You are on page 1of 10

Chemical Engineering Journal 405 (2021) 126636

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Synthesis of propylene glycol methyl ether catalyzed by imidazole polymer T


catalyst: Performance evaluation, kinetics study, and process simulation

Ran Ana,b, Shengxin Chenb, Ruirui Zhangb, Fei Daib, Zhimao Zhoub, Chunhu Lia, Ruixia Liub,c, ,

Weizhong Ana,
a
Department of Chemical Engineering, Ocean University of China, Qingdao 266100, China
b
Beijing Key Laboratory of Ionic Liquids Clean Process, CAS Key Laboratory of Green Process and Engineering, State Key Laboratory of Multiphase Complex Systems,
Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
c
School of Chemical Engineering, University of Chinese Academy of Sciences, Beijing 100049, China

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

• Alysthigh efficiency heterogeneous cata-


was synthesized by crosslinking
copolymerization.
• “Electrophile nucleophile dual activa-
tion” mechanism was attributed to
excellent catalytic performance.
• Detailed kinetic modeling was devel-
oped and kinetic parameters were de-
termined.
• The application of kinetic was made in
a fixed bed using the Aspen Plus si-
mulation.

A R T I C LE I N FO A B S T R A C T

Keywords: An imidazole polymer (divinylbenzene-vinylimidazole) catalyst (P-DVB-VIM) was developed for production of
Propylene oxide the eco-friendly solvent propylene glycol methyl ether (PGME) by the propoxylation reaction of propylene oxide
Kinetics (PO) with methanol (ME). The catalyst was characterized in detail by scanning electron microscopy (SEM),
Mechanism Fourier-transform infrared spectroscopy (FTIR), and Brunauer-Emmett-Teller (BET) analysis. 1H NMR was used
Propoxylation
to characterize the structural features of the catalyst precursor, i.e., the linear polymer of vinylimidazole (PVIM).
Process simulation
An “electrophile nucleophile dual activation” ring-opening mechanism was proposed based on the “cooperative
effect” of the reactant and PVIM. The catalytic performance was investigated and optimized at 100–130 ℃ using
an autoclave batch reactor. The results reveal that the P-DVB-VIM shows excellent catalytic performance, ex-
hibiting the highest turnover frequency (TOF) number at 120 °C within 45 min. The reaction kinetics was
systematically studied using the pseudohomogeneous (PH) kinetic model. The kinetic parameters were de-
termined by non-linear regression analysis of the experimental data. Validation of the model showed that the
experimental data were in good agreement with the predicted outcomes. Moreover, based on the proposed
kinetic model, steady-state simulations of the production of PGME in a fixed-bed reactor were carried out to
determine the optimal design parameters, thus providing a useful guide for the design and production of PGME
on the industrial scale.


Corresponding authors at: Beijing Key Laboratory of Ionic Liquids Clean Process, CAS Key Laboratory of Green Process and Engineering, State Key Laboratory of
Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China (R. Liu).
E-mail addresses: rxliu@ipe.ac.cn (R. Liu), awzhong@ouc.edu.cn (W. An).

https://doi.org/10.1016/j.cej.2020.126636
Received 19 March 2020; Received in revised form 5 August 2020; Accepted 9 August 2020
Available online 12 August 2020
1385-8947/ © 2020 Elsevier B.V. All rights reserved.
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

Nomenclature Xi Conversion of i, %

Notation Abbreviations

Ai Pre-exponential factor of step i, L·mol−1·min−1 AIBN a, a’-Azoisobutyronitrile


bij, bji, cij Binary interaction parameters cat Catalyst
Ei Activation energy of step i, kJ·mol−1 DPGME Dipropylene glycol monomethyl ether
ki Reaction rate constant of step i, L·mol−1·min−1 DVB 1, 4-Divinylbenzene
k0s Initiation constant for propoxylation (primary hydroxyls), EAC Ethyl acetate
L·mol−1·min−1 LHSV Liquid hourly space velocity
k0p Initiation constant for propoxylation (secondary hydro- ME Methanol
xyls), L·mol−1·min−1 TOF Turnover frequency, mol/(h·g)
ri Reaction rate of step i, mol·L -1·min−1 TPGME Tripropylene glycol methyl ether
r0s Reaction rate for propoxylation (primary hydroxyls), PGME Propylene glycol methyl ether
mol·L -1·min−1 PH Pseudo-homogeneous
r0p Reaction rate for propoxylation (secondary hydroxyls), PO Propylene oxide
mol·L -1·min−1 PVIM Polymer of vinylimidazole
Si Selectivity of i, % VIM 1-Vinylimidazole
T Reaction temperature, K 1PO1 2-Methoxy-1-propanol
t Reaction time, h 1PO2 1-Methoxy-2-propanol
u0 Space velocity, h−1 iPOj Oligomerization products

