N. R.; Barker J. F. Persistence of persulfate in uncontaminated aquifer materials. Environ. Sci. Technol. 2010, 44 83098–3104. [PubMed] [Google Scholar]
  • Ahmad M.; Teel A. L.; Watts R. J. Persulfate activation by subsurface minerals. J. Contam. Hydrol. 2010, 115 i34–45. [PubMed] [Google Scholar]
  • Teel A. L.; Ahmad M.; Watts R. J. Persulfate activation by naturally occurring trace minerals. J. Haz. Mater. 2011, 196, 153–159. [PubMed] [Google Scholar]
  • Schwertmann U.; Cornell R. Grand.. Iron oxides in the laboratory: preparation and characterization; VCH Publisher, Inc.: New York, 1991; pp 64–71. [Google Scholar]
  • Liang C.; Huang C. F.; Mohanty N.; Kurakalva R. M. A rapid spectrophotometric conclusion of persulfate anion in ISCO. Chemosphere 2008, 73 91540–1543. [PubMed] [Google Scholar]
  • McElroy W. J.; Waygood South. J. Kinetics of the reactions of the Then4 radical with SO4 , Due southiiO8 2-, H2O and Iron2+ . J. Chem. Soc. Faraday Trans. 1990, 86 142557–2564. [Google Scholar]
  • Yu X. Y.; Bao Z. C.; Barker J. R. Free radical reactions involving Cl•, Cl2 •, and And then4 •– in the 248 nm photolysis of aqueous solutions containing South2O8 two- and Cl . J. Phys. Chem. A 2004, 108 two295–308. [Google Scholar]
  • Lin S.; Gurol M. D. Catalytic decomposition of hydrogen peroxide on fe oxide: kinetics, machinery, and implications. Environ. Sci. Technol. 1998, 32 x1417–1423. [Google Scholar]
  • Anipsitakis G. P.; Dionysiou D. D. Radical generation by the interaction of transition metals with common oxidants. Environ. Sci. Technol. 2004, 38 thirteen3705–3712. [PubMed] [Google Scholar]
  • Schuchmann H. P.; Deeble D. J.; Olbrich M.; von Sontag C. The SO4 •– -induced chain reaction of 1,3- dimethyluracil with peroxodisulphate. Int. J. Radiat. Biol. 1987, 51, 441–453. [PubMed] [Google Scholar]
  • Das T. Northward. Reactivity and function of SO5 •– radical in aqueous medium concatenation oxidation of sulfite to sulfate and atmospheric sulfuric acid generation. J. Phys. Chem. A 2001, 105 forty9142–9155. [Google Scholar]
  • Herrmann H.; Reese A.; Zellner R. Time-resolved UV/VIS diode-array absorption spectroscopy of SOx (x=3, 4, v) radical anions in aqueous solution. J. Mol. Struct. 1995, 348, 183–186. [Google Scholar]
  • Neta P.; Huie R. E.; Ross A. B. Rate constants for reactions of inorganic radicals in aqueous solution. J. Phys. Chem. Ref. Data. 1988, 17 three1027–1284. [Google Scholar]
  • Zuo Z. H.; Cai Z. L.; Katsumura Y.; Chitose N.; Muroya Y. Reinvestigation of the acrid-base equilibrium of the (bi)carbonate radical and pH dependence of its reactivity with inorganic reactants. Radiat. Phys. Chem. 1999, 55 115–23. [Google Scholar]
  • Pan X.; Schuchmann M. North.; von Sontag C. Oxidation of benzene by the OH radical. A production and pulse radiolysis report in oxygenated aqueous solution. J. Chem. Soc. Perkin Trans. ii. 1993, 3, 289–297. [Google Scholar]
  • Sedlak D. L.; Hoigne J. The role of copper and oxalate in the redox cycling of fe in atmospheric waters. Atmosph. Env. Part A. Gen. Top. 1993, 27 fourteen2173–2185. [Google Scholar]
  • Neta P.; Madhavan V.; Zemel H.; Fessenden R. West. Rate constants and mechanism of reaction of sulfate radical anion with aromatic compounds. J. Am. Chem. Soc. 1977, 99 ane163–164. [Google Scholar]
  • Peyton Chiliad. R. The gratis-radical chemical science of persulfate-based full organic carbon analyzers. Mar. Chem. 1993, 41 i91–103. [Google Scholar]
  • Jayson G.; Parsons B.; Swallow A. J. Some simple, highly reactive, inorganic chlorine derivatives in aqueous solution. Their formation using pulses of radiations and their role in the machinery of the Fricke dosimeter. J. Chem. Soc., Faraday Trans. one. 1973, 69, 1597–1607. [Google Scholar]
  • Grigorev A. E.; Makarov I. E.; Pikaev A. K. Formation of Cl2 in the bulk solution during the radiolysis of full-bodied aqueous-solutions of chlorides. High Energy Chem. 1987, 21 two99–102. [Google Scholar]
  • Zehavi D.; Rabani J. Oxidation of aqueous bromide ions by hydroxyl radicals. A pulse radiolytic investigation. J. Phys. Chem. 1972, 76 3312–319. [Google Scholar]

jordancoun1994.blogspot.com

Source: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4151705/

0 Response to "A Review of in Situ Chemical Oxidation and Heterogeneity"

Post a Comment

Iklan Atas Artikel

Iklan Tengah Artikel 1

Iklan Tengah Artikel 2

Iklan Bawah Artikel