Vous êtes sur la page 1sur 7

Chemical Engineering Journal 284 (2016) 233239

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Kinetics study of acrylic acid polymerization with a microreactor


platform
Lu Qiu, Kai Wang , Shan Zhu, Yangcheng Lu, Guangsheng Luo
The State Key Lab of Chemical Engineering, Department of Chemical Engineering, Tsinghua University, Beijing 100084, China

h i g h l i g h t s

g r a p h i c a l a b s t r a c t

 Continuous free-radical

polymerization of acrylic acid was


carried out.
 Reaction time was easily adjusted by
assembling different reaction tubes.
 The platform is fit for fast
polymerization at relatively high
temperatures.

a r t i c l e

i n f o

Article history:
Received 24 March 2015
Received in revised form 9 July 2015
Accepted 14 August 2015
Available online 22 August 2015
Keywords:
Poly acrylic acid
Microreactor
Kinetics
Free radical polymerization

a b s t r a c t
A microreactor platform with easily assembled reaction tubes is developed to study the kinetics of acrylic
acid polymerization in aqueous solutions. In this platform, two micromixers are used to start and terminate the polymerization reaction, and several coiled reaction tubes embedded in water baths control the
residence time. Operated at a total flow rate greater than 40 mL/min, the reaction is independent of the
monomer and initiator mixing. Experimental data showed that the orders of acrylic acid and potassium
persulfate were 1.5 and 0.5, respectively, with an activation energy of 67.4 kJ/mol, which is in accord with
the classical kinetic mode reported in the literature. The molecular weight of polyacrylic acid
(1.03  1051.76  105) also indicated that the microreactor platform works well with macromolecules.
Therefore, this technique will expand the viable methods applicable for kinetic studies.
2015 Elsevier B.V. All rights reserved.

1. Introduction
The application of microreactors in chemistry and chemical
engineering has garnered significant popularity over the last
15 years [1,2], since microreactors are superior to traditional reactors in many aspects. Characterized with a high surface-to-volume
ratio and short mass transfer distance, microreactors enable efficient heat exchange and reactant mixing [3,4], where rapid and
strongly exothermic reactions can be conducted isothermally [5].
The ability of microreactors to ensure the safety of dangerous

Corresponding author.
http://dx.doi.org/10.1016/j.cej.2015.08.055
1385-8947/ 2015 Elsevier B.V. All rights reserved.

chemical processes has also been exploited [6,7]. Free radical polymerization is a reaction with a high reaction enthalpy (e.g.,
DH = 66.9 kJ/mol for the polymerization of acrylic acid) [8] and
ultra-fast chain propagation rate [9]. Generally, most free radical
polymerizations are carried out in batch reactors, which often
gives products with broad molecular weight distributions if living
radical control methods are not used, such as atom transfer radical
polymerization (ATRP), reversible additionfragmentation chain
transfer polymerization (RAFT), and nitroxide mediated living radical polymerizations (NMP) [2]. The poor temperature control and
bad mixing performance in batch reactors has been sufficiently
demonstrated; thus, it is promising to use more efficient reactors
for polymerization reactions [10,11].

