Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

pubs.acs.

org/IECR Article

Microwave Fixed-Bed Reactor for Gas-Phase Glycerol Dehydration:


Experimental and Simulation Studies
Qinglong Xie, Tongbo Pan, Gaoji Zheng, Yuqiang Zhou, Shangzhi Yu, Ying Duan, and Yong Nie*
Cite This: Ind. Eng. Chem. Res. 2022, 61, 10723−10735 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV AUTONOMA METROPOLITANA on April 5, 2024 at 23:45:05 (UTC).

ABSTRACT: Conventional fixed-bed reactors have the issue of uneven temperature distribution within the catalyst bed, which
influences product selectivity and catalyst stability. Microwave irradiation can realize uniform and bulk heating due to its unique
penetration capability. Here, a microwave fixed-bed reactor was developed and applied in a gas-phase glycerol dehydration reaction.
Nine microwave-absorbing binary metal oxide catalysts were used to verify the versatility of microwave reactors in gas−solid
reactions. Compared with an electric-heating reactor, much better catalyst stability was obtained in a microwave reactor for all of the
catalysts. Moreover, no obvious change in catalyst stability was observed with increasing reactor size. Numerical simulation was
adopted to explore the intensification mechanism of microwave heating. The largest temperature difference in the microwave reactor
was around 10 °C, much smaller than that in the electric-heating reactor which reached 57 °C. The electric field distribution within
the penetration depth of microwave irradiation was homogeneous, resulting in uniform temperature distribution in the microwave
reactor and hence good catalyst stability.

1. INTRODUCTION frequency electromagnetic field to generate heat.7 Different


Fixed-bed reactors are widely used in gas−solid and liquid− from conventional heating modes, microwave heating can
solid catalytic reactions due to their simple construction and realize the bulk heating of feedstock through the dissipation of
high surface-area-to-volume ratio.1 However, the issue of microwave energy inside the dielectric.8 Consequently,
uneven temperature distribution within the catalyst bed usually compared with conventional heating, the temperature gradient
exists in fixed-bed reactors because of the conventional within the feedstock can be effectively reduced under
conduction heating method and poor thermal conductivity of microwave heating to achieve a more uniform temperature
most catalysts.2 To improve the heat transfer, a multistage distribution.
fixed-bed reactor system was developed that uses interstage Microwave heating has been widely used in various organic
heat exchange and reduces the overall temperature variation of reactions and proved to have an obvious intensification effect
the catalyst bed, and yet the radial temperature gradient within by significantly reducing the reaction time.9 Recently, micro-
the bed still exists.3 The temperature distribution is enhanced wave-assisted biomass pyrolysis for biofuel production has
in a multitubular reactor through the reduction of tube attracted much attention.10,11 Rapid and efficient conversion of
diameter.4 However, a very small tube size causes the wall
effect resulting in uneven airflow in the catalyst bed and hence
restricted heat and mass transfer between the fluid and catalyst Received: April 4, 2022
particles.5 In addition, the reduction of tube and catalyst size Revised: July 2, 2022
leads to an increase in the pressure drop and thus energy Accepted: July 11, 2022
consumption for the fixed-bed reactor.6 Published: July 20, 2022
Microwave irradiation can induce the rapid movement and
friction of polar molecules (dipoles) in the presence of a high-

