Вы находитесь на странице: 1из 10

Article

pubs.acs.org/IECR

Kinetic Investigations of p‑Xylene Oxidation to Terephthalic Acid


with a Co/Mn/Br Catalyst in a Homogeneous Liquid Phase
Meng Li,†,§ Fenghui Niu,§ Daryle H. Busch,‡,§ and Bala Subramaniam*,†,§

Department of Chemical and Petroleum Engineering, ‡Department of Chemistry, and §Center for Environmentally Beneficial
Catalysis, University of Kansas, Lawrence, Kansas 66045, United States
*
S Supporting Information

ABSTRACT: Kinetic investigations of the liquid phase oxidation of p-xylene (pX) to terephthalic acid (TPA) with Co/Mn/Br
catalyst were performed in a stirred 50 mL Parr reactor at 200 °C and 15 bar pressure under conditions wherein product
precipitation is avoided. The oxidant (O2) was introduced by sparging into the liquid phase at constant gas-phase O2 partial
pressure. Apparent kinetic rate constants, estimated by regressing experimental conversion data to a pseudo-first order lumped
kinetic model, are at least an order of magnitude greater than those reported in the literature for similar catalytic reactions. We
attribute this difference to the presence of gas−solid and liquid−solid mass transfer resistances in the previous studies wherein
the TPA product precipitates as it forms, trapping intermediate products and slowing down their oxidation rates. Our results also
indicate that it is not possible to completely eliminate the gas−liquid mass transfer limitations associated with the fast
intermediate oxidation steps, even when operating without solids formation and at high stirrer speeds. Other types of reactor
configurations are therefore needed to better overcome gas−liquid mass transfer limitations. Systematic studies of bromide
concentration effects show that the observed reaction rates become zero order in bromide concentration at sufficiently high
bromide levels where the elimination of intermediate 4-(bromomethyl)benzoic acid by oxidation is favored. Further, the rate
constants do not show any statistically significant dependence on pX concentration as suggested in other reports involving the
traditional three-phase gas−liquid−solid reaction system. This again confirms that the formation of a solid phase hinders the
overall oxidation rate, resulting in much smaller apparent rate constants.

1. INTRODUCTION of which are chain terminators in the polymerization process


Terephthalic acid (TPA) is a commercially important aromatic for PET production. A hydrogenation purification step is
therefore required to upgrade the crude TPA to high purity
compound used mainly as a major precursor of polyethylene
polymer-grade TPA.9 The hydrogenation of 4-CBA is typically
terephthalate (PET) polymer, which is produced by poly-
carried out in a fixed-bed reactor containing a carbon-supported
condensation of ethylene glycol with TPA. More than 90% of
palladium catalyst to quantitatively convert 4-CBA to p-TA at
the worldwide TPA production goes to make PET, consumed
275−290 °C and 70−90 bar.7,9 Polymer-grade TPA product is
primarily for the manufacture of polyester fibers (dominant use
recrystallized from solution leaving the p-TA behind in
in Asia), solid-state or bottle-grade resins (dominant use in
solution. Currently, the main commercial TPA technology
North America and Europe) and polyester film. Worldwide license holders and licensors are BP, DuPont, Dow Chemical,
demand for TPA is steadily growing at a rate of 7−8%, with Mitsubishi Chemical, Eastman Chemical, Mitsui Chemicals,
current worldwide capacity exceeding 50 million metric tons Interquisa, Hitachi and Grupo Petromex. BP is currently the
per annum.1 The Asia region accounts for about 70% of the world’s largest TPA producer with an annual capacity of 7.5
current world TPA capacity.2 million metric tons in 2011.10
Most of the commercial TPA processes are based on the core We recently reported a spray reactor concept as a greener
technology originally developed by Mid-Century Corpora- alternative to the conventional MC process to produce high-
tiona liquid phase bromine-promoted catalytic oxidation of purity TPA (<25 ppm 4-CBA) in a single step at 200 °C and 15
p-xylene (pX) that uses either air or molecular oxygen.3 The bar pressure.11−13 The spray reactor is configured to disperse
well-known Amoco (acquired by BP in 1997) Mid-Century the reaction mixture as fine droplets (by means of a nozzle)
(MC) process is currently the leading TPA technology.4 It into a continuous gas phase containing a stoichiometric excess
involves a stirred oxidation reactor in which air is dispersed into of O2 in CO2 and saturated acetic acid vapor. The spray reactor
the liquid phase containing pX and Co/Mn/Br based catalyst enhances the gas/liquid interfacial mass transfer area compared
dissolved in aqueous acetic acid.5 The pX oxidation is typically to the stirred reactor and eliminates O2 starvation in the liquid
performed at 190−205 °C and 15−30 bar.6 The reactors are
lined with titanium to resist corrosion by hydrobromic acid
Special Issue: Massimo Morbidelli Festschrift
present in the reaction mixture. The crude TPA solid from the
oxidation reactor contains 1000−4000 ppm of 4-carboxyben- Received: October 13, 2013
zaldehyde (4-CBA) due to incomplete oxidation.7 Most Revised: December 16, 2013
polyester applications require TPA that contains less than 25 Accepted: December 16, 2013
ppm 4-CBA and less than 150 ppm p-toluic acid (p-TA),8 both Published: December 16, 2013

