You are on page 1of 6

Chemosphere 75 (2009) 1015–1020

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Hydrolysis and H2O2-assisted UV photolysis of 3-chloro-1,2-propanediol


Amanda M. Nienow 1, Irene C. Poyer 2, Inez Hua, Chad T. Jafvert *
Purdue University, School of Civil Engineering, 550 Stadium Mall Drive, West Lafayette, IN 47907, United States

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

Article history: 3-Chloro-1,2-propanediol (3-MCPD) is a chlorinated alcohol that is often formed as a by-product in the
Received 31 October 2008 manufacturing of food products. In addition, 3-MCPD may be a disinfection by-product from wastewater
Received in revised form 15 January 2009 treatment by chlorine and may be present in drinking waters from purification plants using epichlorohy-
Accepted 16 January 2009
drin-linked cationic polymer resins as flocculants. Due to concerns about the toxicity of 3-MCPD and its
Available online 17 March 2009
potential presence in water samples, the removal of 3-MCPD from water should be addressed and exam-
ined. For the first time a systematic examination of the removal of 3-MCPD via hydrolysis and photolysis
Keywords:
processes is presented. 3-MCPD is shown to undergo hydrolysis at near neutral pH values, but at much
3-Chloro-1,2-propanediol
3-MCPD
slower rates than can be obtained by UV/H2O2 processes. 3-MCPD does not undergo rapid direct photol-
Oxidation ysis. Re-evaluation of temperature and pH dependent hydrolysis rate data indicates that hydrolysis is first
Photochemistry order with respect to [OH].
Pollution control Ó 2009 Elsevier Ltd. All rights reserved.
AOP

1. Introduction fertility agent (S-enantiomer) and to have effects on the kidneys


(R-enantiomer) (Hamlet et al., 2002). In bacterial assays, 3-MCPD
3-Chloro-1,2-propanediol (3-MCPD) is a chlorinated alcohol and has been shown to be a mutagen (Matthew and Anastasio, 2000).
a chlorohydrin. 3-MCPD is formed as a by-product in the manufac- Due to the toxic effects of 3-MCPD, the EU and the Joint FAO/
turing of hydrolyzed vegetable proteins, soy sauces, and heat-trea- WHO Expert Committee on Food Additives set a tolerable daily in-
ted cereal products (Hamlet et al., 2002; Nyman et al., 2003; take at 2 lg/kg body weight in hydrolyzed vegetable proteins and
Hamlet, 2008). 3-MCPD also is the hydrolysis product of epichloro- soy sauces (JECFA, 2001b; Xing et al., 2005). This limit was set in
hydrin, a commercially produced chemical used in resins, textiles, 2001, and more recent reports issued by the EU have supported
and paper products (Boden et al., 1997; Sarzanini et al., 2000; Dow, the limit (Chung et al., 2002; Hamlet et al., 2002). The United States
2006). The annual global production of epichlorohydrin is esti- Food and Drug Administration has set a limit of 1 ppm for M-3CPD
mated to be 2 billion pounds per year (Dow, 2006). Epichlorohy- in soy sauces (Masten, 2005).
drin-linked cationic polymer resins are sometimes used as In addition to the concern caused by 3-MCPD in food products,
flocculants in water purification, and 3-MCPD has been found to other concerns about drinking water safety have recently arisen. In
be a low-level contaminant in drinking water from these purifica- 2005, the US Army Center for Health Promotion and Preventative
tion plants (Codex Committee on Food Additives, 2001a; Masten, Medicine published an article examining 1756 chemicals, includ-
2005). In addition, other chlorohydrins, such as 1,3-dichloro-2-pro- ing toxic industrial chemicals (TICs), with immediate and severe
panol, have been identified in some water samples that have been adverse health impacts (Hauschild and Bratt, 2005). A portion of
treated with chlorine, chloramines, or a combination of ozone and this article focused on TICs with the potential for food or water
chlorine, and it is possible that 3-MCPD also may be a disinfection contamination. The compounds on the list of potential drinking
by-product (Matthew and Anastasio, 2000). water contaminants included organophosphates (e.g., methamida-
The presence of 3-MCPD in a variety of food products, water phos), inorganic compounds (e.g., cyanide and arsenic), and ‘mis-
sources, and other consumer products is of concern due to the tox- cellaneous’ organics. The latter category included phenol,
icity of the compound. 3-MCPD has been shown to be a male anti- acrolein, acrylonitrile, and 3-MCPD. Although the measured con-
centrations of 3-MCPD in drinking water sources tend to be low,
it is possible that 3-MCPD will be a contaminant in drinking waters
* Corresponding author. Tel.: +1 765 494 2196; fax: +1 765 496 1107. from water purification plants using epichlorohydrin-linked cat-
E-mail address: jafvert@ecn.purdue.edu (C.T. Jafvert). ionic polymer resins or in waste waters.
1
Present address: Gustavus Adolphus College, 800 W College St., St. Peter, MN
To our knowledge, the removal of 3-MCPD from water samples
56082, United States.
2
Present address: ACT I Research and Analytical Laboratory, 316 Kelly Drive.,
has not been investigated to date. In this paper, both hydrolysis
Waco, TX 76710, United States. and photolysis processes are examined as means of removing 3-