234

L. Qiu et al. / Chemical Engineering Journal 284 (2016) 233239

Recently, microreactors have been proved to be suitable for


many polymerization processes, including free radical [12], cationic [13], anionic [14], and polycondensation [15]. In 2005, Iwasaki
and Yoshida conducted a study on the free radical polymerization
of many monomers in a capillary microreactor, including butyl
acrylate (BA), benzyl methacrylate (BMA), methyl methacrylate
(MMA), vinyl benzoate (VBz), and styrene (St) [16]. Narrow molecular weight distributions were obtained for some polymers with
high reaction enthalpies, such as BA, BMA, and MMA, which gave
important evidence for the strong heat-removing ability of
microreactors. Besides using the capillary wall as a heat exchange
medium, multi-phase microreactors are also good devices for
highly controllable suspension polymerizations which utilize
microdroplets as reaction units [17]. There are also heated discussions regarding living radical polymerizations such as ATRP [18]
and RAFT [19] in microreactors. Besides the polymerization
method, the microreactor structure is also critical for implementing polymerization. Parida et al. improved the residence time distribution of reactants using coiled reaction tubes [20]. An
increase in monomer conversion and a decrease in the polydispersity index (PDI) of the molecular weight of the product were
observed. Other detailed works including monitoring methods
[21], heat and transfer capacities [22] were also reported.
Since free radical polymerization without radical dormant species is a fast reaction, the reactor design and control strongly
depend on the kinetic parameters of the reaction. For tubular
microreactors, the reaction rate directly determines the length of
the reaction tubes and the measured kinetics also shed light on
the mixing performances in the microreactors [23]. Recently, several methods have been developed to investigate the kinetics of
small molecular reactions using microreactors. For example,
McMullen and Jensen suggested several automatic microreactor
platforms for the rapid determination of reaction kinetics using
statistical analysis [24,25]. Wang et al. investigated the reaction
kinetics of aniline and benzoyl chloride using capillary microreactors [26]. These studies show that microreactors are good tools for
intrinsic reaction kinetics studies under good operating conditions,
when the effects of reactant mixing have been eliminated and the
reaction time is well controlled [27,28]. Comparing with the small
molecular reactions, there are few kinetic studies regarding free
radical polymerization in microreactors. Previous studies conducted in batch reactors indicated that the kinetics of polymerization determined the properties of the resulting polymers [29];
thus, if the reaction rate in a microreactor is not in accord with
intrinsic reaction kinetics, it is hard to believe that the microreactor has any advantages in regard to mixing and temperature control. Confined by the mixing and sampling techniques in batch
reactors, reported kinetic data of acrylic acid polymerization
primarily entails relatively mild operating conditions (lower
than 70 C) on a minute time scale [30,31]. Therefore, a simple
and robust method to measure the kinetics of free radical
polymerization reactions in microreactors will offer significant
advantages for the development of polymerization techniques
and equipment.
In this work, a tubular microreactor platform with easily assembled reaction tubes was designed to investigate the free radical
polymerization of acrylic acid (AA) in aqueous solutions; this system was selected as a model owing to its well understood kinetics
parameters [31]. The effects of flow rate, AA monomer concentration, initiator concentration, and reaction temperature on the AA
conversion were carefully examined. Compared to the reported
kinetic models [31], the polymerization reaction rate in the
microreactor was very similar to the intrinsic reaction rate under
high flow rate conditions, and the advantages of the microreactor
were obvious in regard to the kinetics measurements and the
molecular weight of the reaction products.

2. Reaction mechanism of AA polymerization


Free radical polymerization primarily consists of three elementary reactions including chain initiation, propagation, and termination. In the chain initiation step, an initiator produces two free
radicals, which combine with monomers to form monomer radicals. Usually, the chain initiation is the rate controlling step [32],
and the initiation step can be expressed as

K2 S2 O8 ! 2KSO4 ; CH2 @CHX KSO4 ! RCH2 CH  X

where R denotes initiator residue and X denotes ACOOH. In the


chain propagation step, monomers quickly combine to form chains,
and heat is released.

RCH2 CH  X CH2 @CHX ! RCH2 CHXCH2 CH  X


RCH2 CHXCH2 CH  X CH2 @CHX !    ! RAAn CH2 CH  X

where (AA)n denotes the units of repeated AA. In the chain termination step, the free radicals combine with each other, forming dead
polymers.

RAAn CH2 CH  X RAAm CH2 CH  X ! RAAmn2 R

For the free radical polymerization of AA, the overall polymerization rate can be expressed as

Rp 

dAA
k  AAn1 KBSn2
dt

where [AA] is the concentration of acrylic acid and [KBS] is the concentration of K2S2O8. The purpose of this work was to determine
kinetic parameters k, n1, and n2 using a microreactor platform.
3. Method and experimental setup
3.1. Microreactor system
In kinetics measurements, it is necessary to change reaction
time. Adjustment of the flow rate of the reactant solutions in a single tube length helps change the residence time, but it will also
affect the mixing performance in micromixers [33,34]. To eliminate the effect of flow rate variation on reactant mixing, the reaction time should be adjusted using tubes of different lengths. A
new microreactor platform facilitating tube length variation with
selective switch-on valves was developed in this study, and a schematic is shown in Fig. 1. The microreactor platform consists of 5
individual parts in water baths, including a mixing system used
to pre-heat and mix the monomer and initiator solutions in water
bath 1, two tubular systems used to adjust the reaction time in
water baths 2 and 3, an additional tubular system to extend the
reaction time in water bath 4, and a quenching system used to terminate the reaction in water bath 5.
In the mixing system, a T-junction micromixer (316-L stainless
steel) with three 1 mm diameter holes is used to mix the monomer
and initiator solutions. The micromixer is connected to the tubes to
pre-heat the reactants and implement the polymerization reaction
with 1/16 in. outer diameter capillaries, whose inner diameters (dc)
are also 1 mm (Fig. 1). The distance from the T-junction to the end
of the 1/16 in. capillary is 4.55.5 cm. Although this is negligible
compared to the length of the subsequent reaction tubes, it is sufficient for the mixing of the aqueous solutions, according to our
previous conclusions on the mixing performances of this type of
micromixer [33,34]. The temperature of the reactant solutions is
controlled by the pre-heating tubes (2 mm inner diameter and
3 m length) in water bath 1 from 80 to 95 C. The quenching system has a similar structure as the mixing system (Fig. 1), which
contains a 1 mm inner diameter micromixer and a pre-cooling