© 2013 American Chemical Society 9017 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026
Industrial & Engineering Chemistry Research Article

Figure 1. Experimental apparatus for kinetic investigations.

phase caused by the rapid reaction rates. Further, the reduced workers22,25 developed a fractional kinetic model based on the
backmixing is also beneficial for obtaining high purity products. complex radical chain reaction mechanism and determined that
Consequently, almost complete pX oxidation to TPA is the apparent reaction order (based on pX oxidation to TPA)
achieved in the spray reactor. The reduced residence times in was ≪1 based on experiments performed under industrial
the stirred reactor also impede solvent burning that is observed oxidation conditions. Model parameters were determined from
in conventional stirred reactors. The potential to produce both batch and semicontinuous experimental results. Yan and
polymer-grade dicarboxylic acids in one step eliminates the co-workers23,26 extended the power law model and the
need for a hydrogenation reactor making the spray process a fractional kinetic model to correlate and re-estimate the kinetic
greener and more sustainable process that reduces capital/ parameters from the laboratory reactor (e.g., the fractional
operating costs as well as adverse environmental impacts. kinetic model of Wang et al.22) based on the sample data
For the rational design, optimization, and scale-up of the collected from the industrial CSTR using a data mining
spray reactor, reliable kinetic data are essential. Investigations of approach based on an artificial neural network (ANN). Sun et
pX → TPA oxidation kinetics have been the subject of several al.17,27 developed a simplified free-radical reaction kinetic
reports.14−23 The catalytic oxidation of pX to TPA with Co/ model in which peroxy radicals were involved in the lumped
Mn/Br catalyst system is known to proceed via a free radical reaction equations.
chain mechanism,24 involving a large number of radical and To the best of our knowledge, all the reported kinetic studies
molecular species. A detailed kinetic scheme including all the of pX oxidation to TPA were performed at pX concentrations at
elementary reactions and intermediates has not been possible which the resulting TPA product precipitates from the reaction
primarily because of the challenges in detecting and monitoring mixture resulting in a gas−liquid−solid reaction system. If TPA
the highly reactive radicals in real time. Hence, “lumped” precipitates as a solid, it could trap some of the reaction
kinetic schemes and models17−23 have been adopted to simplify intermediates into its solid structure, as evidenced by the need
this complex reaction by taking into account only relatively for a secondary hydrogenation step in the MC process to
stable and easily detected product species such as p- reduce the 4-CBA concentration in the crude TPA solid. The
tolualdehyde (TALD), p-TA, 4-CBA, and TPA. trapped intermediates would then react at a slower rate due to
Reported kinetic models for pX oxidation to TPA are either restricted O2 diffusion into the solid matrix. The kinetic
based on the power law assumption or derived from the parameters obtained under such reaction conditions would
simplified radical chain reaction mechanisms. Cao and co- clearly be subject to interphase mass transfer limitations. By
workers18−21 investigated the pX oxidation kinetics at low employing lower initial pX concentrations that prevent TPA
temperatures (80−130 °C) and ambient pressure (1 atm) using precipitation, the formation of solids in the liquid phase is
methyl benzoate as the solvent, and cobalt naphthenate and avoided, thus eliminating gas−solid and liquid−solid mass
TALD as the catalyst and promoter, respectively. The rate of transfer steps. Further, the various intermediate products can be
each oxidation step was assumed to be zero order in oxygen more accurately sampled and analyzed.
(when the oxygen partial pressure was greater than 50−100 In this work therefore, the intrinsic kinetics of the liquid-
Torr) and first-order with respect to the reactants being phase pX oxidation to TPA was investigated in a stirred reactor
oxidized. A semibatch reactor model that accounts for inter- under homogeneous conditions employing low pX concen-
and intraphase gas−liquid mass transfer processes was trations under which product precipitation is avoided even at
presented. The agreement between model predictions and total pX conversion to TPA. A first order kinetic model is
experimental data was shown to be satisfactory. Wang and co- assumed based on a lumped reaction scheme in which the
9018 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026
Industrial & Engineering Chemistry Research Article

