You are on page 1of 8

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: ACS Omega 2019, 4, 8274−8281 http://pubs.acs.org/journal/acsodf

Kinetic Study on the Preparation of Fumaric Acid from Maleic Acid


by Batch Noncatalytic Isomerization
Wangmi Chen,† Xiaoting Chen,‡ and Shouzhi Yi*,‡

College of Marine and Environmental Sciences and ‡College of Chemical Engineering and Materials Science, Tianjin University of
Science and Technology, Tianjin 300457, China
*
S Supporting Information

ABSTRACT: In recent years, the increase in demand for


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

fumaric acid from industry has resulted in an increased need


for a high-selectivity process for the conversion of maleic acid
to fumaric acid. A highly selective conversion of fumaric acid
was achieved without a catalyst by a simple one-step
hydrothermal reaction. In addition, the competitive con-
Downloaded via 179.48.51.56 on June 1, 2023 at 21:43:38 (UTC).

version of maleic acid, fumaric acid, and malic acid was first
systematically investigated in detail without using a catalyst.
The products were characterized by X-ray diffraction and
Fourier transform infrared, which demonstrated that the
product was fumaric acid. The highly selective conversion of
fumaric acid was achieved, and the yield of fumaric acid could
reach 92%. Furthermore, a reaction kinetic model was put forward to study the competitive transformation process. The kinetic
model predictions were found to agree well with the experimental data. The kinetic parameters were used to explain the changes
in the content of every substance at different reaction temperatures and reaction times. In addition, the initial maleic acid
concentration in the reaction was also considered as an influencing factor. These results can facilitate the conditional control
and product control of industrial processes for the production of fumaric acid or malic acid using latter without a catalyst.

■ INTRODUCTION
Fumaric acid, a tetra-carbon unsaturated dicarboxylic acid
new biotechnologies, such as alkali pretreatment,33 microwave
technology,34 immobilization technology,26,35 and biorefinery
classified as an organic acid, is widely distributed in nature. In technology,36 have been developed to improve the yield of
2004, fumaric acid was listed as the most promising building fumaric acid. Nevertheless, biosynthesis methods have suffered
block derived from biomass by Werpy and Petersen due to its from the inherent disadvantages of low yield and efficiency and
excellent performance.1,2 Fumaric acid has traditionally been a long cycle in the process of production. Thus, researchers
used in the production of food acidity regulators,1 feed have tried to develop isomerization methods to improve the
preservatives,3−5 and other high-potency chemical products.6−8 synthesis efficiency. Several noteworthy catalysts have been
Recently, fumaric acid has been applied in metal−organic reported to enhance the reaction conditions and increase the
frameworks,9,10 polymer composites,11,12 composite anodes of productivity of fumaric acid, including bromate ion,37,38
high-energy lithium batteries,13 co-crystals,14,15 and micro- vanadium pentoxide,39 hydrochloric acid,40 and poly(4-vinyl
environment tablets.16 Thus, the demand for fumaric acid has pyridine).41 In a further study, researchers have introduced
dramatically increased with the expansion of the application microwaves to reduce the reaction time of maleic acid and
scope for fumaric acid, which has made it necessary to study malic acid for assisted synthesis.42 By using the chemical
effective fumaric acid production methods. isomerization method, the yield of fumaric acid can reach over
Currently, methods for preparing fumaric acid mainly 80%, which is far higher than that realized by using a
include chemical isomerization and biosynthesis.17,18 Chemical biosynthesis method.
synthesis methods for fumaric acid were investigated in the To further study the chemical isomerization method of
1940s and laid the foundation for the industrial production of fumaric acid synthesis, more researchers have used kinetic
fumaric acid. 19 With the continuous development of methods to explore the mutual transformation between
biotechnology and the rapid growth of fossil (petroleum) fumaric acid and other acids under certain conditions. For
based chemical production costs, investigations on the example, Wojcieszak et al.43 and Delhomme et al.44 discussed
preparation of fumaric acid by biosynthetic methods have the feasibility of and conditions for the mutual conversion of
become more popular. Rhizopus,20−23 Torulopsis glabrata,24
Filamentous fungal strains,25−27 Saccharomyces cerevisiae,28,29 Received: February 2, 2019
Escherichia coli,30,31 and other significant microorganisms32 Accepted: April 30, 2019
have been applied to produce fumaric acid. In addition, some Published: May 8, 2019

