Вы находитесь на странице: 1из 8

Chemical Engineering Journal 243 (2014) 43–50

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Isomer-specific oxidation of nonylphenol by potassium permanganate


Zhijiang Lu ⇑, Jay Gan
Department of Environmental Sciences, University of California, Riverside, CA 92521, United States

h i g h l i g h t s

 This is the first report on isomeric selectivity in the advanced oxidation of nonylphenol.
 The oxidation of nonylphenol isomers by KMnO4 is efficient and isomer-specific.
 Nonylphenol isomers with a-dimethyl substituents have higher reactivity.
 Nonylphenol isomers with a-dimethyl substituents have higher pH dependence.
 This isomer-specificity may influence on water safety, human health and its risk assessment.

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

Article history: Nonylphenol (NP), the metabolite of non-ionic surfactant nonylphenol ethoxylates, is a well-known envi-
Received 24 October 2013 ronmental endocrine disruptor. Nonylphenol is frequently detected in treated wastewater effluent,
Received in revised form 30 December 2013 impacted surface water and even in drinking water due to its heavy use. Nonylphenol has been treated
Accepted 2 January 2014
as a single compound in most studies to date, but in reality it is a mixture of numerous isomers. Recent
Available online 13 January 2014
studies showed that different NP isomers displayed different estrogenicity and biodegradability in the
environment. In this study, we examined, for the first time, the potential isomeric selectivity in the oxi-
Keywords:
dation of NP. The reaction kinetics of 19 NP isomers, including 6 diastereoisomers, with potassium per-
Surfactant metabolite
Advance oxidation process
manganate was measured at pH 5, 7 and 9. At pH 7 with 10 mg L 1 of KMnO4 and 50 lg L 1 technical
Alkylphenol mixture of NP, the half-lives of 19 isomers varied from 4.8 to 6.3 min. In general, the reaction followed
Potassium permanganate the order: a-dimethyl > a-ethylmethyl  a-methylpropyl  a-iso-propylmethyl. The oxidation rate of
NP isomers with KMnO4 increased with increasing pH and a greater dependence on pH was observed
for isomers with a-dimethyl substituents. This study suggested that a-substituents may be the dominant
factor regulating the reactivity and persistence of NP isomers during advanced oxidation processes.
Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction Nonylphenol is commonly treated as a single compound in the


evaluation of its environmental occurrence, fate and transport,
Nonylphenol (NP) is the metabolite of the heavily used nonyl- treatment removal and toxicity [1,2,5]. However, technical nonyl-
phenol ethoxylate surfactants [1,2]. Due to the wide use of nonyl- phenol (tNP) is in fact a mixture of more than 100 isomers [6,7].
phenol ethoxylates in agricultural, industrial and domestic A few recent studies showed that NP isomers have different estrog-
products, NP is frequently detected in surface water, effluents of enicity or biodegradability in the environment [6,8–10]. For exam-
wastewater treatment plants and even in drinking water [2,3]. ple, Gabriel et al. [8] reported that if the average estrogenicity of
For example, levels up to 48 lg L 1 have been reported in water tNP in the yeast estrogen assays was set to be 1, the value for
from Hudson River Estuary, New York [4]. In a recent study, NP NP93 was 1.87 while the value for NP112 was only 0.60. Biodegrada-
was found in 55 samples out of 62 drinking water samples from tion of NP by Sphingobium xenophagum Bayram was also isomer
31 major cities across China with a median concentration of specific, with 99.7% of NP128 degraded after 9 d while only 31% of
27 ng L 1, whereas and in all 62 source water samples with a NP193 dissipated under the same conditions [8]. The isomer-
median concentration of 123 ng L 1 [3]. Nonylphenol is known as specific biodegradation was expected to result in the selective
a potent endocrine disruptor and the potency of NP relative to enrichment of recalcitrant and possibly highly estrogenic NP iso-
17b-estrodiol ranges from 0.001 to 0.05 with arithmetic mean of mers in the environment [8]. This prediction was corroborated
0.023 in in vivo bioassays [1]. by the profile of NP isomers found in river water and sea water
[9,11]. Such isomer selectivity arises from differences in physico-
chemical properties due to changes in spatial hindrance and elec-
⇑ Corresponding author. Tel.: +1 (951) 827 4416; fax: +1 (951) 827 3993. tron distribution among the isomers [8], and is thus an intrinsic
E-mail address: zhijiang.lu@email.ucr.edu (Z. Lu). property of NP that has not been adequately understood.

1385-8947/$ - see front matter Ó 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2014.01.007
44 Z. Lu, J. Gan / Chemical Engineering Journal 243 (2014) 43–50