Table 1. Operating Conditions Used in Kinetics Studies (Varied Parameters Are Shown in Bold)
run no. stirring speed (rpm) T (°C) PO2a (bar) CpX,0 (M) CCo (M) CMn (M) CBr (M)
1−5 600, 800, 1000, 1200, 1400 200 3.4 0.025 0.0125 0.0125 0.035
6−10 1000 200 3.4 0.025 0.0125 0.0125 0.016, 0.009, 0.005, 0.002, 0.001
11 1000 200 3.4 0.035 0.0125 0.0125 0.016
12 1000 200 3.4 0.014 0.0125 0.0125 0.016
industrial conditions28 NAb 195 0.5 3.8 0.006 0.017 0.01
Wang et al.22 800 191 0.4 0.9 0.3 0.3 0.9
a
The total pressure is 15 bar, the makeup gas being CO2. bNA: Not available.

Figure 2. Typical HPLC chromatogram of a product sample after 15 s of pX oxidation, showing initial product + intermediates.

oxygen concentration is maintained constant in stoichiometric in acetic acid. This homogeneous mixture was pressurized with
excess (for pX oxidation to TPA). Intrinsic kinetic rate CO2 to 2.5 bar, and then heated to the desired temperature.
constants, estimated by nonlinear regression of the exper- After the temperature and pressure stabilized, O2 was added to
imental data, are significantly greater than those reported the reactor to initiate the oxidation reaction. The O2 consumed
previously for this catalytic system suggesting that the in the liquid phase by reaction was continuously replenished
avoidance of solid formation in our studies may have from an external reservoir such that O2 partial pressure (and
significantly alleviated interphase transport limitations. therefore reactor pressure) was maintained constant. To avoid
sampling errors due to product flashing and the TPA sticking to
2. EXPERIMENTAL SECTION the inner wall of the tubing, the liquid phase was directly
2.1. Materials. The following chemicals were used as sampled into 15 mL DMF (N,N-dimethylformamide) solvent.
obtained from Sigma-Aldrich without further purification: pX The titanium sampling line was heated to 200 °C to avoid TPA
(substrate), cobalt(II) acetate tetrahydrate (catalyst), precipitation. The unheated line was flushed by DMF after each
manganese(II) acetate tetrahydrate (catalyst), hydrobromic sampling to remove residual reactants and products from the
acid (48% in water) (promoter), acetic acid (purity > 97 wt %) line. Table 1 summarizes the operating conditions employed in
(solvent), biphenyl (internal standard). Ultrahigh-purity the pX oxidation experiments. Typical industrial conditions28
(UHP) grade oxygen and industrial grade CO2 were purchased and operating conditions used for the reported kinetic study are
from Airgas and Linweld, respectively. also listed for comparison.
2.2. Apparatus and Procedure. The experimental 2.3. Analytical Techniques. The reactant (pX), inter-
apparatus employed in this work is shown in Figure 1. The mediates [TALD, p-TA, 4-(bromomethyl)benzoic acid
50 mL titanium stirred vessel (Parr Instrument Company, (BPTA), and 4-CBA] and desired product (TPA) were
series 4560 mini Bench Top Reactor) is equipped with a analyzed by HPLC using the gradient elution technique
magnetically driven impeller for stirring, a dual probe J-type described by Viola and Cao.29 Specifically, the mobile phase
thermocouple, and an OMEGA transducer for measuring the consisted of three eluents: aqueous phase (0.1 wt % phosphoric
reactor temperature and pressure, respectively. The temper- acid) and an organic phase consisting of 7 parts acetonitrile and
ature and pressure are continuously recorded and controlled 2 parts methanol by volume. The gradient elution program was
with the help of a LabVIEW data acquisition system. The as follows: at 0 min, 100 vol % aqueous phase; from 0 to 5 min,
reactor was modified to accommodate multiple inlet/outlet the eluent mixture was changed linearly with time to 95 vol %
ports for (a) introducing O2 into the liquid phase via a dip aqueous phase and 5 vol % organic phase; from 5 to 10 min,
tube/sparger, (b) sampling the liquid phase via a 1/16″ the mixture composition was changed linearly with time to 40
titanium dip tube, (c) withdrawing the vapor phase, and (d) vol % aqueous phase and 60 vol % organic phase; from 10 to 24
providing a pressure relief valve. min, the mixture composition was kept constant. The total flow
In a typical experimental run, the reactor was initially charged rate of the mobile phase was 1 mL/min. A UV detector was
with a 35 mL solution of pX, Co/Mn/Br catalyst, and biphenyl used to quantify the products using biphenyl as internal
9019 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026
Industrial & Engineering Chemistry Research Article