0045-6535/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.chemosphere.2009.01.053
1016 A.M. Nienow et al. / Chemosphere 75 (2009) 1015–1020

MCPD from water. Pseudo-first order rate constants are obtained experiments and solution preparations. A 133.2 lM 3-MCPD solu-
for hydrolysis and indirect photolysis, with hydrogen peroxide act- tion was prepared by dissolving 7.365 mg 3-MCPD in 500 mL of pH
ing as the direct photochemical reactant (at a wavelength of 8.9 borax buffer. Five milliliter aliquots of this solution were added
254 nm) and oxidizing agent. In addition, by-products formed dur- to a series of 15 mL tubes, and over time duplicate tubes were sac-
ing photolysis are identified and quantified, with the derivatization rificed for analysis by adding 1 mL PBA solution, heating and sub-
of 3-MCPD with phenylboronic acid (PBA) examined in some detail. sequently treating as described above in the derivatization
procedure prior to GC analysis.

2. Materials and methods 2.4. Photolysis experiments

2.1. Chemical preparation A 100 mL stock solution of 9.48 mM of 3-MCPD was prepared in
20% NaCl. For each photolysis experiment, 5 mL of this solution was
Neat (±)-3-chloro-1,2-propanediol was purchased from Sigma– dilute 1:100 with either de-ionized water (pH 5), pH 7 phosphate
Aldrich, with a >98% purity. Hexanes solvent (99.5% purity) used buffer, or pH 8.9 borax buffer for an initial concentration of 94.8 lM
for extraction was purchased from VWR Inc. Phenylboronic acid 3-MCPD for each experiment. Photochemical reactions were carried
(>97% pure) and sodium chloride (>99% pure), used in the deriva- out in series on each 500 mL solution in a Rayonet RPR-100 reactor
tization process, were purchased from Sigma–Aldrich. A 30% (Southern New England Ultraviolet Company). To accomplish this, a
weight by volume hydrogen peroxide solution, used to catalyze 500 mL solution was transferred to a 660 mL quartz reaction vessel
the photolysis reaction was obtained from Mallinkrodt. Potassium after addition of the desired amount of H2O2 and the vessel was
dihydrogen phosphate (KH2PO4), potassium hydrogen phosphate held in the center of the RPR-100 reactor directly over a magnetic
(K2HPO4), and sodium tetraborate decahydrate, use in preparing stir plate. A glass magnetic stir bar mixed the solution throughout
buffers, were obtained from Sigma–Aldrich Chemical Co. Phos- the irradiation time. Samples were irradiated with eight 35 W low
phate buffer at pH 7 was prepared by adding 4.20 g K2HPO4 and pressure mercury lamps that emit nearly monochromatic light at
2.93 g KH2PO4 to 8-L de-ionized water to give final concentrations 253.7 nm. The lamps were uniformly distributed around the vessel
of 0.00230 M and 0.00269 M, respectively. A borax buffer of pH 8.9 and each sample was irradiated for 620 min. The photon flux to the
was prepared by adding small amounts of 0.1 M HCl to a 0.025 M reaction vessel was measured by chemical actinometry to be
sodium tetraborate decahydrate solution until the desired pH 7.04  106 Einstein s1 (Nienow et al., 2008). At each sampling
was obtained. ACS reagent grade potassium oxalate monohydrate time, the reactor lamps were turned off for 1 min, and a 5 mL sam-
(99%), 1,10-phenanthroline (P99%), sulfuric acid (95–98%), ACS re- ple was removed with a volumetric pipette and transferred to a
agent grade salicylic acid (P99%), iron (III) sulfate hydrate (97%), 15 mL test tube. Radical reactions were immediately quenched by
ACS reagent grade glycine (P98.5%), and SigmaUltra grade EDTA addition of 1 mL PBA solution. After collecting the last sample, all