Advanced oxidation with potassium permanganate is a promis- reaction solution (50 mL) with 50 lg L 1 tNP was constantly stir-
ing process for removing NP [5,12] and also other micropollutants red at 480 rpm. The reaction was initiated by adding KMnO4 to ar-
[13]. A recent study showed that KMnO4 was more effective at rive at the initial concentration of 1, 2 or 10 mg L 1. At selected
removing endocrine disrupting chemicals, including 4-n-NP (i.e., time intervals (0, 2, 4, 6, 8, 10, 15, 20, 30, 40 min), 0.5 mL of
NP1), than ozone, ferrate or chlorine in wastewater [5]. At pH 7, 200 g L 1 L-ascorbic acid was added to quench the reaction. A con-
80% of NP (1.0 mg L 1) was degraded in 120 min by 3.2 mg L 1 trol group was included with 50 lg L 1 tNP at pH7 but without
KMnO4 [12]. When KMnO4 concentration was increased to 16 or KMnO4. All reactions were carried out in 4 replicates. Additional
32 mg L 1, 100% NP removal was achieved in 100 min. In another groups at pH 5, 7 and 9 with 50 lg L 1 tNP and 10 mg L 1 or
study, with 9.48 mg L 1 of KMnO4 at pH 7, the half-life of 4-n-NP 1 mg L 1 KMnO4 were used to monitor the change in temperature,
(33 lg L 1) was only 3.0 min [5]. However, to date no study has pH or KMnO4 concentration during the reaction. During the 40 min
considered the isomer specificity in the chemical oxidation of NP of reaction, there was no detectable change in pH or KMnO4 con-
with permanganate or any other oxidants [14,15]. centration, while the temperature increased by 2.0 °C, likely due
The objectives of this study were to investigate the oxidation of to the constant stirring.
tNP by KMnO4, to obtain isomer-specific oxidation rates and to
understand the influence of side chain length and a-substituents. 2.3. Experiments with reclaimed water
The oxidation kinetics of each isomer was measured under differ-
ent KMnO4 concentrations and pH conditions. Statistics were used Reclaimed water (COD 21 mg L, NO3 -N 3 mg L, NH3-N
to analyze the relationship between the reaction rate and struc- 0.062 mg L, o-phosphate-P 0.06 mg L) was collected from Michel-
tural properties including the side chain length and types of a- son Water Reclamation Plant, Irvine, CA and filtered with 0.7 lm
substituents. glass fiber membranes. The reaction was initiated by adding
0.5 mL of 1000 mg L 1 KMnO4 to 50 mL filtered reclaimed water
2. Materials and methods containing 50 lg L 1 tNP maintained at pH 7 using 0.01 mol L 1
of phosphate buffer. The other conditions were the same as above.
2.1. Chemicals The detailed reaction conditions of the batch experiments with re-
agent and reclaimed water are listed in Table 1.
Technical NP (tNP), a mixture of NP isomers with branched side
chains (CAS No. 84852-15-3) was purchased from TCI America 2.4. Chemical analysis
(Portland, OR). Isomers NP36, NP37, NP119, NP128 and NP194
(10 mg g 1 in hexane; >95% purity) were obtained from ChemCol- The quenched reaction solutions were extracted twice with hex-
lect GmbH (Remscheid, Germany). Isomers NP38 (48.2 mg g 1), ane (15 and 10 mL, respectively) by liquid-liquid extraction. The
NP65 (59.5 mg g 1), NP111 (49.5 mg g 1) and NP112 (48.52 mg g 1) combined extracts were condensed to about 1 mL under a gentle
in methanol were kindly provided by Prof. Rong Ji at Nanjing Uni- nitrogen stream and then 10 lL of 100 mg L 1 4-tert octylphenol
versity, China. Potassium permanganate (KMnO4, 99+%), 4-tert- was added as the internal standard. The samples were analyzed
octylphenol (>97%), sodium phosphate dibasic (Na2HPO4, 99+%), on an Agilent 6890 GC coupled with 5973 mass spectrometer
sodium tetraborate decahydrate (Na2B4O710H2O, >99.5%) and (Agilent Technologies, Santa Clara, CA). Samples (2 lL) were intro-
L-ascorbic acid (>99.0%) were from Sigma-Aldrich (St. Louis, MO). duced into the inlet at 250 °C and the separation was achieved on a
Sodium acetate trihydrate (99.0–101.0%) and gas chromatography DB-5MS Ultra Inert capillary column (60 m  0.25 mm  0.25 lm,
(GC) grade hexane were from Fisher Scientific (Fair Lawn, NJ). Agilent, Wilmington, DE). The carrier gas was helium (99.999%)
Reagent water (18.3 MX-cm resistivity) was prepared using a and the flow rate was set at 1.0 mL min 1. The column temperature
Barnstead Nanopure water system (Barnstead, Dubuque, IA). The was programmed initially at 80 °C for 0 min, then ramped to 160 °C
stock solution of 100 mg L 1 tNP in methanol was stored at at 10 °C min 1 and held for 42 min, and further ramped to 300 °C at
20 °C before use. The stock solution of 1000 mg L 1 KMnO4 was 30 °C min 1 and held for 3 min. The mass spectrometer was oper-
prepared with boiled reagent water and stored at 4 °C. The KMnO4 ated in the electron ionization – selective ion monitoring (SIM)
stock solution was used within a month after preparation and be- mode at 70 eV. The ion with m/z 135 was used for the quantification
fore each use, its concentration was determined on a UV–Vis spec- of the internal standard and ions with m/z 121, 135, 149, 163, 177
trometer at 525 nm after dilution to 20 mg L 1. The results showed and 191 were used for the quantification of different NP isomers.
there was no change in KMnO4 under these conditions. For the identification of NP isomers, both a full scan mode (m/z
40 to m/z 800) and the SIM mode (m/z 107, 121, 135, 149, 163,
2.2. Reaction setup with reagent water 177, 191 and 220) were applied. The recovery of NP isomers was
97 ± 3%.
Batch experiments were conducted in 125-mL wide-mouth am-
ber borosilicate glass bottles with Teflon-faced caps at room tem- 2.5. Data analysis
perature (22 ± 0.5 °C). Reaction solutions prepared with reagent
water were maintained at the desired pH using 0.01 mol L 1 of All statistical analyses, including linear regression, one-way
the following pH buffers: pH 5, sodium acetate; pH 7, sodium ANOVA and two-way ANOVA, were carried out using SigmaPlot
phosphate dibasic; pH 9, sodium tetraborate decahydrate. The 11 (San Jose, CA) at the significant level of 0.05. The Holm-Sidak

Table 1
Matrix, pH, KMnO4 and nonylphenol concentrations in the batch experiments.