© 2019 American Chemical Society 8274 DOI: 10.1021/acsomega.9b00316


ACS Omega 2019, 4, 8274−8281
ACS Omega


Article

maleic anhydride, maleic acid, fumaric acid, and malic acid. EXPERIMENTAL SECTION
Wang et al.40 successfully converted maleic acid and fumaric Materials. All chemicals were guaranteed reagents and used
acid into malic acid in an acidic aqueous solution at 120 °C. without further purification. The raw material (maleic acid)
Ortiz et al.42 studied the kinetics of the transformation from was an analytical reagent purchased from China Commeo
fumaric acid to maleic acid and malic acid at four different Chemical Co., Ltd. Distilled water was used in the isomer-
temperatures in a homogeneous nonisothermal batch reactor ization and purification stages. For the analysis and testing
by using hydrochloric acid as a catalyst. The reason for and phases, fumaric acid (analytical reagent), maleic acid
processes of the reaction could be explained by the reaction (analytical reagent), malic acid (biological grade), and
kinetics, so it was very meaningful to introduce kinetic studies hydrochloric acid (analytical reagent) were purchased from
into the related reaction processes. Our previous work has China Commeo Chemical Co., Ltd.
demonstrated that the isomerization of maleic acid can Preparation of Samples. The experiments were per-
produce fumaric acid in the absence of a catalyst. 45 formed in a 50 mL Teflon-lined autoclave. Maleic acid and
Nevertheless, no detailed kinetic studies have been reported distilled water were added to the inner liner and stirred to
in the available literature about the mutual conversion of malic achieve a uniform solution. Then, the Teflon-lined autoclave
acid, maleic acid, and fumaric acid, especially the isomerization was sealed and placed in a blast drying oven. Once the
of maleic acid to fumaric acid, although such a kinetic study is temperature of the mixture reached the preset temperature, the
significant, regardless of the use of a catalyst. timing was started. After heating for a preset time, the products
Researchers have proposed many kinetic models that help to were cooled to room temperature and removed. The
study the conversion process between maleic acid, fumaric experiments were carried out at 190, 200, 210, and 220 °C
acid, and malic acid. Rozelle et al.46 studied the hydration for 0−4 h. The initial reaction concentration of maleic acid
kinetic and thermodynamic models of fumaric acid in a (mass fraction 35−100%) on the selectivity of fumaric acid was
concentrated hydrochloric acid solution at 125−200 °C, but also examined. The reaction experiments were repeated twice.
they only explored the mutual conversion process between After adding a certain amount of distilled water, the reaction
fumaric acid and malic acid under the condition of hydro- mixture was stirred and washed for 1 h. Further, insoluble
chloric acid catalysis. In addition, Jwo et al.47 investigated the matter and the washing solution were separated by suction
kinetics of the conversion of maleic acid to fumaric acid under filtration, and the filter cake was dried for 2 h by a blast drying
the action of bromine and strontium ions, while the model oven at 100 °C to obtain the filtered product. The washing
mainly focused on the one-way conversion of maleic acid to solution was prepared for HPLC.
fumaric acid with a catalyst. Furthermore, Ortiz et al.48,49 Characterization. Fourier transform infrared (FT-IR)
converted a low concentration of fumaric acid to maleic acid spectra were measured on a TENSOR27 Fourier infrared
and malic acid under the action of hydrochloric acid or no spectrometer (Brook Spectrometer, Germany). The dried
catalysts. This study only carried out the conversion reaction of samples were characterized by the potassium bromide tableting
fumaric acid rather than that of maleic acid. However, kinetic method, and the spectra were recorded between 400 and 4000
cm−1.
studies on the production of fumaric acid and malic acid using
The X-ray diffraction (XRD) patterns of the samples were
maleic acid, especially in the absence of a catalyst, have not
detected by an XRD-6100 (Shimadzu Corporation, Japan),
been conducted. In addition, previous theoretical studies are
with graphite monochromated Cu Kα (λ = 1.54178 Å)
not fully applicable to this process.
radiation operating at 40 kV and 30 mA in the angular range of
In this study, the selectivity of fumaric acid in the mutual
2θ = 5−80° at room temperature. The scanning speed and the
conversion of fumaric acid, maleic acid, and malic acid was scanning step size were 7° min−1 and 0.02°, respectively. The
enhanced by conditional control in the absence of a catalyst, sample needed to be dried, ground, and placed in a fixed mold
and a more comprehensive model of the reaction kinetics was for compaction before testing.
established to explain the reaction process for the first time. In Quantitative Analysis of Samples. All samples were
previous studies on the mutual conversion of fumaric acid, analyzed by HPLC (Agilent Technologies 1260, Inc.). The
maleic acid, and malic acid, the strong selectivity of specific column used was a Kromasil C18 column (5 μm × 250 mm ×
acids, especially fumaric acid, has not been studied on the basis 4.6 mm, Japan). The mobile phase was an HCl solution in
of strengthening the investigation of the experimental ultrapure water (pH 2) at a flow rate of 1 mL min−1 and an
influencing factors and determining better reaction conditions. injection volume of 5 μL. The sample detection wavelength
In addition, the reaction mechanism of the selective was 210 nm, and the column temperature was 42 ± 0.8 °C.
production of fumaric acid and the mechanism of competition The water samples were diluted 100 times before testing. The
between various substances were also analyzed. It should be dilute samples were analyzed by HPLC to determine the
noted that to obtain a higher yield of fumaric acid in a short content of each component, providing the necessary
time, the mutual conversion between fumaric acid, maleic acid, parameters for the kinetic analyses.
and malic acid was fully considered. In addition, the content of The conversion rates of fumaric acid, malic acid, and maleic
each substance in the reaction was detected by high- acid were calculated by the following equations
performance liquid chromatography (HPLC), and the relevant
kinetic parameters were calculated. The kinetic parameters CF n
fumaric acid yield(%) = × 100 = F × 100
explained the reasons for the change in the selectivity of each CMx0 nMx0 (1)
substance in the reaction as a result of various influencing
factors and provided guidance for the subsequent control of CM n
conditions for the production of fumaric acid and malic acid malic acid yield(%) = × 100 = M × 100
without a catalyst. CMx0 nMx0 (2)

8275 DOI: 10.1021/acsomega.9b00316


ACS Omega 2019, 4, 8274−8281
ACS Omega Article

maleicacidconversionrate(%) acid has a higher solubility in water (100 °C, solubility of