1. Introduction alternative reaction media, with the potential for catalytic application
in industry [20]. However, the imidazole moiety is hydrophilic; thus,
Propylene glycol methyl ether (PGME) is widely used as an eco- PVIM dissolves in aqueous medium to form a homogenous mixture.
friendly solvent in the production of detergents, cosmetics, and emul- Using a crosslinker to lock the imidazole moiety within the methanol or
sifiers due to its negligible toxicity. PGME is generally produced by the PGME is a prospectively feasible approach for obtaining heterogeneous
propylene oxide (PO) method, where the relevant reactions are de- catalysts with sufficiently high catalytic activity.
picted in Scheme 1. In these reactions, the acid catalyst facilitates the Obtaining reliable kinetic data is vital for in-depth understanding of
formation of 2-methoxy-1-propanol (1PO1) and the base catalyst pre- the reaction behavior, particularly for the design of the reactor and for
ferentially leads to the formation of 1-methoxy-2-propanol (1PO2) be- optimizing the design parameters. Kinetics studies of the propoxylation
cause of the asymmetry of propylene oxide [1]. 1PO2 is much less toxic of methanol over homogeneous catalysts such as sodium hydroxide
than 1PO1. Therefore, the synthesis of 1PO2 with high efficiency and [21], potassium hydroxide [22,23], and [N4444] [butyl][8] have been
selectivity in the presence of base catalysts has attracted much atten- widely investigated. However, little information on the kinetics of the
tion. Note that PO may further react with 1PO1 or 1PO2 to generate heterogeneous catalytic reaction is available in the current literature.
different oligomerization products, resulting in low selectivity in these The kinetic parameters can be determined by monitoring the con-
reactions. centration over time in a batch reactor by applying the “pseudo-
Many kinds of catalysts have been developed for this reaction. homogeneous” (PH) model. For the heterogeneously catalyzed reaction,
Among the homogenous acid/base catalysts, common catalysts such as the PH model is not adequate for describing the heterogeneous kinetics
NaOH [2], BF3, amines [3], H2SO4 [4], and ionic liquids [5–9] have considering the effect of adsorption of the reactants on the catalyst
been proven as efficient catalysts. Nevertheless, these catalysts still [24]. Nevertheless, when the phase behavior does not significantly
suffer from many issues such as low isomer selectivity, difficulty in impact the observed reaction kinetics, the simplified PH model can
separating the catalyst from the post-reaction medium, and strong provide a reasonably accurate description of the process [25–28].
corrosivity. To overcome these disadvantages, heterogeneous catalysts It is noted that investigation of the catalytic performance using a
have been developed and have attracted widespread attention recently. continuous reactor is preferred over the use of batch reactor types,
Such catalysts include metal oxides [10–12], mixed-oxides [13–15], where the main products can be removed from the reaction mixture in
alumina‐pillared clays [16], molecular sieves [17,18], Zr, Al‐pillared the former [29]. Therefore, a continuous fixed-bed reactor is employed
clays [19], etc. However, heterogeneous catalysts still suffer from herein. Using the kinetic parameters obtained from a specific experi-
drawbacks such as low reactant activity and target product selectivity, ment, simulation using the Aspen Plus software package can provide
and the need for a high reaction temperature and longer reaction times, further insight to guide the process design for the industrial-scale re-
limiting their application in continuous technology, such as fixed-bed action.
and microflow reactors, etc., which restricts further industrialization This report presents the synthesis and characterization of an imi-
for large-scale application. In studies of catalytic mechanisms, it was dazole polymer catalyst. This is the first study of the use of a polymer
shown that catalysts dissolved in a solvent could control and modify the catalyst to produce PGME. Kinetic study in batch reactors is also per-
reaction by the solvent effect [9], which can dictate the catalytic per- formed and a PO ring opening mechanism is proposed. Furthermore,
formance of the coupling reaction of PO with methanol. This may ac- the kinetic parameters are determined by measuring the PO con-
count for why the performance of homogeneous catalysts is sometimes centration. Finally, the continuous fixed-bed reaction process is
better than that of heterogeneous catalysts. Therefore, greater effort
should be devoted to developing efficient heterogeneous catalysts for
achieving high PO activity and 1PO2 yield under mild conditions. Al-
though many researchers have developed heterogeneous catalysts, only
a limited variety of catalysts including metal oxides, metal clays, and
molecular sieves has been widely investigated. Radical polymerization
of vinylimidazole (PVIM) has received signification attention using Scheme 1. Reaction pathway for synthesis of propylene glycol methyl ethers.

2
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

investigated to extend the experiment scale and to predict the process


behavior based on computer simulations.
2
2. Experimental
9
2.1. Preparation and characterization of catalyst 3

The diblock copolymer P-DVB-VIM was prepared by cross-


linking copolymerization as follows: 4.6 g of 1-vinylimidazole (VIM) 8
was mixed with 0.01 g of α, α'-azoisobutyronitrile (AIBN) and 30 mL of N2
ethyl acetate (EAC) in a 100 mL three-necked flask. After flushing the 5
flask three times with nitrogen, the reaction was allowed to proceed for
1 h at 90 °C with magnetic stirring at 400 rpm. Ethyl acetate solution
containing 4.6 g of 1,4-divinylbenzene (DVB) was then dropped into the PO 7
solution at 90 °C and allowed to react for 24 h; the resultant solid was 6
dried overnight at 80 °C. Finally, the prepared P-DVB-VIM was ground 1 4
and sieved (20–40 mesh) to afford a white solid powder. The synthesis
of P-DVB-VIM and the reaction conditions are presented in Scheme 2. Fig. 1. Schematic of batch reactors used for the kinetics investigation (1 – N2
cylinder; 2 – pressure gauge; 3 – valve; 4 – PO tank; 5 – piston pump; 6 –
The Brunauer-Emmett-Teller (BET) method was used to determine
stainless-steel autoclave batch reactor; 7 – magnetic stirrer; 8 – withdraw; 9 –
the specific surface area and pore volume of the catalysts by using a
temperature control recorder).
Micromeritics ASAP 2460 analyzer. Infrared (IR) spectra were recorded
on a Vertex 70 infrared spectrometer in the region of 4000–500 cm−1.
The morphology of the catalysts was observed using a Hitachi S-4800 was heated to the reaction temperature with magnetic stirring at a rate
scanning electron microscope (SEM). 1H NMR spectra were acquired at of 600 rpm, and propylene oxide was sent to the reactor by a piston
ambient temperature in D2O solution using a 600 MHz AVANCE III pump. The effect of the temperature, methanol/PO molar ratio, and
NMR spectrometer. catalyst concentration was studied. The reaction conditions for each
experiment are displayed in Table 1.
2.2. Reagents
2.5. Application of kinetics and simulation tests
PO (≥99%) and methanol (≥99%) were obtained from Aladdin
Reagent Co., Ltd. (Shanghai, China). All chemicals were of analytical The continuous reaction process was carried out in a fixed-bed
reagent grade and were used without further purification. tubular stainless-steel reactor with an internal diameter of 10 mm. The
catalytic bed was diluted with quartz particles (20–40 mesh) to mini-
2.3. Analytical methods mize the temperature gradients; the particles occupied a length of ap-
proximately 40% of the reaction zone. Before the experiments, the re-
The mixtures after the reaction were analyzed by gas chromato- actants were pre-heated in a preheater ahead of the reactor and the
graphy (GC, Shimadzu GC‐2010 plus) using a system equipped with a catalyst was pre-heated in the catalyst bed. The inlet feed arrived from
flame ionization detector and a capillary column (HP‐INNOWax). the bottom of the reactor, and the outlet products exited from the top of
the reactor. A sketch of the fixed-bed reactor is presented in Fig. 2.
2.4. Kinetics experimental procedures The simulation tests were carried out by considering the fixed-bed
reactor as a pseudo-homogenous plug flow reactor. Aspen Plus (version
Kinetics runs were performed in a stainless-steel autoclave batch 11.0) was used to conduct the process simulation with a plug flow re-
reactor with an inner volume of 100 mL. Before each of the kinetics actor (PFR). For the organic system considered here, the NRTL models
tests, the system was purged with nitrogen flow. Samples (1 mL) were were used to calculate the thermodynamic properties [30]. Table 2
withdrawn in vials at different reaction times. After removal from the presents the values of the NRTL interaction parameters (taken from the
reactor, the reaction products were immediately cooled in cold water. A Aspen properties database) for 10 binary mixtures used in the study.
centrifugal tube was used to collect the liquid product, which was The operating conditions of the fixed-bed are listed in Table 3.
subsequently used for analysis. A sketch of this reactor is presented in
Fig. 1. 3. Results and discussion
Before performing the experiments, 34.56 g of fresh methanol so-
lution and the prepared catalyst were placed into the reactor. After 3.1. Characterization of P-DVB-VIM
purging 3 times with N2 flow, the mixture in the reactor was pressur-
ized at 0.8–1.2 MPa to keep all reagents in the liquid phase. The reactor Fig. 3 shows the FTIR spectra of P-VIM, DVB and P-DVB-VIM. The