© 2022 American Chemical Society https://doi.org/10.1021/acs.iecr.2c01176


10723 Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

biomass can be achieved using a premixed method in which


feedstock is mixed with a microwave absorbent or a non-
premixed method, i.e., feedstock is poured onto the heated
microwave absorbent bed.11 Our group previously reported
methyl ricinoleate pyrolysis by microwave heating coupled
with atomization feeding and obtained high product yields.12,13
However, the above microwave-assisted processes mainly
involved liquid−liquid, liquid−solid, and solid−solid reactions
in tank reactors, and few studies on gas−solid catalytic
reactions in fixed-bed reactors under microwave heating have
been reported.
Gas-phase glycerol dehydration is usually adopted for the
production of acrolein, which is an important chemical
intermediate to produce value-added chemicals such as acrylic, Figure 1. Experimental setup of a microwave fixed-bed reactor. (1)
1,3-propanediol, and methionine.14 The dehydration reaction Infrared radiation thermometer, (2) microwave cavity, (3) fixed-bed
is usually performed in a fixed-bed reactor packed with solid reactor, (4) catalyst bed, (5) microwave constraint chamber, (6)
magnetron, and (7) waveguide.
acid catalysts such as heteropoly acids, phosphates, molecular
sieves, and metal oxides.15 However, the glycerol dehydration
process belongs to complex reactions that are highly sensitive located in the middle of the microwave cavity. The
to temperature.16 Parallel side reaction of glycerol polymer- temperature of the catalyst bed was measured by an infrared
ization and series side reaction of acrolein polymerization are radiation thermometer, with the signal feedback controlling the
favored at lower and higher temperatures, respectively, to form magnetron to be on or off. A microwave constraint chamber
coke on the surface of the catalyst.17 Thus, rapid catalyst with the height same as the catalyst bed was connected with
deactivation ascribed to coke deposition is often observed in the waveguide, with the purpose of increasing the power
conventional fixed-bed reactors due to the existence of a density of microwave energy. In addition, a microwave
temperature gradient. The acrolein yield obtained in the gas- detector (DT-2G, CEM Corporation) was used to monitor
phase glycerol dehydration was reported to reach around 60%, microwave leakage for safety purposes.
yet the deactivation of most catalysts occurred within 10 h.18 2.2. Catalyst Preparation and Characterization. Nine
In our previous work, microwave-assisted glycerol dehy- typical binary metal oxide catalysts for glycerol dehydration
dration to acrolein was conducted over a coated microwave- were prepared with SiC of 500 nm particle size as the
absorbing catalyst WO3−ZrO2@SiC.19 Higher acrolein yield microwave absorbent. Among these, four catalysts, i.e.,
and better catalyst stability were achieved under microwave Nb2O5−ZrO2@SiC, Nb2O5−Al2O3@SiC, WO3−ZrO2@SiC,
heating compared with the electric-heating process. However, and WO3−Al2O3@SiC, were obtained using an excessive
the physicochemical properties of the catalytic materials may impregnation method, which were labeled as Nb−Zr, Nb−Al,
influence the distributions of electromagnetic and temperature W−Zr, and W−Al, respectively. A certain amount of SiC was
fields in the microwave reactor. Therefore, the versatility of a first dispersed in deionized water, followed by the addition of
microwave fixed-bed reactor in gas−solid catalytic reactions ZrOCl2 or Al(NO3)3 solution. Afterward, diluted aqueous
over different catalysts should be explored. Moreover, the ammonia was added dropwise to the solution with a pH of
mechanism of microwave heating in the improvement of 9.0−10.0, which was then aged for 1 h. The precipitate was
product yield and catalyst stability has not been investigated. filtered, exhaustively washed, dried at 110 °C for 12 h, and
In this work, a microwave fixed-bed reactor was developed finally calcined at 550 °C for 4 h. Subsequently, the obtained
for the gas-phase glycerol dehydration reaction. The versatility powder was impregnated with ammonium metatungstate or
of the microwave reactor was verified by being packed with niobium ammonium oxalate solution. The mixture was stirred
nine typical catalysts for glycerol dehydration. The catalyst and exposed to infrared radiation at 50 °C for 24 h, followed
stability in microwave-assisted glycerol dehydration was by being calcined at 600 °C for 6 h yielding the microwave-
experimentally studied and compared with the conventional absorbing binary metal oxide catalysts.
heating method. To explore the intensification mechanism of The co-precipitation method was used to prepare the other
microwave heating, the temperature distribution of the catalyst five catalysts, i.e., 10%SnO2−ZrO2@SiC (SnO2:(SnO2 +
bed and the electric field distribution within the reactor were ZrO2) = 10 wt%, similarly hereinafter), 30%SnO2−ZrO2@
determined by numerical simulation. SiC, 50%SnO2−ZrO2@SiC, TiO2−ZrO2@SiC, and TiO2−
SnO2@SiC, which were labeled as 10%Sn−Zr, 30%Sn−Zr,
2. EXPERIMENTAL SECTION 50%Sn−Zr, Ti−Zr, and Ti−Sn, respectively. A certain amount
2.1. Experimental Setup of the Microwave Fixed-Bed of SiC was first dispersed in deionized water, followed by the
Reactor. The experimental setup of the microwave fixed-bed addition of SnCl4, TiCl4, and ZrOCl2 in ethanol solution.
reactor is shown in Figure 1. The microwave irradiation was Afterward, diluted aqueous ammonia was added dropwise to
provided in a self-manufactured microwave cavity equipped the solution with a pH of 9.0−10.0, which was then aged for 1
with a water-cooled magnetron (LG 2M246) at a power of h. The precipitate was filtered, exhaustively washed, dried at
1000 W and a frequency of 2450 MHz. The microwave power 110 °C for 12 h, and finally calcined at 600 °C for 6 h.
can be controlled by adjusting the anode current of the The catalysts were characterized to determine their
magnetron using a potentiometer. The microwaves generated physicochemical properties. The specific surface area and
by the magnetron were transmitted into the cavity through a pore size, morphology, crystalline structures, and acid proper-
standard BJ26 waveguide (L × W = 86 mm × 43 mm). A ties of catalysts were analyzed by Brunauer−Emmett−Teller
vertical quartz tubular reactor packed with catalysts was (BET) method, transmission electron microscopy (TEM)
10724 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 2. TEM and HAADF-STEM images and the corresponding elemental mapping of (a) W−Zr and (b) 50%Sn−Zr (the red rectangle is the
area for elemental mapping).

coupled with elemental mapping, X-ray diffraction (XRD), conditions were consistent with those of the microwave-
Fourier-transform infrared spectroscopy of pyridine (pyridine- heating process.
FTIR), and thermogravimetry-differential scanning calorimetry 2.4. Product Analysis. For the methods for product
(TG-DSC), respectively. The details of these characterization analysis, we can refer to our previous report.19 Briefly, glycerol
techniques can be found in our previous work.19 In addition, in the liquid was analyzed using a Shimadzu LC-20AT high-
thermogravimetric analysis/infrared (TG-IR) spectrometry performance liquid chromatograph (HPLC) equipped with a
analysis of the used catalyst was performed using a NETZSCH XAmide column and a refractive index detector (RID), with
TG55 thermal analyzer and a Thermo Scientific NicoletiS50 the glycerol conversion calculated by eq 2. Acrolein and
spectrometer. The sample was placed on an alumina crucible byproducts (mainly acetaldehyde and hydroxyacetone) were
and heated from room temperature to 800 °C at a rate of 10 analyzed using an Agilent 7890A gas chromatograph (GC)
°C/min under the air atmosphere. Solid-state C13 nuclear with a DB-WAX capillary column and a flame ionization
magnetic resonance (NMR) spectra of used catalysts were detector (FID). The selectivity and yield of products were
obtained on a Bruker 400 M spectrometer equipped with a 4 determined using eqs 3 and 4.
mm probe at a spinning rate of 40 kHz. The complex
permittivity of catalysts was measured in the frequency range glycerol conversion (%)
from 1 to 18 GHz at 150 °C by means of the coaxial method ij the amount of glycerol in liquid product yzz
= jjj1 z
using an Agilent PNA-N5234A vector network analyzer. The j
k the amount of glycerol in feedstock zz{
thermal diffusivity and specific heat capacity of catalysts were
determined on a NETZSCH LFA 457 MicroFlash laser-flash × 100% (2)
apparatus, which were then used for the calculation of the
thermal conductivity of catalysts. product selectivity (%)
2.3. Catalytic Tests. The gas-phase glycerol dehydration the amount of glycerol converted to a product
reactions were performed in the microwave fixed-bed reactor at = × 100%
various diameters (10, 15, and 20 mm). Prior to microwave the amount of reacted glycerol
heating, N2 at a flow rate of 30 mL/min was introduced to (3)
remove the air in the reactor. The catalysts absorbed the
microwaves and were heated under microwave irradiation. product yield (%)
Once the reaction temperature was reached, an aqueous the amount of glycerol converted to a product
= × 100%
solution of glycerol (20 wt%), after being preheated and the amount of glycerol in feedstock
vaporized, was fed into the reactor by a peristaltic pump. (4)
During the reaction, the temperature of the catalyst bed
monitored by the infrared thermometer was maintained stable
by controlling the magnetron on or off by the signal feedback. 3. RESULTS AND DISCUSSION
The reaction products were condensed and collected for 3.1. Characterization of the Catalysts. The specific
subsequent analysis. The yield of the liquid product was surface area and pore structure of nine microwave-absorbing
calculated using eq 1. catalysts are listed in Table S1. Compared with the SiC alone,
the specific surface area and pore volume were obviously
liquid product yield (wt%) increased for the binary metal oxide catalysts. This indicated
that certain pore structures were formed during the catalyst
the weight of liquid product preparation, which could favor the dispersion of active
= × 100%
the weight of feedstock (1) components and hence improve the activity of catalysts.
Figure 2 shows the TEM images of W−Zr and 50%Sn−Zr
In addition, glycerol dehydration under conventional electric that were obtained using the excessive impregnation and co-
heating was conducted and compared with the microwave- precipitation methods, respectively (TEM images of other
heating method. The experimental procedures and reaction catalysts are displayed in Figure S1). The platelet-shaped
10725 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 3. XRD patterns of binary metal oxide catalysts prepared by (a) excessive impregnation method and (b) co-precipitation method.

