You are on page 1of 10

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

pubs.acs.org/est

Quantification of Phenolic Antioxidant Moieties in Dissolved Organic


Matter by Flow-Injection Analysis with Electrochemical Detection
Nicolas Walpen, Martin H. Schroth, and Michael Sander*
Institute of Biogeochemistry and Pollutant Dynamics (IBP), Department of Environmental Systems Science, ETH Zurich, Zurich,
Switzerland 8092
*
S Supporting Information

ABSTRACT: Phenolic moieties in dissolved organic matter


(DOM) play important roles as antioxidants in oxidation
processes in natural and engineered systems. This work
presents an automated and highly sensitive flow injection
analysis (FIA) system coupled to both spectrophotometric and
electrochemical detection to quantify electron-donating
phenolic moieties in DOM by determining the number of
electrons that these moieties transfer to an added chemical
oxidant, the radical cation of 2,2′-azino-bis(3-ethylbenzothiazo-
line-6-sulfonic acid) (ABTS •+ ). The FIA system was
successfully validated using Trolox as a redox standard. Highest
method sensitivity was attained when combining the FIA with
chronoamperometric detection, resulting in limits of quantifi-
cation of picomolar amounts of Trolox and nanogram amounts
of DOM (corresponding to solutions with <1 mg carbon per liter). The analysis of DOM isolates showed a strong linear
correlation between the number of electrons donated and their titrated phenol contents, supporting oxidation of phenols by
ABTS•+. The broad application spectrum of the FIA system to dilute natural DOM samples was illustrated by analyzing water
samples collected from northern peatlands and by monitoring the oxidation of phenols in one peat sample upon incubation with
a phenol oxidase. The superior analytical capability of the FIA system allows quantifying phenols and monitoring phenol
dynamics in dilute DOM samples.

■ INTRODUCTION
Dissolved organic matter (DOM) is a complex mixture of low
particular research interest: carbon sequestration in these
systems is considered to be linked to low activities of phenol
molecular weight and macromolecular organic compounds that oxidases, leading to an accumulation of dissolved phenols that
originate from the incomplete degradation of plant and inhibit the activity of hydrolases and microorganisms.22−26
microbial precursors.1,2 Among the various functional groups Understanding the role of phenols in all of these processes
in DOM are phenolic moieties (i.e., mono- or polyhydroxylated requires analytical methods to quantify these moieties and to
benzene units).3−8 While their contributions to the acid−base monitor changes in their concentrations during reactions.
and the metal complexation chemistries of DOM have been Phenolic moieties in DOM can be directly quantified by
extensively studied, comparatively little is known about their potentiometric acid−base titrations of DOM solutions.27−30
electron-donating (= antioxidant) properties in oxidation However, this approach requires large amounts of DOM (gram
reactions. Yet, recent studies showed that the antioxidant quantities). Furthermore, because the approach is relatively
properties of phenolic moieties play key roles in both pollutant labor intensive and time-consuming, it is ill suited for a fast and
and biogeochemical redox reactions in engineered as well as automated analysis. An alternative approach are reduction-
natural systems.9−12 For instance, in water-treatment systems, based assays (reviewed in refs 31−34), which are widely used in
phenols in DOM consume chemical oxidants that are added to physiology, pharmacology, food chemistry, and biogeochemis-
the water to remove micropollutants and pathogens.13−16 In try.32−36 These assays quantify the reduction of an added
sunlit natural aquatic systems, phenolic moieties in DOM can chemical oxidant by phenols in the analyzed samples. The most
decelerate indirect photolysis of pollutants and biomolecules by commonly used assay uses the radical cation of 2,2′-azino-
donating electrons to radical cations of these molecules formed bis(3-ethylbenzothiazoline-6-sulfonic acid) (i.e., ABTS•+) as an
via DOM-sensitized photochemical oxidation, thereby reverting
the radical intermediates back to the parent compounds.17−21 Received: March 4, 2016
In natural systems, oxidative processing of DOM may result in Revised: May 17, 2016
a depletion of phenolic moieties.1,9,11 The oxidation of phenolic Accepted: May 26, 2016
moieties in DOM from northern peatlands has received Published: May 26, 2016