Scheme 2. Synthesis of P-DVB-VIM.

3
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

Table 1
List of experiments for the kinetics study.
Experiment T (℃) P-DVB-VIM (wt.%) Mole ratio of ME/PO

1 120 0.33 6
2 120 0.55 6
3 120 0.77 6
4 120 0.99 6
5 120 0.55 3
6 120 0.55 9
7 120 0.55 12
8 100 0.55 6
9 110 0.55 6
10 130 0.55 6

2 8

3
4 6
11
Fig. 3. FTIR spectra of P-DVB-VIM.
N2 PO ME 9

10 1594 cm−1, 1508 cm−1, 1484 cm−1. Obviously, the C = C bond vi-
bration of DVB was removed after the copolymerization of DVB and
7
5 VIM. The final polymer of P-DVB-VIM has more vibration of phenyl
1 12 group at 1604 cm−1, 1447 cm−1, 1409 cm−1 than P-VIM, confirming
the copolymer of P-DVB-VIM. And the vibration of imidazolium group
has a little shift that characteristic bands showed at 3109 cm−1,
2926 cm−1, and 1495 cm−1 in P-DVB-VIM. The inductive electronic
effect of the N atom from the imidazole group conferred strong elec-
Fig. 2. Schematic of continuous fixed-bed reactor for kinetics study (1 – N2
trophilic properties to the hydrogen at the C-2 position, while the
cylinders; 2 –pressure gauge; 3 – valves; 4 – PO tank; 5 – piston pump; 6 – ME
tank; 7 – preheater; 8 – tubular reactor and catalyst bed; 9 – quartz particles; 10
imidazole ring has a lone pair of electrons, making the group even more
– catalyst; 11 – condenser; 12 – product tank). nucleophilic. Therefore, the structure of P-DVB-VIM has a suitable
electrophile and nucleophile, which makes it highly active as an effi-
cient active center. The FTIR spectra of PO, methanol, and the product
Table 2 after the reaction are shown in Fig. S1.
NRTL binary interaction parameters for different component pairs.
To obtain a heterogeneous polymer material, DVB was used as a
Comp. i Comp. j bij (K) bji (K) cij (K) crosslinker to connect the linear PVIM units. P-DVB-VIM was formed as
a heterogeneous polymer by controlling the ratio of DVB and VIM.
PO ME 718.841 –233.753 0.3
PO 1PO2 −472.592 683.9679 0.3
Fig. 4 presents the SEM micrograph of P-DVB-VIM. As shown in Fig. 4,
PO 2PO2 −719.038 1311.214 0.3 the morphology comprises a structural network with numerous pores
ME 1PO2 −482.828 693.0063 0.3 and channels. The catalyst also consists of many nanoparticles with
PO 1PO1 −247.157 294.6229 0.3 diameters of 200–500 nm on top of the network. The numerous pores,
1PO2 2PO2 113.5571 −102.663 0.3
channels, and particles can expose more catalytically active centers for
1PO2 1PO1 −265.86 519.2139 0.3
2PO2 ME 1153.8 −659.228 0.3 improving the reaction efficiency. Additionally, the catalyst particles
2PO2 1PO1 −343.565 609.4743 0.3 with sizes of 0.1–1 μm were well dispersed.
ME 1PO1 260.491 −161.886 0.3 The nitrogen adsorption–desorption isotherms and pore size

Table 3
Operating conditions for fixed-bed.
State variable Value

Reaction volume (mL) 10


Catalyst loading (mL) 6.6
Volume ratio of ME / PO 8
Feed volume flow (mL·h−1) 100
Space velocity (h−1) 10
Reaction pressure (MPa) 1.2
Reaction temperature (℃) 120

characteristic bands at 3112 cm−1 (stretching of C2-H bond),


2952 cm−1 (stretching of C4,5-H bond), 1643 cm−1 (stretching of
C = N bond) and 1500 cm−1 (stretching of C = C bond) suggested the
formation of the P-VIM. In the FTIR spectra of DVB, the crosslinker in
the results polymer, showed the C = C bond stretching vibration of the
vinyl group at 1822 cm−1 and vibration of phenyl group at 1629 cm−1,
Fig. 4. SEM image of P-DVB-VIM sample.