standard (215 nm for pX and 254 nm for all other products). → TPA reaction). Thus, the O2 concentration in the liquid
The HPLC samples were prepared from those collected during phase dominated by acetic acid is maintained constant.
the reaction, and 15 mL of acetic acid was added to each Accordingly, we adopted the pseudo-first-order kinetic model
sample. Figure 2 shows a typical chromatogram obtained from for the sequential reactions involving O2. The model equations
HPLC analysis of a sample. The carbon balance estimated from based on these assumptions are given below:
only the concentrations of the aromatic compounds in the dc1
liquid phase was found to be >95%. = −k1c1
dt
3. LUMPED KINETIC SCHEME AND BATCH REACTOR
dc 2
MODEL = k1c1 − k2c 2
dt
None of the reported lumped kinetic models for pX oxidation
to TPA account for 4-(bromomethyl)benzoic acid (BPTA), dc 3
significant amounts of which are detected in our experiments. = k 2c 2 − k 3c3 − k5c3
dt
Considering its possible important role in the process, the
formation and disappearance of BPTA are also included in our dc4
proposed lumped kinetic scheme as a parallel reaction, as = k5c3 − k6c4
dt
shown in Figure 3. BPTA is formed by the substitution reaction
dc5
= k 3c3 + k6c4 − k4c5
dt
dc6
= k4c5
dt
with the following initial conditions:
c1 = c10 , c 2 = c3 = c4 = c5 = c6 = 0
where the subscripts 1−6 denote pX, TALD, p-TA, BPTA, 4-
CBA, and TPA, respectively, as shown in Figure 3. Athena
Visual Studio30 software was used to estimate the kinetic
parameters. A nonlinear least-squares optimization method was
implemented in the software and encoded using Fortran. The
convergence criteria are as follows:
Figure 3. Proposed lumped kinetic scheme for pX oxidation to TPA.
Sik − Sik + 1
involving p-TA and HBr, and is oxidized in the presence of ≤ 10−3
Sik
oxygen to 4-CBA. The proposed chemical reaction equations
associated with the formation and oxidation of BPTA are n
shown in Figure 4. Si ∑ (ci ,E − ci ,p)2
1

where Si is the sum of squares of component concentration


residuals; cE and cP are experimental and predicted concen-
trations, respectively; n is the number of experimental data
points (n = 10, in this case); and k is the iteration number.
The following observations, based on published literature,
provide valuable guidance for our experimental and modeling
studies. It is generally believed that the p-TA → 4-CBA reaction
is the slowest step in the overall oxidation sequence.21,22 If this
is indeed the case, then the estimated first-order rate constant
for this step (k3) should be several-fold less than the other rate
constants in the sequential oxidation, and the overall oxidation
rate would be the rate of the slowest step. In such a case, it is
Figure 4. Scheme for formation and oxidation of BPTA. possible to assess the presence or absence of gas−liquid mass
transfer limitations associated with the slowest step from the
The reported power law kinetic models for pX oxidation to influence of stirrer speed on the estimated rate constant for that
TPA mainly assume pseudo-first-order oxidation kinetics (in step. It must be recognized that the estimated rate constants
excess of O2). Cao et al.19 applied the first order kinetic model associated with the nonrate determining steps could be merely
at low pX conversions wherein no solid phase was formed; that mass transfer coefficients, since the O2 transport into the liquid
is, the system was considered homogeneous. Wang et al.25 also phase is also a first-order rate process.
deduced from their fractional kinetic model that when the
reactant (pX) concentration was low enough compared to O2, 4. RESULTS AND DISCUSSION
the first order kinetic model was valid. In our experiments, TPA 4.1. Temperature and Pressure Profiles. Figure 5 shows
precipitation is avoided and the O2 concentration was the temperature and pressure profiles corresponding to run 3
maintained in excess (at least 5-fold with respect to the pX (Table 1). Since low enough initial pX concentration (0.025
9020 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026
Industrial & Engineering Chemistry Research Article