microstructures can be observed for all of the catalysts. In


addition, the distributions of Si, O, and metal elements on the
catalysts were quite uniform, as seen from the elemental
mapping acquired using high-angle annular dark-field STEM
(HAADF-STEM) with an energy-dispersive X-ray (EDX)
spectrometer.
The powder XRD patterns of the nine microwave-absorbing
binary metal oxide catalysts are displayed in Figure 3. The
characteristic peaks of SiC and most metal oxides can be
obviously found for catalysts prepared by both excessive
impregnation and co-precipitation methods, indicating the
existence of corresponding crystalline phases in the catalysts.
However, since TiO2 after being calcined at 600 °C basically
existed in the amorphous phase,20 no characteristic peaks of
TiO2 were observed in the XRD spectrum of the Ti−Zn or
Ti−Zr catalyst.
The Brønsted and Lewis acid sites on the catalysts were Figure 4. Pyridine-FTIR profiles at a desorption temperature of 200
analyzed according to the pyridine-FTIR spectra obtained at °C for nine binary metal oxide catalysts.
200 and 350 °C. As shown in Figures 4 and S2, similar FTIR
profiles and desorption peaks were observed for the nine
binary metal oxide catalysts at both desorption temperatures. centered at around 1490 cm −1 were ascribed to the
The bands resulting from Brønsted acid sites existed at around combination of Brønsted and Lewis acid sites.22 The density
1546 and 1640 cm−1, with those attributed to Lewis acid sites of Brønsted and Lewis acid sites on the catalysts was
appearing at ca. 1445, 1546, and 1610 cm−1.21 The peaks determined from the integrated areas of corresponding peaks
10726 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

using the extinction coefficients,23 as listed in Table 1. The obtained over various catalysts were significantly different at
amounts of Brønsted and Lewis acid sites and the ratio of both dehydration temperatures, which was primarily attributed
to the catalyst acidity.24 It is accepted that the existence of
Table 1. Acidity of Nine Binary Metal Oxide Catalysts as Brønsted acid sites enhances the glycerol dehydration to
Determined by the Pyridine-FTIR Technique acrolein by protonating the secondary hydroxyl group of
glycerol.25 More Lewis acid sites on the catalyst would cause
density of acid sites by pyridine-FTIR (μmol/gcat.)
an increase in the formation of byproducts such as
catalysts B200a L200a B200/L200 B350a L350a B350/L350 hydroxyacetone (Table S2) and hence the reduction in
W−Zr 119.4 20.6 5.8 103.2 10.0 10.3 acrolein selectivity.26 However, according to Tables 1 and 2,
Nb−Zr 121.3 29.9 4.1 51.4 8.9 5.8 the acrolein selectivity exhibited no obvious correlation with
W−Al 101.4 20.8 4.9 69.8 6.0 11.6 the density of Brønsted or Lewis acid sites on the catalysts.
Nb−Al 87.2 21.0 4.2 16.4 4.6 3.5 Figure 5 displays the relationship between acrolein selectivity
10Sn−Zr 171.1 90.0 1.9 97.5 39.7 2.5 and the Brønsted-to-Lewis-acid density ratio, with a certain
30Sn−Zr 84.6 52.7 1.6 48.7 15.6 3.1 positive correlation observed. It indicated that the acrolein
50Sn−Zr 69.0 45.4 1.5 47.9 8.9 5.4 selectivity in the glycerol dehydration process largely depended
Sn−Ti 145.6 95.4 1.5 42.5 22.1 1.9 on the ratio of Brønsted to Lewis acid density on the catalyst
Zr−Ti 101.4 51.9 1.9 36.2 7.7 4.7 rather than their respective amounts. In addition, as shown in
a
B: Brønsted acid sites and L: Lewis acid sites. The number after B Tables 2 and S3, higher glycerol conversion was obtained at
and L refers to the pyridine desorption temperature (in °C). the higher temperature for all of the catalysts mainly because of
the higher reaction rate of glycerol dehydration. On the other
Brønsted acid density to Lewis acid density were very different hand, side reactions such as glycerol polymerization were
for various binary metal oxide catalysts, which would lead to promoted at the lower temperature,27 which resulted in lower
their difference in catalytic performance in glycerol dehy- acrolein selectivity.
dration. 3.3. Catalyst Stability. The stability of the nine binary
3.2. Glycerol Dehydration Tests in the Microwave metal oxide microwave-absorbing catalysts in gas-phase
Fixed-Bed Reactor. The tests of glycerol dehydration in a glycerol dehydration using the microwave fixed-bed reactor
microwave fixed-bed reactor packed with nine binary metal was examined and compared with that in a conventional
oxide catalysts were conducted at 250 and 300 °C. As shown in electric-heating reactor. The stability of the catalysts prepared
Table 2, the glycerol conversion and acrolein selectivity by excessive impregnation and co-precipitation methods is
shown in Figures 6 and 7, respectively. The stability
Table 2. Glycerol Dehydration in the Microwave Fixed-Bed performance varied for different catalysts. Overall, the catalysts
Reactor Packed with Various Binary Metal Oxide Catalysts with a higher Brønsted-to-Lewis-acid density ratio exhibited
at 250 and 300 °C better stability than those with a lower ratio. This was mainly
liquid product glycerol conversion acrolein selectivity because more byproducts such as hydroxyacetone and
yield (wt%) (%) (%) acetaldehyde were generated when the Lewis acid density
catalysts 250 °C 300 °C 250 °C 300 °C 250 °C 300 °C was relatively high, which enhanced the formation of coke
W−Zr 97.0 97.5 98.5 99.7 68.1 73.2
deposits on the catalyst and hence reduced the catalyst
Nb−Zr 99.8 99.1 51.0 86.2 36.7 58.4
stability.
W−Al 98.4 96.9 90.9 99.5 52.9 59.0
For all of the catalysts and both dehydration temperatures,
Nb−Al 99.1 99.7 27.1 94.0 45.7 62.3
better catalyst stability was observed in the microwave fixed-
10Sn−Zr 94.1 96.6 42.3 87.6 15.6 20.7
bed reactor compared with the conventional electric-heating
30Sn−Zr 92.2 98.2 35.7 78.1 9.7 17.3
reactor. The main possible reason for this lies in the unique
50Sn−Zr 99.4 97.9 22.6 75.2 18.9 22.7 penetration capability and bulk heating of microwave
Sn−Ti 92.4 96.6 43.0 78.8 20.2 20.0 irradiation, which resulted in more uniform temperature
Zr−Ti 98.4 99.4 68.7 97.4 30.0 39.1 distribution of the catalyst bed. By contrast, since the
temperature was monitored and controlled at the outer wall