235

L. Qiu et al. / Chemical Engineering Journal 284 (2016) 233239

Fig. 1. Schematic of the microreactor platform and pictures of main elements.

tube of water (2 mm inner diameter and 2 m length), connected


with 1/16 in. capillaries. The reaction is quickly stopped in the last
coil tube (2 mm inner diameter and 3 m length) using cold water
(less than 5 C).
The tubular systems are used to control the residence time. The
tubular systems in water baths 2 and 3 are comprised of four 2- or
3-meter-long coiled stainless steel tubes (316-L), whose inner and
outer diameters are 2 and 3 mm, respectively, as shown in Fig. 1.
Two 3-way and two 5-way switching valves are used to connect
the coiled tubes, which enable the tube lengths to be altered during
the experiment. Since the 3-way and 5-way valves can only switch
on two connectors and close the other connectors, 9 tube volumes
are used in the kinetic study experiment (V1V9 in Table 1) with
the coupling of these values. The total length of the tubes is
22.2 m and the percentage of straight tubes is less than 4%. Therefore, the reaction is mainly carried out in the coiled tubes. The
additional tubular system in water bath 4, shown by dashed lines
in Fig. 1, can further extend the reaction time under high reactant
flow rates (40 and 60 mL/min). The length of the coiled tube is
16 m and with the combination of the coiled tubes in water baths
2 and 3, five more tube volumes are obtained, namely V10V14 in
Table 1. The additional tubular system is used to prove that the
reactants have been suitably mixed under the high flow rate, but
it is not used in the kinetic measurements, since the tubes in water
baths 2 and 3 provide AA conversions from 0% to 80%. The dark line
in Fig. 1 shows a condition with 4 consecutive tubes. The reaction

temperature in the tubular systems is controlled by the water


baths from 80 to 95 C. According to an analysis by Lorber et al.
[35], microreactors are suitable for kinetic studies at reaction temperatures higher than 70 C. Lower temperatures were not considered in this study, as they are more suitable for common batch
reactors. The reaction temperature is also monitored by on-line
sensors (PT-100) placed between water baths; the temperature difference between the sensors and the water bath control instrument was less than 0.5 C. All tubes as well as valves and sensors
outside the water baths were coated with polyvinylchloride
(PVC) foam to prevent heat loss.
3.2. Materials and equipment
The T-junction micromixers in the microreactor system were
purchased from Valco Instruments Company Inc., USA. The coiled
tubes and valves were purchased from Beijing Xiongchuan Co.,
Ltd., China. The reactant solutions were fed by metering piston
pumps, which were provided by Beijing Satellite Co. Ltd., China.
Analytically pure AA, K2S2O8 (KBS), hydroquinone, KBr, KBrO3,
HCl, KI, Na2S2O3, K2CO3, K2Cr2O7, were utilized, which were purchased from Beijing Modern Eastern Fine Chemical Co., Ltd., China.
99.9% + N2 was provided by Beijing Huayuan Gas Co., Ltd., China.
Ultrapure water was obtained from a machine purchased from
Lab Instrument and Technology Co., Ltd., USA. A rotary evaporator
purchased from Heidolph Laborota, Germany was used to purify

Table 1
Tubes volumes in the microreactor platform.*
Tube volumes from the first to the second T-junction (mL)

V1

V2

V3

V4

V5

V6

V7

6.06
V8

12.1
V9

18.4
V10

24.7
V11

30.9
V12

40.8
V13

50.6
V14

60.3

69.7

79.1

88.5

98.0

110.5

123.1

The volumes were calibrated by feeding water into the microreactor platform and then measuring the water volume, which was pressed out by N2 under each valve
condition.

