Science of The Total Environment: Klara Rusevova Crincoli Patrick K. Jones Scott G. Huling

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Science of the Total Environment 734 (2020) 139435

Contents lists available at ScienceDirect

Science of the Total Environment

journal homepage: www.elsevier.com/locate/scitotenv

Fenton-driven oxidation of contaminant-spent granular activated carbon


(GAC): GAC selection and implications
Klara Rusevova Crincoli a, Patrick K. Jones b, Scott G. Huling b,⁎
a
National Research Council, Robert S. Kerr Environmental Research Center, 919 Kerr Lab Dr., Ada, OK 74820, USA
b
U.S. Environmental Protection Agency, Office of Research and Development, National Risk Management Research Laboratory, Robert S. Kerr Environmental Research Center, 919 Kerr Lab Dr., Ada,
OK 74820, USA

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Treatment of contaminant-spent granu-


lar activated carbon (GAC); chemical
oxidation.
• 31 GACs evaluated in the adsorption ox-
idation treatment process.
• The adsorption oxidation treatment
process is GAC-dependent.
• Success of GAC raw materials was bitu-
minous coal N wood N coconut N peat.
• GAC selection is a key factor in the suc-
cess of adsorption oxidation technology.

a r t i c l e i n f o a b s t r a c t

Article history: Raw materials, activation methods, and post-activation treatment used in manufacturing granular activated car-
Received 15 January 2020 bon (GAC) results in a spectrum of physicochemical characteristics that potentially impact the adsorption oxida-
Received in revised form 27 April 2020 tion treatment process. A comprehensive study is lacking that assesses the effect of GAC characteristics on
Accepted 12 May 2020
adsorption oxidation treatment of contaminant spent-GAC. Consequently, it is inherently assumed the treatment
Available online 15 May 2020
process is GAC-independent. Here, GACs (n = 31) were characterized and used in the hydrogen peroxide
Editor: Ching-Hua Huang (H2O2)-based adsorption oxidation treatment of 2-chlorophenol (2CP)-spent GAC. The GACs exhibited a range
in surface area, pore volume distribution, metals content, surface functionality, and H2O2 reaction. Chloride re-
Keywords: covery, the treatment metric for 2CP oxidation, indicated a wide range in oxidation (0–49.2%) where
Adsorption bituminous- and wood-based GAC performed best. A selected subset of GACs (n = 12), amended with iron,
Oxidation methyl tert-butyl ether (MTBE), and H2O2, exhibited a range in oxidative treatment (1.1–57.9%). Correlations
GAC treatment were established between GAC surface functionality, H2O2 reactivity, adsorption, and MTBE oxidation indicating
GAC characterization multiple parameters play a collective and compounding role. The order of GACs successfully used in the treat-
MTBE
ment process is bituminous-based coal N wood N coconut N peat. Results showed adsorption oxidation treatment
2-chlorophenol
is GAC-dependent, and therefore, GAC selection is a key factor in the success of this technology.
Published by Elsevier B.V.

⁎ Corresponding author.
E-mail addresses: rusevova.klara@epa.gov (K.R. Crincoli), jones.kyle@epa.gov (P.K. Jones), huling.scott@epa.gov (S.G. Huling).

https://doi.org/10.1016/j.scitotenv.2020.139435
0048-9697/Published by Elsevier B.V.
2 K.R. Crincoli et al. / Science of the Total Environment 734 (2020) 139435

1. Introduction temperature steam or through a high temperature and/or chemical


treatment process. During thermal activation, the internal porosity in-
1.1. Adsorption oxidation process description creases and begins to separate the layered stacks resulting in high sur-
face area. The internal porosity exhibits a range in pore volume
The oxidative treatment of large volumes of contaminated water is distribution including micropores (b2 nm in diameter), mesopores
limited by low aqueous concentrations of target contaminants resulting (2–50 nm), and macropores (N50 nm) which impacts molecular size-
in slow and/or inefficient reaction rates (Georgi and Kopinke, 2005). Al- dependent diffusional access of the contaminant, H2O2, or other re-
ternatively, adsorption followed by oxidative treatment agents used in the adsorption oxidation process.
(i.e., “adsorption oxidation”) involves the combined, synergistic use of Carbon activation impacts the formation of functional groups found
two reliable and well-established treatment technologies; contaminant at the edges of the GAC's graphene planes. These functional groups can
adsorption onto granular activated carbon (GAC), and subsequent be categorized based upon their oxygen content and include acidic sur-
Fenton-driven oxidation of the sorbed contaminants. During the ad- face oxide (ASO) functional groups, generally categorized into four
sorptive step, environmental contaminants are immobilized and con- types: carboxyls, lactones, phenols, and carbonyls (Boehm, 1994); ex-
centrated on the GAC; during oxidative treatment, contaminants are hibit high oxygen content, cation exchange capacity, and are unreactive
transformed by hydroxyl radicals (•OH) formed through the Fenton with H2O2 (Kinoshita, 1988). ASOs increase the surface polarity of GAC
mechanism and are oxidized on or near the GAC surfaces. Mass transfer resulting in a decrease in the carbon surface affinity for some com-
of contaminants from the aqueous phase to the GAC in this manner pounds. Basic surface oxide (BSO) functional groups, including
eliminates the need to carry out chemical oxidation of large volumes pyrone- and chromene-type structures, exhibit low oxygen content,
of water; consequently, treatment is directed towards a reduced vol- anion exchange capacity, and are reactive with H2O2. The distribution
ume of water in a GAC slurry involving enhanced rates of reaction and of ASOs and BSOs in GAC influences surface chemistry, contaminant ad-
treatment efficiency. The objectives of the treatment process are to sorption, and H2O2 reactivity (Morena-Castilla et al., 1995; Zeid et al.,
transform the contaminants into less toxic byproducts, increase the use- 1995; Choma et al., 1999; Karanfil and Kilduff, 1999).
ful life of the GAC, and reduce costs for GAC regeneration and water or The metals content, including iron and manganese, can vary signifi-
air treatment (Huling et al., 2005). cantly in the raw materials used to produce GAC. Post-activation acid
The raw materials used in GAC production, the activation process, washing treatment of the GAC may be carried out to reduce metals,
and post-activation treatment significantly alter GAC physical and ash, and other water-soluble metals from leaching out of the GAC
chemical characteristics that potentially impacts the adsorption oxida- when placed into service. Through these processes, the final GAC prod-
tion treatment process. Studies involving adsorption oxidation treat- uct exhibits a range in physical and chemical characteristics including
ment of GAC often do not differentiate the GAC type used in the particle size (i.e., mesh size), pH (DeSilva, 2001), metals content, surface
experiments and therefore, inherently assume the treatment process area, pore volume, pore size distribution, and surface functionality. Fur-
is GAC-independent. Further, the adsorption oxidation treatment of ther, the final GAC product exhibits significant variability in adsorptive
GAC (Huling et al., 2005; Huling et al., 2000a; Kommineni et al., 2003) properties. For example, five types of physical adsorption curves can
potentially involves many key factors in the treatment process including be described that result from variability in the microstructure of the
oxidant type, physical and chemical treatment conditions, oxidant, GAC (Chiang et al., 2001). The pore volume and specific surface area
method of oxidant activation, contaminant (Huling et al., 2005; Huling are still considered the most important factors regarding the adsorption
et al., 2000a; Kommineni et al., 2003; Huling and Hwang, 2010; of volatile organic compounds (VOCs) (Chiang et al., 2001). Overall, it is
Anfruns et al., 2013; Kan and Huling, 2009; Huling et al., 2009; Huling likely that these physical and chemical characteristics directly or indi-
et al., 2012; Kim et al., 2015; De Las Casas et al., 2006; Chiu et al., rectly impact the adsorption oxidation treatment process. Thus, it is
2012; Liu et al., 2018; Toledo et al., 2003; He et al., 2017; Bach and proposed that a key factor in the success of the adsorption oxidation
Semiat, 2011; Liang and Chen, 2010; Huling et al., 2011; Hutson et al., treatment technology is functionally dependent on GAC selection.
2012; Horng and Tseng, 2008). However, the role of GAC type has
never been investigated and a comprehensive study does not exist in- 1.3. Oxidative treatment overview
volving GAC raw materials, their physical and chemical characteristics,
and adsorption oxidation results. Currently, it is unclear whether GAC H2O2-driven processes serve as the basis for different oxidative
behaves similarly, or whether adsorption oxidation treatment is GAC- treatment applications including the chemical oxidation regeneration
dependent. of contaminant-spent GAC (Georgi and Kopinke, 2005; Huling et al.,
The purpose of this study was to systematically measure the chem- 2005; Huling et al., 2009; Huling et al., 2007). Oxidative systems based
ical and physical characteristics of GAC that potentially influence the on iron (Fe)-activated hydrogen peroxide (Fe-AHP) (R1-R3; Table 1)
treatment process, to assess whether treatment is GAC-dependent, generate •OH, which are powerful and indiscriminate free radicals that
and if so, to identify top GAC candidates for the adsorption oxidation react with a wide array of environmentally relevant contaminants
treatment process. Here, a range of commercially available GACs (n = (Buxton et al., 1988; Haag and Yao, 1992). Naturally occurring Fe pres-
31) consisting of several raw materials (bituminous coal, lignite coal, co- ent in the GAC matrix, amended Fe present as iron oxides fixed to the
conut shell, wood, peat) and particle sizes were tested in the adsorption GAC surfaces, and/or soluble or colloidal iron in solution react with
oxidation treatment of 2-chlorophenol (2CP) using activated hydrogen H2O2 in the Fe-AHP mechanism (R4a, R4b; Table 1).
peroxide (H2O2). A subset (n = 12) of the original GACs were selected Oxidative transformation of two environmentally relevant and
for further testing of methyl tert-butyl ether (MTBE) adsorption structurally distinct probe compounds was investigated in this study;
oxidation. 2CP and MTBE. 2CP (R5a; Table 1) and MTBE (R5b; Table 1) react with
•OH at high rates (Buxton et al., 1988; Getoff and Solar, 1986); are rela-
1.2. Variability in physical and chemical characteristics of GAC tively hydrophilic, poorly volatile, and exhibit different sorption affini-
ties towards GAC (Supporting Information Table S.1) (US EPA, 2008;
GAC is produced from various raw materials including bituminous Liu and Chang, 1997).
coal, lignite coal, coconut shells, wood, peat, sawdust, and nut shells Inefficiency in the oxidative treatment processes is attributed to two
(Chiang et al., 2001). The carbon materials are composed of particles main mechanisms, radical scavenging and nonproductive reactions
consisting of small, irregularly ordered layer stacks (i.e., graphene layers (NPR). The rates of •OH reaction with non-target chemical species
that exhibit a preferential orientation parallel to that of the particle sur- (i.e., scavengers, Si) in the aqueous phase have been determined for
face) (Boehm, 1994). Carbon particles are activated either by high some aqueous systems (Kwan and Volker, 2003; Grebel et al., 2010;
K.R. Crincoli et al. / Science of the Total Environment 734 (2020) 139435 3