Figure 5. Correlation between acrolein selectivity and Brønsted-to-Lewis-acid density ratio for various binary metal oxide catalysts at (a) 250 °C
and (b) 300 °C.

10727 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 6. Glycerol conversion with time on stream in the microwave fixed-bed reactor (filled symbols with solid lines) and an electric-heating
reactor (open symbols with dashed lines) packed with catalysts prepared by the excessive impregnation method including (a) W−Zr, (b) Nb−Zr,
(c) W−Al, and (d) Nb−Al at temperatures of (●) 250 °C and (■) 300 °C.

of the electric-heating reactor, the temperature gradient caused catalyst bed can be achieved as long as the reactor size was
by the thermal resistance of catalyst particles and heat within the penetration depth of microwaves. From this
consumption of the glycerol dehydration reaction led to perspective, the microwave fixed-bed reactor has more
lower temperatures within the catalyst bed. Thus, side potential for reactor scale-up.
reactions, especially glycerol polymerization, were promoted, 3.4. Coking of Catalysts. According to our previous study
which increased the coke formation for catalyst deactivation.28 and other reports, coking of catalyst was considered the main
This can be confirmed by the better catalyst stability obtained reason for the reduction of catalyst stability during glycerol
at the temperature of 300 °C than that at 250 °C for all of the dehydration.17,19 Characterization techniques including TG-
catalysts. Since various binary metal oxide catalysts with DSC, TG-IR, and NMR were used to explore the mechanism
different physicochemical properties were used, the results of coke formation. Figure 9a displays TG-DSC curves of a
proved the universal superiority of microwave fixed-bed typical catalyst after a 16 h reaction at 250 °C in the electric-
reactors over electric-heating reactors in gas−solid catalytic heating reactor. DTG and DSC peaks of the used catalyst at
reactions. around 460 °C can be clearly observed. According to the
The advantage of a microwave fixed-bed reactor over a corresponding TG-IR curves, as shown in Figure 9b, the
conventional electric-heating reactor became more obvious weight loss at 460 °C resulted in the formation of CO2. Thus,
with increasing reactor size. As shown in Figure 8, the catalyst the TG-DSC-IR results verified the existence of carbonaceous
stability in the electric-heating reactor was continuously deposits on the used catalyst. In addition, the weight loss
reduced when the reactor diameter was increased from 10 to increased from 6.8 to 11.5% for the used catalysts in electric-
20 mm. The glycerol conversion obtained in the 20 mm heating reactors with diameters increasing from 10 to 20 mm.
reactor was lower than 80% after a 16 h run over the W−Zr This indicated the generation of more coke deposits at a larger
catalyst at 250 °C. By contrast, no significant difference in reactor size, which caused more rapid catalyst deactivation as
catalyst stability was observed for microwave fixed-bed reactors presented in Figure 8.
of various sizes. The glycerol conversion was still higher than The typical 13C NMR spectrum of the used catalyst at 250
95% after the 16 h reaction. For common fixed-bed reactors °C as shown in Figure S3 exhibited two main peaks at σ = 71
with the heat transfer mainly through conduction, the and 73 ppm as well as two minor peaks at σ = 20 and 132 ppm.
temperature difference within the catalyst bed would be The main peaks corresponding to branched oxygen-containing
greater at a larger reactor size. Consequently, the even lower groups were attributed to polyglycol compounds.29 This
temperature enhanced the side reactions for the coking of confirmed that the primary reason for the coking of catalysts
catalysts. Because of the special heating mechanism of was glycerol polymerization, which usually occurs at lower
microwave irradiation, the bulk and uniform heating of the temperatures. The peak at σ = 20 ppm was associated with
10728 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 7. Glycerol conversion with time on stream in the microwave fixed-bed reactor (filled symbols with solid lines) and an electric-heating
reactor (open symbols with dashed lines) packed with catalysts prepared by the co-precipitation method including (a) 10Sn−Zr, (b) 30Sn−Zr, (c)
50Sn−Zr, (d) Sn−Ti, and (e) Zr−Ti at temperatures of (●) 250 °C and (■) 300 °C.

saturated carbon atoms of terminal chains bound to the adopted to determine the temperature distributions using the
oligomeric compounds that had not yet cyclized, with that COMSOL Multiphysics software.
centered at σ = 132 ppm ascribed to polyaromatic 3.5.1. Model Simplification. The simulation involving an
compounds.29 These minor carbonaceous deposits may be electromagnetic field coupled with heat transfer is a highly
produced from aldol condensation of acrolein, hydroxyacetone, coupling and nonlinear process. To reduce the difficulty of
and acetaldehyde.30 simulation and save simulation time, some reasonable
3.5. Numerical Simulation. According to the experimen- assumptions are proposed to simplify the model. (1) The
tal results, the temperature distribution of the catalyst bed is a microwave absorption capacity of air and glass is zero, (2) the
key factor influencing the catalyst stability in the fixed-bed catalyst bed is the only heat source and the reaction heat of the
reactor. Thus, it is essential to determine and compare the endothermic glycerol dehydration reaction is considered as
temperature distributions within the electric-heating and equivalent heat loss from the whole catalyst bed, (3) the
microwave-heating reactors. However, it is difficult to monitor catalyst particles within the reactor are spatially homogeneous
all of the temperature points within the catalyst bed, especially with the whole catalyst bed simplified as a porous medium for
under microwave irradiation.31 In this work, a numerical simulation, and (4) the dielectric and thermal properties of the
simulation based on the finite element method (FEM) was catalysts are constants with temperature.
10729 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Table 3. Dimensions of the Main Components of the