disodium salt dihydrate (99%) were all obtained from Sigma–Al- of the samples were subjected to the derivatization procedure prior
drich Chemical Co. and were used as received in experiments mea- to GC analysis. In addition to conducting experiments under light at
suring the photon flux. the three pH values, control experiments were conducted at each
pH value under the following conditions: (i) no irradiation and no
2.2. Standards and derivatization procedure added H2O2 (i.e., no treatment), (ii) irradiation without H2O2 addi-
tion (no added oxidant), and (iii) no irradiation with H2O2 addition
A stock solution of 3.34 mM 3-chloro-1,2-propanediol (3-MCPD) (no irradiation).
was prepared in de-ionized water. From this stock solution, 3-MCPD
standards were prepared by adding 0.1–1.25 mL stock into three 2.5. GC analysis
sets of five 25-mL volumetric flasks. The three sets of five flasks were
filled to the mark with either pH 8.9 borax buffer, pH 7 phosphate All sample extracts were analyzed on a Shimadzu GC17A gas
buffer, or de-ionized water. This yielded solutions at pH 8.9, 7, and chromatograph equipped with a flame ionization detector (GC–
5 at concentrations ranging from 0.0134 mM to 0.167 mM. FID) set to 300 °C, an AS 20 autosampler, and 30 m  0.25 mm 
Phenylboronic acid (PBA) was used as a derivatizing agent, fol- 0.10 lm film thickness (J&W Scientific DB-1 column). The carrier
lowing the methods of Rodman and Ross (1986), and Plantinga gas was helium at 2 mL/min flow rate with 35 mL/min nitrogen
et al. (1991). PBA was prepared by dissolving 52.64 g in 190 mL makeup gas. The 1075 split/splitless injector was held at 280 °C
acetone and 10 mL water to yield a 2.16 M solution. This solution and 1 lL of sample was injected. Splitless sampling time was set
was diluted with acetone to a concentration of 1.08 M. To deriva- to 0.80 min with a septa purge flow of 25 mL/min. During each
tize 3-MCPD in each standard or hydrolyzed/photolyzed sample, analysis, the oven was held at 50 °C for 2 min, ramped to 250 °C
5 mL of sample or standard and 1.0 mL PBA solution were added at 15 °C/min, and held at the final temperature for 1 min. Data
to a 15 mL test tube. The tubes were sealed, placed in a heating was processed using linear regression analysis based on concentra-
block preheated to 90 °C, and heated for 20 min. Upon cooling tion of analyte to peak area response.
slightly, 0.4 ± 0.05 g NaCl was added to each tube, and tubes were
re-sealed. After cooling to room temperature, 3.0 mL of hexanes 2.6. GC/MS analysis
were added to each tube. The tubes were mixed by vortex for
30 s, and after complete separation of the organic and water layers, Derivatized 3-MCPD was analyzed on a Thermo-Finnigan Po-
GC autosampler vials were filled with the hexane solutions. The laris Q Octapole/Ion-Trap GC/MS equipped with a J&W Scientific
concentrations of the three series of standards in hexane were DB5MS column (dimensions: 30 m  0.32 mm  0.25 lm). Carrier
0.0223–0.278 mM. The derivatization procedure will be discussed gas was helium set at 1.2 mL/min. The column temperature pro-
in more detail below. gram was set at an initial temperature of 35 °C, held for 4 min,
ramped to 280 °C with a final hold for 5 min. The 1075 split/split-
2.3. Hydrolysis experiments less injector was held at 250 °C, transfer line at 300 °C, and the ion
source at 200 °C. Splitless sampling time was set to 1 min with a
All hydrolysis experiments were conducted in a temperature septa purge flow of 25 mL/min. The MS was run in full scan mode
controlled room maintained at 24.1 ± 0.8 °C over the course of all (50–600 m/z).
A.M. Nienow et al. / Chemosphere 75 (2009) 1015–1020 1017