4
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

distributions of P-DVB-VIM are shown in Fig. 5. Clearly, the samples where the respective values were maintained at 98.0% and 2%.
exhibit typical type-IV isotherms with H3-type hysteresis loops (P/
Po > 0.2), characterized by a steep increase in the relative pressure 3.2.6. Comparison of the catalytic performance
range of 0.8 < P/Po < 0.95. The isotherms clearly indicate the for- The catalytic efficacy of P-DVB-VIM for the synthesis of PGME was
mation of mesopores in these samples [31]. A broad peak centered compared with that of various catalysts reported in the literature. The
around 40 nm was observed in the pore size distributions, further main results are listed in Table 4. The catalyst efficiency was estimated
confirming the mesoporous structure. P-DVB-VIM possesses a relatively by comparing the turnover frequency (TOF) data given that the ex-
high specific surface area of up to 485.43 m2·g−1 and a large pore perimental conditions employed in the studies are different [13]:
volume of around 0.99 cm3·g−1.
XPO
TOF =
3.2. Reaction behavior t × Wcat (1)

A PO conversion of 97.7% and 1PO2 selectivity of 97.8% were


3.2.1. Elimination of mass transfer resistance achieved under the optimal conditions. It should be noted that PO is a
Before conducting the kinetic measurements, the influence of the very reactive, flammable, and explosive reactant, and the PO ring-
external mass transfer limitations was determined by varying the stir- opening process is a strongly exothermic reaction. Thus, high me-
ring rate in the experiment. Fig. 6 was obtained by varying the stirring thanol/PO molar ratios (>3:1, excess MeOH) and an appropriate
rate while keeping the other operating conditions unchanged. The re- temperature (<30 °C) were used in the reaction to reduce the danger
sults indicate that there is no significant external transfer effect for and prevent the formation of oligomerization products. Note that the
stirring rates > 600 rpm. Therefore, at the stirring rate of 600 rpm TOF of P-DVB-VIM under the conditions used herein was 236.8%·(h·wt
mass transfer resistance was avoided; thus, this rate was chosen for the %)−1, which is one to two orders higher than those reported in the
remaining experiments. In addition, when the catalyst was ground and literature. Compared with P-DVB-VIM, the conversion of PO reached up
sieved to a powder (around 30 mesh) before the experiment, the in- to 97–99% with other catalysts at 120 °C, but a longer reaction time was
ternal mass transfer resistance could be ignored. required and the selectivity for 1PO2 was lower at similar conversion
[5,10,16,18].
3.2.2. Effect of catalyst amount
Fig. 7 shows that by increasing the catalyst amount (wt% of the 3.3. Ring opening mechanism
substrate) the PO conversion could be increased, reaching 97.7% with
0.55 wt% of catalyst at 45 min, which could be attributed to an increase To achieve a high activity of PGME, it is desirable to study the
in the concentration of active sites. When the catalyst loading was in- mechanism and kinetics of the reaction over the P-DVB-VIM catalyst.
creased beyond 0.55 wt%, the PO conversion increased smoothly. Because P-DVB-VIM does not dissolve in D2O, the linear polymer vi-
However, the selectivity for 1PO2 declined to 96.6% and 90.1% when nylimidazole (PVIM) that provides the active sites in P-DVB-VIM was
the catalyst loading was increased to 0.77 and 0.99 wt%, respectively. selected to allow dissolution in D2O and to further explore the reaction
The selectivity for 1PO2 decreased, suggesting that an excess of the mechanism. Note that the DVB block is used to maintain the hetero-
catalyst facilitates ring-opening of the epoxide at the primary alcohol geneous nature of the catalyst, while the VIM block provides the active
groups. This demonstrates that the excess catalyst enhances undesirable sites. Fig. 11 shows the 1H NMR spectrum of PVIM, which was used to
side-reactions. investigate the interactions between the catalyst and reactant (PO and
methanol) before and after the reaction. The details of the full 1H NMR
3.2.3. Effect of reactant molar ratios spectra are provided in the Supporting Information (Fig. S2). Fig. 11
Fig. 8 illustrates the influence of the inlet ME/PO molar ratio on the shows that with the addition of 1 equivalent of PO or methanol, the
PO conversion and 1PO2 selectivity. With variation of the ME/PO resonance of both C2-H1 protons of PVIM shifted up-field from
molar ratio, the conversion of PO remained almost unchanged. A slight d = 7.14 to 7.07 ppm, which indicates the formation of a hydrogen
increase in the 1PO2 selectivity was observed when the ME/PO molar bond between PVIM and both PO or methanol. The shift indicates that
ratio was varied from 3 to 6, but when the ME/PO molar ratio was PVIM interacts with the carbon atom of PO and methanol. This may be
increased to 12, the selectivity remained stable. This suggests that the attributed to the increased electron density around the carbon atom due
ME/PO molar ratio of 6:1 is optimal for improving the selectivity for
1PO2.

3.2.4. Effect of reaction temperature


Fig. 9 shows the effect of the reaction temperature on the PO con-
version and 1PO2 selectivity. It can be clearly seen that an increase in
temperature enhances the reaction rate, which can contribute to the
kinetic factors. The maximum PO conversion of 99.4% was achieved
when the temperature exceeded 130 ℃. A smooth decrease in the 1PO2
selectivity from 100.0% to 95.4% was observed when the temperature
was increased from 100 ℃ to 130 ℃. This finding shows that high
temperature can exert a positive impact on the PO conversion, but a
slightly negative impact on the 1PO2 selectivity.

3.2.5. Effect of reaction time


Fig. 10 shows the catalytic performance obtained within the reac-
tion time range of 5 to 50 min. The PO conversion approached 97.7%
when the reaction was carried out for 45 min. The PO conversion in-
creased slightly when the reaction time was prolonged to 50 min,
suggesting that the kinetically controlled part of the reaction was al-
most completed at that time. The selectivity for 1PO2 and 1PO1 was Fig. 5. N2 adsorption–desorption isotherms and pore size distributions (inset)
almost constant during the reaction time in the range of 5 to 50 min, of P-DVB-VIM.