236

L. Qiu et al. / Chemical Engineering Journal 284 (2016) 233239

AA prior to the experiments. An automatic titrator (T50) purchased


from Mettler Toledo, Germany was used to measure the concentration of AA in the aqueous solutions. A gel permeation chromatograph (GPC) and standard polyacrylic acid (PAA) samples
purchased from Waters Co., Ltd., USA were used to measure the
molecular weight distribution of PAA.

where [Br2] and VBr2 are the concentration and volume of Br2; [Na2S2O3] is the concentration of Na2S2O3; Vexperiment and Vblank are the
volumes of the Na2S2O3 titrating samples and blank solution; Vsamples is the volume of the samples collected from the microreactor
platform, and QAA, QKBS, and QCW are the flow rates of AA, K2S2O8
solution, and cold water. According to the measured AA concentrations, the AA conversion can be expressed as

3.3. Operation and analysis



AAQ AA Q KBS
 100%
X 1
AA0 Q AA

At the beginning of the experiment, 5 g hydroquinone was


added to 500 g crude AA and then the mixture was vaporized at
70 C in the rotary evaporator. The purified AA vapor was condensed in the collecting bottle of the rotary evaporator and was
dissolved in ultrapure water, which was deoxygenated by bubbling
N2 for 1 h. K2S2O8 was used as initiator, which was also dissolved in
deoxygenated water. The initial concentrations of AA ranged from
0.35 to 1.40 mol/L and the initial concentrations of K2S2O8 were
varied from 0.028 to 0.112 mol/L. The viscosity of the AA solution
ranged from 0.91 to 0.93 mPa s and the viscosity of the K2S2O8
solution was around 0.90 mPa s at room temperature 25 C. Therefore the mixing of the reactants solutions in the micromixer should
be similar to the aqueous solutions we used in previous works to
investigate the mixing performance [33,34]. Guided by these conclusions, solutions of AA and K2S2O8 that were pumped into the
first T-junction micromixer could be sufficiently mixed in the
1 mm flow channel downstream of the T-junction. The polymerization reaction began in the tubular systems. In the coiled tube, the
reaction time was controlled by the flow rate and tube volume.
According to Serra et al. [36], there are strong circulating secondary
flows in coiled tubes, which maintain a uniform concentration distribution in the cross-section of the tube; this is not the case in
straight reaction tubes. Serra et al. also found that coiled reaction
tubes provided a uniform concentration field for free radical polymerization reactions [37]. Based on this analysis, the residence
time is used as the reaction time in this study. At the end of the
platform, the reacting solution and 10 mL/min cold water (5 C,
not deoxygenated) were mixed in the second micromixer. Through
rapid mixing with cold water, the solution temperature could be
reduced from 95 to 76 C based on the energy conservation of fluids; with the help of the fast heat exchange between the tube and
the cooling water in water bath 5, the polymerization was quickly
terminated downstream of the micromixer. The fluid temperature
out of the cooling system was less than room temperature (25 C).
The platform was flushed for at least 10 min with deoxygenated
water before the reactant was pumped in; a stabilization time that
was 3 times the residence time was required when the tube volume was changed. Samples were collected at the outlet after the
stabilization time was reached.
The variation in AA concentration was used to characterize the
polymerization kinetics in this study; it was measured using an
automatic titrator. The concentration of the AA monomer was
quantified by iodimetry. Prior to the titration, 7 mL of KBrKBrO3
solution (0.05 mol/L KBrO3 and 0.34 mol/L KBr) was added to the
samples (1 mL), and was acidified by addition of HCl (about
6 mol/L) and kept for 30 min in the dark. During this process, the
generated Br2 reacted with the double bond in AA. Afterwards,
3 mL KI (1 mol/L) was added to replace the remaining Br2, then
the generated I2 was quantified using the Na2S2O3 (0.1 mol/L) solution. Since unreacted K2S2O8 can react with KI, blank experiments
without KBrKBrO3 must be carried out. Samples were treated
with the KI solution directly and then titrated with Na2S2O3. Therefore, the concentration of AA in the reacting solution could be
expressed as:

AA

5KBrO3 V KBrO3  0:5Na2 S2 O3 V experiment  V blank


V sample Q AA Q KBS =Q AA Q KBS Q CW

where [AA]0 is the acrylic acid concentration in the original solution. The molecular weight of the PAA product was analyzed by
GPC. Ultrapure water was used as the eluent at a flow rate of
1 mL/min. The column temperature was 40 C. Standard PAA samples with average molecular weights (Mn) from 103 to 106 were
used for the calibration.