(beyond the range of uncertainty) and tends to level off around


1000 rpm. In contrast, the other rate constants in the oxidation
sequence (k1, k2, and k4) do not vary in this stirrer speed range.
These observations suggest that while it is possible to overcome
gas−liquid mass transfer limitations associated with the p-TA
→ 4-CBA step in the stirred reactor, the kinetics associated
with other rates are indeed much faster as reported in previous
studies. In addition, the lowest rate constant (k6) is invariant in
this stirrer rpm range suggesting that it is not influenced by
gas−liquid mass transfer limitations. Note that there is an
apparent decrease in rate constants with a further increase in
the stirring rate to 1400 rpm. This is probably due to the
Figure 5. Temporal temperature and pressure profiles during intrinsic formation of a deep meniscus at the higher agitation speed that
kinetic studies of pX oxidation in a stirred reactor. decreases the liquid level below that of the sparger (either
partly or fully) used for introducing O2 into the reactor. Hence,
M) was charged in the reactor, the reactor temperature rise we conclude that the maximum rate of TPA formation in this
(caused by the heat of reaction) was within 1 °C, which allowed stirred reactor occurs beyond 1000 rpm. Figure 6 shows the
us to assume that the reactor was operated under isothermal experimentally observed concentrations at various stirrer
conditions. By replenishing O2 as needed from an external speeds. The fitted concentration profiles at 1000 rpm (solid
reservoir, the reactor pressure was maintained constant (15 line) match the experimental data quite well.
bar) except for brief, insignificant pressure fluctuations during Gas−liquid mass transfer coefficient for the pX oxidation
sampling. system at conditions corresponding to Run No. 4 in Table 1 (at
4.2. Effect of Stirring Rate. The intrinsic kinetic rates of 1200 rpm) was estimated. Details of estimation are provided in
the sequential steps during pX oxidation under MC process the Supporting Information. The product of the gas−liquid
conditions (200 °C, 15 bar) are very fast and tend to be limited mass transfer coefficient (kL) and the gas−liquid interfacial area
by gas−liquid mass transfer resistance. The experiments (Runs per unit volume (av), kLav, was estimated to be 65 min−1. The
1−5 of Table 1) were carried out in the stirred reactor to fact that the mass transfer rate constant (kLav) is roughly an
evaluate the effect of stirring speed on the reaction rates. Figure order of magnitude greater than the rate constant associated
6 shows the temporal product distributions at various stirring with the slowest oxidation step (k3) suggests that gas−liquid
speeds. mass transfer limitations associated with the p-TA → 4-CBA
Effective rate constants, estimated from the reactant and step are eliminated at these operating conditions. However, the
product concentration profiles as explained previously, are gas−liquid mass transfer rate constant is of the same order of
shown in Table 2. Clearly, the slowest step in the series reaction magnitude as other rate constants in the oxidation sequence
sequence is the formation of 4-CBA from p-TA, and that in the (k1, k2, and k4), suggesting that O2 saturation of the liquid phase
parallel reaction is the oxidation of BPTA to 4-CBA. At all might not have been achieved even at high stirring speeds. In
stirrer speeds ranging from 600 to 1400 rpm, the rate constants such a scenario, any further increase in the overall TPA
associated with these steps (k3 and k6) are several-fold lower formation rate may be achieved only with a reactor
than those associated with the other steps. The uncertainties configuration that provides enhanced gas−liquid mass transfer
associated with the larger rate constants are greater, as might be area such that O2 saturation of the liquid phase is not rate-
anticipated with the measurement of faster reaction rates. For limiting. This is achieved in a spray reactor in which the liquid
instance, because the first two reactions in the oxidation phase is dispersed as fine droplets into a continuous gas
sequence are so fast, the sampling errors associated with TALD phase.13
and p-TA will be substantially more. In contrast, the 4.3. Comparison of Estimated Rate Constants with
uncertainties associated with the rate constants of the slowest Literature Data. Table 3 compares the estimated reaction rate
steps (k3, k5, and k6) are much lower. For the sequential constants in this work with the reported data from literature.
oxidation steps, the rate constant associated with 4-CBA Even though different reaction conditions and kinetic models
formation from p-TA (k3) increases from 600 to 800 rpm are employed, the relative rate constants of the various steps