5
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

Fig. 6. External diffusion tests performed at different stirring rates. Reaction Fig. 9. Effect of temperature on synthesis of 1PO2. Reaction conditions—PO:
conditions: reactant 45 g, ME/PO mole ratio: 6, catalyst loading: 0.55 wt%; 0.18 mol; methanol: 1.08 mol; reaction time: 45 min; catalyst concentration:
pressure: 1.2 MPa; temperature: 100 ℃. 0.55 wt%.

Fig. 7. Effect of catalyst concentration on the synthesis of 1PO2. Reaction Fig. 10. Effect of reaction time on synthesis of 1PO2. Reaction conditions-PO:
conditions: PO (0.18 mol), methanol (1.08 mol), reaction time (45 min), tem- 0.18 mol; methanol: 1.08 mol; catalyst concentration: 0.55 wt%; temperature:
perature (120 ℃). 120 ℃.

to shielding, leading to the up-field shift of the C2 resonance.


For the synthesis of PGME from PO and methanol, the solid basic
catalysts promoted the (1) deprotonation of CH3OH [12,32] and (2)
ring-opening of PO [2]. In the process of dissociation of CH3OH into a
proton and methoxide ion, SN2 attack of the methoxide on PO to open
the ring framing C-O bond precede proton transfer to the negatively
charged species [10]. The rate-determining step of the reaction is the
ring-opening of PO. Note that solid catalysts such as metal oxides are
not readily dissolved, but the imidazole group is hydrophilic and may
form a solvent layer on the interface. In the solvent layer, the imidazole
moiety partly contained in the solvent could promote the reaction via
the solvent effect [9]. By controlling the proportion of VIM and DVB,
high activity was obtained and the heterogeneous nature of the catalyst
was maintained. Thus, the activity of P-DVB-VIM is higher than that of
other solid catalysts. An “electrophile nucleophile dual activation”
mechanism was proposed with P-DVB-VIM as a catalyst in the synthesis
of PGME from PO and methanol. P-DVB-VIM is proposed to facilitate
ring-opening because of the drag effect of the hydrogen bonding in-
Fig. 8. Effect of methanol/PO molar ratio on the synthesis of 1PO2. Reaction teraction (Scheme 3). The “cooperative effect” in the ring-opening of
conditions—Catalyst concentration: 0.55 wt%; reaction time: 45 min; tem- PO was assumed to be responsible for the excellent catalytic perfor-
perature: 120 ℃. mance [33].

6
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

Table 4
Comparison of the catalytic performance of P-DVB-VIM versus reported catalysts.
Catalyst Reaction conditions XPO (%) S1PO2 (%) TOF (%·(h·wt%)−1) Ref.

ME:PO (mol/mol) T (℃) Cat. (wt.%) t (h)

P-DVB-VIM 6 120 0.55 0.75 97.7 97.8 236.8 This work


[P4444][Buty] 3 80 8.26 6 97.8 95.9 2.0 [5]
KAcMIMOH 20 60 16.67 4 94.2 99.1 1.41 [6]
CH 6 55 5.91 4 95.5 94.6 4.0 [7]
Fe2O3 37 160 0.57 8 97.7 83.0 21.4 [10]
MgO 5 120 Not determined 5 71.1 89.2 Not determined [11]
Brucite 4 120 1.80 3 68.1 93.8 12.6 [13]
Mg,Al-LDH 4 140 1.20 6 92.4 96.5 12.8 [14]
Zn,Mg,Al-LDH 5 120 1.80 3 95.5 90.9 17.7 [15]
Al-PILCs 8–10 60 3.00 1–6 98.7 61.2 32.9 (t = 1 h) [16]
MCM-41 3 110 4.95 3 98.4 93.0 6.6 [18]
Zr, Al-MM 8 60 2.57 6 74–75 60–66 4.9 (XPO = 75%) [19]

Fig. 11. 1H NMR spectra of PVIM, PVIM + ME, PVIM + PO.

3.4. Kinetics study Scheme 3. Proposed mechanism for ring opening of PO, catalyzed by P-DVB-
VIM.

Due to the asymmetric nature of PO, the epoxide ring may open at
either of the C-O bonds to generate a primary or secondary alcohol, (3) The rate equation for the initial reactant in the pathways was as-
respectively. A propoxylation kinetics scheme was developed, which sumed to be first-order as methanol is in excess.
includes in situ formation of the catalyst, initiation reactions, and
propagation reactions. Because the mechanism of the propoxylation Oligomerization products were detected with the use of some tra-
reaction is the same as that reported by Serio [23], for which the ditional catalysts like BF3, KOH and NaOH [21–23]. To avoid formation
substrate was ethylene glycol, the mechanism model is not discussed in of the oligomerization products, excess methanol is generally employed
detail herein. Note that in the literature [23], the initial reaction to in the experiments [8–19]. For the catalyst described herein, oligomeric
generate a primary alcohol is negligible, where 1PO1 and the oligo- products were negligible under the experimental conditions. Fig. 12
meric product were not considered. Theoretically, each adduct can shows that the main reaction products were 1PO2 and a small amount
undergo two possible reactions involving the primary or secondary al- of 1PO1 because the propagation rate was very slow and only trace
cohol groups. For example, dipropylene glycol methyl ether (DPGME) amounts of the different oligomerization products (iPOj, i ≥ 2) were
occurs as four isomers, tripropylene glycol methyl ether (TPGME) oc- detected. Therefore, we can simplify the kinetic model taking into ac-
curs as eight isomers. On this basis, we proposed a reaction pathway count the negligible formation of oligomerization products in this work.
including complete propagation reactions, as shown in Scheme 4. Under the above-mentioned assumptions, the rate of change of the
Based on the reaction pathways in Scheme 4, the reaction kinetics concentration of the liquid-phase components can be expressed as fol-
was analyzed. Several assumptions were considered to simplify the lows:
model:
d([1PO1])
= r0s = k 0s [PO]
dt (2)
(1) Reaction system heterogeneity can be neglected and the system can
be treated as ‘‘pseudo-homogeneous”.
(2) The power-law model was chosen to describe the reaction rate d([1PO2])
= r0p = k 0p [PO]
considering the reaction in the absence of the gas-phase. dt (3)

7
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

Scheme 4. Reaction pathways for propoxylation of methanol.