4. Results and discussions


4.1. Effect of flow rate on mixing
In most microreactors, mixing performance is determined by
the reactant flow rate [33,34], because of the passive mixing mechanism. Poor mixing would result in a flow-rate variable conversion
of monomer, which results in uncertain kinetic measurements.
However, the reactant flow rate cannot be increased excessively.
An excessive flow rate would lead to massive consumption of the
monomer and a larger tube volume would be needed to finish
the reaction. Both mixing performance and residence time should
be taken into account and the influence of flow rate should be evaluated first. In Fig. 2, the molar conversions of AA at different residence times (t = Vi/QT, i = 115) are shown at different total flow
rates (QT = QAA + QKBS). Each result was repeated 3 times and the
absolute deviations were less than 3%. It is obvious that the polymerization reaction looked slower at total flow rates lower than
20 mL/min. When the total flow rate of the monomer and initiator
solutions reached 40 mL/min (the average velocity, 4QT/pd2c , was
0.85 m/s in 1 mm inner diameter channel downstream of the
T-junction), the reaction rate no longer increased with increasing
flow rates, thus the reaction is no longer limited by mixing. The
time dependent AA conversions at QT = 40 mL/min were varied
from 0% to 80% with residence times ranging from 0 to 120 s,
which is wide enough to establish the kinetic model given in
Eq. (4). The maximum tube volume is almost 70 mL under these

Fig. 2. Conversions of acrylic acid with varied residence times and different total flow
rates. The experiments were carried out at [AA]0 = 0.7 mol/L, [KBS]0 = 0.056 mol/L,
QAA:QKBS = 3:1, T = 95 C, and QCW = 10 mL/min (cold water for quenching).

237

L. Qiu et al. / Chemical Engineering Journal 284 (2016) 233239

Fig. 3. (a) Conversions of acyclic acid with various residence times at different
initial acyclic acid concentrations; (b) the fitting of log Rp with log[AA]. Other
experimental details are [KBS]0 = 0.056 mol/L, QAA = 30 mL/min, QKBS = 10 mL/min,
T = 95 C, and QCW = 10 mL/min.

conditions, and the consumption of the reactant solution is about


1.2 L in each test under all residence times.

Fig. 4. (a) Concentrations of acyclic acid with varied residence times at different
initial K2S2O8 concentrations; (b) linear fitting of log[KBS]0M with log k[KBS]0Mn2.
Other experimental details are [AA]0 = 1.40 mol/L, QAA = 30 mL/min, QKBS = 10 mL/
min, T = 95 C, and QCW = 10 mL/min.

suggests that the [KBS] obeyed an n2 = 0.50 order, consistent with


the result in the literature [31,35]. In Fig. 4a, the y-axis is [AA]0.5,
which was obtained from solving Eq. (4):

4.2. Reaction kinetics measurements

n1  1AAn1 1 KBSn2
After carefully controlling the mixing performance in the
microreactor platform, the reaction kinetics were studied. The
half-life of K2S2O8 is 12 min at 95 C, and the blank titration in
the analysis of AA concentration indicated that the K2S2O8 concentration decreased less than 5% after the reaction; thus, the reduction of K2S2O8 in the microreactor platform was negligible.
Fig. 3a shows the AA conversions at different initial concentrations.
The conversion increased quickly at higher initial concentrations,
indicating that the order of [AA] is higher than 1. Fig. 3b shows
the rates of polymerization (log Rp), and the linear fitting in this figure gives an order of 1.41, which is different from other free radical
polymerization kinetics such as styrene, whose monomer order is
1.0 theoretically [38]. Previous studies suggested that the order
of AA is 1.5 in theory, owing to the secondary monomerenhanced decomposition of the initiator [31]. The measured order
of AA in this study almost fits well with the intrinsic kinetic
parameter.
Based on the approximation that the exact order of AA is
n1 = 1.5, the effects of initiator concentration were evaluated.
According to classical principles [32], the chain initiation is the rate
determining step. As expected, polymerization speeds up at higher
initiator concentrations. Since the concentration of K2S2O8 is
almost constant during the reaction, different K2S2O8 concentrations would reveal the kinetic parameter of the initiator. Fig. 4

AA0:5

1
AA0:5
0M

1
n1 1
2
n1  1AA0M
KBSn0M

kt

k
2
KBSn0M
t
2

where [AA]0M and [KBS]0M are the concentrations of acrylic acid and
K2S2O8 after mixing ([AA]0M = [AA]0QAA/(QAA + QKBS), [KBS]0M =
[KBS]0QKBS/(QAA + QKBS)). With the slopes in Fig. 4a, the n2 value
was confirmed. The reaction rate constant could be drawn on the
basis of the above results. As shown in Table 2, the calculated
results fit with each other, proving that the results were reliable.
Here, the reaction temperature can be easily changed by regulating hot-water baths 14. As shown in Fig. 5, the reaction rate
increases significantly with elevated temperatures. In this case,

Table 2
Reaction rate constants at 95 C.
[AA]0 (mol/L)