392.6/100 g water, 33.84 mol/L) and a higher solubility at
C − CMx
= Mx0 × 100 high temperatures, much higher than the concentration used in
CMx0 the kinetic experiments.50 At the same time, Ortiz et al.48
n − nMx confirmed that the solubility of fumaric acid increased
= Mx0 × 100
nMx0 (3) exponentially with increasing temperature (4.84 mol/L at
189 °C). Therefore, it can be considered that the detected
−1
where CMx0 (mol L ) is the initial concentration of maleic acid reaction is carried out under homogeneous conditions.
and CF, CM, and CMx (mol L−1) are the concentrations of The effect of temperature on the reaction rate constant was
maleic acid and fumaric acid at time t in the respective calculated by the Arrhenius equation, and the activation energy
solutions. nMx0 (mol) is the initial molar mass of maleic acid for the reaction was determined
and nF, nM, and nMx (mol) are the molar masses of maleic acid
and fumaric acid. k = Ae−Ea / RT (8)
Reaction Kinetic Model. The reaction model in Figure 1 −1
where T (K) is the absolute temperature, k (h ) is the
was established to study the kinetics of this series of reaction rate constant at the preset temperature, Ea (kJ mol−1)
is the activation energy, R (8.314 J (mol K)−1) is the general
gas constant, and A (h−1) is a pre-exponential factor. Ea was
estimated based on linear regression analysis of a plot of ln(ki)
and 1/Ti (i = 190, 200, 210, and 220 °C).
The Gibbs free energy of every reaction was calculated by
using the following equation
ki
Kj =
k −i (9)

ΔG = −RT ln(Kj) (10)


−1
Figure 1. Simplified reaction model for the isomerization and where ΔG (J mol ) is the Gibbs free energy at the preset
hydration of maleic acid, fumaric acid, and malic acid. temperature, Kj (j = 1, 2, 3) is the reaction equilibrium
constant, and ki (i = 1, 2, 3) is the reaction rate constant (h−1)
isomerization reactions and hydration reactions, which can for each reaction.
explain the mutual transformation between fumaric acid, malic
acid, and maleic acid. In this context, Mx, F, and M represent
maleic acid, fumaric acid, and malic acid, respectively.
■ RESULTS AND DISCUSSION
Characterization of Products. The products were
According to the model with the assumption of a first-order identified from the measured XRD pattern (Figure 2a). The
reaction, the following set of eqs 4−6 was obtained.
dCMx
rMx = = −(k1 + k 2)CMx + k −1C F
dt (4)

dC F
rF = = k1CMx − (k −1 + k −3)C F + k 3CM
dt (5)

dC M
rM = = k 2CMx + k −3C F − k 3CM
dt (6)
The following two boundary conditions could be determined

l
by the analyses of the experimental process.
o
Figure 2. (a) XRD patterns of the products, standard samples, and

o
o
o
other research products.51 (b) FT-IR spectra of the products,

oCMx = CMx0
o
t=0
o
o
standard samples, and other research products.

m
o
o
o
o
o
o
o
diffraction pattern showed six major diffraction peaks at 2θ =
o
CF = 0

nC M = 0
18.4, 22.8, 28.7, 29.4, 31.2, and 35.8°. Compared with the
peaks in the pattern of fumaric acid standard and fumaric acid
CMx0 = CMx + C F + CM reported in the references,51 the peaks in the XRD pattern of
(7)
the products were consistent. This could indicate that the
It is worth noting that the reaction involving malic acid was a products had the same crystal structure as fumaric acid and
pseudo-first-order reaction because water was present in large could be preliminarily identified as fumaric acid. Since there
amounts during the reaction. Although water was involved in were no other peaks present in the pattern, the purity of the
some reactions, the amount of water in the system was products was relatively high.
considered to be constant, and the progress of every reaction FT-IR analysis was further carried out to determine the
was only related to the concentrations of fumaric acid, maleic structure of the product. As shown in Figure 2b, the
acid, and malic acid. characteristic peak at 3081 cm−1 can be assigned to the
The kinetic experiments were carried out using a maleic acid stretching vibration of CC−H, while the peaks at 2863 and
solution having a mass fraction of 60% (12.94 mol/L). Maleic 1670 cm−1 can be ascribed to the stretching vibrations of
8276 DOI: 10.1021/acsomega.9b00316
ACS Omega 2019, 4, 8274−8281
ACS Omega Article

Table 1. Reaction Rate Constants for Every Reaction Process at Different Temperatures
reaction rate constant (h−1)
T (K) k1 k−1 k2 k3 k−3
463 2.502862914 0.060560596 0.561579371 0.257186356 0.176258077
473 3.550348358 0.099529637 0.983973071 0.338702133 0.199889328
483 5.42534805 0.165669838 1.504939419 0.409023597 0.234727444
493 8.713134203 0.312128031 2.578914071 0.518292295 0.278754867