Reaction conditions A B C D E
Matrix Reagent water Reagent water Reagent water Reagent water Reclaimed water
pH 5 7 7 9 7
1
KMnO4 concentration (mg L ) 10 10 2 1 10
tNP concentration (lg L 1) 50 50 50 50 50
Z. Lu, J. Gan / Chemical Engineering Journal 243 (2014) 43–50 45

Table 2
Names, structures and pseudo first-order rate constants of 19 nonylphenol isomers separated and identified in this study.
1
Isomer Structure Side chain a Group Rate constant (min )
length Substituents
A B C D E
36a 6 Me, Me 1 0.0564 ± 0.0016 0.1319 ± 0.0024 0.0556 ± 0.0024 0.2120 ± 0.0082 0.258 ± 0.016

HO

Xb 5 Me, Me 1 0.0489 ± 0.0015 0.1261 ± 0.0022 0.0566 ± 0.0022 0.2242 ± 0.0081 0.261 ± 0.016

128a 5 Me, Me 1 0.0552 ± 0.0017 0.1382 ± 0.0024 0.0628 ± 0.0024 0.2285 ± 0.0087 0.277 ± 0.016
HO

38a 6 Me, Me 1 0.0552 ± 0.0016 0.1232 ± 0.0017 0.0543 ± 0.0022 0.1860 ± 0.0069 0.241 ± 0.014

HO

37a 6 Me, Me 1 0.0543 ± 0.0016 0.1300 ± 0.0018 0.0589 ± 0.0023 0.2025 ± 0.0074 0.248 ± 0.013

HO

119a 5 Me, Me 1 0.0561 ± 0.0014 0.1478 ± 0.0029 0.0614 ± 0.0024 0.2135 ± 0.0067 0.280 ± 0.015

HO

35 6 Me, Me 1 0.0579 ± 0.0017 0.1443 ± 0.0023 0.0617 ± 0.0027 0.2134 ± 0.0080 0.258 ± 0.017

HO

9 7 Me, Me 1 0.0576 ± 0.0015 0.1268 ± 0.0025 0.0480 ± 0.0025 0.1648 ± 0.0089 0.177 ± 0.008

HO

65a 6 Me, Et 2 0.0586 ± 0.0014 0.1184 ± 0.0018 0.0521 ± 0.0024 0.1400 ± 0.0083 0.185 ± 0.014

HO

112a 5 Me, Et 2 0.0525 ± 0.0015 0.1114 ± 0.0014 0.0529 ± 0.0021 0.1557 ± 0.0059 0.203 ± 0.010

HO

111aa 5 Me, Et 2 0.0572 ± 0.0014 0.1146 ± 0.0014 0.0553 ± 0.0023 0.1648 ± 0.0062 0.212 ± 0.010

HO

111ba 5 Me, Et 2 0.0601 ± 0.0014 0.1172 ± 0.0014 0.0572 ± 0.0024 0.1695 ± 0.0067 0.212 ± 0.012

HO

110a 5 Me, Et 2 0.0582 ± 0.0013 0.1174 ± 0.0014 0.0554 ± 0.0025 0.1535 ± 0.0049 0.175 ± 0.009

HO

110b 5 Me, Et 2 0.0585 ± 0.0014 0.1182 ± 0.0013 0.0549 ± 0.0025 0.1526 ± 0.0052 0.203 ± 0.011

HO

193a 4 Me, Pr 3 0.0581 ± 0.0012 0.1177 ± 0.0016 0.0544 ± 0.0024 0.1488 ± 0.0053 0.185 ± 0.009

HO

193b 4 Me, Pr 3 0.0598 ± 0.0013 0.1224 ± 0.0016 0.0560 ± 0.0026 0.1551 ± 0.0055 –*

HO

(continued on next page)


46 Z. Lu, J. Gan / Chemical Engineering Journal 243 (2014) 43–50

Table 2 (continued)
1
Isomer Structure Side chain a Group Rate constant (min )
length Substituents
A B C D E

152 5 Me, Pr 3 0.0555 ± 0.0013 0.1130 ± 0.0015 0.0505 ± 0.0022 0.1289 ± 0.0089 0.188 ± 0.012

HO

194a 4 Me, Pr 3 0.0577 ± 0.0013 0.1098 ± 0.0016 0.0492 ± 0.0018 0.1536 ± 0.0052 0.189 ± 0.010

HO

143 5 Me, i-Pr 4 0.0582 ± 0.0013 0.1159 ± 0.0014 0.0557 ± 0.0024 0.1419 ± 0.0067 0.191 ± 0.010