© 2016 American Chemical Society 6423 DOI: 10.1021/acs.est.6b01120


Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

oxidant, because it has a high water solubility and a standard Organic Matter Samples. Humic substances (HS) and
reduction potential (i.e., Eh0 (ABTS•+/ABTS) = 0.70 V) natural organic matter (NOM) samples were purchased from
sufficiently high to oxidize phenols. In these assays, ABTS•+ the International Humic Substances Society (IHSS, St. Paul,
is obtained through either chemical, enzymatic, or electro- MN) and included Suwanee River II Standard humic acid (HA)
chemical one-electron oxidation of ABTS.37−39 The reduction (SRHA), Elliot Soil Standard HA (ESHA), Pahokee Peat
of ABTS•+ to ABTS by an analyte can be quantified either Standard HA (PPHAS), Leonardite Standard HA (LHA),
electrochemically37,38 or spectrophotometrically39 as ABTS•+ Suwannee River I Standard fulvic acid (FA) (SRFA I),
has a strong absorbance in the red (λmax = 728 nm), whereas Suwannee River II Standard FA (SRFA II), Pahokee Peat
ABTS is colorless.40 Reference HA (PPHAR), Nordic Lake Reference HA (NLHA),
We recently used the ABTS•+/ABTS redox couple in Nordic Lake Reference FA (NLFA), Pony Lake Reference FA
mediated electrochemical oxidation (MEO), a constant- (PLFA), Suwannee River I Aquatic NOM (SRNOM I),
potential amperometric technique, to quantify phenolic Suwannee River II Aquatic NOM (SRNOM II), and Upper
moieties in DOM.8 In these measurements, the addition of Mississippi River Aquatic NOM (UMNOM).
DOM samples to electrochemical cells resulted in the reduction Stock solutions of HS and NOM (100 mg of HS or NOM
of preoxidized ABTS•+ to ABTS, which subsequently was L−1) were prepared by dissolving the respective material in 0.1
reoxidized at the working electrode (WE) of the cell to M phosphate buffer (pH 7), followed by filtration through
ABTS•+. Integration of the resulting oxidative current peaks syringe filters (polypropylene, 0.45 μm, Pall, Switzerland). The
directly provided the electron-donating capacity (EDC) of the HS/NOM samples were diluted 20- to 50-fold to 2−5 mg HS
DOM sample (i.e., the amount of electrons transferred per or NOM L−1 prior to analysis on the FIA system.
amount of DOM analyzed). The EDC of several humic Natural dissolved organic matter (DOM) samples were
substances correlated strongly with their titrated phenol collected in July 2014 from three peat bogs in Värmland,
contents, supporting the hypothesis that mainly phenolic Sweden (i.e., Lungsmossen (LM), Norra Romyren (NR), and
groups donated electrons to ABTS•+. Likstamossen (LK); see Figure S1 and Table S1 in the
While the EDC values of as little as a few micrograms of Supporting Information for bog locations and exact sampling
DOM could be quantified by MEO,6 the sensitivity of the coordinates). DOM samples were collected from the peat pore
method was insufficient to analyze dilute DOM samples (i.e., water (three locations per peat, 125 cm below the peat surface)
below approximately 10 mgC L−1). Furthermore, MEO relies and from open water pools in the peats (three to five pools per
on manual sample addition to the electrochemical cells with peat, 40 cm (NR) or 120 cm (LM and LK) below the water
little opportunity for automation. Finally, the MEO setup has surface). All samples were immediately syringe filtered
limited portability and hence cannot easily be used for (polypropylene, 0.45 μm), frozen, and stored at −25 °C until
measurements outside the laboratory. These limitations of analysis.
MEO may, in principle, be overcome by implementing the Solutions. All solutions were prepared in ultrapure water
reduction-based approach on a flow-injection analysis (FIA) (resistivity >18.2 MΩ cm) from a Barnstead NANOpure
platform. Such platforms have been successfully used to Diamond system. All solutions used in electrochemical
quantify antioxidants in other disciplines, including food measurements contained 0.1 M KCl. The FIA reagent and
chemistry.37−39,41−43 Compact, robust, and field-deployable carrier solutions (see details below) were pH-buffered with
FIA systems can be designed that allow for automated sample 0.001 M acetate (pH 5) and 0.1 M phosphate (pH 7),
analysis. respectively.
The objective of this work was to develop and validate a FIA Characterization of Organic Matter Samples. UV−
system for the accurate and precise quantification of antioxidant visible light-absorbance spectra of all samples were collected
phenolic moieties in dilute DOM samples. The system used from 200 to 800 nm on a Varian Cary 100 Bio. The
ABTS•+ as chemical oxidant and spectrophotometric and nonpurgable organic carbon (NPOC) was measured on a
electrochemical flow cells to detect the extents of ABTS•+ Shimadzu total organic carbon (TOC-L) analyzer calibrated
reduction. The system response was validated using Trolox, a with a TOC standard (Sigma-Aldrich, Switzerland). Titrated
redox standard with a known EDC, and using a series of model phenol contents for the IHSS HS/NOM samples were
DOM isolates with published phenol contents and EDC values obtained from the literature.29,30,44
that were previously quantified by MEO. Finally, the FIA MEO. Mediated electrochemical oxidation of SRHA was
system was used to quantify the EDC values of dilute DOM conducted in a glassy carbon cylinder (volume, 9 mL; Sigradur
samples collected from three ombrotrophic bogs and, for one of G, HTW, Germany) that served as the WE.45 The cylinder was
the peat samples, to monitor changes in the phenol contents of filled with pH 7 buffer (0.1 M phosphate and 0.1 M KCl) and
the peat DOM during incubation with a phenol oxidase. was polarized to Eh = +0.7 V. A platinum wire was used as a


counter electrode and separated from the WE compartment by
a porous glass frit. An Ag/AgCl reference electrode was used
MATERIALS AND METHODS (ALS, Japan). All potentials are referenced versus the standard
Chemicals. ABTS (≥98%), (±)-6-hydroxy-2,5,7,8-tetrame- hydrogen electrode (SHE). The EDC values were quantified by
thylchromane-2-carboxylic acid (Trolox, 97%), potassium integrating the oxidative current peaks according to
chloride (≥99%), acetic acid (≥99.8%), and potassium
phosphate dibasic (≥99%) were obtained from Sigma-Aldrich
(Switzerland). Potassium dihydrogen phosphate was obtained
EDC =
1
mSRHA
∫ FI dt (1)
from Merck (Switzerland). Laccase from Trametes versicolor was
from Fluka (activity, 22.4 U mg−1) and used as received. The where mSRHA is the mass of analyzed SRHA, I (A) is the
laccase stock solution (0.745 mg mL−1) was prepared in pH baseline-corrected current, and F (= 96 485 C mol−1) is the
4.75 acetate buffer (1 mM). Faraday constant.
6424 DOI: 10.1021/acs.est.6b01120
Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

Figure 1. Scheme of the flow-injection analysis (FIA) system for quantification of the electron-donating capacity (EDC) of dissolved organic matter
(DOM) samples. The system contains two separate lines for the reagent (i.e., the oxidant ABTS•+) solution (green syringe) and the carrier solution
(blue syringe). The DOM samples are injected into the carrier-solution stream using an injector valve with a 100 μL sample loop. A third syringe
(red syringe) is used to load the sample loop. Up to 10 samples can be automatically injected using a 11-port/10-position selector valve. Following
mixing of the reagent and carrier streams in a mixing tee, the solution is passed through a 10-m long knitted open-tubular reaction coil. Reduction of
the oxidant ABTS•+ to ABTS by electron-donating moieties in DOM (bottom left panel) are detected either via current responses in an
electrochemical flow cell (bottom, middle panel) operated in chronoamperometry mode or via absorbance loss at a wavelength of λ = 728 nm in an
optical flow cell (bottom, right panel). The absorbance loss reflects the reductive decolorization of ABTS•+ (absorbance in the red) to ABTS
(colorless).

Enzymatic Oxidation Experiments. One DOM sample reduction potential of Eh = +0.82 V. The counter and reference
from LM bog was thawed and split into six subsamples of 8 mL electrodes were the same as used in MEO of SRHA (see
each. The pH was adjusted to 4.75 by addition of concentrated above). The bulk electrolysis was terminated when 60% of the
acetate buffer (0.1 M acetate; final acetate concentration, 1 added ABTS was oxidized (i.e., CABTS•+/(CABTS + CABTS•+) =
mM). This pH falls within the pH range of highest laccase 0.6).
activity46 and was close to the pH of the original sample (pH The total volumetric flow rate (qV = qreagent + qcarrier) was set
4.0). Two subsamples were kept as negative controls (no to 90 μL min−1 unless indicated differently. A volumetric
addition of laccase). Laccase was added to four of the aliquots mixing ratio of qcarrier/qreagent = 5 was used in all analyses to
(final activity of 0.5 U mL−1; initial ratio of enzyme activity to minimize sample broadening that results from mixing the
NPOC (excluding the acetate buffer) of 13.4 U mgC−1 (with carrier and the reagent solutions. All samples were adjusted to
NPOC = 37.28 mgC L−1)). Two of these aliquots also received have the same pH and electrical conductivity as the carrier
ABTS (final concentration, 7.5 μM). All samples were solution prior to injection.
incubated on a horizontal shaker (300 rpm, 25 ± 1 °C) in Sample Selection and Injection. The carrier solution was
amber vials for 3 days. Aliquots of 150 μL were repeatedly delivered through an injector valve connected to a 10-position
withdrawn from the subsamples and mixed with phosphate selector valve (both Cetoni, Germany) used for automated
buffer (final concentration 0.1 M phosphate, pH 7, 0.1 M KCl) sample injection. The injection loop (volume, 100 μL) was
to inactivate the laccase.46 The quenched aliquots were stored filled with sample solutions using a third syringe pump (Cetoni,
at 4 °C until analysis on the FIA system (Eh = 0.71 V; pH 7). Germany). The samples were introduced to the carrier stream