2.7. Data analysis OH OH O Cl


Cl OH + B B
To compare the effects of the different reaction variables, a OH O
pseudo-first order rate constant was calculated for each experi-
ment via weighted linear least squares regression, regressing Scheme 1.

ln[3-MCPD] against time. Due to the natural log transformation


of the variable, 1=r2y is the appropriate weighting factor, where acted phenylboronic acid; therefore lower concentrations were
r2y is the variance on [3-MCPD]. Because this term is proportional assessed for reaction yield. A series of 5 mL 3-MCPD samples
to the square of the 3-MCPD concentration, [3-MCPD]2 was used (0.204 mM in 20% NaCl) were prepared and 1 mL of PBA solution
as the weighting factor on each datum point (Green and Margeri- in acetone was added. The concentrations of PBA varied from
son, 1977). The use of these weighting factors effectively reduces 0.054 to 2.16 M. These samples were then derivatized as described
the ‘‘weight” of later time points where the concentration of 3- above. Fig. 2a shows the relative yield of 3CPDP in hexane as a
MCPD is low and where the associated error on the transformed function of PBA concentration. Based on these results, all kinetic
variable is large. experiments were performed with 1.08 M phenylboronic acid as
described in the derivatization procedure section.
3. Results and discussion According to Plantinga et al., addition of sodium chloride to the
aqueous derivatized compound increases the extraction efficiency
3.1. Derivatization (Plantinga et al., 1991). In that work, a concentration of 12–20% by
weight NaCl was deemed necessary for optimal extraction effi-
The derivatization of 3-MCPD with phenylboronic acid was ad- ciency. Because we changed the PBA concentration for derivatiza-
justed to maximize the recovery of the derivatized compound in tion, the extraction efficiency as a function of NaCl concentration
the hexane layer using GC–MS analysis to verify identity and pur- was re-examined. For this purpose, five 9.49  102 mM 3-MCPD
ity. GC–MS chromatograms of the derivatized 3-MCPD indicated samples were derivatized with two samples receiving either 0.2%
only one peak, and the mass spectra of this peak (Fig. 1) corre- or 20% NaCl prior to the addition of PBA. For the other three sam-
sponds to 3-chloropropanediol phenylboronate (3CPDP) (Scheme ples the NaCl (5%, 10% and 15% NaCl) was added after derivatiza-
1) and matches that observed by Rodman and Ross (1986). The tion and before extraction with hexane. Fig. 2b shows the
parent compound peak is observed at the m/z ratio of 196. The base relative recovery of 3CPDP as a function of NaCl concentration.
peak at 147 corresponds to a loss of CH2Cl. Fig. 2b indicates that good recovery occurred at NaCl concentra-
Phenylboronic acid concentrations used in our early experi- tions P10% and that the order of NaCl addition does not dramati-
ments (2.16 M) resulted in overloading the FID detector with unre- cally affect recovery. For the hydrolysis experiments, 0.4 ± 0.05 g

147.1
100
O +
95 B
O
90

85

80

75

70

65

[C7H7]+
Relative Abundance

60

55
91.1
50

45

40 O +
Cl
B
35
O
30 146.1 196.0
25
+
B
20 -
O
15
104.1
10 105.2 148.1 198.0
77.1 195.0
92.1
5 50.9
65.2 89.2 117.2 123.2 149.1 165.1
181.1 199.0 221.2 243.2 266.2 277.2
0
60 80 100 120 140 160 180 200 220 240 260 280
m/z

Fig. 1. GC–MS of derivatized product (i.e., 3-chloropropanediol phenylboronate). Peak at 196.0 m/z corresponds to the molecular weight peak.
1018 A.M. Nienow et al. / Chemosphere 75 (2009) 1015–1020

a 350000
Table 1
Hydrolysis rate constants and calculated hydroxide ion concentrations.