[KBS]0 (mol/L)

k [L/(mol s)]

0.35
0.70
1.05
1.40
1.40
1.40

0.056
0.056
0.056
0.028
0.056
0.112

0.2577
0.2471
0.2524
0.2650
0.2647
0.2650

238

L. Qiu et al. / Chemical Engineering Journal 284 (2016) 233239

was in accord with the reported data, which ranged from 40 and
98 kJ/mol. Lorber et al. [35] gave an explanation for the varied activation energies based on the complexity of polymerization. They
reported that the reaction rate might become monomer
diffusion-limited with the generation of polymers at the end of
the reaction process. Since the AA concentrations in this study
were much lower those reported previously (24 mol/L) [39,30],
the results in this study should be more accurate.
As a brief summary, the kinetic equation for the AA polymerization reaction under the above experimental conditions is expressed
as

!
dAA
6:74  104
8
AA1:5 KBS0:5

8:0  10 exp
dt
RT

The advantages of the microreactor in measuring polymerization kinetics are obvious. Usually, it is very difficult to collect samples from batch reactors in less than 1 min, where the reactant
mixing is not complete. However, a microreactor can easily provide
data on the second time scale, since the reactant mixing is complete within several milliseconds [34]. Most kinetic parameters
are collected from low temperature batch reactors, but with the
microreactor platform it is much easier to obtain the kinetic
parameters for fast polymerization reactions at high temperatures,
where the reactant concentration changes drastically.
4.3. Molecular weight distributions

Fig. 5. (a) Concentrations of AA and K2S2O8 at different temperatures; (b) the linear
fitting of ln k with T1. Other experimental details are [AA]0 = 0.70 mol/L,
[KBS]0 = 0.056 mol/L, QAA = 30 mL/min, QKBS = 10 mL/min, and QCW = 10 mL/min.

Table 3
GPC results of the PAA products at the longest residence time in each test.*

[AA]0 (mol/L)

[KBS]0 (mol/L)

T (C)

Mn

PDI

1.40

0.028
0.056
1.112

95

162,099
129,016
106,445

2.10
2.30
2.32

0.70
1.05
1.40

0.056

95

103,326
129,016
171,934

2.42
2.30
2.34

1.40

0.056

80
85
90
95

176,052
145,781
140,415
129,016

2.03
2.09
2.08
2.30

QAA = 30 mL/min, QKBS = 10 mL/min.

immense reaction heat is released in a short time, which might


result in uncontrollable reactions in conventional batch reactors.
The advantage of microreactors is that they can be operated near
the boiling point of water and researchers do not need to worry
about solvent or monomer evaporation in the enclosed space of a
microreactor. The lowest temperature evaluated in this work was
80 C, and at temperatures less than 80 C, the maximum acrylic
acid conversion would be lower than 50%. This low reaction rate
requires longer reaction tubes, but it is outside the scope of minimizing chemical reactors. Thus, the microreactor platform is more
suitable for measuring fast polymerization reactions at relatively
high temperatures. Data collected at different temperatures helped
us calculate the activation energy. Using the Arrhenius equation,
we found that the activation energy was 67.4 kJ/mol and the preexponential factor was 8.0  108 L/(mol s). This activation energy

One of the other advantages of using microreactors is the wellcontrolled molecular weight of the product, which has been
reported previously [2,10]. GPC analyses of the PAA samples at
the longest reaction time (out of tube V9) summarize the variation
in PAA molecular weight. As shown in Table 3, the number average
molecular weights (Mn) ranged from 1.03  105 to 1.76  105. The
microreactor platform works well for this reaction with macromolecules, which may open new applications in kinetics research.
The polydispersity indexes (PDI) ranged from 2.0 to 2.4 in this
experiment and were mainly affected by the reaction temperature.
5. Conclusions
The reaction kinetics of AA polymerization were carefully investigated using a microreactor platform. The microreactor platform
could suitably mix the monomer and imitator solutions at flow
rate greater than 40 mL/min. An advantage of this microreactor
platform is the easily assembled reaction tubes, when combined
with the switch-on valves, which can provide variable reaction
times. The highly exothermic free radical polymerization was conducted isothermally at temperatures ranging from 80 to 95 C. The
kinetic orders of AA and potassium persulfate, 1.5 and 0.5 respectively, were in accord with the reported kinetic parameters. The
measured activation energy was 67.4 kJ/mol, which was also in
accord with the literature value. The molecular weights of PAA produced in the platform ranged from 1.03  105 to 1.76  105 with
PDIs from 2.0 to 2.4. In addition, the advantages of the microreactor for studying the kinetics of free radical polymerization reactions, including accurate reaction time and temperature control,
as well as rapid mixing of reactants, were clarified.
Acknowledgement
We gratefully acknowledge the National Natural Science Foundation of China (91334201), the Foundation for the Author of
National Excellent Doctoral Dissertation of the Peoples Republic
of China (FANEDD 201349) and the Tsinghua University Initiative
Scientific Research Program for their support of this work.