HO

a
Isomers confirmed with authentic standards.
b
The structure of this nonylphenol isomer is not fully identified; it may be NP92, NP93 or NP95 and the structure may be , and
HO HO

, respectively.
HO
*
Measurement was not possible due to interference.

method was used in the pairwise multiple comparison procedure two methyl groups at the a position and no dimethyl group at the
in the one-way and two-way ANOVA. other positions or ethyl group or propyl group. Therefore, the isomer
NPx may be NP92, NP93 or NP95 and its systematic name may be
4-[1,1,2,3-tetramethylpentyl]phenol, 4-[1,1,2,4-tetramethylpentyl]
3. Results and discussion phenol or 4-[1,1,3,4-tetramethylpentyl]phenol.
According to their a substituents, the 19 identified isomers are
3.1. Separation and identification of nonylphenol isomers grouped in 4 groups: Group 1 with two methyl substituents; Group
2 with one methyl and one ethyl substituents; Group 3 with one
A total of 19 isomers, including 6 diastereoisomers (Tables 2 and methyl and one propyl substituents; and Group 4 with one methyl
S1), were separated and identified under the used analytical condi- and 1 iso-propyl substituent. A similar number of NP isomers were
tions by comparing the retention time, peak intensity and fragmen- separated in Gabriel et al. [8]; however, more authentic NP isomer
tation patterns against the available authentic NP isomer standards standards were used in this study to confirm the identification.
and/or chromatographic and mass spectrometric characteristics of With more sophisticated instruments, for example, two dimen-
NP isomers in published studies [8,16]. All identifications agreed sional GC, a better separation may be achieved [17].
with the NP isomer fragmentation mechanism in Thiele et al. [16].
In Tables 1 and S1, an NP isomer labeled as NPX was not fully identi-
fied. This isomer had the dominant fragment at m/z 135 and also a 3.2. Reaction kinetics of NP isomers
few minor fragments at m/z 107 (9%), 121 (5%) and 220 (3%) but
an absence of fragments at m/z 149, 163, 177 and 191. According The oxidation of NP isomers by KMnO4 was consistently rapid
to the fragmentation mechanism [16], this isomer should have (Table 2 and Fig. 1). For example, at pH 7 with 10 mg L 1 KMnO4,

1
Fig. 1. Isomer specific reaction rates in 10 mg L KMnO4 at pH 7 with reagent water. (a) Removal of NP194 and NP119; and (b) pseudo first-order rate constants of
nonylphenol isomers.
Z. Lu, J. Gan / Chemical Engineering Journal 243 (2014) 43–50 47

Table 3
1
Comparison of pseudo first-order rate constants among nonylphenol isomers with 10 mg L KMnO4 at pH 7 in reagent water.

Isomers X 128 38 37 119 35 9 65 112 111a 111b 110A 110B 193A 193B 152 194 143
36 Y Y Y N Y Y Y Y Y Y Y Y Y Y Y Y Y Y
X Y N N Y Y Y Y Y Y Y Y Y N Y Y Y
128 Y Y Y Y Y Y Y Y Y Y Y Y Y Y Y Y
38 Y Y Y N Y Y Y Y Y Y Y N Y Y Y
37 Y Y N Y Y Y Y Y Y Y Y Y Y Y
119 N Y Y Y Y Y Y Y Y Y Y Y Y
35 Y Y Y Y Y Y Y Y Y Y Y Y
9 Y Y Y Y Y Y Y N Y Y Y
65 Y N N N N N N Y Y N
112 N Y Y Y Y Y N N Y
111a N N N N Y N Y N
111b N N N Y N Y N
110a N N Y N Y N
110b N N Y Y N
193a Y Y Y N
193b Y Y Y
152 N N
194 Y Y

Y: significant difference between two NP isomers.


N: no significant difference between two NP isomers.