Table 1 to remove fines, dried in an oven (105 °C), and stored in a desiccator
H2O2 activation and related reactions. until used. All GAC samples were subjected to a 2CP adsorption oxida-
H2O2 + Fe2+ → Fe3+ + OH- + •OH k1=76 1/M×s Walling, 1975 R1 tion test where 2CP was adsorbed and subsequently oxidized by the ad-
H2O2 + Fe3+ → Fe2+ + •O-2 + 2H+ R2 dition of H2O2. A subset (n = 12) of the 31 original GACs used in the 2CP
•O-2 + Fe3+ → Fe2+ + O2 R3 oxidation tests were selected and used in another adsorption oxidation
H2O2 + ≡Fen-oxide → ≡Fen+1-oxide + R
test where Fe, MTBE, and H2O2 were amended to the GAC to assess
•OH + OH- 4a
H2O2 + ≡Fen+1-oxide → ≡Fen-oxide + R MTBE adsorption oxidation.
HO2• + H+ 4b
10
•OH + 2CP → products k5a = 1.2×10 1/M×s Getoff and R 2.3. 2CP adsorption oxidation screening test
Solar, 1986 5a
•OH + MTBE → products k5b = 1.6×109 1/M×s Buxton et al., R
1988 5b
The 2CP screening test involved the adsorption and oxidation of 2CP
•OH + Σni=1 [Si] → products R6 under optimal pH conditions (pH 3.0) (Pignatello et al., 2006). Repre-
•OH + H2O2 → HO2•- + H2O k7 = 2.7×107 1/M×s Buxton et al., R 7 sentative GAC samples (1 g) collected using a soil riffle were placed in
1988 a reactor vessel (125 mL Erlenmeyer flask) and amended with an
Mn- and some Fe-oxides
H2SO4 solution (28.4 mM) to acidify the GAC suspension (pH 3.0;
2H2O2 → O2 + 2H2O R8
C+]OH- + HOOH → C+]OOH- + H2O Kinoshita, 1988 R9 24 h). The final pH was measured, and the acidic solution was decanted.
C+]OOH- + H2O2 → C+]OH- + H2O + Kinoshita, 1988 R The 2CP stock solution was analyzed (n = 3) and added to the acidified
O2 10 GAC in the flask reactor (40 mL; 5.0–5.8 mM). Flasks were wrapped in
Note: Σni=1 [Si] is the pseudo-first-order rate constant (kS 1/s) for •OH scavenging by all aluminum foil, labeled, para-filmed, and placed on an orbital shaker
constituents in solution except the target compound. (N24 h) to allow 2CP adsorption. The post-sorption 2CP solution was
sampled and analyzed (n = 2) for 2CP and anions (Cl−, Br−). The source
Rusevova et al., 2012). These calculations use second-order rate con- of Cl− could be released from the oxidation of 2CP and/or the GAC
stants (Buxton et al., 1988; Haag and Yao, 1992) and scavenger concen- (Huling et al., 2000b); the source of Br− could be released from the ox-
trations ([Si]). Si can be naturally occurring (e.g., HCO− 2−
3 , CO3 ), and can idation of the GAC. The remaining solution was decanted, and oxidation
be from anthropogenic sources (e.g. H2O2) (R6-R7; Table 1). •OH scav- was accomplished by amending the GAC with H2O2 (50 mL; 20 g/L).
enging by solid surfaces has only recently been measured for some min- Aqueous samples were collected over time to determine the H2O2 deg-
erals (alumina, silica, montmorillonite) (Rusevova Crincoli and Huling, radation rate (5 ≤ n ≤ 10); and analyzed for [H2O2] (n = 2). After the
2020). It is probable that GAC surfaces will also scavenge radicals and H2O2 was fully reacted, the spent H2O2 solution was analyzed for anions
treatment will be GAC-dependent. Therefore, oxidation results may em- (Cl−, Br−) and the pH was measured. 2CP-free control reactors were
pirically reflect GAC-dependent radical scavenging. Manganese and prepared for each GAC type (1 g GAC) and amended with H2O2. These
some iron oxides are involved in NPR and deplete the source of oxidant 2CP-free controls were used to quantify Cl− and Br− released from
(i.e., H2O2) without formation of •OH. These reactions diminish radical the oxidation of GAC. Background Cl− measured in the 2CP-free GAC
production and cause treatment inefficiency (R8; Table 1). BSO func- was used to correct the total Cl− measured in the post-oxidation 2CP-
tional groups (C+]OH−) involving H2O2 disassociation reactions (R9- amended GAC sample. 2CP oxidation evaluation was based on a Cl−
R10; Table 1) (Kinoshita, 1988) also cause NPR and contribute to treat- mass balance (1 mol Cl− /mol 2CP) involving the recovery of Cl− re-
ment inefficiency. leased from the oxidation of 2CP in the GAC suspension relative to the
total amount of Cl− amended to the GAC as 2CP (Eq. 1). Minor amounts
2. Methods, materials, and analytical of Cl− were measured in the post-sorption solution phase. Conse-
quently, the post-sorption mass of Cl− amended was adjusted for the
Overall GAC characterization and testing summary: GAC (n = 31) N2 pre-treatment Cl− released during the sorption phase.
BET surface area, pore volume distribution; average concentration of  
metals (Fe, Mn, Cu) and bromide in GAC; 2CP adsorption and oxidation MCl−;POST OXID:;2CP−AMENDED ; −; MCl− ;POST OXID:;2CP−FREE ð100Þ
2CP Loss ð%Þ ¼  
treatment, H2O2 reaction rate, and Br− residuals; GAC (n = 12) func- M Cl− ;PRE OXID:;2CP−AMENDED − MCl− ;POST SORP:;2CP−AMENDED
tionality measurements (acidic and basic functional groups), MTBE ad- ð1Þ
sorption and oxidation treatment.
where,
2.1. Analytical methods MCl-, POST OXID., 2CP-AMENDED = (VH2O2 × [2CP]POST OX × L/
1000 mL × SF)/MW2CP.
The analytical methods used to characterize GAC (surface area and MCl-, POST OXID., 2CP-FREE = (VH2O2 × [Cl−] × L/1000 mL)/MW− Cl
pore volume, metals and nonmetals content, surface functional groups) MCl-, PRE OXID., 2CP-AMENDED = V2CP × [2CP]INITIAL × MW2CP × SF
and to measure oxidation performance (i.e., 2CP, MTBE, H2O2, iron, an- MCl-, POST SORP., 2CP-AMENDED = V2CP × [Cl−]POST SORPTION × MW− Cl × SF
ions [i.e., Cl−, Br−] and pH) are documented in the Supporting Informa- VH2O2 = 50 mL
tion (S.1 Analytical Methods). All analytical measurements included V2CP = 40 mL
duplicate analyses. [2CP]POST OX = Post oxidation concentration of 2CP (mg/L)
[Cl−] POST SORPTION = Post 2CP sorption [Cl−], before oxidation
2.2. GAC samples (mg/L)
MW2CP = 228.6 (mg 2CP/mmol)
Thirty-one (31) commercially available GACs were acquired from MW− −
Cl = 35.45 (mg Cl /mmol)
different manufacturers (Table 2). These GACs were manufactured for SF (stoichiometric factor) = 1 mmol Cl−/1 mmol 2CP
water treatment applications, derived from a range of raw materials,
and activated and processed under various conditions. The GACs were 2.4. MTBE oxidation screening test
analyzed for physicochemical parameters and used in screening tests
to evaluate their role in the adsorption oxidation treatment process. An MTBE oxidation screening test was conducted on a subset of GAC
Analyses were performed to establish baseline characteristics of each samples (n = 12) selected from the original set of GAC samples (n =
GAC and to potentially correlate results with adsorption oxidation re- 31). A ranking method was developed to select GAC samples that exhib-
sults. GAC samples (200 g) were rinsed with deionized water (DIW) ited characteristics favorable to adsorption oxidation treatment.
4 K.R. Crincoli et al. / Science of the Total Environment 734 (2020) 139435