■ FLOW-INJECTION ANALYSIS SYSTEM


General Setup. The FIA system setup is shown schemati-
by redirecting the carrier stream through the injection loop.
Mixing and Reaction. The carrier and reagent solutions
were combined in a mixing tee (Vici, Switzerland). The mixed
cally in Figure 1 and contains four units in sequential solution had a pH of 7 due to the higher buffering capacity of
alignment: solution delivery, sample selection and injection, the carrier than the reagent solution. Following the mixing tee,
mixing and reaction, and detection. the solution passed through a PTFE knitted open tubular
Solution Delivery. Reagent and carrier solutions were reaction coil (length, 10 m; Biotech, Sweden) that served to
continuously delivered through the system by computer- increase the reaction time in the system prior to detection.
controlled syringe pumps (Cetoni, Germany). The carrier Detection. Two complementary detection methods were
solution was buffered to pH 7 (0.1 M phosphate and 0.1 M used to quantify the extents of ABTS•+ reduction: absorbance
KCl). The reagent solution contained the oxidant ABTS•+ measurements in a spectroscopic flow cell and chronoamper-
(0.75 mM) and was buffered to pH 4 (1 mM acetate; 0.1 M ometry in an electrochemical flow cell. Spectrophotometric
KCl), because the stability of the ABTS•+ in solution was found detection was conducted in a Z-flow cell with a 1 cm path
to decrease with increasing pH (Supporting Information, length (FIAlab Instruments, Seattle, WA), an HL-2000 light
Figure S2). The reagent solution was obtained by bulk source, and an STS-Vis detector (both Ocean Optics, Dunedin,
electrolysis of the respective ABTS solution in a 25 mL glassy FL). We detected the reductive decolorization of ABTS•+ to
carbon WE cylinder (Sigradur G, HTW, Germany) at a the colorless ABTS. The absorbance was monitored at 728 nm,
6425 DOI: 10.1021/acs.est.6b01120
Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

Figure 2. (a) Change in the solution absorbance at 728 nm in the spectrophotometric detector resulting from triplicate injections of Trolox
standards with amounts ranging from 6 to 1 nmol. Inset: Linear increase in the amounts of ABTS•+ reduced, Δn ABTS•+, with increasing amounts of
injected Trolox, nTrolox. (b) Linear increase in the heights, h, of the negative absorbance peaks (shown in panel a) with increasing amounts of injected
Trolox, nTrolox, between 6 and 1 nmol. (c) Oxidative current responses in chronoamperometric detection resulting from triplicate injections of Trolox
standards with amounts ranging from 4.6 to 0.02 nmol. (d) Linear increase in the heights h of the oxidative current peaks (shown in panel c) with
increasing amounts of injected Trolox, nTrolox, between 4.6 and 0.02 nmol. LOQ and σ correspond to the limits of quantification and the
measurement noises, respectively.

the absorbance maximum of ABTS•+ (Supporting Information, analyzer (CH Instruments, Austin, TX). All reduction
Figure S3a). The EDC value of Trolox, a redox standard, was potentials (Eh) applied to the WE are reported against the
obtained by integrating the negative absorbance peak using standard hydrogen electrode. The chronoamperometric
OriginPro (version 9.1, OriginLab, Northhampton, MA) responses were analyzed using OriginPro. The EDCj value
according to (mmole‑ gC−1) of an organic matter sample j was obtained from
qV the heights of its oxidative current peak(s), following
1 A(728 nm)
EDCTrolox =
n Trolox l
∫ ε(728 nm)
dt
(2)
calibration of the detector with Trolox standards:

hj 1
where nTrolox (molTrolox) is the injected amount of Trolox, qV EDCj = EDCTrolox
(mL min−1) is the total volumetric flow rate, ε(728 nm) (= mj a Trolox (3)
14 000 M−1 cm−1) is the molar absorption coefficient of
ABTS•+ (Supporting Information, Figure S3b), l (cm) is the where hj (nA) is the height of the oxidative current peak for
path length in the Z-flow cell, and A(728 nm) is the measured sample j, mj (gC,j) is the injected amount of carbon in j
absorbance. determined by NPOC measurements (see above), aTrolox (nA
The electrochemical cross-flow cell consisted of a glassy mmolTrolox−1) is the slope of the linear calibration curve of
carbon WE, a steel tube counter electrode, and a Ag/AgCl oxidative current peak heights of Trolox standards versus the
reference electrode (all ALS, Japan) and was run in respective injected Trolox amounts, and EDCTrolox (= 2 mmole‑
chronoamperometry mode using a 630D electrochemical mmolTrolox−1) is the electron donating capacity of Trolox.
6426 DOI: 10.1021/acs.est.6b01120
Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