Table 2. Tuned Model Parameters


rate constant linear equation of ln(ki) and 1/T R2 Eai (kJ mol−1) ΔHj (×107 J kmol−1)
k1 ln(k1) = −9495.4 × (1/T) + 21.386 0.9927 78.94475 −2.39493
k−1 ln(k−1) = −12376 × (1/T) + 23.886 0.9940 102.89406
k2 ln(k2) = −11408 × (1/T) + 24.069 0.9975 94.84611
k3 ln(k3) = −5231.5 × (1/T) + 9.9526 0.9962 43.49469 1.43932
k−3 ln(k−3) = −3500.3 × (1/T) + 5.8087 0.9924 29.10149

saturated C−H and CO, respectively. The peak at 1424 [λ1 , λ 2]


cm−1 was caused by the in-plane bending vibration of the
saturated C−H. The three peaks at approximately 1276 cm−1 − (k ′ + k ″ ) ± (k′ + k″)2 − 4(k′k″ − k −1k‴)
resulted from the stretching vibration of C−O. The peaks at =
2
922 and 643 cm−1 were attributed to the out-of-plane bending (19)
vibration of O−H and the deformation vibration of OC−O,
where ki (i = 1, 2, 3, −1, −3) (h−1) was the reaction rate
respectively. Further, the single peak at 1008 cm−1 was
constant for each reaction stage.
attributed to the trans-substituted characteristic peak. The FT-
The content of each substance was determined by HPLC,
IR analysis further indicated that the product was fumaric acid,
and these concentration equations were used to determine the
which was in accordance with the XRD result.
rate constant under each condition. The Solver function in
Kinetic Study. To determine the rate-limiting step of the
Microsoft Excel was used to find the best values for the
isomerization reaction, the following concentration equations
reaction rate constants. Using the Arrhenius form for the rate
were obtained by analyzing the above differential equation
equations, the activation energy (Ea) and pre-exponential
system (4−6)
factor (A) were calculated for each individual reaction
CMx k 3k −1 temperature using the slope and intercept of ln(ki) (i = 1, 2,
= C1 e λ1t + C2 e λ2t + 3, −1, −3) versus 1/T charts. The specific calculation results
CMx0 k′k″ − k −1k‴ (11)
are shown in Tables 1, 2, and 4.
CF C C As shown in Table 2, R2 was greater than 0.99, which
= 1 (k′ + λ1)e λ1t + 2 (k′ + λ 2)e λ2t suggested that ln(ki) had a good linear relationship with 1/T,
CMx0 k −1 k −1
and the obtained reaction rate constants satisfied the Arrhenius
k 3k′ expression. Compared with the values reported in the
+
k′k″ − k −1k‴ (12) literature, the results from the literature and those currently
adjusted from the experimental kinetic data reported in Tables
CM C C 1, 2, 3, 4, 5, and 6 were almost the same; however, the value of
= − 1 (k′ + k −1 + λ1)e λ1t − 2 (k′ + k −1 + λ 2)
CMx0 k −1 k −1
Table 3. Comparison of the Enthalpy of Reaction (1)
k k + k k ′
e λ 2 t + 2 −1 −3 Calculated in This Study with the Median Value from a
k′k″ − k −1k‴ (13) Previous Study
where data sources ΔH1 (×107 J kmol−1) reference
this research −2.39493

C1 =
λ2 ( k 3k −1
k ′ k ″ − k −1k ‴ )
− 1 − k′ previous study
literature report
−2.83
−2.28
48
52
λ1 − λ 2 (14)

Table 4. Pre-Exponential Factor and Reaction Equilibrium


C2 =
λ1 ( k 3k −1
k ′ k ″ − k −1k ‴ )
− 1 − k′
Constant
λ 2 − λ1 (15)
Kj
k ′ = k1 + k 2 −1
(16) rate constant A (s ) 463 K 473 K 483 K 493 K
k1 5.39 × 105 41.33 35.67 32.75 27.92
k ″ = k −1 + k 3 + k −3 (17) k−1 6.56 × 106
k2 7.88 × 106
k ‴ = k1 − k 3 (18) k3 5.84 1.46 1.69 1.74 1.86
and k−3 9.26 × 10−2

8277 DOI: 10.1021/acsomega.9b00316


ACS Omega 2019, 4, 8274−8281
ACS Omega Article

Table 5. Gibbs Free Energy of Reactions (1) and (3) slightly with temperature, but they did not reach the elevated
level of reaction (1). It was indicated that the high selective
ΔG1 of reaction (1) ΔG3 of reaction (3)
T (K) (×107 J kmol−1) (×107 J kmol−1) conversion of fumaric acid could be well achieved by adjusting
the temperature.
463 −1.432565246 −0.145449544
As indicated in Table 5, with increasing temperature, the
473 −1.40561925 −0.207384381
absolute value of the Gibbs free energy in reaction (1)
483 −1.401000403 −0.223008914
decreased, while that in reaction (3) showed an increasing
493 −1.364562253 −0.254210396
trend. It was suggested that the spontaneous inhibition of the
conversion from maleic acid to fumaric acid was weakened,
Table 6. Selectivities of Fumaric Acid as Defined by eq 20 at
while the spontaneous inhibition of the conversion from malic
Different Reaction Temperatures
acid to fumaric acid increased with increasing temperature.
T (K) tmax (h) S This meant that the conversion from maleic acid to fumaric
463 0.818 6.58 acid and the conversion from fumaric acid to malic acid were
473 0.582 12.57 inhibited by the increase in temperature. This echoed the
483 0.473 14.39 conclusion of Ortiz et al. that a temperature rise reduces the
493 0.314 23.40 selectivity for malic acid.48
Discussion on Product Selectivity. For practical
ΔH1 was better (Table 3) than that reported in other studies.48 purposes, the selectivity of the conversion of maleic acid to
The difference between the reaction rate constants and other fumaric acid and malic acid remains to be studied. The reason
parameters could be explained as being due to the presence of for this is that the rapid conversion of maleic acid to fumaric
a catalyst, different temperatures, and different substrate acid with high selectivity is conducive to the development of a
concentrations. production process for fumaric acid. It is particularly
The calculated data were used to plot the changes in the noteworthy that when the temperature increases from 463 to
three organic acid concentrations over time at each temper- 493 K, the maximum selectivity time for fumaric acid
ature, and the curves were combined with experimental data to decreases. At the same time, the selectivity for malic acid
form Figure 3a−d. It was found that the figures drawn by the decreased first and then increased with reaction time.
Selectivity is determined by using the following equation.
CF
S=
CMx (20)
Analysis of Influencing Factors. Effect of Initial
Reactant Concentration. To evaluate the effect of the initial
concentration of maleic acid on the reaction, the intersub-
stance conversion was investigated. As shown in Figure 4, the