Table 2
Granular activated carbon types1, mesh size, raw materials, activation, and general information; , , , peat raw materials.
Manufacturer Mesh Activation
GAC Type Size Method Raw Materials Handling Apparent Density
Calgon Carbon
CGB 8×30 Steam select grades of coal 0.49 g/mL
CGB MOD1 8×30 Steam select grades of coal 0.49 g/mL
CAX 8×30 Steam coconut shell acid washed 0.453 g/mL
BCP MOD12 8×30 Steam bituminous coal 0.366 g/mL
URV-MOD12 8×30 Steam bituminous coal 0.49 g/mL
Filtrasorb 400 12×40 Steam bituminous coal 0.54g/mL, tamped
Filtrasorb 600 12×40 Steam bituminous coal 0.62g/mL, tamped
CarboChem
LQ-900S 8×30 Steam bituminous coal 0.50 g/mL
LQ-1240 12×40 Steam bituminous coal 0.50 g/mL

MeadWestvaco 14×35 Steam wood 0.24-0.30 g/mL


Nuchar WV-BV 20×50 Steam wood 0.26-0.32 g/mL
NucharWV-BV 1500
Bio-Nuchar 120 14×18 Steam wood 0.21-0.28 g/mL
Nuchar RGC 40 12×40 Steam wood 0.23-0.30 g/mL
Norit Americas
C Granular 8×30 Steam + H3PO4 wood 0.23 g/mL
Darco 12×40 12×40 Steam lignite coal HCl washed 0.40 g/mL
HydroDarco 3000 8×30 Steam lignite coal 0.38 g/mL
HydroDarco 4000 10×30 Steam lignite coal 0.38 g/mL
Norit GAC 830 8×30 Steam bituminous coal
Norit GAC 830 Plus 8×40 Steam bituminous coal acid washed 0.51 g/mL
Norit GAC 1240 12×40 Steam bituminous coal 0.5 g/mL
Norit GAC 1240 Plus 12×40 Steam bituminous coal HCl washed 0.5 g/mL
Norit ROX 0.8 Extruded Steam peat HCl washed particle size < 60 mm;
0.2-0.6 g/mL
Waterlink-
Barnebey Sutcliffe
3007 20×50 Steam coconut shell HNO3 washed 3
3034 12×30 Steam coconut shell acid washed 3
4040 12×30 Steam coconut shell acid washed 3
HR5-AW 12×40 Steam coconut shell acid washed 3
206C-AW 12×40 Steam bituminous coal acid washed 0.48 g/mL 3
207A-AW 12×40 Steam bituminous coal acid washed 0.50 g/mL 4
DCL 12×40 Steam wood acid washed 4
HI PUR 8×30 Steam bituminous coal acid washed 0.48 g/mL 4
HI PUR PLUS 8×30 Steam bituminous coal acid washed 0.48 g/mL 4
1
Some GACs have been discontinued by the manufacturer; replaced by other GAC;
2
Specialty carbon;
3
The extent of steam activation and acid/water washing varied between coconut carbons producing a range in ash
content and/or activity;
4
The extent of steam activation, acid washing, and screening varies between coal carbons producing a range in ash
content and/or pore volume distribution.