■ RESULTS AND DISCUSSION


Spectrophotometric Detection and Validation of the
current only changes if the ABTS•+/ABTS ratio changes). The
noise in the baseline current was σ = 0.39 nA (Supporting
Information, Figure S4d).
FIA System. The baseline absorbance A in the spectrophoto-
Injections of standard solutions with decreasing amounts of
metric cell at 728 nm was relatively constant over several hours
Trolox from 4.6 to 0.02 nmol resulted in decreasing heights of
at values of A = 1.07 (Figure 2a). Yet, the baseline absorbance
the oxidative current responses (Figure 2c). The highest Trolox
showed systematic oscillations that resulted in a relatively large
amount of 4.6 nmol was chosen such that the ABTS•+/ABTS
measurement noise of σ = 0.011 (see baseline in Figure 2a
ratio after reaction with Trolox did not decrease below 0.33
labeled “0 nmol” and Figure S4a in the Supporting molABTS•+ molABTS−1 and, therefore, that the solution Eh
Information). These oscillations possibly reflected variations decreased by at most ΔEh = 0.04 V from its initial value (i.e.,
in the volumetric mixing ratios of the reagent and carrier between 0.71 and 0.72 V; see the Supporting Information,
solutions in the mixing tee, which resulted in periodic changes Figure S5 for details). The heights of the oxidative current
in the absolute ABTS•+ concentration entering the spectro- peaks linearly increased with increasing injected amounts of
photometric cell. No attempts were made to eliminate these Trolox, nTrolox (Figure 2d). The linear response range covered
oscillations because they were absent in the more sensitive more than 2 orders of magnitude in nTrolox.
electrochemical detection and, therefore, did not interfere with We note that the measured current responses resulted from
the electrochemical quantification of EDC values (see below). the re-equilibration of the ABTS•+-ABTS redox couple only in
We validated the FIA system based on absorbance loss the solution volume that was in direct contact with the WE and
resulting from injections of six different Trolox standard not the entire solution volume in the detector. We
solutions. Trolox was chosen because it is rapidly and independently determined that for the electrochemical flow
completely oxidized to its quinone structure by ABTS•+ in a cell used in this work, approximately 15% of the ABTS formed
two-electron transfer reaction.8,47,48 The spectrophotometric during reaction of Trolox with ABTS•+ was reoxidized to
detection was used for validation, because, in contrast to the ABTS•+ while passing through the cell (Supporting Informa-
electrochemical detection, it allows quantifying the total tion, Figure S6). For this reason, a series of Trolox standards
amount of ABTS•+ molecules that were converted to ABTS was analyzed in all subsequent DOM analyses to calibrate the
(eq 2). Triplicate injections of decreasing amounts of Trolox electrochemical detection (eq 3).
from 6 to 1 nmol resulted in decreasing heights of the negative Determination of the Limits of Quantification. The
absorbance peaks (Figure 2a). Integration of the peaks yielded limits of quantification (LOQ) of the FIA method for both the
an EDC value of 2.06 ± 0.03 mole‑ molTrolox−1 (inset in Figure spectrophotometric and electrochemical detections were
2a), which was in very good agreement with the expected EDC determined based on the heights of negative absorbance and
value of 2 mole‑ molTrolox−1.8 These values correspond to a oxidative current peaks of Trolox standards, using a statistical
Trolox recovery of 103 ± 2%, thereby validating the FIA approach:49−51
system.
Chronoamperometric Detection. Electron transfer from LOQ = kσ /RF (4)
electron-donating molecules in the injected samples to ABTS•+ where k (= 10) is a constant, σ is the noise in the baseline
lowered the ratio of ABTS•+/ABTS and, as a consequence, the readings (i.e., σ = 0.011 and 0.39 nA for the spectrophotometric
Eh of the solution volume that contained the injected sample. and electrochemical detection, respectively), and RF is the
The decrease of the solution Eh relative to the constant Eh response factor (i.e., RF = 0.114 molTrolox−1 and 178.65 A
applied to the WE in the electrochemical detector resulted in molTrolox−1 for the spectrophotometric and electrochemical
current peak responses relative to the baseline current detection, respectively).
measured in the absence of injected samples. To maximize The LOQ of the FIA system with spectrophotometric
sensitivity, the WE was polarized to the open-circuit potential detection was approximately 980 pmol Trolox (= 1.96 nmol of
(OCP) that was measured for the ABTS•+/ABTS ratio in the transferred electrons). The LOQ for the electrochemical
mixed reagent and carrier solutions delivered to the electro- detection was 22 pmol Trolox (= 44 pmol of transferred
chemical cell at the onset of each FIA analysis. We note that the electrons) and thus approximately 50 times lower than for the
OCP values were between +0.71 V and +0.72 V for all spectroscopic detection. The much lower LOQ obtained by
experiments, demonstrating high reproducibility in the electro- chronoamperometry reflected the significantly smaller noise in
chemical oxidation of ABTS used to generate the ABTS•+ the electrochemical detection. Because of the superior
reagent. As expected, the baseline currents increased with sensitivity of the electrochemical detection, it was used for all
increasing difference between the Eh applied to the WE and the subsequent analyses.
OCP (Supporting Information, Figure S4b,c). Polarizing the Analysis of SRHA by FIA and MEO. We further validated
WE to the OCP values resulted in very low baseline currents of the FIA method by quantifying the EDC of Suwannee River
Ibaseline < 10 nA in the subsequent chronoamperometric humic acid (SRHA) as a model DOM. In a first step, we
detection. While the current baselines typically showed small systematically assessed the effect of volumetric flow rate (varied
drifts (<50 nA) toward more oxidative values over the course of between 90 and 270 μL min−1) on the current responses.
several hours of continuous analysis, these drifts were always These flow rates corresponded to reaction times in the FIA
baseline corrected and hence did not affect EDC quantification. system from 25 min down to 8.5 min, respectively. Volumetric
The drifts likely resulted from a slow loss of ABTS•+ in the flow rates above 270 μL min−1 were not tested because they
reagent solution (despite its low pH of 4) over time. resulted in very short reaction times and an unacceptably high
In contrast to the spectrophotometric detection, no baseline backpressure (>1 bar) in the system. Volumetric flow rates
oscillations were observed in the electrochemical detection, below 90 μL min−1 were omitted because they resulted in
suggesting that mixing of the reagent and carrier solutions broader current peaks and increased the intervals between
resulted in constant ratios of ABTS•+ to ABTS (note that the injections, thereby minimizing sample throughput. For each
6427 DOI: 10.1021/acs.est.6b01120
Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

Figure 3. (a) Oxidative current responses to injections of different amounts of Suwannee River Humic Acid (SRHA) into the flow injection analysis
(FIA) system at total volumetric flow rates of 270 μL min−1 (left; reaction time tReaction= 8.5 min)) and 90 μL min−1 (right; tReaction= 25 min). (b)
Electron donating capacity (EDC) values for SRHA obtained by FIA analysis at different flow rates (from 90 to 270 μL min−1) and by mediated
electrochemical oxidation (MEO). The EDC values are given in absolute values (left ordinate) and normalized to the EDC values obtained by MEO
for integration over 60 min. Inset: Overlaid baseline-corrected oxidative current responses in MEO to triplicate additions of 10 μg of SRHA. (c)
Oxidative current responses in the FIA system to injections of Trolox calibration standards (first five peaks) and SRHA (triplicate injections each) at
amounts between 600 and 5 ng of SRHA. (d) Linear correlation between the heights of the oxidative current peaks (determined from data in panel
c) and the amount of injected SRHA, mSRHA. LOQ and σ correspond to the limits of quantification and the measurement noise, respectively.