Microwave Fixed-Bed Reactor
length width height
(mm) (mm) (mm)
microwave cavity 300 320 320
waveguide 86 86 43
microwave constraint chamber 180 160 60
catalyst bed 50

device. The microwave cavity, waveguide, and microwave


constraint chamber were polished 304 steel, with the reactor
being made of quartz. The properties of catalysts were used
according to their measurements. The dielectric and thermal
Figure 8. Glycerol conversion with time on stream using reactors with properties of steel, quartz, and typical catalysts used for
various diameters. The reaction temperature was 250 °C, and the simulation work are displayed in Table 4.
catalyst was WO3−ZrO2@SiC. 3.5.4. Governing Equations. The electromagnetic field
distribution in the microwave cavity is solved using Maxwell’s
equations. The governing equation of the electric field is
expressed as follows
ij j yzz
× 1
( ×E) k 02jjj zzE = 0
r j r z
k 0{ (5)
where μr is the relative permeability of materials, E represents
the electric field, k0 is the wave number in vacuum, εr is the
relative permittivity of materials, σ is the electrical conductivity
in S/m, ω is the angular frequency in Hz, and ε0 denotes the
vacuum dielectric constant (8.85 × 10−12 F/m).
Fourier’s equation describing the heat transfer is used to
solve the temperature distribution in the microwave cavity.
The governing equation of the temperature field is shown as
follows
T
CP = · (k T ) + Q
t (6)
3
where ρ is the density of materials in kg/m , Cp represents the
specific heat capacity in J/(kg·K), T is the temperature in K, k
denotes the thermal conductivity in W/(m·K), and Q refers to
the dissipated microwave energy in the microwave-absorbing
catalyst bed in W, which is expressed by the following equation
1
Q= ( + 0 )E 2
2 (7)
where ε″ represents the imaginary part of the complex relative
permittivity, which is also known as the loss factor.
3.5.5. Boundary Conditions. The impedance boundary
Figure 9. (a) Typical TG-DSC curves of the used (solid lines) and condition refers to the surface of the microwave cavity and
fresh (dashed lines) catalyst and (b) the corresponding TG-IR curves waveguide (Figure S4b), which is expressed using the following
of the used catalyst. equation
0 r
n×H+E (n·E)n = (n·Es)n Es
3.5.2. Geometry. The geometric model of the microwave 0 r j (8)
fixed-bed reactor in the COMSOL graphic interface was
developed based on the experimental apparatus, as shown in where μ0 is the relative permeability in vacuum, H is the
Figure S4. The dimensions of the main components including magnetic field, and Es is the electric field on the surface of the
microwave cavity, waveguide, microwave constraint chamber, microwave cavity and waveguide.
and catalyst bed are listed in Table 3. In addition, the Microwaves at a frequency of 2450 MHz are excited through
percentage of voids in the catalyst bed was measured to be a rectangular port in the TE10 mode. The boundary of the port
65.2%. describes the area of microwave irradiation entering the
3.5.3. Material Properties. The material properties of the microwave cavity, as shown in Figure S4c. The propagation
geometric model were defined according to the experimental constant (β) of microwave is defined as follows
10730 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Table 4. Dielectric and Thermal Properties of Materials Used for Simulation Work
properties steela quartza W−Zr Nb−Al Sn−Zr Sn−Ti
relative permeability 1.0135 1 7.72 1.92 2.50 2.07
relative permittivityb 1 3.8 1.9332−0.2142*j 0.6233−0.9075*j 1.1333−0.0523*j 0.6067−0.8733*j
electrical conductivity (S/m) 4.032 × 106 1 × 10−14 0 0 0 0
specific heat capacity (J/(kg·K)) 741 699 925 724 980
density (kg/m3) 2485 3059 2052 2467 2667
thermal conductivity (W/(m·K)) 1.4 0.399 0.911 0.682 0.373
a
The properties are taken from the COMSOL built-in material library. bFor relative permittivity, j is the imaginary number i.

Figure 10. Temperature distributions in electric-heating fixed-bed reactors packed with various binary metal oxide catalysts including (a) W−Zr,
(b) Nb−Al, (c) Sn−Zr, and (d) Sn−Ti.

2 2 2 as shown in Figure S4d. The number of region, edge, and


= c
c0 (9) vertex elements were 341 435, 23 434, and 1053, respectively.
3.5.7. Solver. The solver based on the generalized minimum
where c0 is the velocity of light and ν and νc refer to the residual method (GMRES) was used to solve the wave
frequency and cutoff frequency of microwave in Hz, equations in the electromagnetic field, which belongs to
respectively. frequency domain analysis. The parallel direct solver
In the microwave fixed-bed reactor, the temperature of the (PARDISO) was selected to analyze the heat transfer, which
catalyst bed was monitored by an infrared radiation is a transient solution process.
thermometer that was located on the left side of the microwave 3.5.8. Model Validation. To verify the reliability and
cavity. To keep the simulation consistent with the experimental accuracy of the model, the temperature of the catalyst bed
setup, the temperature measurement point in the simulation under microwave heating monitored by the infrared radiation
was set on the left side of the catalyst bed, as displayed in thermometer was recorded and compared with that acquired in
Figure S4a. The temperature of this point was defined as the the simulation work. As shown in Figure S5, the simulated
integral of temperature at the point that was labeled as temperature profile agreed well with the experimental data.
intop(T). In the simulation work, the heating process of the This indicated that the simulation process was close to the
microwave fixed-bed reactor is controlled by the following actual experimental conditions. The slight difference between
command: experimental and simulated results was probably due to the
if (in top(T ) 300 [deg C], emw. Q , 0) (10)
assumption that the dielectric and thermal properties of the
catalysts are constant with temperature. In fact, these
which means that the excited microwave power is Q if the parameters such as thermal conductivity and specific heat
temperature of the measurement point is equal to or lower capacity of the catalysts slightly change with temperature,32
than the set value (taking 300 °C as an example), otherwise which would cause a small deviation of the simulation results.
zero. In addition, the temperature of the catalyst bed in the
3.5.6. Mesh. To guarantee the simulation accuracy, the electric-heating reactors with various diameters was also
largest size of the mesh elements for the whole geometry was detected by a thermocouple and compared with the simulated
restricted to be one-tenth of the microwave wavelength. The values. The results showed that the deviation of simulated
part of the catalyst bed was even performed a finer subdivision, temperatures from the actual measurements at different
10731 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 11. Temperature distributions of the catalyst bed in electric-heating fixed-bed reactors with various diameters of (a) 10 mm, (b) 15 mm,
and (c, d) 20 mm. The temperature of the measurement point for (a−c) is 250 °C and that for (d) is 300 °C.