Parameters included surface area, meso+macro pore volume, [Fe]GAC, The MTBE solution (0.12 L; 1.83 mg/L) was added to the Fe-amended
H2O2 half-life, post-sorption and post-oxidation [2CP]. The Cl− released GAC (3 d) for the MTBE adsorption step. The post-sorption solution
was used as an indicator of 2CP transformation. The MTBE oxidation test was sampled, analyzed for MTBE (n = 2), and the pH was adjusted to
involved pre-oxidation iron-amendment to the GAC, followed by MTBE 3.0 with H2SO4. The H2O2 solution was amended to the Fe- and MTBE-
adsorption and oxidation (pH 3.0). The objectives were to assess the ex- amended GAC and the reactors were placed on an orbital shaker
tent to which the naturally occurring iron could be augmented and (100 rpm) until the H2O2 was fully reacted. The H2O2 dosage was
immobilized on the GAC, and to quantify the extent of MTBE adsorption based on cost for practicality (commercial 50% H2O2 cost ~ $0.34/L; den-
and oxidation. Post-oxidation MTBE residuals on the GAC were ex- sity of 50% H2O2 = 1.18 g/mL; density of 30% H2O2 = 1.11 g/mL), where
tracted and quantified, and results were used to quantify treatment per- $0.1 H2O2/lb. GAC ($0.22/kg GAC) was applied to the GAC in 3 sequen-
formance. Pre- and post-oxidation [Fe]GAC and N2 BET surface area were tial treatments using 30% H2O2 (1.91 mL 30% H2O2; 1.11 g/mL; [H2O2]INI-
also measured. This set of treatment objectives was selected to identify TIAL = 15.8 g/L). H2O2 re-amendment occurred after the H2O2 was fully
an optimal pool of GACs, widen the range of contaminants tested, im- reacted. The GAC suspension was maintained at pH 3.0 during
prove oxidative treatment via Fe amendment, utilize a method of con- oxidation.
taminant extraction and analysis for oxidative treatment performance The post-oxidation solution was sampled and analyzed for MTBE
evaluation, assess Fe retention efficiency, and the overall impact of (n = 2), anions (Cl−, Br−) and the remaining solution was decanted.
treatment on surface area. Subsamples (n = 2) of the treated GAC were collected and extracted
Representative GAC samples (5 g) were collected using a soil riffle, with methanol (MeOH) (20 mL) and the MeOH extract was analyzed
added to the test reactor (250 mL Erlenmeyer flask), and amended for MTBE. The remaining wet GAC was weighed, dried, and re-
with an Fe solution (15 mL; FeSO4•7H2O; 3.12 g/L as Fe; 55.9 mM). weighed to determine the moisture content and surface area (n = 2).
After sufficient contact time (3 d) the solution was decanted and FeT, Subsamples of the dried GAC were crushed and analyzed for metals
Br−, and Cl− were measured. Fe amendment in GAC using these proce- via ICP (n = 3). MTBE adsorption oxidation treatment performance
dures may result in Fe penetration and immobilization within a short was determined by percent loss from the initial MTBE immobilized on
transport distance into the GAC (Huling et al., 2009). Consequently, the GAC (Eqs. 2–4).
the reactive zone in the GAC where aggressive MTBE oxidation occurs  
may be within a narrow interval on the periphery of the GAC particle. MTBE Loss ð%Þ ¼ MMTBE;INITIAL –MMTBE;FINAL =MMTBE;INITIAL  100 ð2Þ
K.R. Crincoli et al. / Science of the Total Environment 734 (2020) 139435 5