E0s = 80.9 kJ·mol−1 and E0p = 77.9 kJ·mol−1, respectively. The pre-
exponential factors were respectively determined as A0s = 1.08 × 108
L·mol−1·min−1 and A0p = 1.33 × 109 L·mol−1·min−1. The value of the
activation energy (E0p) is less than E0s and the pre-exponential factor for
step A0p is larger than that of step A0s, indicating that step r0p dominates
in the reaction. This observation shows that the kinetic model presents
good agreement (R2 = 0.987 and 0.993) between the calculated and
experimental data.

3.5. Simulation and optimization of fixed-bed process

Once the kinetic parameters have been determined against the ex-
perimental data, the continuous production process can be simulated
using Aspen Plus to provide guidance for process optimization and
parameter design for industrial scale-up. The results obtained from the
experiment and the simulation for the fixed-bed reactor are compared
in Fig. 15. The simulation results also matched well with the experi-
mental results. The residence time distribution (RTD) was determined
by monitoring the molar composition of each component. As shown in
Fig. 12. Oligomer mass fractions and related simulation vs. time profiles (lines Fig. 15, the PO concentration decreases along the reaction residence
and symbols represent the simulation and experimental data, respectively).
time profile and the highest conversion of PO of 91.5% was achieved at
Reaction conditions: reactant 45 g, ME/ PO mole ratio: 6, catalyst: 0.55 wt%;
5.4 min. These results show that the PO conversion and 1PO2 se-
pressure: 1.2 MPa; temperature: 120 ℃.
lectivity can be changed by tuning the residence time. Besides, a longer
residence time is favorable for increasing the consumption of the re-
d([PO]) actants and thus leads to higher PO conversion. However, the PO
- = r0p + r0s = (k 0p + k 0s )[PO]
dt (4)
conversion was much lower than 95% under these conditions. The re-
where [PO] and [iPOj] represent the concentration of PO and iPOj. sults suggest that the residence time is inadequate and detailed opti-
Eqs. (2)-(4) can be solved to obtain the liquid-phase concentrations of mization of the process parameters is needed to ensure a PO conversion
the reactants and products as a function of time using the Runge-Kutta exceeding 95%.
method. The rate parameters were evaluated by simulating the ex- After validating the reaction kinetics model, the PFR reactor model
perimental PO concentration–time data by non-linear least-squares re- was applied to predict the behavior of the propoxylation reaction. The
gression analysis. The reaction activation energy was then calculated by influence of the liquid hourly space velocity (LHSV) and ME/PO molar
linear fitting using the Arrhenius equation (Eq. (5)). ratio on the PO conversion and 1PO2 selectivity was studied at
T = 120 °C; the results are presented in Figs. 16 and 17. Fig. 16 shows
Ea that as the ME/PO molar ratio increased, the PO conversion increased
ki = A0 exp ⎛− i ⎞
⎝ RT ⎠ (5) because of the increased methanol concentration. Increasing the ME/
Here, A0 is the pre-exponential factor; Eai is the activation energy; R PO molar ratio led to an enhancement of the 1PO2 selectivity initially,
is the universal gas constant; and T is the absolute temperature. but when the molar ratio of ME/PO in the system was too high, the
Kinetic experiments were conducted in the temperature range of 1PO2 selectivity smoothly decreased, indicating that a high ME/PO
100 ℃ to 130 ℃ and the kinetic parameters are summarized in Table 5.
Table 5
The k0p values were about 40 times larger than k0s. The effect of the
Kinetic parameters obtained by mathematical regression analysis.
reaction time on the PO concentration at different temperatures from
the experimental evaluation and prediction are compared in Fig. 13. An T (℃) Parameter Optimal estimate 95% Confidence intervals
obvious enhancement in the PO consumption rate was achieved by -4
100 k0s 3.72 × 10 1.22 × 10-4
increasing the temperature from 100 °C to 130 °C, demonstrating good k0p 1.59 × 10-2 2.00 × 10-4
agreement between the simulations and the results. 110 k0s 6.78 × 10-4 1.89 × 10-4
Units: activation energies E0s, E0p = kJ·mol−1; pre-exponential k0p 2.58 × 10-2 3.37 × 10-4
factor A0s, A0p = L·mol−1·min−1. 120 k0s 1.29 × 10-3 5.79 × 10-4
k0p 5.48 × 10-2 1.48 × 10-3
The temperature-dependence of rate the constants and the correla-
130 k0s 2.68 × 10-3 6.47 × 10-4
tion coefficient (R2) are displayed in Fig. 14, and the fitted results are k0p 9.92 × 10-2 1.93 × 10-3
presented in Table 6. The obtained reaction activation energies were

8
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

Fig. 15. Comparison of results from experiment and simulation using fixed-bed
Fig. 13. PO consumption and related simulation vs. time profiles at different (for reaction conditions, see Table 3).
temperatures (lines and symbols represent the simulation and experimental
data). Reaction conditions: reactant 45 g, ME/PO ratio: 6, catalyst: 0.55 wt%;
pressure: 1.2 MPa; agitation: 600 rpm).

Fig. 16. Effect of ME/PO molar ratio on PO conversion and 1PO2 selectivity.
Reaction conditions: LSHV = 6 h−1.

Fig. 14. Temperature-dependence of rate constants.