Figure 3. Experimental (symbol) and calculated (solid line) variations


in the concentration of three organic acids with reaction time at Figure 4. Concentrations of maleic acid, malic acid, and fumaric acid
temperatures (a) 190 °C, (b) 200 °C, (c) 210 °C, and (d) 220 °C. vary with the initial maleic acid concentration in the reaction (1 h).
Black for maleic acid; red for fumaric acid; and blue for malic acid.
amount of fumaric acid increased with increasing initial maleic
equations were better matched with the experimental data acid concentration. At the same time, the conversion rate of
points. The trends for the three organic acids were consistent maleic acid also increased with increasing fumaric acid
with that of the graph line, which further proved the accuracy concentration. In addition, the yield of malic acid first
of the experimental data. increased and then tended to balance out. Because the
As shown in Table 1, as the temperature increased, the reduction in water content was not favorable for the
forward reaction rate constant of reaction (1) (the reaction conversion of maleic acid and fumaric acid to malic acid,
rate constant of conversion of maleic acid to fumaric acid) more maleic acid could be converted to fumaric acid, and the
increased continuously from 2.5 to 8.71 h−1. In addition, at selectivity of fumaric acid was enhanced. Furthermore, once
each temperature, the forward reaction rate constant of the temperature was too high, the products were oxidized,
reaction (1) was the largest, and the reverse reaction rate thereby leading to the possibility of coking products being
constant was relatively less than 0.32 h−1. This was because cis- formed in the reaction vessel.
maleic acid was easily attacked on account of its small steric Effect of Reaction Temperature and Time. The effects of
effect and was transformed into malic acid and fumaric acid.40 reaction temperature and reaction time on the mutual
The reaction rate constants of reactions (2) and (3) increased transformation of materials were investigated. A maleic acid
8278 DOI: 10.1021/acsomega.9b00316
ACS Omega 2019, 4, 8274−8281
ACS Omega Article

solution with an initial mass fraction of 60% was selected and (3) to produce fumaric acid was enhanced with increasing
was subject to an experiment from 0 to 4 h at 190, 200, 210, temperature, the reaction rate constant for conversion to
and 220 °C. As shown in Figure 5a,b, the rate of fumaric acid obtain fumaric acid was small, and the reaction was not
affected much.

■ CONCLUSIONS
Highly selective conversion from maleic acid to fumaric acid
was achieved by a simple one-step hydrothermal synthesis at
four different temperatures without any catalyst, and the
kinetics of the mutual conversion of maleic acid, fumaric acid,
and malic acid were studied. The highly selective process had
higher stability and fumaric acid yield and could be used to
obtain the desired fumaric acid product without a catalyst.
Figure 5. Variation in fumaric acid yield with reaction temperature Through the mutual confirmation of the data spectra and the
and reaction time. (a) Fumaric acid yield changes with temperature. kinetic parameters, the mutual transformation of the three
(b) Fumaric acid yield changes with time. organic acids under hydrothermal conditions was studied
systematically and completely for the first time and reached a
level superior to that obtained in previous reports. Based on
production was faster when the reaction temperature was the kinetic data obtained experimentally, 20 reaction rate
increased. This was because the reaction rate constant for the constants were determined, and the Ea values of 5 reactions
formation of fumaric acid became larger due to the increase in were calculated. It was found that the time to reach the highest
temperature. At the same time, the yield of fumaric acid yield point of fumaric acid could be effectively reduced with
increased with increasing time in the initial stage, while it increasing temperature and that a high concentration of
decreased when the reaction time was over 1 h. This was fumaric acid in the conversion process can convert fumaric
because the concentration of fumaric acid increased as the acid into malic acid, which provided the guidance on the
reactions proceeded, thereby promoting the conversion of development of related production processes. The change in
fumaric acid to maleic acid and malic acid. the Gibbs free energy in each reaction reflected the fact that an
Further analysis of Figure 5a showed that the amount of increase in temperature was not favorable for the formation of
fumaric acid increased significantly as the temperature fumaric acid from maleic acid, but instead promoted the
increased within 0.75 h. However, as the reaction time formation of fumaric acid from malic acid and reduced the
continued to increase, the effect of temperature on the reaction selectivity of malic acid. This study showed that the highly
became less pronounced. When the reaction time was over 1 h, selective conversion of fumaric acid can be achieved by a
the conversion of fumaric acid decreased with increasing hydrothermal method and that the conversion of malic acid
temperature. This was because the content of fumaric acid can also be achieved by conditional control without a catalyst.
increased with the reaction time, while the content of maleic Therefore, the hydrothermal synthesis of fumaric acid and
acid continued to decrease, which caused the main reactions to malic acid without a catalyst has huge commercial potential.