MMTBE;INITIAL ¼ ½MTBESTOCK –½MTBEFINAL  VSTOCK ð3Þ respectively. Naturally occurring Fe associated with the GAC can acti-
vate H2O2, forming •OH, which is needed to oxidize contaminants on
 or near the surfaces of the GAC (R1-R4, R5; Table 1). Conversely,
MMTBE;FINAL ¼ ½MTBEEXTRACT  VEXTRACT =1000 ð4Þ lower concentrations of Mn in GAC are favorable to the oxidative treat-
ment process by limiting NPR (R8; Table 1), a well-known source of
where, treatment inefficiency.
MMTBE, INITIAL = pre-oxidation MTBE mass initially immobilized in
GAC (mg). 3.2. 2CP oxidation screening
MMTBE, FINAL = post-oxidation MTBE mass extracted from treated
GAC (mg). The [2CP]Initial amended to the GAC suspension ranged in concentra-
[MTBE]STOCK = MTBE concentration in stock solution amended to tion (4970–5770 μM) and the post-sorption [2CP]FINAL was b32 μM in all
GAC (mg/L). samples, resulting in N99% 2CP removal for all GAC (Table S.4). The mass
[MTBE]FINAL = post-sorption MTBE concentration (mg/L). of 2CP adsorbed to the GAC was used to estimate the initial mass of Cl−
VSTOCK = volume of stock solution amended to GAC (0.12 L). adsorbed. The percentage of Cl− released from the initial Cl− mass
[MTBE]EXTRACT = MTBE concentration in MeOH extract solution adsorbed was used as an indicator of 2CP transformation, and as a per-
amended to GAC (μg/L). formance metric to assess 2CP oxidation treatment. Fourteen (14) GAC
VEXTRACT = volume of MeOH used to extract post-oxidation GAC samples resulted in 0% release of the Cl− from the 2CP; and 8 that re-
(0.02 L). sulted in N5% Cl− release. The average percent loss of Cl− from 2CP at-
tributed to oxidative treatment for coal- (n = 19), wood- (n = 6),
3. Results and discussion coconut shell- (n = 5), and peat-based (n = 1) raw materials was 5.2,
4.6, 0.7, and 0%, respectively. It is noted, however, that transformation
3.1. GAC characterization of either the parent chlorinated compound or chlorinated byproducts
may proceed without simultaneous Cl− release. Further, complete Cl−
All GACs were activated using steam, and many of the GACs were release does not assure complete mineralization of the parent com-
acid washed to reduce metals leaching and ash content (Table 2). The pound as indicated by intermediates measured in such systems
BET surface area of the GAC ranged widely from the 650 m2/g (Huling et al., 2000a).
(HydroDarco 3000 and 4000, lignite-based) to 2200 m2/g (Nuchar The first-order H2O2 degradation rate constants (kH2O2) and H2O2
WV-B 1500, wood-based). The average BET surface area for coal (n = half-lives were determined using H2O2 concentrations measured during
19), wood (n = 6), coconut (n = 5), and peat (n = 1) raw materials the 2CP adsorption oxidation treatment (Table S.4). During oxidative
was 990, 1540, 1050, and 980 m2/g, respectively (i.e., as calculated treatment, the H2O2 undergoes intraparticle diffusion-reactive trans-
using values in Table S.2). In general, higher BET surface area of GAC re- port into the GAC. H2O2 reactants in the GAC slurry include transition
flects a greater potential for interactions with chemicals including con- metals (Fe, Mn, Cu), BSOs, and •OH. Ideally, the longer H2O2 persists
taminant adsorption, H2O2 reactivity, and anchoring the iron amended (i.e., lower kH2O2; longer H2O2 half-life), the greater the potential for
to the GAC. intraparticle H2O2 diffusive transport and penetration distance into
Similarly, a broad range in the GAC micropore volume and the GAC. This allows a larger intraparticle treatment zone critical in ox-
meso+macropore volume was measured (Table S.2). For example, the idative treatment of the spent GAC (Kan and Huling, 2009; Huling et al.,
ratio of micropore volume to the meso+macropore volume ranged 2012).
from 0.32 (HydroDarco 4000, lignite coal) to 11.4 (HR5-AW, coconut Post-oxidation solutions were analyzed for Br− in all post-oxidation
shell). The average total pore volume in GAC involving coal (n = 19), solutions and was elevated in several of the GAC slurries (Table S.4). Br−
wood (n = 6), coconut (n = 5), and peat based (n = 1) raw materials was measured as a precautionary step to assess whether residuals may
was 0.56, 0.91, 0.51, and 0.61 mL/g, respectively. The average ratio of introduce post-treatment disposal issues. Under oxidative conditions,
micropore volume to total pore volume (i.e., total pore volume = sum Br− residuals from the oxidized GAC may potentially form bromate, a
of micropore and meso+macropore volume) was greatest for coconut disinfectant byproduct regulated by the US EPA (maximum concentra-
(0.89, n = 5), followed by coal (0.69, n = 19), peat (0.6, n = 1), and tion limit = 0.01 mg/L). Although post-oxidation solutions were not an-
wood (0.48, n = 6). It is proposed that the ideal GAC used in the adsorp- alyzed for bromate, these data suggest that H2O2 treatment of GAC has
tion oxidation process exhibits a balance between micropores and the potential to produce this residual and should be monitored prior
meso+macropores depending on the molecular size of the contaminant to disposal of the regenerant solution.
undergoing adsorption. The surface area generally associated with mi- Overall, there does not appear to be a correlation between the H2O2
cropore volume is ideal for the adsorption of smaller molecules; while degradation rate and either the GAC BET surface area, the GAC pore vol-
larger pores are needed for larger molecules. A factor of 2.5× the kinetic ume, the transition metals concentration of the GAC, or the Cl− released
diameter of adsorbates is favorable for solute transport to inner pores from 2CP oxidation (Figs. S.1 – S.4). Further, a consistent lack in correla-
(Quinlivan et al., 2005). Given the atomic radii of 2CP and MTBE tion was observed between Cl− released and [Fe]T, micropore volume,
(Table S.1), these adsorbates would have access to the surface area meso+macropore volume, micropore/meso+macropore volume, sur-
within the micropore volume. Larger pores would also be favorable for face area, and post-sorption [2CP] (data not shown). The lack of correla-
intraparticle transport of Fe ions and H2O2 into the GAC particle and tion between these parameters and the H2O2 degradation rate and Cl−
thus beneficial to the oxidative treatment process. release suggests that multiple parameters collectively play a
The concentration of transition metals known to react with H2O2 compounding role in the reaction of H2O2 with GAC, and 2CP oxidation
was variable in the GAC (Table S.3). For example, low concentrations treatment efficiency. However, empirical observations involving these
of Fe (30 mg/kg) and Mn (5 mg/kg) were measured in C Granular, a treatment metrics and GAC parameters can be used to narrow down
steam and phosphoric acid activated wood material; while very high the broad range of commercially available GAC investigated here to
concentrations of Fe (21,650 mg/kg) and Mn (510 mg/kg) were mea- identify good candidates to use in the adsorption oxidation process.
sured in LQ-900S, a steam activated bituminous coal. The average [Fe]
GAC in coal (n = 19), coconut (n = 6), wood (n = 5), and peat based 3.3. GAC selection for further testing
(n = 1) raw materials was 3290, 860, 200, and 144 mg/kg, respectively;
the average [Mn]GAC in coal (n = 19), coconut (n = 6), wood (n = 5), GAC parameters and treatment metrics favorable to adsorption oxi-
and peat based (n = 1) raw materials was 54, 20, 6, and 9 mg/kg, dation were identified and used in an empirical ranking method. This
6 K.R. Crincoli et al. / Science of the Total Environment 734 (2020) 139435

method was used to identify the top 12 GACs selected for further testing by a factor of 1.5–7. In all cases, the total [Fe]GAC measured by ICP anal-
in the MTBE adsorption oxidation treatment process. The ranking pa- ysis was 2–3 orders of magnitude greater than the [Mn]GAC (Table S.3)
rameters were BET surface area, meso+macro pore volume, [Fe]GAC, indicating that Fe dominated the reaction with H2O2. The Fe-
H2O2 half-life, [2CP]FINAL, and Cl− released. The following criteria were amendment pre-treatment method has a potential disadvantage;
used in conjunction with the ranking parameters; high surface area while it limits the relative contribution of NPR attributed to Mn and
(i.e., reactive surfaces) that favor MTBE adsorption and intraparticle Fe BSOs, faster H2O2 reaction potentially limits H2O2 diffusive transport
immobilization; high meso+macro pore volume to accommodate and intraparticle penetration.
intraparticle transport of H2O2, Fe, and post-oxidation residuals; high The post-sorption aqueous phase concentrations of MTBE (Pre-Ox
naturally occurring [Fe]GAC to facilitate Fe activation of H2O2 and •OH [MTBE]AQ) reflect 91% removal of MTBE in CGB MOD 1 and Bio-
formation; long H2O2 half-lives allowing longer duration for H2O2 diffu- Nuchar 120, and 97–99% removal of MTBE in the remaining GACs
sive transport and intraparticle penetration distance; lower [2CP]FINAL to (Table S.6). Oxidative losses between the pre- and post-oxidation
achieve water treatment objectives; and high Cl− released as a general [MTBE]GAC indicated 1.1% to 57.9% removal of MTBE (Table S.6; Fig. 1).
indicator of overall 2CP oxidation. The top 12 GACs for these ranking pa- The average loss of MTBE attributed to oxidative treatment for coal-
rameters were assigned a value of 1; the ranking parameters were (n = 6), wood- (n = 5), and coconut shell-based (n = 1) raw materials
summed up for an overall ranking (Table S.5). The following 12 GACs was 23.4, 12.3, and 1.1%.
were selected for additional testing where BCP MOD1 and URV MOD1 The average post-oxidation reduction in BET N2 surface area was
were ranked the highest (5); the next highest grouping (4) was Nuchar greatest for wood GAC (13.5%), followed by coal GAC (9.9%), and coco-
WV-BV, Nuchar WV-BV 1500, Bionuchar 120, Nuchar RCG 40, and Norit nut shell GAC (2.8%) (Table S.5). Reduction in surface area could be at-
GAC 1240; followed by the next highest grouping (3) CGB, CGB MOD1, tributed to blockage of sorption sites or pore throats by the amended
CAX, C Granular, 207A-AW, and HIPUR. This selection process has limi- Fe (Huling et al., 2007), or oxidative effects (Huling et al., 2005;
tations, however: CGB was eliminated to broaden the scope of materials Karanfil and Kilduff, 1999).
tested; Norit GAC 1240 was selected because it exhibited 4 characteris-
tics favorable for adsorption oxidation yet no Cl− release was measured; 3.5. Role of GAC functional groups
Waterlink-Barneby Sutcliff 3007 exhibited only 2 characteristics favor-
able to adsorption oxidation, including 2.6% Cl− release, but was not se- The GACs used in the 2CP and MTBE adsorption oxidation tests ex-
lected for further testing. hibited a wide range in ASOs and BSOs (Fig. 2). The Norit Americas C
Granular exhibited the highest ASO content, and no BSOs were mea-
3.4. MTBE oxidation results sured (i.e., 0.0 μeq/m2). The hydrophilicity, or surface polarity of the
GAC increases with increasing ASOs, resulting in a decrease in contam-
Fe was amended to the GAC in a pre-treatment step to undermine inant adsorption (Quinlivan et al., 2005). This explains why C Granular,
the role of NPR caused by the naturally occurring BSO and Mn content exhibiting the greatest amount of ASOs, exhibited the lowest sorption
of the GAC. Augmenting the naturally occurring Fe enhances the reac- capacity of 2CP (i.e., highest [2CP]FINAL) (Table S.4). Similarly, correla-
tion between Fe and H2O2, limits NPR, and helps to limit treatment inef- tions between high ASOs and high [2CP]FINAL are also observed for
ficiency. The amended Fe immobilized in the GAC ([Fe]GAC ~ 900 mg/kg; CGB MOD1, Nuchar WV-BV, Nuchar WV-BV 1500 and Bio-Nuchar 120.
n = 12) increased the total [Fe]GAC over background [Fe]GAC (Table S.3) Given that BSOs were measured in all other GAC samples, measurement