Table 6
Estimated optimal parameters with 95% confidence intervals.
Parameter Optimal estimate 95% Confidence intervals

E0s 80.9 17.0


E0p 77.9 22.1
A0s 1.08 × 108 1.93 × 102
A0p 1.33 × 109 9.46 × 102

molar ratio promotes side reactions. It is intuitively deduced from


Fig. 17 that the 1PO2 selectivity increased and the PO conversion de-
creased with increasing LHSV, because higher LHSV led to short re-
sidence times, which influences the PO conversion as well as the pro- Fig. 17. Effect of LHSV on PO conversion and 1PO2 selectivity. Reaction con-
duct selectivity. To enhance the product separation and production ditions: ME/PO molar ratio = 6.
yield, the appropriate conditions employing a molar ratio = 10 and
LHSV = 5.5 were adopted and XPO = 98.8% and S1PO2 = 98.1% were
analysis shows that P-DVB-VIM possesses a structure with pores and
obtained.
channels, and BET analysis proves that the sample has a large specific
surface area with mesoporous features. P-DVB-VIM exhibits out-
4. Conclusion
standing catalytic performance with 97.7% conversion of PO and 97.8%
selectivity for 1PO2 at 120 ℃ within 45 min. The experimental results
P-DVB-VIM was successfully synthesized and used as a catalyst for
indicate that the TOF values of P-DVB-VIM are the highest compared
the synthesis of PGME via the reaction of PO with methanol. SEM