Figure 12. Temperature distributions in microwave fixed-bed reactors packed with various binary metal oxide catalysts including (a) W−Zr, (b)
Nb−Al, (c) Sn−Zr, and (d) Sn−Ti.

locations in the catalyst bed was within ±5 °C. This in large thermal resistance and hence a large temperature
demonstrated that the parameter settings in the simulation gradient within the catalyst bed under the heat transfer mode
work and the assumptions to simplify the simulation are of conduction. The temperature distributions of the catalyst
reasonable. Overall, the established model can be used to bed in electric-heating fixed-bed reactors with various
describe the temperature distributions in both microwave- diameters are displayed in Figure 11. The temperature
heating and electric-heating fixed-bed reactors. difference between the center and the wall of the reactor
3.5.9. Temperature Distributions in Fixed-Bed Reactors. increased from 19 °C at the diameter of 10 mm to 44 °C at 20
Figure 10 shows the radial temperature distributions of the mm. This meant that the lowest temperature at the center of
catalyst bed in the electric-heating fixed-bed reactor with a the 20 mm reactor was only 206 °C. Side reactions during the
diameter of 20 mm. The temperature difference between the glycerol dehydration process, especially glycerol polymer-
center and the wall of the reactor was up to around 50 °C for ization, would occur at such low temperatures, resulting in
various catalysts. This was mainly due to the low thermal the formation of coke deposits on the catalyst. In addition, as
conductivity of the binary metal oxide catalysts, which resulted shown in Figure 11d, the largest temperature difference for a
10732 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 13. Temperature distributions of the catalyst bed in microwave fixed-bed reactors with various diameters of (a) 10 mm, (b) 15 mm, and (c,
d) 20 mm. The temperature of the measurement point for (a−c) is 250 °C and that for (d) is 300 °C.

Figure 14. Electric field distributions within the catalyst bed in microwave fixed-bed reactors with various diameters of (a) 10 mm, (b) 15 mm, and
(c, d) 20 mm. The temperature of the measurement point for (a−c) is 250 °C and that for (d) is 300 °C.

20 mm reactor with the temperature of a measurement point at shown in Figure 13, no obvious change in the temperature
300 °C reached up to 57 °C. Overall, the simulation results distribution of the catalyst bed was observed for reactors with
confirmed the existence of a large temperature gradient in a increasing diameters from 10 to 20 mm. The largest
conventional electric-heating fixed-bed reactor, which was the temperature difference for the 20 mm reactor at the control
main reason for rapid catalyst deactivation. temperature of 300 °C was only 15 °C. Thus, uniform and
By comparison, a more uniform temperature distribution bulk heating of the catalyst bed achieved in the microwave
was achieved in the microwave fixed-bed reactor. As presented fixed-bed reactor owing to the unique penetration capability of
in Figure 12, a small temperature gradient within the whole microwave irradiation was verified by the simulation results.
catalyst bed can be observed for various catalysts, with the According to Figure 13c, the central temperature of the 20 mm
largest temperature difference being only 7−16 °C. The slight reactor with the temperature of a measurement point at 250
difference in temperature gradient for various catalysts was was 263 °C. The slightly higher temperature within the catalyst
attributed to their different dielectric properties including bed could reduce the occurrence of side reactions and hence
relative permittivity and relative permeability. Moreover, as improve the catalyst stability.
10733 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

To further understand the intensification mechanism of TEM and HAADF-STEM images and the corresponding
microwave heating, the electric field distributions within the elemental mapping of various binary metal oxide
reactors were examined in the simulation work. As displayed in catalysts including Nb−Zr, W−Al, Nb-Al, 10Sn−Zr,
Figure 14, for microwave fixed-bed reactors with different sizes, 30Sn−Zr, Sn−Ti, and Zr−Ti; pyridine-FTIR profiles at
the electric field distributions within the reactors were basically a desorption temperature of 350 °C for nine binary
uniform. This can be explained by the homogeneity of the metal oxide catalysts; typical 13C NMR spectrum of used
electromagnetic field within the penetration depth (Dp) of catalyst; geometric model of the microwave fixed-bed
microwave irradiation. According to eq 11, the Dp of 2450 reactor in the COMSOL graphic interface; comparison
MHz microwave into the microwave-absorbing binary metal between experimental and simulated temperature
oxide catalysts prepared in this work was calculated to be 2−40 profiles of the catalyst bed under microwave heating;
cm. The larger Dp than the reactor diameters led to the specific surface area and pore size of SiC and nine binary
uniform electric field distributions within the reactors and metal oxide catalysts; hydroxyacetone and acetaldehyde
hence the uniform temperature distributions of the catalyst selectivities in microwave and electric-heating fixed-bed
bed. It can be noted that for W−Zr, which is the most effective reactors packed with various binary metal oxide catalysts
catalyst in glycerol dehydration to acrolein in this study, the Dp at 250 and 300 °C; and glycerol dehydration in the
is 12.7 cm, which even reaches 34 cm if the industrial electric-heating fixed-bed reactor packed with various
magnetron at the microwave frequency of 915 MHz is used. binary metal oxide catalysts at 250 and 300 °C (PDF)
Therefore, the microwave fixed-bed reactor is a promising
alternative to the conventional reactors for glycerol dehy-
dration and other gas−solid catalytic reactions, especially as a
multitubular reactor for industrial scale-up. ■ AUTHOR INFORMATION
Corresponding Author
c0 Yong Nie − Biodiesel Laboratory of China Petroleum and
Dp = ÄÅ ÉÑ
ÅÅ 2 ÑÑ Chemical Industry Federation, Zhejiang Provincial Key
Å Ñ
2 2 ÅÅÅ 1 + () 1ÑÑÑ Laboratory of Biofuel, and College of Chemical Engineering,
ÅÅ ÑÑ
ÅÇ ÑÖ (11) Zhejiang University of Technology, Hangzhou 310014,
China; orcid.org/0000-0002-1671-7669;
where ε′ refers to the real part of the complex relative Email: ny_zjut@zjut.edu.cn/nieyongnieyyucheng@
permittivity. gmail.com