Fig. 1. MTBE loss in GAC-specific Fe-AHP oxidative treatment system.


K.R. Crincoli et al. / Science of the Total Environment 734 (2020) 139435 7

Acidic or Basic Functional Groups (meq/100g)


2000

Acidic Functional Groups GAC Raw Materials


1600 Coal
Basic Functional Groups
Wood
Coconut
1200

800

400

Fig. 2. Acid surface oxide (ASO) and basic surface oxide (BSO) functional groups on specific GACs used in the 2CP adsorption oxidation screening test.

of 100% functionality as ASOs in C Granular was likely attributed to the for adsorption oxidation treatment. Coconut-based GACs are typically
phosphoric acid activation step. Further, C Granular exhibited the characterized as having greater micropore volume than
highest H2O2 half-life (111.8 h; Table S.4) and was attributed to the meso+macropore volume. The ratio of micropore volume and
lack of BSOs and the very low concentrations of Fe and Mn (Table S.3) meso+macropore volume in coconut-based GACs in this study ranged
resulting from phosphoric acid treatment. from 4.2 to 11.4 (Table S.5) indicating predominantly micropore vol-
General correlations can be made between the ASO/BSO ratio and ume. Micropore volume is unlikely to limit the intraparticle transport
H2O2 reaction, and the ASO/BSO ratio and fraction of Cl− released or of 2CP or MTBE due to molecular size (Table S.1; Section 3.1) However,
MTBE removed (Fig. 3). H2O2 reaction is generally correlated to the
BSO content of the GAC (Kinoshita, 1988). Consequently, the extent of 0.6 0.8
2CP oxidative treatment is inversely proportional to the BSO content
and was attributed to NPR. Prior to the oxidation of 2CP in the screening
study, the GAC was acidified (pH 3.0) which reduced the BSO content of 0.7
the GAC by converting some of the BSOs to ASOs (Jones, 2007). The acid- 0.5
ification step limited BSOs and NPR, and increased 2CP oxidative treat-
ment. Similarly, amendment of an acidic Fe solution, and the GAC 0.6
Fraction Cl- released, or MTBE Removed (%)

Fraction Cl- loss from 2CP


acidification step (pH 3.0), were also responsible for suppressing BSO-
related NPR and improving MTBE oxidative treatment. 0.4 Fraction MTBE Loss

kH2O2 (hr-1) 0.5


3.6. Selection of GAC for adsorption oxidation

A wide range of treatment was achieved in 2CP-spent GAC (n = 31; increasing ASOs increasing BSOs

kH2O2 (hr-1)
0.3 0.4
0–49.2% Cl− removal) (Table S.4), and MTBE-spent GAC (n = 12;
1.1–57.9% MTBE removal) (Fig. 1). While the results were variable, in
general, the coal-based raw materials used in GAC manufacturing re-
0.3
sulted in the best 2CP and MTBE oxidative treatment performance,
0.2
followed by wood, coconut, and peat materials. Three GACs produced
by steam activation of bituminous coal (CGB MOD1, BCP MOD1, URV
0.2
MOD1) resulted in the greatest release of Cl− from oxidative treatment
of 2CP (Table S.4) and the greatest removal of MTBE in 3 of the top 4
0.1
GACs tested (Fig. 1). However, it is noted that the lack of 2CP treatment
0.1
was also measured (i.e., 0% removal of Cl−) by all the raw materials used
to make GAC including bituminous and lignite coal, wood, coconut, and
peat (Table S.4). The screening results reported in this study do not nec-
0 0
essarily reflect the level of treatment that could be achieved under opti- 0.0 1.0 2.0 3.0 4.0 5.0 6.0
mized conditions. For example, optimal conditions may involve Fe
amendment conditions (Huling et al., 2007), regeneration temperature ASO/BSO Ratio
(Kan and Huling, 2009), Fe type (Huling and Hwang, 2010), particle size
Fig. 3. General correlation between the ratio of acidic and basic surface oxide (ASO, BSO)
(Huling et al., 2009) and additional applications of H2O2. functional groups in the GAC and (1) the pseudo-first order H2O2 degradation rate
2CP and MTBE removal in the coconut-based GAC was limited sug- constant (kH2O2), and (2) the oxidative treatment of 2CP measured as a percentage of
gesting GACs made from this raw material are generally less favorable Cl− released from 2CP, and percent MTBE removed.
8 K.R. Crincoli et al. / Science of the Total Environment 734 (2020) 139435