tested flow rate, four Trolox calibration standards with amounts Figure 3b also shows the baseline-corrected oxidative current
between 2 and 0.1 nmol were injected prior to SRHA. responses in MEO to triplicate injections of 10 μg of SRHA
The heights of the oxidative current peaks obtained from (inset) and the corresponding cumulative numbers of electrons
injections of SRHA solutions decreased linearly with decreasing transferred obtained by integrating of the current peaks up to
injected SRHA masses (from 400 to 300 and 200 ng) at all 60 min after SRHA injections. The good agreement of EDC
tested flow rates (shown for the highest and lowest tested flow values obtained by FIA and MEO confirmed accurate
rates of 270 and 90 μL min−1 in Figure 3a). The broadening of quantification of the EDC values of DOM by the FIA system.
the current peaks from 270 to 90 μL min−1 was due to Oxidation at a slightly higher Eh in the FIA than MEO systems
increased time for longitudinal dispersion of ABTS and ABTS•+ (i.e., + 0.71 to +0.72 V vs +0.70 V, respectively) likely
during advective transport through the FIA system. While contributed to the slightly faster reaction in FIA system
sharper current peaks are preferred because of their higher (reaction time, 25 min at 90 μL min−1) than in the MEO cell.
signal-to-noise ratios, the increase in EDC values of SRHA with The quantification of EDC values fundamentally differed in
decreasing flow rates (Figure 3b) demonstrated that electron the FIA and MEO systems in that ABTS•+ reduction is
transfer from electron donating phenolic moieties in SRHA to detected after it has occurred in the reaction coil whereas it is
ABTS•+ was incomplete at the higher tested flow rates. directly monitored over time in MEO. This difference results in
Conversely, the EDC value of SRHA obtained at the lowest very different shapes of the current responses in FIA and MEO.
flow rate was in good quantitative agreement with the EDC The current peaks in FIA were relatively narrow (peak width of
value of SRHA obtained by MEO (i.e., EDCFIA ≈ 0.96 Δt < 10 min) and symmetrical (Figure 3a,c). These peak
EDCMEO), which has until now been the state-of-the art shapes resulted from small longitudinal dispersion during
approach to determine EDC values of DOM. For comparison, advective transport of the samples through the FIA system.
6428 DOI: 10.1021/acs.est.6b01120
Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

Figure 4. (a) Electron donating capacities (EDC) of selected humic and fulvic acid (HA and FA) and natural organic matter (NOM) isolates used as
models for dissolved organic matter (DOM), as determined by flow injection analysis (EDCFIA; Eh= 0.71 V, pH 7) and mediated electrochemical
oxidation (EDCMEO; Eh= 0.73 V, pH 7). The EDCMEO data was taken from a previous publication.8 The brown and the blue symbols represent
DOM from terrestrial and aquatic systems, respectively. Error bars represent standard deviations of triplicate measurements. (b) EDC values of
model DOM isolates and DOM collected from three ombrotrophic bogs versus their specific UV absorbance values at 254 nm (SUVA254). The peat
DOM samples were collected from the peat pore water at 125 cm depth (dark green symbols) and from open water pools within the bog at 40 (NR)
or 120 cm (LK, LM) depth (light green symbols). Error bars represent standard deviations of triplicate measurements. (c) Comparison of EDC
values from tested DOM samples after normalization to SUVA254. The error bars for HS/NOM isolates (blue and brown bars) represent standard
deviations of triplicate measurements. For peat DOM sample data, the error bars represent the standard deviations among the EDC averages shown
in panel b (either n = 3 or 5 average EDC values, as specified on the data bars). (d) Changes in the EDC values of a selected peatland DOM (peat
pore water from LM) incubated at pH 4.75 without laccase (red symbols; control), with laccase (blue symbols) and with laccase and ABTS as
electron transfer mediator (green symbols). Shown also are selected oxidative current peak responses for samples from each of the incubations for
four selected incubation times. All EDC values were calibrated with five injections of Trolox standards. Error bars represent ranges of duplicate
incubations.

Conversely, the current responses in MEO were much wider current peaks resulted from Tolox injections and served to
(width of Δt = 60 min) and asymmetric, a fast initial increase in calibrate the system (from 2 to 0.092 nmol of Trolox). The
the current to maximum values was followed by prolonged, LOQ of SRHA was approximately 19 ngSRHA (calculated with σ
approximately exponential decays in the currents over time = 0.39 nA, and RF = 0.206 A gSRHA−1 in eq 4 (Figure 3d)),
(Figure 3b). The prolonged current decay challenges accurate which corresponds to an approximate NPOC concentration of
integration of small peaks that result from the analysis of small
0.4 mgC L−1 (using the measured carbon content of SRHA of
analyte amounts.
We estimated the LOQ of the FIA system for DOM by 0.53 gC gSRHA−1 and an injection volume of 100 μL). These low
triplicate injections of SRHA at amounts between 600 and 5 DOC values imply that the FIA system is sufficiently sensitive
ngSRHA (Figure 3c). Each of the triplicate injections resulted in to determine the EDC values of dilute DOM samples collected
nearly identical current responses, highlighting the high from natural systems. Such dilute DOM solutions cannot be
reproducibility of the FIA analysis. Note that the first five accurately analyzed using MEO. In fact, the addition of 600 ng
6429 DOI: 10.1021/acs.est.6b01120
Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