300000 pH T (°C) kobsa (s1) [OH]b (M) kobsc (M1 s1)


5.5 85 4.06  106 9.86  108 3.71  106
250000 7.0 85 9.52  105 3.12  106 11.7  105
8.0 85 1.24  103 3.12  105 1.17  103
Area Count

9.0 85 9.79  103 3.12  104 11.7  103


200000 9.0 65 1.13  103 1.25  104 1.08  103
9.0 45 6.59  105 4.04  105 6.72  105
150000 a
Reported by Dolezal and Velisek (1992).
b
Calculated from pH with the temperature-appropriate Kw value calculated with
100000 Eq. (4).
c
Calculated from pH, and the calculated values of Kw (T), A, and Ea.
50000

0 indicating the pH was measured at the experimental temperature


0 0.5 1 1.5 2 2.5 (Dolezal and Velisek, 1992). Because pH (i.e., hydrogen ion activity)
Concentration of Phenylboronic Acid (M) was experimentally measured and not [OH], we assumed that the
non-unity dependence on pH found by Hamlet and Sadd is largely
b 160000 due to the temperature dependence on the dissociation constant
(i.e., ion product) of water, Kw, and therefore the data was re-eval-
140000
uated assuming the following elementary reaction rate law holds,
120000 d½3CPD
¼ kOH  ½OH   ½3CPD ð2Þ
dt
100000
Area Count

where the ‘‘observed” pseudo-first order rate constant at constant


80000 pH and temperature is defined by kobs = kOH  [OH], or
logðkobs Þ ¼ logðkOH Þ þ log½OH  ð3Þ
60000
Because no buffers were apparently used by Dolezal and Velisek,
40000 and 3-MCPD is a neutral compound, it was further assumed that
the ionic strength of their solutions was low, such that activity coef-
20000 ficient corrections are not necessary (Dolezal and Velisek, 1992).
Harned and Owen (1958) calculated the following temperature
0
0 5 10 15 20 25 (T) dependence on Kw:
Concentration of NaCl (% weight) log K w ¼ 4470:99=T þ 6:0875  0:01706T ð4Þ
Fig. 2. (a) Area counts from FID detector for 3-chloro-1,2-propanediol phenylbor- where T has units of Kelvin (Harned and Owen, 1958). Because
onate as a function of phenylboronic acid concentration. Initial concentration of Kw = [H+][OH], the values of [OH] were calculated at each pH
3CPD was 0.204 mM. NaCl (20 wt%) was added during derivatization. (b) Area
counts from FID detector for 3-chloro-1,2-propanediol phenylboronate as a
and temperature combination, and are reported in Table 1. For
function of NaCl. Initial concentration of 3CPD was 9.49  102 mM; 1.079 M PBA Dolezal and Velisek’s data at 85 °C, Fig. 3a shows a plot of log (kobs)
was used during derivatization. versus log [OH] (i.e., Eq. (3)) (Dolezal and Velisek, 1992). Linear
regression, with both the slope and intercept used as fitting param-
eters, results in a slope of 0.977 confirming that the reaction is first-
(8% by weight) was added to the derivatized samples prior to order in [OH]. Forcing the slope to be 1.0 (i.e., [OH]1), gives an
extraction and for the photolysis experiments NaCl was added intercept of 1.57 (= log (kOH)). The resulting one parameter regres-
prior to irradiation. sion line is shown in Fig. 3a.
The values of kobs at the different temperatures at pH 9 can now
3.2. Hydrolysis be used to calculate the values of the pre-exponential factor, A, and
activation energy, Ea. For the temperature dependent data in Table
The degradation of 3-MCPD by hydrolysis was first examined by 1, the values of kOH were calculated from the respective value of
Dolezal and Velisek who determined hydrolysis rate constants at a kobs and Kw. Regression of ln (kOH) versus 1/T, according to the
series of pH values and temperatures (Dolezal and Velisek, 1992). Arrhenius equation,
These pseudo-first order hydrolysis rate constants are provided
lnðkOH Þ ¼ lnðAÞ þ Ea =RT ð5Þ
in Table 1; however, no attempt was made by the original authors
12
to resolve the elementary rate constant from these data. Hamlet results in the line shown in Fig. 3b, where A = 2.27  10 M s1 1