4. CONCLUSIONS Authors
Qinglong Xie − Biodiesel Laboratory of China Petroleum and
In this study, we developed a microwave fixed-bed reactor Chemical Industry Federation, Zhejiang Provincial Key
packed with nine microwave-absorbing binary metal oxide Laboratory of Biofuel, and College of Chemical Engineering,
catalysts with different physicochemical properties for gas- Zhejiang University of Technology, Hangzhou 310014,
phase glycerol dehydration to acrolein. The acrolein selectivity China
exhibited a positive correlation with the Brønsted-to-Lewis- Tongbo Pan − Biodiesel Laboratory of China Petroleum and
acid density ratio on the catalyst. More importantly, for all of Chemical Industry Federation, Zhejiang Provincial Key
the catalysts at dehydration temperatures of 250 and 300 °C, Laboratory of Biofuel, and College of Chemical Engineering,
better catalyst stability was observed in the microwave fixed- Zhejiang University of Technology, Hangzhou 310014,
bed reactor compared with the conventional electric-heating China
reactor. In addition, the catalyst stability in the electric-heating Gaoji Zheng − Biodiesel Laboratory of China Petroleum and
reactor was continuously reduced with increasing reactor size; Chemical Industry Federation, Zhejiang Provincial Key
yet this showed almost no difference in the microwave reactor. Laboratory of Biofuel, and College of Chemical Engineering,
The characterization of used catalysts confirmed that the rapid Zhejiang University of Technology, Hangzhou 310014,
catalyst deactivation in the electric-heating reactor was mainly China
attributed to coke formation caused by glycerol polymerization Yuqiang Zhou − Biodiesel Laboratory of China Petroleum and
at lower temperatures. The simulation of temperature Chemical Industry Federation, Zhejiang Provincial Key
distribution indicated the largest temperature difference of Laboratory of Biofuel, and College of Chemical Engineering,
57 °C within the 20 mm electric-heating reactor, which was Zhejiang University of Technology, Hangzhou 310014,
only around 10 °C for microwave reactors with various China
diameters. The more uniform temperature distribution in the Shangzhi Yu − Biodiesel Laboratory of China Petroleum and
microwave reactor resulted from the homogeneous electric Chemical Industry Federation, Zhejiang Provincial Key
field distribution within the penetration depth of microwave Laboratory of Biofuel, and College of Chemical Engineering,
irradiation, which contributed to better catalyst stability. Zhejiang University of Technology, Hangzhou 310014,
Overall, the microwave fixed-bed reactor demonstrates its China
effectiveness and versatility in gas−solid catalytic reactions, Ying Duan − Biodiesel Laboratory of China Petroleum and
with great potential in reactor scale-up. Chemical Industry Federation, Zhejiang Provincial Key