of SRHA into the MEO cell did not result in quantifiable ABTS by laccase (i.e., ABTS is a known substrate for laccase56
current responses (Supporting Information, Figure S7). and is oxidized by the copper atoms in the active site of the
FIA Analysis of a Diverse Set of DOM Samples. Figure enzyme57,58). The final decrease in EDC values in the presence
4a shows good agreement in the EDC values of 13 HS and of ABTS and laccase (i.e., 38%) was, however, only slightly
NOM isolates quantified by FIA (y-axis) and MEO (x-axis).8 larger than in the systems containing laccase but no ABTS,
This result demonstrates that the FIA method can be used to suggesting that most phenolic moieties in the DOM were
quantify EDC values of DOM from very different sources. The directly oxidizable in the active site of the enzyme.
EDC values of the HS/NOM determined by FIA also showed a The considerable EDC values that remained after laccase
strong linear correlation (R2 = 0.94) with their titrated phenol treatment likely resulted from the significantly lower solution
contents29,30,44 (Supporting Information, Figure S8a), strongly pH during incubation (pH 4.75) than in the FIA system for
supporting that phenolic moieties in DOM donated electrons EDC quantification (pH 7.0). Because increasing pH favors
to ABTS•+.8 phenol oxidation both thermodynamically and kinetically,59 the
The EDC values of the model DOM quantified by FIA are DOM likely contained a subpool of phenolic moieties that were
replotted in Figure 4b versus their specific UV absorbance oxidizable by ABTS•+ at pH 7 but not oxidizable by laccase or
values at 254 nm (SUVA254). This parameter is linearly ABTS•+ at pH 4.75. We note that the FIA analysis could not be
correlated to the reported aromaticity values of these DOM (R2 conducted at pH 4.75 because laccase from the incubation
= 0.97, Figure S8b). Combining this information with the linear samples would have oxidized ABTS at this pH and hence
correlation of EDC values to the phenol contents, Figure 4b,c interfered with the EDC analysis. Conversely, laccase is inactive
suggests that electron-donating phenolic moieties in DOM per at pH 7 and hence did not interfere with the EDC
unit aromaticity (i.e., SUVA254) decrease from aquatic to quantification at this pH.
terrestrial materials. This trend was previously reported and
explained by a depletion of phenolic moieties in the older, more
oxidatively processed terrestrial DOM than in the younger, less
■ IMPLICATIONS
This work presents a highly sensitive FIA system with
processed aquatic DOM.8,52 chronoamperometric detection for the automated quantifica-
DOM samples collected from the pore waters and open- tion of electron donating phenolic moieties in DOM. The very
water pools of three ombrotrophic bogs had comparatively high low LOQ of this method (i.e., around 20 ngDOM) allows
EDC values between 5.34 mmole‑ gC−1 and 7.14 mmole‑ gC−1 quantifying EDC values for natural samples with NPOC
(green points, Figure 4b). When normalized to SUVA254 values concentrations below 1 mgC L−1. As such, this work largely
(from 3.45 to 4.02 L mgC−1 m−1; see Table S2 for DOC and advances the analytical capabilities to characterize the redox
absorbance values used in the calculation of SUVA254 values), properties of DOM. The superior sensitivity of the FIA method
the peat EDC values were higher than those of the aquatic and compared to any previously published method is in large part
terrestrial HS/NOM isolates (Figure 4c). The concentrations due to using an electrochemical flow cell detector with low
of iron and sulfide in the peat samples were too small for these background current noise.
species to have had major contributions to the measured EDC The FIA system allows characterizing the redox properties of
values (Table S2, Supporting Information). The finding of DOM in dilute samples using only very small solution volumes.
comparatively high EDC values for peat DOM is consistent As such, this technique opens new possibilities to monitor
with the overall high contents of phenols in peatlands and their changes in the phenol contents of DOM in oxidation processes
low extents of oxidative transformation due to anoxic at high temporal and spatial resolution. We anticipate that
conditions that prevail in the permanently water-saturated future work will employ the FIA system to monitor changes in
pores of these systems.53−55 Interestingly, the lower SUVA254- the EDC values of DOM during its enzymatic, photochemical,
normalized EDC values of peat-pool than peat-pore water or chemical oxidation. Direct analytical access to the dynamics
possibly reflected more extensive oxidation of DOM in the oxic in DOM phenol pools will contribute to a more holistic
peat pools as compared with the anoxic peat-pore water. The understanding of their role in oxidation processes in both
analysis of peat DOM demonstrates that the FIA system allows natural and engineered systems.


quantifying EDC values of DOM from natural sources.
Furthermore, the data suggests that EDC measurements by ASSOCIATED CONTENT
FIA can be used to monitor the extents of oxidative processing
of DOM in natural systems. *
S Supporting Information

Enzymatic Oxidation of Phenols in Peatland DOM. To The Supporting Information is available free of charge on the
demonstrate that the FIA system can be used to monitor the ACS Publications website at DOI: 10.1021/acs.est.6b01120.
oxidation of phenolic moieties in DOM, we incubated one of Additional information and data on peat sampling
the peat pore-water samples from LM with laccase from locations and peat DOM properties, the FIA system,
Trametes versicolor and monitored the resulting change in EDC the FIA system, the redox properties of the ABTS•+/
values over 3 days. As expected, the EDC values of control ABTS couple, and the EDC values of model HS and
DOM samples (i.e., no laccase added) stayed approximately NOM isolates (PDF)


constant during the incubation (Figure 4d; red squares).
Conversely, the EDC values of the DOM incubated with AUTHOR INFORMATION
laccase decreased by nearly one-third over 2 days and thereafter
remained approximately constant (blue lines). Addition of Corresponding Author
minute amounts of ABTS to samples containing laccase *Phone: +41 (0)44 632-8314; fax: +41 (0)44 633-1122; e-mail:
facilitated the initial decrease in the EDC values of the DOM michael.sander@env.ethz.ch.
(green lines), consistent with mediated oxidation of the Notes
phenolic moieties by ABTS•+ formed via the oxidation of The authors declare no competing financial interest.
6430 DOI: 10.1021/acs.est.6b01120
Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

■ ACKNOWLEDGMENTS
The project was funded by the Swiss National Science
(17) Canonica, S.; Laubscher, H.-U. Inhibitory effect of dissolved
organic matter on triplet-induced oxidation of aquatic contaminants.
Photochemical & Photobiological Sciences 2008, 7, 547−551.
Foundation (Project 200020_159692). We thank ETH Zurich (18) Wenk, J.; Aeschbacher, M.; Salhi, E.; Canonica, S.; von Gunten,
for funding the FIA system via the Scientific Equipment U.; Sander, M. Chemical Oxidation of Dissolved Organic Matter by
Program. We further thank L. Klüpfel, K.H. Jacobson, P. Nauer Chlorine Dioxide, Chlorine, And Ozone: Effects on Its Optical and
(all ETH Zurich), M. Lau (IGB Berlin), D. Cervenka, and R. Antioxidant Properties. Environ. Sci. Technol. 2013, 47, 11147−11156.
Sander for support during DOM sample collection, L. Klüpfel (19) Janssen, E. M. L.; Erickson, P. R.; McNeill, K. Dual roles of
for helpful discussions, and Ashley Brown (Eawag) for the dissolved organic matter as sensitizer and quencher in the photo-
quantification of total iron in the peat DOM samples. oxidation of tryptophan. Environ. Sci. Technol. 2014, 48, 4916−4924.