become the reactions of conversion of fumaric acid to maleic
acid and malic acid. As shown in Figure 5b, the maximum yield
of fumaric acid decreased slightly (approximately 1%) and the ASSOCIATED CONTENT
time of the maximum conversion rate decreased with *
S Supporting Information
increasing temperature. This could be explained by the fact The Supporting Information is available free of charge on the
that the increase in temperature enhanced the reaction rate ACS Publications website at DOI: 10.1021/acsome-
constant for maleic acid, which reduced the reaction time. A ga.9b00316.
slight decrease was because, although the formation of fumaric
acid was inhibited by the increase in temperature, the Typical appearance of the products; SEM image of the
conversion from malic acid to fumaric acid was promoted. products; 1H-NMR spectrum; analytical method of
Furthermore, on the falling phase, the higher the temperature, solution of the system of ordinary differential equations
the faster the fumaric acid decreased. This was because the given by eqs 4−6 (PDF)
reaction rate constants of reactions (3) and (1) increased as
the temperature increased, which caused fumaric acid to be
converted to maleic acid and malic acid more rapidly at a
higher concentration of fumaric acid. The fumaric acid content
■ AUTHOR INFORMATION
Corresponding Author
at which the final fumaric acid concentration reached an
*E-mail: 389085304@qq.com (S.Y.).
equilibrium decreased with increasing temperature, which was
illustrated by the change in the Gibbs free energy and the ORCID
reaction rate constants with increasing temperature. As the Wangmi Chen: 0000-0003-4348-4765
temperature increased, the absolute value of the Gibbs free Author Contributions
energy of reaction (1) decreased, which indicated that the
The manuscript was written through contributions of all
selectivity of the reaction to form fumaric acid decreased,
authors. All authors have given approval to the final version of
causing more fumaric acid to be converted to maleic acid. At
the manuscript.
the same time, since the rate constant of reaction (2) increased
significantly with increasing temperature, more maleic acid was Notes
converted to malic acid. Although the selectivity of reaction The authors declare no competing financial interest.
8279 DOI: 10.1021/acsomega.9b00316
ACS Omega 2019, 4, 8274−8281
ACS Omega


Article

ACKNOWLEDGMENTS (18) Roa Engel, C. A.; Straathof, A. J.; Zijlmans, T. W.; van Gulik,
W. M.; van der Wielen, L. A. Fumaric acid production by
The authors wish to acknowledge laboratory teachers and fermentation. Appl. Microbiol. Biotechnol. 2008, 78, 379−389.
classmates from the Tianjin University of Science and (19) Foster, J. W.; Carson, S. F.; Anthony, D. S.; Davis, J. B.;
Technology for technology support of this study. Jefferson, W. E.; Long, M. V. Aerobic Formation of Fumaric Acid in