and Sadd evaluated these data and derived the following expres- and Ea = 73.9 kJ mol1. To evaluate the accuracy of the overall mod-
sion for the observed first order rate constant as a function of pH: el, corresponding values of kobs calculated at each respective pH –
temperature combination, from A, Ea and the temperature depen-
K ¼ ð107:73347þ0:83775pH ÞðeEa =RT Þ s1 ð1Þ
dent value of Kw are reported in the final column of Table 1. These
1
where Ea, the activation energy, is 119.2 kJ mol (Hamlet and Sadd, rate constants have an average relative error from the correspond-
2002). The data and this rate constant indicate that within this pH ing measured values of 10%, yet range over three orders-in-magni-
range, the elementary reaction is catalyzed by OH, although the tude in absolute value.
non-unity coefficient on pH would suggest the order of the reaction Extrapolating the rate of hydrolysis to our experimental condi-
is not first-order in OH concentration. In each hydrolysis experi- tions of 24.1 °C and pH 8.9, results in an observed rate constant
ment conducted by Dolezal and Velisek, the pH was held constant (kobs) of 1.72  106 s1, or a half-life (t1/2) of approximately
by automatically titrating in NaOH at the designated temperature, 112 h. Fig. 4 compares the rate of decay predicted with this rate
A.M. Nienow et al. / Chemosphere 75 (2009) 1015–1020 1019

a Velisek. Given the differences between the experimental protocols,


-2 and the extent of the extrapolation to our lower temperature value,
the difference of about a factor of 2 in the rate constant is reason-
able. In addition, despite the small differences in rate constant,
-3 both our work and that of Dolezal and Velisek show that 3-MCPD
log (kobs)

in water samples will be removed, albeit slowly, due to hydrolysis.

-4
3.3. Photolysis

-5
Because 3-MCPD does not absorb light at a wavelength of
254 nm, it can be expected that direct photolysis at this wave-
length will not occur. Fig. 5 indicates that this is the case as irradi-
-6 ation of 3-MCPD in the absence of H2O2 at pH 5 results in
-8 -7 -6 -5 -4 -3 essentially no chemical loss over a time period of 15 min. However,
-
log [OH ] addition of H2O2 results in indirect photolysis via photochemical
formation of hydroxyl radicals that react with 3-MCPD and inter-
b 4 mediate products. Hence, experiments were conducted to deter-
mine the rates of indirect photochemical transformation with
added H2O2 at pH values spanning the range of most natural
3 waters. Fig. 5 displays the data collected at pH 5, indicating that
the 0.1 mM 3-MCPD in buffered solutions rapidly decomposed
within 15 min with millimolar addition of H2O2. Also noted is the
ln (kOH)

2 lack of significant decomposition without the addition of H2O2 or


when the experiment was run in the dark. At pH 5 with 1 and
5 mM H2O2, the half-life of 3-MCPD was 1.5 and 0.84 min, respec-
1
tively (corresponding rate constants k = 7.63  103 s1 and
k = 1.37  102 s1). With 5 mM added H2O2, the reaction rate
was not greatly affected by pH over the pH range 5–8.9. A slightly
0
faster reaction rate was measured at pH 7 (k = 1.56  102 s1,
0.0027 0.0029 0.0031 t1/2 = 0.74 min) than at pH 5 (k = 1.37  102 s1, t1/2 = 0.84 min)
or 8.9 (k = 1.07  102 s1, t1/2 = 1.08 min).
1/T
During the indirect photochemical reaction with OH, assuming
Fig. 3. (a) [OH]-dependence of the observed first-order hydrolysis rate constant, the reaction goes to complete mineralization the final products of
and (b) the Arrhenius equation regression of kOH from 45° to 85°. (Data from Dolezal OH attack on 3-MCPD would be CO2, H+, and Cl. To assess
and Velisek (1992)). whether complete oxidation occurs in the UV/H2O2 system,
91.0 lM 3-MCPD with 5 mM H2O2 at an initial pH of 6.2 was irra-
diated at 254 nm. The solution was prepared without buffer so that
constant to our experimental data. As can be seen from the figure,
the pH could be monitored as a function of irradiation time. The
the pseudo first-order rate constant obtained directly from our
sample collected at each sampling time was split into subsamples,
data (kobs = 7.42  107 s 1 or t1/2 = 260 h) is about a factor of 2
with one subsample analyzed via GC–FID to determine the 3-
smaller than the value calculated using the rate law and associated
MCPD concentration, one subsample analyzed via IC to determine
relationships presented in Eqs. (2)–(5) with the data of Dolezal and
the Clconcentration, and the final subsample used to measure pH.