Figure 6. Experimental and simulated product distributions at different stirring speeds [reaction conditions: T = 200 °C, P = 15 bar; initial pX = 25
mM, Co = 12.5 mM, Mn = 12.5 mM, Br = 32.5 mM; O2/CO2 (mol/mol) = 1:1].

9021 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026


Industrial & Engineering Chemistry Research Article

Table 2. Estimated Reaction Rate Constants at Different Stirring Speeds (Initial [pX] = 0.025 M; All Other Experimental
Conditions Are as Shown in Table 1)
stirrer speed (rpm) k1 (min−1) k2 (min−1) k3 (min−1) k4 (min−1) k5 (min−1) k6 (min−1)
a
600 11.7 (1.72) 33.2 (18.9) 3.90 (0.548) 17.5 (4.78) 4.12 (0.568) 0.421 (0.0405)
800 11.8 (1.96) 37.1 (26.0) 4.45 (0.764) 18.6 (5.24) 4.71 (0.799) 0.423 (0.0450)
1000 11.7 (1.85) 39.0 (27.1) 5.89 (1.10) 22.5 (6.83) 6.00 (1.10) 0.414 (0.0406)
1200 13.6 (2.31) 42.1 (32.9) 5.56 (1.03) 23.2 (6.77) 5.81 (1.05) 0.435 (0.0364)
1400 13.9 (2.75) 36.8 (28.9) 4.49 (0.793) 20.5 (5.91) 4.88 (0.844) 0.401 (0.0369)
a
95% confidence level shown in parentheses.

Table 3. Comparison of Estimated Rate Constants with Literature Data


model type k1 (min−1) k2 (min−1) k3 (min−1) k4 (min−1) ref remarks
fractional kinetic 0.176 0.725 0.0361 0.338 22
model 0.732 1.422 0.139 0.873 26 parameter re-estimated for Wang et al.’s model22
power law assumption 1.9(10−3) 7.6(10−2) 6.6(10−4) 1.86(10−2) 19 first order in reactant concentrations
0.38 0.99 0.123 0.44 23 0.65 order in pX concentration, first order in other reactant
concentrations
11.74 39.03 5.889 22.50 this worka run no. 3 in Table 1
a
Stirrer speed = 1000 rpm; initial [pX] = 0.025 M; all other operating conditions are as listed in Table 1.

Figure 7. Color changes observed in the reaction solution due to the formation of cobalt bromides31 at various bromide concentrations in acetic
acid: [Co(II)] = 12.5 mM.

follow the same trend: k3 < k1 < k4 < k2. Our investigations due to mass transfer limitations stemming from intermediate
concur with previous reports that the slowest step is the products being trapped in the solids.
oxidation of p-TA to 4-CBA. However, the estimated reaction 4.4. Effect of Bromide Concentration. Different bromide
rate constants in this work are at least an order of magnitude concentrations (runs 3 and 6−10 of Table 1) were used in the
greater than those reported in the referenced studies. We kinetic study experiments for pX oxidation to TPA to
attribute this to the fact that the reaction was performed under investigate the effect of bromide concentration on the reaction
homogeneous conditions in this work, avoiding product rate. Figure 7 shows the color change of the reaction solution
corresponding to the various bromide concentrations in run 3
precipitation in the liquid phase and thereby eliminating gas−
and runs 6−10 in Table 1, respectively. The solution changes
solid and liquid−solid mass transfer limitations.
color from blue to pink when bromide concentration is
It is worth pointing out that Cincotti et al.21 found that their
changed from 34.8 mM to 0.91 mM. The reason of the color
first order kinetic model fits the experimental data well even at
change is probably associated with the formation of cobalt
initial pX concentrations ([pX] = 4.33−9.75 M) that are higher bromides.31
than industrial conditions. The reaction was performed at low Figure 8 shows the component concentrations with different
temperatures (80−130 °C) to slow down the kinetics, and the bromide concentrations. Surprisingly, high concentrations of
first order kinetic model matches well with experimental data BPTA were detected during the pX oxidation. Although it is
prior to the formation of the solid products. On the basis of this well recognized that the p-TA to 4-CBA conversion is the rate
evidence, it seems plausible that the apparent change in order at determining step, the slowest step is BPTA oxidation according
high pX concentrations reported in some reported studies is to the estimated rate constants (Table 4). The decrease of the
9022 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026
Industrial & Engineering Chemistry Research Article