about 90% of 50 lg L 1 tNP was removed (Fig. 1) while there was Some NP isomers have two chiral carbon centers, and these iso-
no change in the KMnO4-free control. The oxidation reaction was mers may have four stereoisomers. After the non-chiral selective
pseudo-first order as shown in Fig. 1a for NP119 and NP194 at pH GC separation, these four stereoisomers appeared as two peaks in
7 with 10 mg L 1 KMnO4, with the first-order rate constants of diastereomers, e.g., NP111a/b and NP193a/b. In this study, two diaste-
0.1098 ± 0.0016 and 0.1478 ± 0.0026 min 1, or half-lives of 4.8 reomers are noted by alphabets a and b according their retention
and 6.3 min, respectively. The high efficiency and pseudo-first or- times. Unlike the enantiomers, diastereomers are not mirror images
der behavior agreed with previous studies [5,12]. For example, of each other; therefore, they may have different physical proper-
Jiang et al. [5] showed that the oxidation of 0.15 lmol L 1 ties and chemical reactivity. However, in this study, the diastereo-
4-n-NP (i.e., 33 lg L 1 NP1) with KMnO4 was pseudo-first order mers did not show any difference in their oxidation by KMnO4 in
at pH 7; the second-order rate constant was estimated to be most cases. For example, at pH 7 with 10 mg L 1 of KMnO4, the rate
63.9 L mol 1 s 1. Using the reported rate constants, the half-life constants for NP110a and NP110b were 0.1174 ± 0.0014 and
of 4-n-NP in 60 lmol 1 KMnO4 solution (or 9.48 mg L 1) should 0.1182 ± 0.0013 min 1, respectively. The only exception was the
be about 3 min, which agrees well with that observed in this study diastereomers NP193a and NP193b at pH 7 with 10 mg L 1 of KMnO4,
(Table 2). where the rate constants were 0.1177 ± 0.0016 and
The oxidation reaction was obviously isomer specific (Table 2 0.1224 ± 0.0016 min 1, respectively. Limited differences between
and Fig. 1). With higher KMnO4 concentrations or at higher pH, diastereomers were also observed for biodegradation [8]. For
the isomeric difference was much more significant, probably due instance, after 9 d of incubation with Sphingobium xenophagum
to larger pseudo first order rate constants. To clearly identify the Bayram, the removal rates for NP111a and NP111b were 94.0% and
difference among isomers, a one-way ANOVA with Holm-Sidak 90.4%, respectively [8].
pairwise multiple comparisons was carried out to determine if
the differences were statistically significant (Table 3). For example, 3.3. Effect of side chain length and types of a-substituents
at pH 7 with 10 mg L 1 KMnO4, the rate constant of NP119
(0.1478 ± 0.0029 min 1) was significantly different from that of Gabriel et al. reported that a-substitution was the most impor-
NP194 (0.1098 ± 0.0016 min 1). The rate constant of NP38 tant factor controlling the biodegradation rate of NP isomers with
(0.1232 ± 0.0017 min 1) was also significantly different from that Sphingobium xenophagum Bayram [8]. This was also confirmed by
of NP152 (0.1130 ± 0.0015 min 1) (Table 3). However, the isomeric the study on the degradation of four NP isomers in an oxic soil
differences in this study were generally smaller than those ob- [10]. Gabriel et al. also showed that a suitable main chain length
served in biodegradation [8,10]. For instance, in an oxic soil, the (4–6 carbon atoms) was important for estrogenicity of NP isomers
biodegradation half-life was 10.3 d for NP111, 8.4 d for NP112, 5.8 [8]. Moreover, in previous studies, the types of substituents were
d for NP65 and 2.1 d for NP38 [10]. In Gabriel et al. [8], isomeric bio- found to greatly affect the oxidation rates of other substituted phe-
degradation was measured only by the percentages degraded after nols. For instance, the oxidation of mono-substituted phenols with
9 d of incubation with Sphingobium xenophagum Bayram. Over 90% singlet oxygen followed the order of 4-(tert-butyl) phenol > 4-
removal was achieved for more than half of the NP isomers while methylphenol > phenol [21] and the oxidation with manganese
only about 30% degradation was found for NP193a/b. The relative dioxides followed the order 4-methyl phenol > 4-ethylphe-
differences in isomer selectivity are likely due to the fact that bio- nol > phenol [22]. Stone concluded that electron donating substit-
transformation catalyzed by specific enzymes affords more selec- uents enhanced the reaction of substituted phenols with
tivity than chemical reactions. The biodegradation of NP isomers manganese dioxides while the electron withdrawing groups were
was initiated by ipso-hydroxylation process in which a hydroxyl inhibitory [22]. Therefore, we hypothesized that a-substitution
group was attached at the para position [18,19]. The corresponding and side chain length were two most important factors in deter-
enzyme NP ipso-hydroxylase was a member of the flavin monoox- mining the oxidation rate of NP isomers.
ygenase enzyme family [20]. The bulky a-substituents could hin- To test this hypothesis, two-way ANOVA was performed using
der the ipso-hydroxylation sterically and thus resulted in a-substitution type and side chain length as two factors (Table
isomeric biodegradation of NP [8]. 2). The results showed that the side chain length did not have a
48 Z. Lu, J. Gan / Chemical Engineering Journal 243 (2014) 43–50

Fig. 2. Pseudo first-order rate constants of nonylphenol isomers and their ratios in 10 and 2 mg L 1 KMnO4 at pH 7 with reagent water. The red dotted reference line is 0.236.
(For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.).

Fig. 3. Pseudo first-order rate constants of nonylphenol isomers and their ratios in 10 mg L 1 KMnO4 at pH 5 and pH 7 with reagent water. The red dotted reference line
shows the ratio 2.18. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.).

significant influence on the reaction kinetics (p = 0.68). However, isomers in Group 2, 3 or 4 with a few exceptions, i.e., NP38, NPX and
the effect of a-substituents was found to be statistically significant NP9 had no significant difference from NP193b. Therefore, Group 1
(p = 0.005). The pairwise multiple comparisons also showed that isomers generally displayed significant differences from all other
Group 1 isomers (with dimethyl substitution) were significantly groups, although the two-way ANOVA showed no significant dif-
different from Group 2 (with methylethyl substitution) or Group ference between Group 1 and Group 4 isomers due to large varia-
3 (with methylpropyl substitution), while there was no significant tions within Group 1. Group 2 isomers had some variations within
difference among the other groups. the group and in some cases they had no significant differences
As the variation within a group may mask the variation be- from isomers in Group 3 or 4. Group 3 isomers also had some vari-
tween groups, comparisons were further made for isomers within ations within the group and some isomers had significant differ-
the same group and between groups (Table 3). For Group 1 iso- ences from the isomers in Group 4. Therefore, the overall
mers, there were large variations within the group and in many conclusion was that Group 1 isomers consistently showed differ-
cases the isomers in Group 1 differed significantly from each other. ences from isomers in the other groups and the differences of iso-
In addition, all Group 1 isomers showed significant difference from mers among the other groups were less consistent. That is, with
Z. Lu, J. Gan / Chemical Engineering Journal 243 (2014) 43–50 49