(20) Wenk, J.; Canonica, S. Phenolic Antioxidants Inhibit the Triplet-
Induced Transformation of Anilines and Sulfonamide Antibiotics in
REFERENCES Aqueous Solution. Environ. Sci. Technol. 2012, 46, 5455−5462.
(1) Leenheer, J. A.; Croué, J.-P. Peer Reviewed: Characterizing (21) Wenk, J.; Aeschbacher, M.; Sander, M.; von Gunten, U.;
Aquatic Dissolved Organic Matter. Environ. Sci. Technol. 2003, 37, Canonica, S. Photosensitizing and Inhibitory Effects of Ozonated
18A−26A. Dissolved Organic Matter on Triplet-Induced Contaminant Trans-
(2) Bolan, N. S.; Adriano, D. C.; Kunhikrishnan, A.; James, T.; formation. Environ. Sci. Technol. 2015, 49, 8541−8549.
McDowell, R.; Senesi, N. Dissolved Organic Matter: Biogeochemistry, (22) Verhoeven, J. T. A.; Liefveld, W. M. The ecological significance
Dynamics, and Environmental Significance in Soils. Adv. Agron. 2011, of organochemical compounds in Sphagnum. Acta Bot. Neerl. 1997, 46,
110, 1−75. 117−130.
(3) Scott, D. T.; McKnight, D. M.; Blunt-Harris, E. L. Quinone (23) Mellegard, H.; Stalheim, T.; Hormazabal, V.; Granum, P. E.;
moieties act as electron acceptors in the reduction of humic substances Hardy, S. P. Antibacterial activity of sphagnum acid and other phenolic
by humics-reducing microorganisms. Environ. Sci. Technol. 1999, 33, compounds found in Sphagnum papillosum against food-borne
372. bacteria. Lett. Appl. Microbiol. 2009, 49, 85−90.
(4) Ratasuk, N.; Nanny, M. A. Characterization and Quantification of (24) Chiapusio, G.; Jassey, V. E. J.; Hussain, M. I.; Binet, P. Evidences
Reversible Redox Sites in Humic Substances. Environ. Sci. Technol. of Bryophyte Allelochemical Interactions: The Case of Sphagnum. In
2007, 41, 7844−7850. Allelopathy; Springer Berlin Heidelberg: Berlin, Heidelberg, Germany,
(5) Bauer, I.; Kappler, A. Rates and extent of reduction of Fe(III) 2013; pp 39−54.
compounds and O2 by humic substances. Environ. Sci. Technol. 2009, (25) Freeman, C.; Ostle, N.; Kang, H. An enzymic “latch” on a global
43, 4902−4908. carbon store - A shortage of oxygen locks up carbon in peatlands by
(6) Aeschbacher, M.; Sander, M.; Schwarzenbach, R. P. Novel restraining a single enzyme. Nature 2001, 409, 149−149.
Electrochemical Approach to Assess the Redox Properties of Humic (26) Freeman, C.; Ostle, N. J.; Fenner, N.; Kang, H. A regulatory role
Substances. Environ. Sci. Technol. 2010, 44, 87−93. for phenol oxidase during decomposition in peatlands. Soil Biol.
(7) Aeschbacher, M.; Vergari, D.; Schwarzenbach, R. P.; Sander, M. Biochem. 2004, 36, 1663.
Electrochemical Analysis of Proton and Electron Transfer Equilibria of (27) Helburn, R. S.; MacCarthy, P. Determination of Some Redox
the Reducible Moieties in Humic Acids. Environ. Sci. Technol. 2011, 45, Properties of Humic-Acid by Alkaline Ferricyanide Titration. Anal.
8385−8394. Chim. Acta 1994, 295, 263−272.
(8) Aeschbacher, M.; Graf, C.; Schwarzenbach, R. P.; Sander, M. (28) Struyk, Z.; Sposito, G. Redox properties of standard humic
Antioxidant Properties of Humic Substances. Environ. Sci. Technol. acids. Geoderma 2001, 102, 329−346.
2012, 46, 4916−4925. (29) Ritchie, J. D.; Perdue, E. M. Proton-binding study of standard
(9) Rimmer, D. L. Free radicals, antioxidants, and soil organic matter and reference fulvic acids, humic acids, and natural organic matter.
recalcitrance. Eur. J. Soil Sci. 2006, 57, 91−94. Geochim. Cosmochim. Acta 2003, 67, 85−96.
(10) Romera-Castillo, C.; Jaffé, R. Free radical scavenging (30) Driver, S. J.; Perdue, E. M. Acidic Functional Groups of
(antioxidant activity) of natural dissolved organic matter. Mar. Chem. Suwannee River Natural Organic Matter, Humic Acids, and Fulvic
2015, 177, 668. Acids. In Advances in the Physicochemical Characterization of Dissolved
(11) Rimmer, D. L.; Abbott, G. D. Phenolic compounds in NaOH Organic Matter: Impact on Natural and Engineered Systems; Rosario-
extracts of UK soils and their contribution to antioxidant capacity. Eur. Ortiz, F., Ed.; ACS Symposium Series; American Chemical Society:
J. Soil Sci. 2011, 62, 285−294. Washington, DC, 2014; Vol. 1160, pp 75−86.
(12) Schlichting, A.; Rimmer, D. L.; Eckhardt, K.-U.; Heumann, S.; (31) Prior, R. L.; Wu, X. L.; Schaich, K. Standardized methods for the
Abbott, G. D.; Leinweber, P. Identifying potential antioxidant determination of antioxidant capacity and phenolics in foods and
compounds in NaOH extracts of UK soils and vegetation by dietary supplements. J. Agric. Food Chem. 2005, 53, 4290−4302.
untargeted mass spectrometric screening. Soil Biol. Biochem. 2013, (32) Huang, D. J.; Ou, B. X.; Prior, R. L. The chemistry behind
58, 16−26. antioxidant capacity assays. J. Agric. Food Chem. 2005, 53, 1841−1856.
(13) Wenk, J.; von Gunten, U.; Canonica, S. Effect of Dissolved (33) Liu, Z.-Q. Chemical Methods To Evaluate Antioxidant Ability.
Organic Matter on the Transformation of Contaminants Induced by Chem. Rev. 2010, 110, 5675−5691.
Excited Triplet States and the Hydroxyl Radical. Environ. Sci. Technol. (34) Gülçin, I.̇ Antioxidant activity of food constituents: an overview.
2011, 45, 1334−1340. Arch. Toxicol. 2012, 86, 345−391.
(14) Westerhoff, P.; Aiken, G.; Amy, G.; Debroux, J. Relationships (35) Box, J. D. Investigation of the Folin-Ciocalteau Phenol Reagent
between the structure of natural organic matter and its reactivity for the Determination of Polyphenolic Substances in Natural-Waters.
towards molecular ozone and hydroxyl radicals. Water Res. 1999, 33, Water Res. 1983, 17, 511−525.
2265−2276. (36) Apak, R.; Ç apanoğlu, E.; Arda, A. Ü . Nanotechnological
(15) Wert, E. C.; Rosario-Ortiz, F. L.; Snyder, S. A. Effect of ozone Methods of Antioxidant Characterization. In Advances in the
exposure on the oxidation of trace organic contaminants in wastewater. Physicochemical Characterization of Dissolved Organic Matter: Impact
Water Res. 2009, 43, 1005−1014. on Natural and Engineered Systems; Rosario-Ortiz, F., Ed.; ACS
(16) Chon, K.; Salhi, E.; von Gunten, U. Combination of UV Symposium Series; American Chemical Society: Washington, DC,
absorbance and electron donating capacity to assess degradation of 2015; Vol. 1191, pp 209−234.
micropollutants and formation of bromate during ozonation of (37) Chan-Eam, S.; Teerasong, S.; Damwan, K.; Nacapricha, D.;
wastewater effluents. Water Res. 2015, 81, 388−397. Chaisuksant, R. Sequential injection analysis with electrochemical