L. Qiu et al. / Chemical Engineering Journal 284 (2016) 233239

References
[1] V. Hessel, D. Kralisch, N. Kockmann, T. Noel, Q. Wang, Novel process windows
for enabling, accelerating, and uplifting flow chemistry, ChemSusChem 6
(2013) 746789.
[2] A. Abe, K. Lee, L. Leibler, S. Kobayashi, A. Nagaki, J. Yoshida, Controlled
polymerization and polymeric structures, Adv. Polym. Sci. 259 (2013) 150.
[3] T. Illg, P. Lob, V. Hessel, Flow chemistry using milli- and microstructured
reactors from conventional to novel process windows, Bioorg. Med. Chem. 18
(2010) 37073719.
[4] J.I. Yoshida, A. Nagaki, T. Yamada, Flash chemistry: fast chemical synthesis by
using microreactors, Chem. Eur. J. 14 (2008) 74507459.
[5] K. Mae, Advanced chemical processing using microspace, Chem. Eng. Sci. 62
(2007) 48424851.
[6] J.A. Newby, L. Huck, D.W. Blaylock, P.M. Witt, S.V. Ley, D.L. Browne,
Investigation of a lithium-halogen exchange flow process for the preparation
of boronates by using a cryo-flow reactor, Chem. Eur. J. 20 (2014) 263271.
[7] M. Baumann, I.R. Baxendale, L.J. Martin, S.V. Ley, Development of fluorination
methods using continuous-flow microreactors, Tetrahedron 65 (2009) 6611
6625.
[8] W. Fred, J.R. Billmeyer, Textbook of Polymer Science, Wiley, New York, 1984.
[9] K. Matyjaszewski, T.P. Davis, Handbook of Radical Polymerization, Wiley
Online Library, 2002.
[10] C. Tonhauser, A. Nataello, H. Lowe, H. Frey, Microflow technology in polymer
synthesis, Macromolecules 45 (2012) 95519570.
[11] F. Bally, C.A. Serra, V. Hessel, G. Hadziioannou, Micromixer-assisted
polymerization processes, Chem. Eng. Sci. 66 (2011) 14491462.
[12] C. Serra, G. Schlatter, N. Sary, F. Schonfeld, G. Hadziioannou, Free radical
polymerization in multilaminated microreactors: 2D and 3D multiphysics CFD
modeling, Microfluid. Nanofluid. 3 (2007) 451461.
[13] A. Nagaki, T. Iwasaki, K. Kawamura, D. Yamada, S. Suga, T. Ando, M. Sawamoto,
J. Yoshida, Microflow system controlled carbocationic polymerization of vinyl
ethers, J. Chem. Asian J. 3 (2008) 15581567.
[14] A. Nagaki, A. Miyazaki, Y. Tomida, J. Yoshida, Anionic polymerization of alkyl
methacrylates using flow microreactor systems, Chem. Eng. J. 167 (2011) 548
555.
[15] D. Kessler, H. Lowe, P. Theato, Synthesis of defined poly(silsesquioxane)s: fast
polycondensation of trialkoxysilanes in a continuous-flow microreactor,
Macromol. Chem. Phys. 210 (2009) 807813.
[16] T. Iwasaki, J. Yoshida, Free radical polymerization in microreactors. Significant
improvement in molecular weight distribution control, Macromolecules 38
(2005) 11591163.
[17] Z.D. Liu, Y.C. Lu, B.D. Yang, G.S. Luo, Controllable preparation of poly(butyl
acrylate) by suspension polymerization in a coaxial capillary microreactor, Ind.
Eng. Chem. Res. 50 (2011) 1185311862.
[18] D. Parida, C.A. Serra, F. Bally, D.K. Garg, Y. Hoarau, Intensifying the ATRP
synthesis of statistical copolymers by continuous micromixing flow
techniques, Green Process. Synth. 1 (2012) 525532.
[19] C.H. Hornung, C. Guerrero-Sanchez, M. Brasholz, S. Saubern, J. Chiefari, G.
Moad, E. Rizzardo, S.H. Thang, Controlled RAFT polymerization in a continuous
flow microreactor, Org. Process Res. Dev. 15 (2011) 593601.
[20] D. Parida, C.A. Serra, D.K. Garg, Y. Hoarau, F. Bally, R. Muller, M. Bouquey, Coil
flow inversion as a route to control polymerization in microreactors,
Macromolecules 47 (2014) 32823287.