the Mold Rhizopus nigricans: Synthesis by Direct C2Condensation.
REFERENCES Proc. Natl. Acad. Sci. U.S.A. 1949, 35, 663−672.
(20) Deng, F.; Aita, G. M. Fumaric Acid Production by Rhizopus
(1) Ding, Y.; Li, S.; Dou, C.; Yu, Y.; Huang, H. Production of oryzae ATCC 20344 from Lignocellulosic Syrup. BioEnergy Res. 2018,
fumaric acid by Rhizopus oryzae: role of carbon-nitrogen ratio. Appl. 11, 330−340.
Biochem. Biotechnol. 2011, 164, 1461−1467. (21) Gu, S.; Xu, Q.; Huang, H.; Li, S. Alternative respiration and
(2) Werpy, T.; Petersen, G. Top Value Added Chemicals From fumaric acid production of Rhizopus oryzae. Appl. Microbiol.
Biomass: I. Results of Screening for Potential Candidates from Sugars and Biotechnol. 2014, 98, 5145−5152.
Synthesis Gas; U.S. Department of Energy, 2004. (22) Moresi, M.; Parente, E.; Petruccioli, M.; Federici, F. Fumaric
(3) Pérez-Díaz, I. M.; McFeeters, R. F. Preservation of acidified acid production from hydrolysates of starch-based substrates. J. Chem.
cucumbers with a natural preservative combination of fumaric acid Technol. Biotechnol. 2007, 54, 283−290.
and allyl isothiocyanate that target lactic acid bacteria and yeasts. J. (23) Rodríguez-López, J.; Sánchez, A. J.; Gómez, D. M.; Romaní, A.;
Food Sci. 2010, 75, M204−M208. Parajó, J. C. Fermentative production of fumaric acid from Eucalyptus
(4) Gabriele, A.; Alex, B.; Vasileios, B.; et al. Scientific Opinion on globulus wood hydrolyzates. J. Chem. Technol. Biotechnol. 2012, 87,
the safety and efficacy of fumaric acid as a feed additive for all animal 1036−1040.
species. EFSA J. 2013, 11, 3102. (24) Chen, X.; Wu, J.; Song, W.; Zhang, L.; Wang, H.; Liu, L.
(5) Guido, R.; Gabriele, A.; Giovanna, A.; et al. Safety and efficacy of Fumaric acid production by Torulopsis glabrata: engineering the urea
AviMatrix (benzoic acid, calcium formate and fumaric acid) for cycle and the purine nucleotide cycle. Biotechnol. Bioeng. 2015, 112,
chickens for fattening, chickens reared for laying, minor avian species 156−167.
for fattening and minor avian species reared to point of lay. EFSA J. (25) Das, R. K.; Brar, S. K.; Verma, M. Enhanced fumaric acid
2017, 15, 5025. production from brewery wastewater by immobilization technique. J.
(6) Yéramian, N.; Chaya, C.; Suarez Lepe, J. A. L-(-)-malic acid Chem. Technol. Biotechnol. 2015, 90, 1473−1479.
production by Saccharomyces spp. during the alcoholic fermentation (26) Das, R. K.; Lonappan, L.; Brar, S. K.; Verma, M. Bio-conversion
of wine (1). J. Agric. Food Chem. 2007, 55, 912−919. of apple pomace into fumaric acid in a rotating drum type solid-state
(7) Hronska, H.; Tokosova, S.; Pilnikova, A.; Kristofikova, L.; bench scale fermenter and study of the different underlying
Rosenberg, M. Bioconversion of fumaric acid to L-malic acid by the mechanisms. RSC Adv. 2015, 5, No. 104472.
bacteria of the genus Nocardia. Appl. Biochem. Biotechnol. 2015, 175, (27) Moon, S.-K.; Wee, Y.-J.; Yun, J.-S.; Ryu, H.-W. Production of
266−273. Fumaric Acid Using Rice Bran and Subsequent Conversion to
(8) Sun, X.; Shen, X.; Jain, R.; Lin, Y.; Wang, J.; Sun, J.; Wang, J.; Succinic Acid Through a Two-Step Process. Appl. Biochem. Biotechnol.
Yan, Y.; Yuan, Q. Synthesis of chemicals by metabolic engineering of 2004, 115, 0843−0856.
microbes. Chem. Soc. Rev. 2015, 44, 3760−3785. (28) Shah, M. V.; van Mastrigt, O.; Heijnen, J. J.; van Gulik, W. M.
(9) Andrew Lin, K.-Y.; Chang, H.-A.; Hsu, C.-J. Iron-based metal Transport and metabolism of fumaric acid in Saccharomyces
organic framework, MIL-88A, as a heterogeneous persulfate catalyst cerevisiae in aerobic glucose-limited chemostat culture. Yeast 2016,
for decolorization of Rhodamine B in water. RSC Adv. 2015, 5, 33, 145−161.
32520−32530. (29) Xu, G.; Liu, L.; Chen, J. Reconstruction of cytosolic fumaric
(10) Mejia-Ariza, R.; Huskens, J. The effect of PEG length on the acid biosynthetic pathways in Saccharomyces cerevisiae. Microb. Cell
size and guest uptake of PEG-capped MIL-88A particles. J. Mater. Fact. 2012, 11, 24.
Chem. B 2016, 4, 1108−1115. (30) Song, C. W.; Kim, D. I.; Choi, S.; Jang, J. W.; Lee, S. Y.
(11) Gérardy, R.; Winter, M.; Horn, C. R.; Vizza, A.; Van Hecke, K.; Metabolic engineering of Escherichia coli for the production of
Monbaliu, J.-C. M. Continuous-Flow Preparation of γ-Butyrolactone fumaric acid. Biotechnol. Bioeng. 2013, 110, 2025−2034.
Scaffolds from Renewable Fumaric and Itaconic Acids under (31) Song, C. W.; Kim, J. W.; Cho, I. J.; Lee, S. Y. Metabolic
Photosensitized Conditions. Org. Process Res. Dev. 2017, 21, 2012− Engineering of Escherichia coli for the Production of 3-Hydrox-
2017. ypropionic Acid and Malonic Acid through beta-Alanine Route. ACS
(12) Rorrer, N. A.; Vardon, D. R.; Dorgan, J. R.; Gjersing, E. J.; Synth. Biol. 2016, 5, 1256−1263.
Beckham, G. T. Biomass-derived monomers for performance- (32) Liu, Y.; Song, J.; Tan, T.; Liu, L. Production of fumaric acid
differentiated fiber reinforced polymer composites. Green Chem. from L-malic acid by solvent engineering using a recombinant
2017, 19, 2812−2825. thermostable fumarase from Thermus thermophilus HB8. Appl.
(13) Yook, S.-H.; Kim, S.-H.; Park, C.-H.; Kim, D.-W. Graphite− Biochem. Biotechnol. 2015, 175, 2823−2831.
silicon alloy composite anodes employing cross-linked poly(vinyl (33) Li, X.; Zhou, J.; Ouyang, S.; Ouyang, J.; Yong, Q. Fumaric Acid
alcohol) binders for high-energy density lithium-ion batteries. RSC Production from Alkali-Pretreated Corncob by Fed-Batch Simulta-
Adv. 2016, 6, 83126−83134. neous Saccharification and Fermentation Combined with Separated
(14) Bekö, S. L.; Schmidt, M. U.; Bond, A. D. An experimental Hydrolysis and Fermentation at High Solids Loading. Appl. Biochem.
screen for quinoline/fumaric acid salts and co-crystals. CrystEngComm Biotechnol. 2017, 181, 573−583.
2012, 14, 1967. (34) Das, R. K.; Brar, S. K.; Verma, M. Application of calcium
(15) Cherukuvada, S.; Nangia, A. Fast dissolving eutectic carbonate nanoparticles and microwave irradiation in submerged
compositions of two anti-tubercular drugs. CrystEngComm 2012, 14, fermentation production and recovery of fumaric acid: a novel
2579. approach. RSC Adv. 2016, 6, 25829−25836.
(16) Menning, M. M.; Dalziel, S. M. Fumaric acid microenviron- (35) Naude, A.; Nicol, W. Improved continuous fumaric acid
ment tablet formulation and process development for crystalline production with immobilised Rhizopus oryzae by implementation of a
cenicriviroc mesylate, a BCS IV compound. Mol. Pharmaceutics 2013, revised nitrogen control strategy. New Biotechnol. 2018, 44, 13−22.
10, 4005−4015. (36) Zhang, K.; Zhang, B.; Yang, S. T. Production of Citric, Itaconic,
(17) Goldberg, I.; Rokem, J. S. Fumaric Acid Biosynthesis and Fumaric, and Malic Acids in Filamentous Fungal Fermentations; Wiley:
Accumulation; Wiley: Hoboken, New Jersey, 2014. Hoboken, New Jersey, 2013.