9
R. An, et al. Chemical Engineering Journal 405 (2021) 126636

with other catalysts. Because of the formation of a hydrogen bond be- [9] S. Chen, G. Zeng, Y. Li, B. He, R. Liu, S. Zhang, Epoxide ring-opening reaction
tween PVIM and the reactant, excellent catalytic performance was ob- promoted by ionic liquid reactivity: interplay of experimental and theoretical stu-
dies, Catal. Sci. Technol. 9 (20) (2019) 5567–5571.
tained by the “electrophile nucleophile dual activation” mechanism. [10] J. Zhang, Q. Cai, J. Zhao, S. Zang, Nano metal oxides as efficient catalysts for se-
The kinetics of the propoxylation reaction using the P-DVB-VIM catalyst lective synthesis of 1-methoxy-2-propanol from methanol and propylene oxide, RSC
was studied in a batch reactor over the temperature range of 100–130 Adv. 8 (8) (2018) 4478–4482.
[11] W. Zhang, H. Wang, W. Wei, Y. Sun, Solid base and their performance in synthesis
℃. The PH kinetics model and the corresponding parameters were de- of propylene glycol methyl ether, J. Mol. Catal. A: Chem. 231 (2005) 83–88.
termined. The static experiments indicate that the effect of external [12] W. Zhang, Q. Dong, N. Zhao, W. Wei, Y. Sun, The reaction mechanism for the
diffusion can be ignored. The kinetics model provided good agreement synthesis of 1-methoxy-2-propanol on MgO catalyst, Acta Physchim Sin 6 (2005)
653–657.
between the calculated and experimental values. Under the optimal [13] M.N. Timofeeva, A.E. Kapustin, V.N. Panchenko, E.O. Butenko, V.V. Krupskaya,
operation conditions, the PO conversion and 1PO2 selectivity in the A. Gil, M.A. Vicente, Synthetic and natural materials with the brucite-like layers as
continuous process implemented in the fixed-bed reactor reached high active catalyst for synthesis of 1-methoxy-2-propanol from methanol and
propylene oxide, J. Mol. Catal. A: Chem. 423 (2016) 22–30.
98.8% and 98.1%, respectively. This work provides a significant guide
[14] W. Cheng, W. Wang, Y. Zhao, L. Liu, J. Yang, M. He, Influence of acid-base prop-
for the optimization and design of propoxylation reactions. erties of ZnMgAl-mixed oxides for the synthesis of 1-methoxy-2-propanol, Appl.
Clay Sci. 42 (1-2) (2008) 111–115.
Declaration of Competing Interest [15] H.-Y. Zeng, Y.-j. Wang, Z. Feng, K.-y. You, C.e. Zhao, J.-W. Sun, P.-l. Liu, Synthesis
of Propylene Glycol Monomethyl Ether Over Mg/Al Hydrotalcite Catalyst, Catal
Lett 137 (1-2) (2010) 94–103.
The authors declare that they have no known competing financial [16] M.N. Timofeeva, V.N. Panchenko, A. Gil, Y.A. Chesalov, T.P. Sorokina,
V.A. Likholobov, Synthesis of propylene glycol methyl ether from methanol and
interests or personal relationships that could have appeared to influ- propylene oxide over alumina-pillared clays, Appl. Catal. B 102 (2011) 433–440.
ence the work reported in this paper. [17] M.N. Timofeeva, V.N. Panchenko, J.W. Jun, Z. Hasan, O.V. Kikhtyanin,
I.P. Prosvirin, S.H. Jhung, Effect of the acid–base properties of metal phosphate
molecular sieves on the catalytic performances in synthesis of propylene glycol
Acknowledgments methyl ether from methanol and propylene oxide, Microporous and Mesoporous
Materials 165 (2013) 84–91.
This work was supported by the National Nature Science [18] S. Liang, Y. Zhou, H. Liu, T. Jiang, B. Han, Synthesis of Propylene Glycol Methyl
Ether Catalyzed by MCM-41, Synth. Commun. 41 (6) (2011) 891–897.
Foundation of China (21878315 and 21808223), National Key Research
[19] M.N. Timofeeva, V.N. Panchenko, M.M. Matrosova, A.S. Andreev, S.V. Tsybulya,
and Development Program of China (2017YFA0206803), the Key A. Gil, M.A. Vicente, Factors Affecting the Catalytic Performance of Zr,Al-Pillared
Programs of Fujian Institute of Innovation, CAS (FJCXY18020203), the Clays in the Synthesis of Propylene Glycol Methyl Ether, Ind. Eng. Chem. Res. 53
(35) (2014) 13565–13574.
National Science Fund for Excellent Young Scholars (21722610), the
[20] M.H. Allen Jr., S.T. Hemp, A.E. Smith, T.E. Long, Controlled Radical Polymerization
One Hundred Talent Program of CAS, K. C. Wong Education Foundation of 4-Vinylimidazole, Macromolecules 45 (9) (2012) 3669–3676.
(No. GJTD-2018-04). The authors are grateful for the assistance from [21] M.D.S. Lay, Kinetics of the liquid-phase addition reactions initiated by propylene
teachers Hui Wu, Ling Wang and Na Zhou of Analysis and Test Center, oxide and methanol and catalyzed by sodium hydroxide, PhD thesis University of
Michigan, Detroit, 1959.
Institution of Process Engineering, Chinese Academy of Sciences. [22] M. Di Serio, G. Vairo, P. Iengo, F. Felippone, E. Santacesaria, Kinetics of
Sincerely appreciate Prof. Suojiang Zhang (IPE,CAS) for his careful Ethoxylation and Propoxylation of 1- and 2-Octanol Catalyzed by KOH, Ind. Eng.
academic guidance and great support. Chem. Res. 35 (11) (1996) 3848–3853.
[23] M. Di Serio, Kinetics of Ethoxylation and Propoxylation of Ethylene Glycol
Catalyzed by KOH, Ind Eng Chem Res 35 (2002) 3848–3853.
Appendix A. Supplementary data [24] T. Pöpken, L. Götze, J. Gmehling, Reaction Kinetics and Chemical Equilibrium of
Homogeneously and Heterogeneously Catalyzed Acetic Acid Esterification with
Methanol and Methyl Acetate Hydrolysis, Ind. Eng. Chem. Res. 39 (7) (2000)
Supplementary data to this article can be found online at https:// 2601–2611.
doi.org/10.1016/j.cej.2020.126636. [25] H. Chuncheng, Catalytic distillation Experimental study and theoretical modelling,
University of Waterloo, Ontario, 1999.
[26] J. Pecha, L. Šánek, T. Fürst, K. Kolomazník, A kinetics study of the simultaneous
References methanolysis and hydrolysis of triglycerides, Chem. Eng. J. 288 (2016) 680–688.
[27] A. Orjuela, A.J. Yanez, A. Santhanakrishnan, C.T. Lira, D.J. Miller, Kinetics of mixed
[1] C. Zhao, S. Chen, R. Zhang, Z. Li, R. Liu, B. Ren, S. Zhang, Synthesis of propylene succinic acid/acetic acid esterification with Amberlyst 70 ion exchange resin as
glycol ethers from propylene oxide catalyzed by environmentally friendly ionic li- catalyst, Chem. Eng. J. 188 (2012) 98–107.
quids, Chin. J. Catal. 38 (5) (2017) 879–888. [28] L. Shen, L. Wang, H. Wan, G. Guan, Transesterification of Methyl Acetate withn-
[2] R.E. Parker, N.S. Isaacs, Mechanisms Of Epoxide Reactions, Chem. Rev. 59 (4) Propanol: Reaction Kinetics and Simulation in Reactive Distillation Process, Ind Eng
(1959) 737–799. Chem Res 53 (2014) 3827–3833.
[3] H.C. Chitwood, B.T. Freure, The Reaction of Propylene Oxide with Alcohols, J. Am. [29] D.A. González-Casamachin, J. Rivera De La Rosa, C.J. Lucio-Ortiz, D.A. De Haro De
Chem. Soc. 68 (4) (1946) 680–683. Rio, D.X. Martínez-Vargas, G.A. Flores-Escamilla, N.E. Dávila Guzman, V.M.
[4] S. Chen, R. Liu, Y. Li, R. Zhang, C. Zhao, H. Tang, C. Qiao, S. Zhang, Relationship of Ovando-Medina, E. Moctezuma-Velazquez, Visible-light photocatalytic degradation
basicity and hydrogen bond properties of ionic liquids with its catalytic perfor- of acid violet 7 dye in a continuous annular reactor using ZnO/PPy photocatalyst:
mance: Application to synthesis of propylene glycol methyl ether, Catal Commun 96 Synthesis, characterization, mass transfer effect evaluation and kinetic analysis,
(2017) 69–73. Chem Eng J, 373(2019)325-337.
[5] D.-J. Tao, F. Ouyang, Z.-M. Li, N.a. Hu, Z. Yang, X.-S. Chen, Synthesis of [30] A.V. Sulimov, S.M. Danov, A.V. Ovcharova, A.A. Ovcharov, V.R. Flid,
Tetrabutylphosphonium Carboxylate Ionic Liquids and Its Catalytic Activities for O.V. Ugryumov, Vapor–liquid equilibrium in the system of propylene oxide
the Alcoholysis Reaction of Propylene Oxide, Ind. Eng. Chem. Res. 52 (48) (2013) synthesis products, Theor Found Chem Eng 49 (6) (2015) 854–863.
17111–17116. [31] F. Liu, A. Zheng, I. Noshadi, F. Xiao, Design and synthesis of hydrophobic and stable
[6] Y.u. Bai, Q. Cai, X. Wang, B. Lu, Tunable synthesis of propylene glycol ether from mesoporous polymeric solid acid with ultra strong acid strength and excellent
methanol and propylene oxide under ambient pressure, Kinet Catal 52 (3) (2011) catalytic activities for biomass transformation, Appl. Catal. B 136–137 (2013)
386–390. 193–201.
[7] C. De, Q. Cai, X. Wang, J. Zhao, B. Lu, Highly selective synthesis of propylene glycol [32] W. Zhang, H. Wang, Q. Li, Q. Dong, N. Zhao, W. Wei, Y. Sun, The mechanism for the
ether from methanol and propylene oxide catalyzed by basic ionic liquid, J. Chem. synthesis of 1-methoxy-2-propanol from methanol and propylene oxide over mag-
Technol. Biotechnol. 86 (1) (2011) 105–108. nesium oxide, Appl. Catal. A 294 (2) (2005) 188–196.
[8] Y.-M. Liu, Y. Zhou, W.-Q. Gong, Z.-M. Li, C.-L. Wang, D.-J. Tao, Highly efficient [33] S. Denmark, G. Beutner, Lewis Base Catalysis in Organic Synthesis, Angew. Chem.
synthesis of 1-methoxy-2-propanol using ionic liquid catalysts in a micro-tubular Int. Ed. 47 (9) (2008) 1560–1638.
circulating reactor, Green Energy Environ. 5 (2) (2020) 147–153.

10

You might also like