Fig. 4. Pseudo first-order rate constants of nonylphenol isomers and their ratios in reagent water and reclaimed water with 10 mg L 1 KMnO4 at pH 7. The red dotted
reference line shows the ratio 1.77. Measurement was not possible due to interference. (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.).

respect to a-substitution, the reactivity of NP isomers toward increase in KMnO4 concentration from 2 to 10 mg L 1 at pH 7
KMnO4 followed the order of a-dimethyl > a-ethylmethyl  a- resulted in an increase in the first-order rate constant of NP119
methylpropyl  a-iso-propylmethyl (Fig. 1). from 0.0614 ± 0.0024 to 0.1478 ± 0.0029 min 1. When the KMnO4
The substituent effects in this study may originate from induc- concentration was increased 5 fold, the reaction rate constant gen-
tive effects and/or steric hindrance. Previous studies showed that erally increased by a factor of 2.1–2.6 (mean value 2.23) (Fig. 2). If
oxidation rates of phenols were greatly affected by the inductive the reaction order with respect to KMnO4 is 0.5, the ratio should be
capacity of substituents and the effect may be described by the 2.236. No significant differences were found between the mea-
Hammett equation [22–24] depicting that electron donating sured value for each NP isomer and 2.236 after t-test (the smallest
groups increase the reaction while the electron withdrawing p value was 0.11) (Fig. 2), indicating that the reaction order with
groups decrease the reaction. Therefore, the electron donating respect to KMnO4 was 0.5 and that there was a general absence
capacities of alky substituents may have influenced the oxidation of isomer selectivity. Since the order with respect to KMnO4 was
rate of NP. The well recognized Hammett’s constants for para- not 1, in this study, pseudo first-order rate constants instead of sec-
methyl, ethyl, iso-propyl, propyl are 0.17, 0.15, 0.15 and ond-order rate constants were used. The order with respect to
0.13, respectively [23]. Therefore, the para-methyl substitution KMnO4 was found to be 1 in other studies [5,28,29]. The inconsis-
may increase the reaction as compared to the other substitutions. tency may indicate that another oxidative species, for example, Mn
However, in this study, the differences of substituents were at the (III), also reacted with NP. Mn (III) was found to form during
a position instead of the para position and the inductive effects de- KMnO4 treatment and to oxidize bisphenol A that contained two
creased with increasing distance. For example, the Hammett’s con- phenol moieties [28]. Further investigations are needed to better
stants for tert-butyl (C(CH3)3) and triethylmethyl (C(C2H5)3) are understand the cause for the deviation from the second-order
both 0.20 [23]. However, the NP isomers with a-dimethyl substi- reaction kinetics.
tuent had the highest reactivity, indicating that its inductive effect When the solution pH was increased, the reaction between NP
was still greater than the others. The steric hindrance in this study isomers and KMnO4 increased drastically (Table 2 and Fig. 3). For
might be negligible since the phenoxy oxygen was bound with example, when pH increased from pH 5 to pH 7 in the 10 mg L 1
manganese (VII) atom in the complex and the substituents were KMnO4 solution, the reaction rate constant for NP119 increased
far away [24]. In comparison, during biodegradation, the ipso- from 0.0561 ± 0.0014 to 0.1478 ± 0.0029 min 1. When pH further
hydroxylation happened at the para position, and thus the steric increased to pH 9, although the concentration of KMnO4 was only
hindrance was likely more important [8]. The oxidation rate of 1 mg L 1, the rate constant increased to 0.2135 ± 0.0067 min 1
phenols by permanganate under both acidic and alkaline condi- (Table 2). Similar strong pH dependence was found in the
tions was found to be limited by the electron transfer from the oxidation of other phenolic compounds, including bisphenol A,
phenol/phenolate anion to phenoxy radical [25,26]. Therefore, an- 17b-estradiol and dichlorophenol [5]. Deprotonation of phenolic
other parameter to describe the reactivity of organics could be the compounds was believed to be the cause for the pH dependence
oxidation/ionization potential [27]. However, the measured values [5] since deprotonated phenols showed a higher reactivity than
for oxidation and ionization potentials of NP isomers are not avail- protonated phenols [21]. For example, when reacting with singlet
able in the literature, preventing a further elucidation of the poten- oxygen, the second-order rate constants for deprotonated
tial significance of this mechanism. and protonated 4-(tert-butyl) phenol were 1.2 ± 1.3  107 and
3.55 ± 0.23  108 mol 1 s 1, respectively [21]. Therefore, this pH
3.4. Effect of KMnO4 concentration, pH and matrix dependence would be monotonic in pH range 5–9 due to the
monotonic deprotonation, even though the reaction kinetics was
The reaction rate consistently increased with increasing measured only at pH 5, 7 and 9. Moreover, NP isomers with differ-
concentrations of KMnO4 (Table 2 and Fig. 2). For example, the ent capacities for deprotonation may have different dependence on
50 Z. Lu, J. Gan / Chemical Engineering Journal 243 (2014) 43–50