140
1.2

2
[3CPD] Normalized to [3CPD] t=0

120 R = 0.8024 1

100 0.8
[3CPD] /µM

80 0.6

[H2O2] = 0 mM, DC
60 0.4
[H2O2] = 0 mM, UV
[H2O2] = 1 mM
40 0.2 [H2O2] = 5 mM

20 0
0 5 10 15
Irradiation Time / min
0
0 75 150 225 300 375 450 525 Fig. 5. Experimental photolysis data for four experiments conducted at pH 5. In the
Time / h graph, all 3CPD concentrations are normalized to the initial 3CPD concentration for
each respective run. The  represent 3CPD concentrations as a function of time for a
Fig. 4. Hydrolysis of 3CPD. The  represents the experimental data for hydrolysis at direct photolysis experiment and h are the 3CPD concentrations for the same
pH 5 and 24.1 °C, and the solid line is an exponential fit to the data. The dotted line experimental conditions but without the UV light. are the 3CPD concentrations
represents the predicted reaction rates based on the work of Dolezal and Velisek as a function of irradiation time with 1 mM H2O2 added and  are the 3CPD
(1992). concentrations with 5 mM H2O2 added.
1020 A.M. Nienow et al. / Chemosphere 75 (2009) 1015–1020

7 remove the 3-MCPD from water. It has been shown that the UV/
H2O2 process takes 3-MCPD to nearly complete conversion within
6 minutes whereas hydrolysis at reasonable temperatures takes sev-
eral months.
pH

4 Acknowledgements

3
The authors gratefully acknowledge the financial support of Ad-
vanced Concepts and Technologies, International (ACT I), Waco, TX
120 and the US Army Tank Automotive Research, Development and
Engineering Center (TARDEC) for sponsoring this research.
100
References
] /µM