6431 DOI: 10.1021/acs.est.6b01120


Environ. Sci. Technol. 2016, 50, 6423−6432
Environmental Science & Technology Article

detection as a tool for economic and rapid evaluation of total (56) Areskogh, D.; Li, J.; Nousiainen, P.; Gellerstedt, G.; Sipila, J.;
antioxidant capacity. Talanta 2011, 84, 1350−1354. Henriksson, G. Oxidative polymerisation of models for phenolic lignin
(38) Milardovic, S.; Kerekovic, I.; Derrico, R.; Rumenjak, V. A novel end-groups by laccase. Holzforschung 2010, 64, 21−34.
method for flow injection analysis of total antioxidant capacity using (57) Reinhammar, B. R. M. Oxidation-Reduction Potentials of
enzymatically produced ABTS(center dot+) and biamperometric Electron Acceptors in Laccases and Stellacyanin. Biochim. Biophys.
detector containing interdigitated electrode. Talanta 2007, 71, 213− Acta, Bioenerg. 1972, 275, 245−259.
220. (58) Piontek, K.; Antorini, M.; Choinowski, T. Crystal structure of a
(39) Iveković, D.; Milardovic, S.; Roboz, M.; Grabarić, B. S. laccase from the fungus Trametes versicolor at 1.90-angstrom
Evaluation of the antioxidant activity by flow injection analysis method resolution containing a full complement of coppers. J. Biol. Chem.
with electrochemically generated ABTS radical cation. Analyst 2005, 2002, 277, 37663−37669.
130, 708−714. (59) Warren, J. J.; Tronic, T. A.; Mayer, J. M. Thermochemistry of
(40) Scott, S. L.; Chen, W.-J.; Bakac, A.; Espenson, J. H. Proton-Coupled Electron Transfer Reagents and its Implications.
Chem. Rev. 2010, 110, 6961−7001.
Spectroscopic Parameters, Electrode-Potentials, Acid Ionization-
Constants, and Electron-Exchange Rates of the 2,2′-Azinobis(3-
Ethylbenzothiazoline-6-Sulfonate) Radicals and Ions. J. Phys. Chem.
1993, 97, 6710−6714.
(41) Milardovic, S.; Kerekovic, I.; Rumenjak, V. A flow injection
biamperometric method for determination of total antioxidant capacity
of alcoholic beverages using bienzymatically produced ABTS(center
dot+). Food Chem. 2007, 105, 1688−1694.
(42) Magalhaes, L. M.; Segundo, M. A.; Reis, S.; Lima, J. L. F. C.;
Toth, I. V.; Rangel, A. O. S. S. Automatic flow system for sequential
determination of ABTS(center dot+) scavenging capacity and Folin-
Ciocalteu index: A comparative study in food products. Anal. Chim.
Acta 2007, 592, 193−201.
(43) Magalhaes, L.; Santos, M.; Segundo, M.; Reis, S.; Lima, J. Flow
injection based methods for fast screening of antioxidant capacity.
Talanta 2009, 77, 1559−1566.
(44) IHSS (International Humic Substances Society). Acidic
Functional Groups of IHSS Samples, http://www.humicsubstances.
org/acidity.html (accessed January 27, 2016).
(45) Klüpfel, L.; Keiluweit, M.; Kleber, M.; Sander, M. Redox
properties of plant biomass-derived black carbon (biochar). Environ.
Sci. Technol. 2014, 48, 5601−5611.
(46) Leonowicz, A.; Edgehill, R. U.; Bollag, J. M. The effect of pH on
the transformation of syringic and vanillic acids by the laccases of
Rhizoctonia praticola and Trametes versicolor. Arch. Microbiol. 1984,
137, 89−96.
(47) Thomas, M. J.; Bielski, B. H. J. Oxidation and reaction of trolox
c, a tocopherol analog, in aqueous solution. A pulse-radiolysis study. J.
Am. Chem. Soc. 1989, 111, 3315−3319.
(48) Bentayeb, K.; Rubio, C.; Nerin, C. Study of the antioxidant
mechanisms of Trolox and eugenol with 2,2 ′-azobis (2-
amidinepropane)dihydrochloride using ultra-high performance liquid
chromatography coupled with tandem mass spectrometry. Analyst
2012, 137, 459−470.
(49) Keith, L. H.; Crummett, W.; Deegan, J., Jr; Libby, R. A.; Taylor,
J. K.; Wentler, G. Principles of environmental analysis. Anal. Chem.
1983, 55, 2210−2218.
(50) MacDougall, D.; Crummett, W. B. Guidelines for data
acquisition and data quality evaluation in environmental chemistry.
Anal. Chem. 1980, 52, 2242−2249.
(51) Rubinson, K. A.; Rubinson, J. F. Contemporary Instrumental
Analysis; Prentice Hall, Inc.: Upper Saddle River, NJ, 2000; pp 139−
161.
(52) Malcolm, R. L. The uniqueness of humic substances in each of
soil, stream and marine environments. Anal. Chim. Acta 1990, 232,
19−30.
(53) Rasmussen, S.; Wolff, C.; Rudolph, H. Compartmentalization of
Phenolic Constituents in Sphagnum. Phytochemistry 1995, 38, 35−39.
(54) Rasmussen, S.; Wolff, C.; Rudolph, H. 4′-O-beta-D-glucosyl-cis-
p-coumaric acid - A natural constituent of Sphagnum fallax cultivated
in bioreactors. Phytochemistry 1996, 42, 81−87.
(55) Abbott, G. D.; Swain, E. Y.; Muhammad, A. B.; Allton, K.;
Belyea, L. R.; Laing, C. G.; Cowie, G. L. Effect of water-table
fluctuations on the degradation of Sphagnum phenols in surficial peats.
Geochim. Cosmochim. Acta 2013, 106, 177−191.

6432 DOI: 10.1021/acs.est.6b01120


Environ. Sci. Technol. 2016, 50, 6423−6432

You might also like