pH. Fig. 3 shows the ratio of rate constants at pH 5 and pH 7 in the [3] Z. Fan, J. Hu, W. An, M. Yang, Detection and occurrence of chlorinated
byproducts of bisphenol A, nonylphenol, and estrogens in drinking water of
10 mg L 1 KMnO4 solution. Although large variations existed with-
China: comparison to the parent compounds, Environ. Sci. Technol. (2013).
in Group 1 isomers, it is clear that pH exhibited a greater influence [4] J. Dachs, D.A. Van Ry, S.J. Eisenreich, Occurrence of estrogenic nonylphenols in
on the oxidation of Group 1 isomers than isomers in the other the urban and coastal atmosphere of the lower Hudson River estuary, Environ.
groups. A similar trend was also observed when pH was varied Sci. Technol. 33 (1999) 2676–2679.
[5] J. Jiang, S.Y. Pang, J. Ma, H.L. Liu, Oxidation of phenolic endocrine disrupting
from pH 7 to pH 9 (Table 2). Since electron donating substituents chemicals by potassium permanganate in synthetic and real waters, Environ.
decrease the acidity of substituted phenols, a-dimethyl substituent Sci. Technol. 46 (2012) 1774–1781.
should have a larger electron donating effect than the other sub- [6] R.P. Eganhouse, J. Pontolillo, R.B. Gaines, G.S. Frysinger, F.L.P. Gabriel, H.P.E.
Kohler, W. Giger, L.B. Barber, Isomer-specific determination of 4-nonylphenols
stituents in this study. Therefore, Group 1 isomers were more using comprehensive two-dimensional gas chromatography/time-of-flight
deprotonated and have relatively higher reactivity at high pH than mass spectrometry, Environ. Sci. Technol. 43 (2009) 9306–9313.
the other isomers and thus exhibit higher pH dependence. [7] T. Ieda, Y. Horii, G. Petrick, N. Yamashita, N. Ochiai, K. Kannan, Analysis of
nonylphenol isomers in a technical mixture and in water by comprehensive
Nonylphenol isomers generally showed a higher reactivity in two-dimensional gas chromatography-mass spectrometry, Environ. Sci.
the reclaimed water (Fig. 4 and Table 2), and the reaction rates in- Technol. 39 (2005) 7202–7207.
creased by 1.40–2.07 times (average 1.76) as compared to those in [8] F.L.P. Gabriel, E.J. Routledge, A. Heidlberger, D. Rentsch, K. Guenther, W. Giger,
J.P. Sumpter, H.P.E. Kohler, Isomer-specific degradation and endocrine
the reagent water. The increase was not isomer selective. The disrupting activity of nonylphenols, Environ. Sci. Technol. 42 (2008) 6399–
enhancement of KMnO4 catalyzed oxidation of phenolic com- 6408.
pounds in real water (i.e., waste water, surface water, et al.) over [9] Y.S. Kim, T. Katase, Y. Horii, N. Yamashita, M. Makino, T. Uchiyama, Y. Fujimoto,
T. Inoue, Estrogen equivalent concentration of individual isomer-specific 4-
synthetic water was also observed in other studies [5]. Various
nonylphenol in Ariake sea water, Japan, Mar. Pollut. Bull. 51 (2005) 850–856.
hypotheses have been proposed, such as unknown ligands existing [10] J. Shan, B.Q. Jiang, B. Yu, C.L. Li, Y.Y. Sun, H.Y. Guo, J.C. Wu, E. Klumpp, A.
in real water helping to stabilize reactive Mn(III) species, phenoxy Schaffer, R. Ji, Isomer-specific degradation of branched and linear 4-
radicals formed from organic matter, and also formation of Mn(V) nonylphenol isomers in an oxic soil, Environ. Sci. Technol. 45 (2011) 8283–
8289.
and Mn(VI) species during reaction [5]. Except for NP9, Group 1 iso- [11] Y. Horii, T. Katase, Y.S. Kim, N. Yamashita, Determination of individual
mers still displayed a higher reactivity than isomers from the other nonylphenol isomers in water samples by using relative response factor
groups. This further confirmed that intrinsic properties, for exam- method, Bunseki Kagaku 53 (2004) 1139–1147.
[12] Y. Abe, T. Umemura, K. Tsunoda, Decomposition of phenolic endocrine
ple, a-substitution, contributed to the isomer specificity. disrupting chemicals by potassium permanganate, Nippon Kagaku Kaishi
(2001) 239–242.
[13] X.H. Guan, D. He, J. Ma, G.H. Chen, Application of permanganate in the
4. Conclusions
oxidation of micropollutants: a mini review, Front Environ. Sci. Eng. 4 (2010)
405–413.
Nonylphenol, the commonly occurring metabolite from nonyl- [14] G. Bertanza, R. Pedrazzani, M. Dal Grande, M. Papa, V. Zambarda, C. Montani, N.
Steimberg, G. Mazzoleni, D. Di Lorenzo, Effect of biological and chemical
phenol ethoxylate surfactants, encompasses a large number of iso-
oxidation on the removal of estrogenic compounds (NP and BPA) from
mers, and yet its isomer-specific behaviors are relatively unknown. wastewater: an integrated assessment procedure, Water Res. 45 (2011) 2473–