8280 DOI: 10.1021/acsomega.9b00316


ACS Omega 2019, 4, 8274−8281
ACS Omega Article

(37) Castro, A. J.; Ellenberger, S. R.; Sluka, J. P. The photochemical


isomerization of maleic to fumaric acid: an undergraduate organic
chemistry experiment. J. Chem. Educ. 1983, 60, 521.
(38) Chen, Y.-H.; Jwo, J.-J. Isomerization of Maleic Acid to Fumaric
Acid Catalyzed by Bromate Ion and Bromine. J. Chin. Chem. Soc.
1983, 30, 45−54.
(39) Tachibana, Y.; Masuda, T.; Funabashi, M.; Kunioka, M.
Chemical synthesis of fully biomass-based poly(butylene succinate)
from inedible-biomass-based furfural and evaluation of its biomass
carbon ratio. Biomacromolecules 2010, 11, 2760−2765.
(40) Wang, X. P.; Zhao, Y. Q.; Jaglicic, Z.; Wang, S. N.; Lin, S. J.; Li,
X. Y.; Sun, D. Controlled in situ reaction for the assembly of Cu(II)
mixed-ligand coordination polymers: synthesis, structure, mechanistic
insights, magnetism and catalysis. Dalton Trans. 2015, 44, 11013−
11020.
(41) Li, Q.; Tao, W.; Li, A.; Zhou, Q.; Shuang, C. Poly (4-
vinylpyridine) catalyzed isomerization of maleic acid to fumaric acid.
Appl. Catal., A 2014, 484, 148−153.
(42) Ortiz, R. W. P.; de Jesús, B. G.; Franceschi, E.; Dariva, C.;
Cardozo-Filho, L.; Zanoelo, E. F. Microwave-assisted synthesis of
malic acid involving hydrochloric acid as catalyst. React. Kinet., Mech.
Catal. 2017, 122, 793−802.
(43) Wojcieszak, R.; Santarelli, F.; Paul, S.; Dumeignil, F.; Cavani,
F.; Gonçalves, R. V. Recent developments in maleic acid synthesis
from bio-based chemicals. Sustainable Chem. Processes 2015, 3, 9.
(44) Delhomme, C.; Weuster-Botz, D.; Kühn, F. E. Succinic acid
from renewable resources as a C4building-block chemical-a review of
the catalytic possibilities in aqueous media. Green Chem. 2009, 11,
13−26.
(45) Gao, Z.; Chen, W.; Chen, X.; Wang, D.; Yi, S. Study on the
Isomerization of Maleic Acid to Fumaric Acid without Catalyst. Bull.
Korean Chem. Soc. 2018, 39, 920−924.
(46) Rozelle, L. T.; Alberty, R. A. High-Pressure Acid-Catalyzed
Isomerization and Kinetics of the Acid Catalysis of the Hydration of
Fumaric Acid to Malic Acid. J. Phys. Chem. 1957, 61, 1637−1640.
(47) Jwo, J.-J.; Chen, Y.-H.; Chang, E.-F. Isomerization of Maleic
Acid to Fumaric Acid Catalyzed by Cerium(IV) and N-Bromo
Compounds. J. Chin. Chem. Soc. 1983, 30, 103−115.
(48) Ortiz, R. W. P.; Benincá, C.; Cardozo-Filho, L.; Zanoelo, E. F.
High-Pressure Acid-Catalyzed Isomerization and Hydration of
Fumaric Acid in a Homogeneous Nonisothermal Batch Reactor.
Ind. Eng. Chem. Res. 2017, 56, 3873−3879.
(49) Mattar Knesebeck, A.; Ortiz, R. W. P.; Cardozo-Filho, L.;
Zanoelo, E. F. Isomerization and hydration of fumaric acid under
catalytic and noncatalytic conditions. React. Kinet., Mech. Catal. 2018,
125, 521−534.
(50) Lohbeck, K.; Haferkorn, H.; Fuhrmann, W.; Fedtke, N. Maleic
and Fumaric Acids. Anal. Chem. 2000, 1454−1459.
(51) Wang, Y.; Qu, Q.; Liu, G.; Battaglia, V. S.; Zheng, H.
Aluminum fumarate-based metal organic frameworks with tremella-
like structure as ultrafast and stable anode for lithium-ion batteries.
Nano Energy 2017, 39, 200−210.
(52) DIPPR The DIPPR Information And Data Evaluation Manager
(DIADEM), version 1.2; AIChE: New York, 2000.

8281 DOI: 10.1021/acsomega.9b00316


ACS Omega 2019, 4, 8274−8281

You might also like