accumulation of Fe in the pore throats may limit transport of treatment An inverse correlation between the ASO content in GAC, and 2CP and
chemicals (i.e., H2O2) during oxidation (Huling et al., 2007). The lowest MTBE adsorption was observed (i.e., C Granular, CGB MOD1, Nuchar
MTBE removal was measured in CAX, and no Cl− was released in 4 of WV-BV, Nuchar WV-BV 1500 and Bio-Nuchar 120) and was attributed
the other 5 coconut-based GACs. Conversely, CGB MOD1, a coal-based to hydrophilicity of the GAC with increasing ASOs. A direct correlation
GAC exhibiting predominantly micropore volume (Table S.6), as did between H2O2 reaction and the BSO content of the GAC was established.
the coconut GACs, achieved the greatest treatment for both 2CP The NPR between BSOs and H2O2 were responsible for the inverse cor-
(Table S.5) and MTBE (Fig. 1). relation between BSOs and 2CP and MTBE treatment. These results sug-
In another study, both H2O2 reaction and MTBE oxidation were in- gest that multiple GAC-specific parameters played a collective and
versely correlated with GAC particle size and was attributed to shorter compounding role in the treatment process.
intraparticle diffusion transport distances for both H2O2 and MTBE Contaminant adsorption oxidation screening results were useful in
(Huling et al., 2009). In the current study, the greatest 2CP oxidation identifying top candidate GAC materials to be used in the treatment pro-
(i.e., Cl− released) (Table S.4) and MTBE loss (Fig. 1) occurred in GACs cess. The overall order of GAC materials successfully used in the adsorp-
exhibiting the largest particle sizes (8 × 30; avg. particle diam. tion oxidation treatment process is bituminous-based coal N wood N
1.49 mm) including CGB MOD1, BCP MOD1, URV MOD1, HiPur, and C coconut N peat. While the removal of 2CP and MTBE in coal- and
Granular; and (14 × 18; avg. particle diam. 1.2 mm) BioNuchar 120. wood-based GACs indicate that these raw materials are most effective,
While it is apparent that GAC particle size plays a key role in oxidative it is apparent that not all coal- or wood-based GACs are favorable. Con-
treatment when utilizing the same type of GAC; when contrasting oxi- sequently, site- and contaminant-specific screening of candidate GAC
dative treatment results between different GAC types, the effect of par- materials, using a similar testing approach, would provide a strong
ticle size was not found to be a key parameter. Rather, particle size basis for GAC selection. It is concluded that the GAC adsorption oxida-
played a secondary role relative to the collective and compounding ef- tion treatment technology is GAC-dependent, and that GAC selection
fects of multiple parameters. is an important decision and a key factor in the overall success that
Results indicate that chemical oxidation treatment of GAC is func- can be achieved.
tionally dependent on the type of GAC used in the treatment process.
In most studies involving adsorption oxidation treatment of GAC, the Notice
specific type of GAC has not been reported. Consequently, it is difficult
to gather information on the specific characteristics of GAC, compare re- The views expressed in this journal article are those of the authors
sults between studies, and to assess which GACs are favorable or unfa- and do not necessarily represent the views and policies of the U.S. Envi-
vorable to chemical oxidative treatment. Two GACs widely used for ronmental Protection Agency. The U.S. Environmental Protection
adsorption in water treatment processes includes Calgon Carbon Agency, through its Office of Research and Development, funded and
Filtrasorb 400 (F400) and Filtrasorb 600 (F600). These GACs are highly managed the research described here. Any mention of trade names or
effective in contaminant adsorption, and in this study achieved the low- commercial products does not constitute endorsement or recommen-
est post-sorption [2CP]FINAL (Table S.4). However, no Cl− was released dation for use.
during oxidative treatment of the 2CP-spent GAC (Table S.4). It is con-
cluded that contaminant adsorption and contaminant oxidation are
CRediT authorship contribution statement
not necessarily complimentary characteristics in the adsorption oxida-
tion treatment process. In trichloroethylene (TCE)-spent F400 GAC re-
Klara Rusevova Crincoli:Conceptualization, Methodology, Formal
generated using ozone and H2O2, or UV-light catalyzed H2O2, the rate
analysis, Investigation, Resources, Data curation, Writing - review &
of H2O2 reaction was rapid, and the quantity of oxidants required in
editing, Visualization, Supervision, Project administration.Patrick K.
the regeneration process was found to be much more than using homo-
Jones:Conceptualization, Methodology, Formal analysis, Investigation,
geneous advanced oxidation for TCE destruction (Mourand et al., 1995).
Resources, Data curation, Writing - review & editing, Visualization, Su-
These results are consistent with our study where H2O2 reaction in F400
pervision, Project administration.Scott G. Huling:Conceptualization,
and F600 was rapid, and 2CP oxidation was limited. Specifically, poor
Methodology, Formal analysis, Investigation, Resources, Data curation,
oxidative treatment in these GACs was attributed to NPR, and rapid
Writing - review & editing, Visualization, Supervision, Project
H2O2 reaction led to limited intraparticle penetration of H2O2 into the
administration.
GAC.

Declaration of competing interest


4. Conclusions
The authors declare that they have no known competing financial
A range of commercially available GACs (n = 31) composed of sev- interests or personal relationships that could have appeared to influ-
eral raw materials (bituminous coal, lignite coal, coconut shell, wood, ence the work reported in this paper.
peat) and particle sizes were tested in the adsorption oxidation treat-
ment of 2CP. Prior to treatment testing, physical and chemical charac- Acknowledgements
teristics of the GAC were measured indicating a wide range in surface
area, pore volume distribution, and transition metals (i.e., Fe and Mn). The Authors acknowledge Dr. B. Pivetz (CSS, Buffalo, NY) for scien-
Cl− recovery, used as the treatment performance metric for the 2CP ad- tific support.
sorption oxidation testing, indicated a wide range of treatment
(0–49.2% loss) where bituminous coal- and wood-based GAC per- Appendix A. Supplementary data
formed best. Subsequently, GAC parameters and treatment metrics fa-
vorable to adsorption oxidation were used in an empirical ranking Supplementary data to this article can be found online at https://doi.
method to identify the top 12 GACs selected for further testing involving org/10.1016/j.scitotenv.2020.139435.
MTBE adsorption oxidation. In these experiments, Fe was amended to
the GAC to limit the role of treatment inefficiency caused by NPR. A References
wide range in oxidation loss of MTBE in the GAC (1.1–57.9%) was mea-
Anfruns, A., Montes-Morán, M.A., Gonzalez-Olmos, R., Martin, M.J., 2013. H2O2-based ox-
sured where bituminous coal- and wood-based GAC also performed idation processes for the regeneration of activated carbons saturated with volatile or-
best. ganic compounds of different polarity. Chemosphere 91, 48–54.
K.R. Crincoli et al. / Science of the Total Environment 734 (2020) 139435 9