80
Boden, L., Lundgren, M., Stensio, K.E., Gorzynski, M., 1997. Determination of 1,3-
-
[3CPD] or [Cl

dichloro-2-propanol and 3-chloro-1,2-propanediol in papers treated with


60 polyamidoamine–epichlorohydrin wet-strength resins by gas chromatography
mass spectrometry using selective ion monitoring. J. Chromatogr. A 788, 195–
203.
40 Chung, W.C., Hui, K.Y., Cheng, S.C., 2002. Sensitive method for the determination of
1,3-dichloropropan-2-ol and 3-chloropropane-1,2-diol in soy sauce by capillary
[3CPD]
gas chromatography with mass spectrometric detection. J. Chromatogr. A 952,
20 185–192.
[Cl-] Codex Committee on Food Additives and Contaminants (CCFAC), 2001a. Position
Paper on chloropropanols: CX/FAC 01/31, 33rd Session the Hauge, the
0 Netherlands, <http://www.codexalimentarius.net/ccfac33/fa01_01e.htm>
0 2 4 6 8 10 (accessed 18.12.08).
Dolezal, M., Velisek, J., 1992. Kinetics of 3-chloro-1,2-propanediol degradation in
Time / min model systems. Chem. React. Foods II 2, 297–302.
Dow, 2006. Product safety assessment: epichlorohydrin, Form Number 233-00239-
Fig. 6. Photolysis by-products; } illustrates changes in pH, d illustrates the decay KC-0406 http://www.dow.com/PublishedLiterature/dh_0073/0901b80380073-
of 3-chloro-1,2-propanediol, and 4 illustrates the increase in Cl. The remaining fdb.pdf?filepath=productsafety/pdfs/noreg/233-00239.pdf&fromPage=GetDoc.
carbon concentration and speciation were not obtained. Green, J.R., Margerison, D., 1977. Statistical Treatment of Experimental Data.
Elsevier Scientific Publishing Company, New York.
Hamlet, C.G., 2008. Chloropropanols and their esters. In: Senyuva, J.G.A.H. (Ed.),
The results of this experiment are shown in Fig. 6. As in the previ- Bioactive Substances in Foods – Natural and Man-Made. Wiley-Blackwell,
ous UV/H2O2 experiments (e.g., Fig. 5), the 3-MCPD concentration Bognor Regis.
Hamlet, C.G., Sadd, P.A., 2002. Kinetics of 3-chloropropane-1,2-diol (3-MCPD)
dropped rapidly to concentrations below the detection limit. The
degradation in high temperature model systems. Eur. Food Res. Technol. 215,
chloride concentration increased rapidly during the first 7 min of 46–50.
irradiation, and then reached a plateau at a concentration approx- Hamlet, C.G., Sadd, P.A., Crews, C., Velisek, J., Baxter, D.E., 2002. Occurrence of 3-
imately the same as the initial 3-MCPD concentration. This sug- chloro-propane-1, 2-diol (3-MCPD) and related compounds in foods: a review.
Food Addit. Contam. 19, 619–631.
gests that Cl was nearly quantitatively formed over the course Harned, H.S., Owen, B.B., 1958. The Physical Chemistry of Electrolytic Solutions. Van
of the reaction. Similarly, the pH of the solution dropped rapidly Nostrand Reinhold, New York.
to pH 4.0, which is near the predicted pH based on reaction stoi- Hauschild, V.D., Bratt, G.M., 2005. Prioritizing industrial chemical hazards. J. Toxicol.
Environ. Health A 68, 857–876.
chiometry. Hence both H+ and Cl formation indicate nearly com- Masten, S.A., 2005. 1,3-Dichloro-2-propanol [CAS No. 96-23-1], review of
plete conversion to final products within minutes of irradiation. toxiciological literature. A report prepared for National Toxicology Program
Due to limits in time and instrument availability, the remaining (NTP), National Institute of Environmental Health Services (NIEHS), National
Institute of Health, US Department of Health and Human Services. <http://ntp-
carbon species were not definitively identified. It is possible that server.niehs.nih.gov/ntp/htdocs/Chem_Background/ExSumPdf/dichloropropanol.
complete mineralization of 3-MCPD to CO2 did not occur and that pdf> (accessed 18.12.08).
other carbon containing intermediate compounds, such as glycerol Matthew, B.M., Anastasio, C., 2000. Determination of halogenated mono-alcohols
and diols in water by gas chromatography with electron-capture detection. J.
(Hamlet et al., 2002), were formed. As glycerol is nearly non-toxic
Chromatogr. A 866, 65–77.
(LD50 >20 mL/kg; orally in rats) (Merck Index 11th Edition), it can Nienow, A.M., Bezares-Cruz, J.C., Poyer, I.C., Hua, I., Jafvert, C.T., 2008. Hydrogen
be concluded that the conversion of 3-MCPD under these condi- peroxide-assisted UV photodegradation of Lindane. Chemosphere 72, 1700–
tions likely leads to the formation of non-toxic compounds. 1705.
Nyman, P.J., Diachenko, G.W., Perfetti, G.A., 2003. Survey of chloropropanols in soy
sauces and related products. Food Addit. Contam. 20, 909–915.
4. Conclusion Plantinga, W.J., Van Toorn, W.G., Van der Stegen, G.H.D., 1991. Determination of 3-
chloropropane-1, 2-diol in liquid hydrolyzed vegetable proteins by capillary gas
chromatography with flame ionization detection. J. Chromatogr. 555, 311–314.
Due to concerns about chlorinated by-products from wastewa- Rodman, L.E., Ross, R.D., 1986. Gas–liquid chromatography of 3-chloropropanediol.
ter treatment and the recent publication from the US Army Center J. Chromatogr. 369, 97–103.
Sarzanini, C., Bruzzoniti, M.C., Mentasti, E., 2000. Determination of epichlorohydrin
for Health Promotion and Preventative Medicine, (Hauschild and
by ion chromatography. J. Chromatogr. A 884, 251–259.
Bratt, 2005) the removal of 3-MCPD from water samples needs to Summary of the 57th Meeting of the Joint FAO/WHO Expert Committee on Food
be addressed. The work presented here, the first systematic exam- Additives (JECFA), 2001b. Joint FAO/WHO Expert Committee on Food Additives,
p. 20, Rome.
ination of 3-MCPD chemistry in water, shows that 3-MCPD can be
Xing, X., Cao, Y., Wang, L., 2005. Determination of rate constants and activation
removed from water by both hydrolysis and photolysis methods. energy of 3-chloro-1,2-propanediol hydrolysis by capillary electrophoresis with
Of the methods examined, UV/H2O2 provides the quickest way to electrochemical detection. J. Chromatogr. A 1072, 267–272.

You might also like