239

[21] A.K. Yadav, M. Krell, W. Hergeth, J.C. de la Cal, M.J. Barandiaran, Monitoring
polymerization kinetics in microreactors by confocal Raman microscopy,
Macromol. React. Eng. 8 (2014) 543549.
[22] L.S. Mndez-Portillo, C. Dubois, P.A. Tanguy, Free-radical polymerization of
styrene using a split-and-recombination (SAR) and multilamination
microreactors, Chem. Eng. J. 256 (2014) 212221.
[23] K. Wang, Y.C. Lu, Y. Xia, H.W. Shao, G.S. Luo, Kinetics research on fast
exothermic reaction between cyclohexanecarboxylic acid and oleum in
microreactor, Chem. Eng. J. 169 (2011) 290298.
[24] J.P. McMullen, K.F. Jensen, Integrated microreactors for reaction automation:
new approaches to reaction development, Annu. Rev. Anal. Chem. 3 (2010) 19
42.
[25] J.P. McMullen, K.F. Jensen, Rapid determination of reaction kinetics with an
automated microfluidic system, Org. Process Res. Dev. 15 (2011) 398407.
[26] P. Wang, K. Wang, J. Zhang, G. Luo, Kinetic study of reactions of aniline and
benzoyl chloride in a microstructured chemical system, AIChE J. (2015), http://
dx.doi.org/10.1002/aic.14891.
[27] N.R. Beer, K.A. Rose, I.M. Kennedy, Monodisperse droplet generation and rapid
trapping for single molecule detection and reaction kinetics measurement, Lab
Chip 9 (2009) 841844.
[28] Z.Y. Han, W.T. Li, Y.Y. Huang, B. Zheng, Measuring rapid enzymatic kinetics by
electrochemical method in droplet-based microfluidic devices with pneumatic
valves, Anal. Chem. 81 (2009) 58405845.
[29] C. Tonhauser, A. Natalello, H. Lwe, H. Frey, Microflow technology in polymer
synthesis, Macromolecules 45 (2012) 95519570.
[30] R.A. Scott, N.A. Peppas, Kinetic study of acrylic acid solution polymerization,
AIChE J. 43 (1997) 135144.
[31] S.S. Cutie, P.B. Smith, D.E. Henton, T.L. Staples, C. Powell, Acrylic acid
polymerization kinetics, J. Polym. Sci. Part B: Polym. Phys. 35 (1997) 2029
2047.
[32] G. Odian, Principles of Polymerization, John Wiley & Sons, 2004.
[33] J.S. Zhang, K. Wang, Y.C. Lu, G.S. Luo, Characterization and modeling of
micromixing performance in micropore dispersion reactors, Chem. Eng.
Process. 49 (2010) 740747.
[34] Z.D. Liu, Y.C. Lu, J.W. Wang, G.S. Luo, Mixing characterization and scaling-up
analysis of asymmetrical T-shaped micromixer: experiment and CFD
simulation, Chem. Eng. J. 181 (2012) 597606.
[35] N. Lorber, B. Pavageau, E. Mignard, Droplet-based millifluidics as a new
miniaturized tool to investigate polymerization reactions, Macromolecules 43
(2010) 55245529.
[36] M.M. Mandal, C. Serra, Y. Hoarau, K.D.P. Nigam, Numerical modeling of
polystyrene synthesis in coiled flow inverter, Microfluid. Nanofluid. 10 (2011)
415423.
[37] D. Parida, C.A. Serra, D.K. Garg, Y. Hoarau, R. Muller, M. Bouquey, Flow
inversion: an effective means to scale-up controlled radical polymerization
tubular microreactors, Macromol. React. Eng. 8 (2014) 597603.
[38] A.A. Butakov, E.N. Shatunova, Macrokinetic features of the free-radical
polymerization of styrene under nonisothermal conditions, Theor. Found.
Chem. Eng. 44 (2010) 717722.
[39] S.P. Manickam, K. Venkatarao, N.R. Subbaratnam, Peroxo salts as initiators of
vinyl polymerization 2. Kinetics of polymerization of acrylic acid and
sodium-acrylate by peroxodisulphate in aqueous solution effect of pH and
Ag+ catalysis, Eur. Polym. J. 15 (1979) 483487.

Vous aimerez peut-être aussi