■ ASSOCIATED CONTENT
* Supporting Information

Laboratory of Biofuel, and College of Chemical Engineering,
Zhejiang University of Technology, Hangzhou 310014,
China
The Supporting Information is available free of charge at Complete contact information is available at:
https://pubs.acs.org/doi/10.1021/acs.iecr.2c01176. https://pubs.acs.org/10.1021/acs.iecr.2c01176
10734 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Notes (17) Jiang, X. C.; Zhou, C. H.; Tesser, R.; Di Serio, M.; Tong, D. S.;
The authors declare no competing financial interest. Zhang, J. R. Coking of catalysts in catalytic glycerol dehydration to
acrolein. Ind. Eng. Chem. Res. 2018, 57, 10736−10753.
■ ACKNOWLEDGMENTS
The authors express their great appreciation to the National
(18) Choi, Y.; Park, D. S.; Yun, H. J.; Baek, J.; Yun, D.; Yi, J.
Mesoporous siliconiobium phosphate as a pure brønsted acid catalyst
with excellent performance for the dehydration of glycerol to acrolein.
Natural Science Foundation of China (22178313, 21808203, ChemSusChem 2012, 5, 2460−2468.
and 21776261), the Zhejiang Provincial “Jianbing” “Lingyan” (19) Xie, Q.; Li, S.; Gong, R.; Zheng, G.; Wang, Y.; Xu, P.; Duan, Y.;
Research and Development Program of China (2022C01191), Yu, S.; Lu, M.; Ji, W.; Nie, Y.; Ji, J. Microwave-assisted catalytic
and the Natural Science Foundation of Zhejiang Province dehydration of glycerol for sustainable production of acrolein over a
(LQ17B060003) for their financial support for this work. microwave absorbing catalyst. Appl. Catal., B 2019, 243, 455−462.
(20) Tao, L.-Z.; Chai, S.-H.; Zuo, Y.; Zheng, W.-T.; Liang, Y.; Xu,
■ REFERENCES
(1) Karthik, G. M.; Buwa, V. V. Effect of particle shape on fluid flow
B.-Q. Sustainable production of acrolein: Acidic binary metal oxide
catalysts for gas-phase dehydration of glycerol. Catal. Today 2010,
158, 310−316.
and heat transfer for methane steam reforming reactions in a packed (21) Zhu, S.; Zhu, Y.; Hao, S.; Zheng, H.; Mo, T.; Li, Y. One-step
bed. AIChE J. 2017, 63, 366−377. hydrogenolysis of glycerol to biopropanols over Pt-H4SiW12O40/ZrO2
(2) Fratalocchi, L.; Visconti, C. G.; Groppi, G.; Lietti, L.; Tronconi, catalysts. Green Chem. 2012, 14, 2607−2616.
E. Intensifying heat transfer in Fischer-Tropsh tubular reactors (22) Barzetti, T.; Selli, E.; Moscotti, D.; Forni, L. Pyridine and
through the adoption of conductive packed foams. Chem. Eng. J. 2018, ammonia as probes for FTIR analysis of solid acid catalysts. J. Chem.
349, 829−837. Soc., Faraday Trans. 1996, 92, 1401−1407.
(3) Huang, X.; Li, H.; Li, H.; Xiao, W.-D. A computationally efficient (23) Emeis, C. A. Determination of integrated molar extinction
multi-scale simulation of a multi-stage fixed-bed reactor for methanol coefficients for infrared absorption bands of pyridine adsorbed on
to propylene reactions. Fuel Process. Technol. 2016, 150, 104−116. solid acid catalysts. J. Catal. 1993, 141, 347−354.
(4) Fu, J.; Shen, R.; Shu, Z.; Zhou, X.; Yuan, W. Numerical (24) Yadav, G. D.; Sharma, R. V.; Katole, S. O. Selective dehydration
reconstruction of the catalyst bed temperature distribution in a of glycerol to acrolein: Development of efficient and robust solid acid
multitubular fixed-bed reactor by Karhunen-Loeve expansion. Ind. catalyst MUICaT-5. Ind. Eng. Chem. Res. 2013, 52, 10133−10144.
Eng. Chem. Res. 2013, 52, 7818−7826. (25) Katryniok, B.; Paul, S.; Capron, M.; Lancelot, C.; Bellière-Baca,
(5) Tang, D.; Jess, A.; Ren, X.; Bluemich, B.; Stapf, S. Axial V.; Rey, P.; Dumeignil, F. A long-life catalyst for glycerol dehydration
dispersion and wall effects in narrow fixed bed reactors: A to acrolein. Green Chem. 2010, 12, 1922−1925.
comparative study based on RTD and NMR measurements. Chem. (26) Thanasilp, S.; Schwank, J. W.; Meeyoo, V.; Pengpanich, S.;
Eng. Technol. 2004, 27, 866−873. Hunsom, M. Preparation of supported POM catalysts for liquid phase
(6) Pashchenko, D.; Karpilov, I.; Mustafin, R. Numerical calculation oxydehydration of glycerol to acrylic acid. J. Mol. Catal. A: Chem.
with experimental validation of pressure drop in a fixed-bed reactor 2013, 380, 49−56.
filled with the porous elements. AIChE J. 2020, 66, No. e16937. (27) Costa, B. O. D.; Legnoverde, M. S.; Lago, C.; Decolatti, H. P.;
(7) Goyal, H.; Chen, T. Y.; Chen, W. Q.; Vlachos, D. G. A review of Querini, C. A. Sulfonic functionalized SBA-15 catalysts in the gas
microwave-assisted process intensified multiphase reactors. Chem. phase glycerol dehydration. Thermal stability and catalyst deactiva-
Eng. J. 2022, 430, No. 133183. tion. Microporous Mesoporous Mater. 2016, 230, 66−75.
(8) Guler, M.; Dogu, T.; Varisil, D. Hydrogen production over (28) Vieira, L. H.; Carvalho, K. T. G.; Urquieta-González, E. A.;
molybdenum loaded mesoporous carbon catalysts in microwave Pulcinelli, S. H.; Santilli, C. V.; Martins, L. Effects of crystal size,
heated reactor system. Appl. Catal., B 2017, 219, 173−182. acidity, and synthesis procedure on the catalytic performance of
(9) Stankiewicz, A. Energy matters − Alternative sources and forms gallium and aluminum MFI zeolites in glycerol dehydration. J. Mol.
of energy for intensification of chemical and biochemical processes. Catal. A: Chem. 2016, 422, 148−157.
Chem. Eng. Res. Des. 2006, 84, 511−521. (29) Rodrigues, M. V.; Vignatti, C.; Garetto, T.; Pulcinelli, S. H.;
(10) Arpia, A. A.; Chen, W. H.; Lam, S. S.; Rousset, P.; de Luna, M. Santilli, C. V.; Martins, L. Glycerol dehydration catalyzed by MWW
D. G. Sustainable biofuel and bioenergy production from biomass zeolites and the changes in the catalyst deactivation caused by
waste residues using microwave-assisted heating: A comprehensive porosity modification. Appl. Catal., A 2015, 495, 84−91.
review. Chem. Eng. J. 2021, 403, No. 126233. (30) Park, H.; Yun, Y. S.; Kim, T. Y.; Lee, K. R.; Beak, J.; Yi, J.
(11) Zhang, Y.; Cui, L.; Liu, S.; Fan, L.; Zhou, N.; Peng, P.; Wang, Kinetics of the dehydration of glycerol over acid catalysts with an
Y.; Guo, F.; Min, M.; Cheng, Y.; Liu, Y.; Lei, H.; Chen, P.; Li, B.; investigation of deactivation mechanism by coke. Appl. Catal., B 2015,
Ruan, R. Fast microwave-assisted pyrolysis of wastes for biofuels 176-177, 1−10.
production − A review. Bioresour. Technol. 2020, 297, No. 122480. (31) Zhao, Z.; Shen, X.; Li, H.; Liu, K.; Wu, H.; Li, X.; Gao, X.
(12) Nie, Y.; Duan, Y.; Gong, R.; Yu, S.; Lu, M.; Yu, F.; Ji, J. Watching microwave-induced microscopic hot spots via the
Microwave-assisted pyrolysis of methyl ricinoleate for continuous thermosensitive fluorescence of europium/terbium mixed-metal
production of undecylenic acid methyl ester (UAME). Bioresour. organic complexes. Angew. Chem. 2022, 134, No. e202114340.
Technol. 2015, 186, 334−337. (32) Liu, H.; Tian, H.; Cheng, H. Dielectric properties of the SiC
(13) Yu, S.; Duan, Y.; Mao, X.; Xie, Q.; Zeng, G.; Lu, M.; Nie, Y.; Ji, fiber-reinforced SiC matrix composites in the temperature range from
J. Pyrolysis of methyl ricinoleate by microwave-assisted heating 25 to 700 °C at frequencies between 8.2 and 18 GHz. J. Nucl. Mater.
coupled with atomization feeding. J. Anal. Appl. Pyrolysis 2018, 135, 2013, 432, 57−60.
176−183.
(14) Yan, W.; Suppes, G. J. Low-pressure packed-bed gas-phase
dehydration of glycerol to acrolein. Ind. Eng. Chem. Res. 2009, 48,
3279−3283.
(15) Ding, J.; Wang, L.; Zhang, Z.; Zhao, S.; Zhao, J.; Lu, Y.; Huang,
J. Microstructured ZSM-11 catalyst on stainless steel microfibers for
improving glycerol dehydration to acrolein. ACS Sustainable Chem.
Eng. 2019, 7, 16225−16232.
(16) Chai, S.-H.; Wang, H.-P.; Liang, Y.; Xu, B.-Q. Sustainable
production of acrolein: Investigation of solid acid−base catalysts for
gas-phase dehydration of glycerol. Green Chem. 2007, 9, 1130−1136.

10735 https://doi.org/10.1021/acs.iecr.2c01176
Ind. Eng. Chem. Res. 2022, 61, 10723−10735

You might also like