Bach, A., Semiat, R., 2011. The role of activated carbon as a catalyst in GAC/iron oxide/ Hutson, A., Ko, S., Huling, S.G., 2012. Persulfate regeneration of spent granular activated
H2O2 oxidation process. Desalination 273, 57–63. carbon – oxidative, acidic, and sulfur accumulation effects on MTBE sorption.
Boehm, H.P., 1994. Some aspects of the surface chemistry of carbon blacks and other car- Chemosphere 89, 1212–1223.
bons: review article. Carbon 32 (5), 759–769. Jones, P.K., 2007. Impact of Surface Functionality Modification on Fenton Regeneration of
Buxton, G.V., Greenstock, C.L., Helman, W.P., Ross, A.B., 1988. Critical review of rate con- Activated Carbon. M.S. Thesis. Univ. Oklahoma (62 pgs).
stants for reactions of hydrated electrons, hydrogen-atoms and hydroxyl radicals Kan, E., Huling, S.G., 2009. Effects of temperature and acidic pre-treatment on Fenton-
(•OH/O•−) in aqueous-solution. J. Phys. Chem. Ref. Data 17, 513–886. driven oxidation of MTBE-spent granular activated carbon. Environ. Sci. Technol. 43
Chiang, Y.C., Chiang, P.C., Chang, E.E., 2001. Effects of surface characteristics of activated (5), 1493–1499.
carbons on VOC adsorption. J. Environ. Eng. 127 (1), 54–62. Karanfil, T., Kilduff, J.E., 1999. Role of granular activated carbon surface chemistry on the
Chiu, C., Hristovski, K., Huling, S.G., Westerhoff, P., 2012. In-situ regeneration of phenol- adsorption of organic compounds. 1. Priority pollutants. Environ. Sci. Technol. 33
adsorbed granular activated carbon by iron oxide nanocatalyst. Wat. Res. 47 (4), (18), 3217–3224.
1596–1603. Kim, J.R., Huling, S.G., Kan, E., 2015. Adsorption and oxidative degradation of bisphenol A
Choma, J., Burakiewicz-Mortka, W., Jaroniec, M., Li, Z., Klinik, J., 1999. Monitoring changes on surface modified iron-amended activated carbon: effects of temperature on ad-
in surface and structural properties of porous carbons modified by different oxidizing sorption and Fenton oxidation. Chem. Eng. J. 262, 1260–1267.
agents. J. Col. Inter. Sci. 214 (2), 438–466. Kinoshita, K., 1988. Electrochemical behavior of carbon. Carbon: Electrochemical and
De Las Casas, C., Bishop, K., Bercik, L., Johnson, M., Potzler, M., Ela, W., Sáez, A.E., Huling, S., Physicochemical Properties. John Wiley and Sons, pp. 369–371.
Arnold, R., 2006. In-place regeneration of GAC using Fenton’s reagents. In: Clark, C., Kommineni, S., Ela, W.P., Arnold, R.G., Huling, S.G., Hester, B.J., Betterton, E.A., 2003.
Lindner, A. (Eds.), Innovative Approaches for the Remediation of Subsurface- NDMA treatment by sequential GAC adsorption and Fenton-driven destruction.
Contaminated Hazardous Waste Sites: Bridging Flask and Field Scales. ACS Sympo- J. Environ. Eng. Sci. 20 (4), 361–373.
sium Series 940, pp. 43–65. Kwan, W.P., Volker, B.M., 2003. Rates of hydroxyl radical generation and organic com-
De Silva, F.J., 2001. The issue of pH adjustment in acid-washed carbons. Water Condition- pound oxidation in mineral-catalyzed Fenton-like systems. Environ. Sci. Technol. 37
ing and Purification May 2001, pp. 40–44. (6), 1150–1158.
Georgi, A., Kopinke, F.-D., 2005. Interaction of adsorption and catalytic reactions in water Liang, C., Chen, Y.J., 2010. Evaluation of activated carbon for remediating benzene contam-
decontamination processes part I. Oxidation of organic contaminants with hydrogen ination: adsorption and oxidative regeneration. J. Haz. Mat. 182, 544–551.
peroxide catalyzed by activated carbon. J. Appl. Cat. B 58 (1–2), 9–18. Liu, J.C., Chang, P.S., 1997. Solubility and adsorption behaviors of chlorophenols in the
Getoff, N., Solar, S., 1986. Radiolysis and pulse radiolysis of chlorinated phenols in aqueous presence of surfactants. Water Sci. Technol. 35, 123–130.
solutions. Int. J. Radiation Applications and Instrumentation, Part C, Radiation, Phys. Liu, Z., Meng, H., Zhang, H., Cao, J., Zhou, K., Lian, J., 2018. Highly efficient degradation of
and Chem., Oxford, UK 28 (5–6), 443–450. phenol wastewater by microwave induced H2O2-CuOx/GAC catalytic oxidation pro-
Grebel, J.E., Pignatello, J.J., Mitch, W.A., 2010. Effect of halide ions and carbonates on or- cess. Sep. Purif. Technol. 193, 49–57.
ganic contaminant degradation by hydroxyl radical-based advanced oxidation pro- Morena-Castilla, C., Ferro-Garcia, M.A., Joly, J.P., Bautista-Toledo, I., Carrasco-Marin, F.,
cesses in saline waters. Environ. Sci. Technol. 44 (17), 6822–6828. Rivera-Utrilla, J., 1995. Activated carbon surface modifications by nitric acid, hydro-
Haag, W.R., Yao, C.C.D., 1992. Rate constants for reaction of hydroxyl radicals with several gen peroxide, and ammonium persulfate treatments. Langmuir 11 (11), 4386–4392.
drinking water contaminants. Environ. Sci. Technol. 26 (5), 1005–1013. Mourand, J.T., Crittenden, J.C., Hand, D.W., Perram, D.L., Notthakun, S., 1995. Regeneration
He, X., Elkouz, M., Inyang, M., Dickenson, E., Wert, E.C., 2017. Ozone regeneration of gran- of spent adsorbents using homogeneous advanced oxidation. Water Environ. Res. 67
ular activated carbon for trihalomethane control. J. Haz. Mat. 326, 101–109. (3), 355–363.
Horng, R., Tseng, L.C., 2008. Regeneration of granular activated carbon saturated with ac- Pignatello, J.J., Oliveros, E., MacKay, A., 2006. Advanced oxidation processes for organic
etone and isopropyl alcohol via a recirculation process under H2O2/UV oxidation. contaminant destruction based on the Fenton reaction and related chemistry. Crit.
J. Haz. Mat. 154, 366–372. Rev. Environ. Sci. Technol. 36 (1), 1–83.
Huling, S.G., Hwang, S., 2010. Iron amendment and Fenton oxidation of MTBE-spent gran- Quinlivan, P.A., Li, L., Knappe, D.R.U., 2005. Effects of activated carbon characteristics on
ular activated carbon. Wat. Res. 44 (8), 2663–2671. the simultaneous adsorption of aqueous organic micropollutants and natural organic
Huling, S.G., Arnold, R.G., Sierka, R.A., Jones, P.K., Fine, D., 2000a. Contaminant adsorption matter. Water Res. 39, 1663–1673.
and oxidation via Fenton reaction. J. Environ. Eng. 126 (7), 595–600. Rusevova Crincoli, K., Huling, S.G., 2020. Hydroxyl scavenging rate constants for solid
Huling, S.G., Arnold, R.G., Sierka, R.A., Jones, P.K., Fine, D., 2000b. Contaminant adsorption phase mineral surfaces in oxidative treatment systems. Wat. Res. 169, 115240.
and oxidation via Fenton reaction. J. Environ. Engin. 126 (7), 595–600. Rusevova, K., Kopinke, F.D., Georgi, A., 2012. Nano-sized magnetic iron oxides as catalysts
Huling, S.G., Jones, P.K., Ela, W.P., Arnold, R.G., 2005. Fenton-driven chemical regeneration for heterogeneous Fenton-like reactions - influence of Fe(II)/Fe(III) ratio on catalytic
of MTBE-spent granular activated carbon. Wat. Res. 39, 2145–2153. performance. J. Hazard. Mater. 241-242, 433–440.
Huling, S.G., Jones, K.P., Lee, T., 2007. Iron optimization for Fenton-driven oxidation of Toledo, L.C., Silva, A.C.B., Augusti, R., Lago, R.M., 2003. Application of Fenton’s reagent to
MTBE-spent granular activated carbon. Environ. Sci. Technol. 41 (11), 4090–4096. regenerate activated carbon saturated with organochloro compounds. Chemosphere
Huling, S.G., Kan, E., Wingo, C., 2009. Fenton-driven regeneration of MTBE-spent granular 50, 1049–1054.
activated carbon - effects of particle size and iron amendment procedures. J. Appl. US EPA, 2008. Regulatory Determinations Support Document for Selected Contaminants
Cat. B Environ. 89, 651–657. from the Second Drinking Water Contaminant Candidate List, OGWDW, EPA Report
Huling, S.G., Ko, S., Park, S., Kan, E., 2011. Persulfate-driven oxidation of MTBE- and 815-R-08-012.
chloroform-spent granular activated carbon. J. Haz. Mat. 192 (3), 1484–1490. Walling, C., 1975. Fenton’s reagent revisited. Acc. Chem. Res. 8, 125–131.
Huling, S.G., Kan, E., Wingo, C., Park, S., 2012. Fenton-driven regeneration of MTBE-spent Zeid, N.A., Nakhla, G., Farooq, S., Osei-Twum, E., 1995. Activated carbon adsorption in ox-
granular activated carbon – a pilot study. J. Haz. Mat. 205–206, 55–62. idizing environments. Water Res. 29 (2), 653–660.

You might also like