Results from this study showed that some isomeric selectivity was 2484.
apparent in the oxidation of NP by KMnO4, although the selectivity [15] Y.P. Zhang, X. Zhou, Y.X. Lin, X. Zhang, Ozonation of nonylphenol and
octylphenol in water, Fresenius Environ. Bull. 17 (2008) 760–766.
appeared to be more limited as compared to that in biodegrada- [16] B. Thiele, V. Heinke, E. Kleist, K. Guenther, Contribution to the structural
tion. In general, NP isomers with dimethyl substitution at the a po- elucidation of 10 isomers of technical p-nonylphenol, Environ. Sci. Technol. 38
sition were more reactive toward KMnO4 than isomers with (2004) 3405–3411.
[17] A. Vallejo, M. Olivares, L.A. Fernandez, N. Etxebarria, S. Arrasate, E. Anakabe, A.
ethylmethyl, methylpropyl or iso-propylmethyl substitution. Oxi- Usobiaga, O. Zuloaga, Optimization of comprehensive two dimensional gas
dation by KMnO4 increased with increasing pH, and the influence chromatography-flame ionization detection-quadrupole mass spectrometry
of pH was more pronounced for NP isomers with a-dimethyl sub- for the separation of octyl- and nonylphenol isomers, J. Chromatogr. A 1218
(2011) 3064–3069.
stitution than those with other substituents. Isomeric specificity in [18] F.L.P. Gabriel, M. Cyris, N. Jonkers, W. Giger, K. Guenther, H.P.E. Kohler,
other oxidation systems, such as ozonation, merits further investi- Elucidation of the ipso-substitution mechanism for side-chain cleavage of
gation. Isomer-dependent reactivity or removal of NP during alpha-quaternary 4-nonylphenols and 4-t-butoxyphenol in Sphingobium
xenophagum Bayram, Appl. Environ. Microbiol. 73 (2007) 3320–3326.
wastewater treatment should be considered when predicting the
[19] B.A. Kolvenbach, P.F.X. Corvini, The degradation of alkylphenols by
discharge of NP into the open environment, as isomeric selectivity Sphingomonas sp. strain TTNP3 – a review on seven years of research, New
in both occurrence and toxicological effects may affect the overall Biotechnol. 30 (2012) 88–95.
[20] A.W. Porter, A.G. Hay, Identification of opdA, a gene involved in biodegradation
environmental risk of NP. Moreover, the isomeric selectivity during
of the endocrine disrupter octylphenol, Appl. Environ. Microbiol. 73 (2007)
drinking water treatment may have a direct implication on human 7373–7379.
exposure and health. [21] P.G. Tratnyek, J. Holgne, Oxidation of substituted phenols in the environment:
a QSAR analysis of rate constants for reaction with singlet oxygen, Environ. Sci.
Technol. 25 (1991) 1596–1604.
Acknowledgement [22] A.T. Stone, Reductive dissolution of manganese(III/IV) oxides by substituted
phenols, Environ. Sci. Technol. 21 (1987) 979–988.
We sincerely acknowledge Prof. Rong Ji for providing us nonyl- [23] C. Hansch, A. Leo, R.W. Taft, A survey of Hammett substituent constants and
resonance and field parameters, Chem. Rev. 91 (1991) 165–195.
phenol isomers. [24] S.V. Khansole, D.D. Gaikwad, S.D. Gaikwad, R.D. Kankariya, Kinetics and
mechanism of acid catalyzed oxidation of phenol and substituted phenols by
isoquinolinium bromochromate, Russ. J. Phys. Chem. A 84 (2010) 2233–2237.
Appendix A. Supplementary material
[25] L. Copolovici, I. Baldea, Kinetics of the phenol oxidation by permanganate in
acidic media, Rev. Roum. Chim. 52 (2007) 1045–1050.
Supplementary data associated with this article can be found, in [26] D.G. Lee, C.F. Sebastian, The oxidation of phenol and chlorophenols by alkaline
the online version, at http://dx.doi.org/10.1016/j.cej.2014.01.007. permanganate, Can. J. Chem.-Revue Can. De Chimie 59 (1981) 2776–2779.
[27] D.H. Evans, One-electron and two-electron transfers in electrochemistry and
homogeneous solution reactions, Chem. Rev. 108 (2008) 2113–2144.
References [28] J. Jiang, S.Y. Pang, J. Ma, Role of ligands in permanganate oxidation of organics,
Environ. Sci. Technol. 44 (2010) 4270–4275.
[1] A. Soares, B. Guieysse, B. Jefferson, E. Cartmell, J.N. Lester, Nonylphenol in the [29] L. Hu, H.M. Martin, O. Arcs-Bulted, M.N. Sugihara, K.A. Keatlng, T.J. Strathmann,
environment: a critical review on occurrence, fate, toxicity and treatment in Oxidation of carbamazepine by Mn(VII) and Fe(VI): reaction kinetics and
wastewaters, Environ. Int. 34 (2008) 1033–1049. mechanism, Environ. Sci. Technol. 43 (2009) 509–515.
[2] G.G. Ying, B. Williams, R. Kookana, Environmental fate of alkylphenols and
alkylphenol ethoxylates – a review, Environ. Int. 28 (2002) 215–226.

Вам также может понравиться