Figure 8. Experimental and simulated product distributions with various bromide concentrations (initial [pX] = 0.025 M; stirring rate =1000 rpm; all
other experimental conditions are as shown in Table 1).

Table 4. Estimated Reaction Rate Constants at Different Bromide Concentrations (Initial [pX] = 0.025 M; Stirring Rate =1000
rpm; All Other Experimental Conditions Are as Shown in Table 1)a
CBr (mM) k1 (min−1) k2 (min−1) k3 (min−1) k4 (min−1) k5 (min−1) k6 (min−1)
a
34.8 11.7 (1.91) 39.0 (27.1) 5.89 (1.10) 22.5 (6.83) 6.00 (1.10) 0.414 (0.0405)
15.8 11.8 (1.60) 39.6 (25.1) 7.75 (1.45) 29.3 (10.3) 7.69 (1.42) 0.307 (0.0227)
9.44 11.6 (2.28) 45.6 (43.2) 5.64 (1.27) 25.4 (12.8) 4.17 (0.955) 0.261 (0.0359)
4.93 8.25 (2.36) 56.3 (48.2) 2.43 (0.648) 11.8 (6.51) 1.17 (0.413) 0.205 (0.0913)
2.40 6.45 (1.70) 17.6 (11.0) 0.927 (0.183) 7.84 (4.08) 0.223 (0.119) 0.152 (0.125)
0.91 7.80 (2.66) 14.5 (8.92) 0.419 (0.0719) 3.87 (1.93) 0.0585 (0.0477) 0.0658 (0.0599)
a
95% confidence level shown in parentheses.

bromide concentration will slow down the BPTA disappearance attains a maximum, increased when lowering the cobalt
rate. Partenheimer32 illustrated the role of bromide in the Co/ concentration.
Mn/Br catalyst system during the oxidation of pX to TPA. It is reported that as much as 99% of the initial inorganic
When bromide is added, there is a large increase in activity and bromide is converted to benzylic bromide and its concentration
selectivity due to the rapid electron transfer from cobalt to remains approximately the same until the oxidation of the
hydrocarbon is nearly complete.34 Partenheimer,35,36 as well as
bromide. Kamiya et al.33 investigated the effect of bromine/
Saha and Espenson,37 also found that a decrease in Co/Mn/Br
metals ratio on the reaction rate and found that at a cobalt catalytic activity is attributed to the benzylic bromide formation
concentration of 0.05 M, the pX → TPA oxidation rate rapidly and that benzylic bromide, unlike the inorganic bromide, has
increased to a maximum value as the molar Br/Co ratio was virtually no activity or promotional effect in Co/Mn/Br
increased to 1 and remained constant up to a ratio of 9. The autoxidations. The mechanism of BPTA oxidation has not
optimum Br/Co ratio, at which the pX → TPA oxidation rate been well studied. Metelski et al.34 developed a proposed
9023 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026
Industrial & Engineering Chemistry Research Article

mechanism for benzyl bromide oxidation (Figure 9). According invariant when the confidence intervals are taken into account.
to this mechanism, benzyl bromide is oxidized in the presence This invariance is to be expected in the case of intrinsic rate
constants. However, as shown in Figure 10, the rate constants

Figure 10. Effect of initial pX concentration on the estimated rate


constants23

are reported to decrease with the increase of initial pX


concentration in a three-phase reaction system.23 This again
confirms that the formation of a solid phase introduces mass
transfer resistances, thereby lowering the rates and associated
rate constants.

5. CONCLUSIONS
The kinetics of the liquid phase oxidation of p-xylene (pX) to
terephthalic acid (TPA) was investigated in a stirred reactor
Figure 9. Proposed scheme for the autoxidation of benzyl bromide under homogeneous conditions (i.e., low pX concentrations
(Adapted from ref 34. Copyright 2000 American Chemical Society). employed to avoid TPA precipitation). A first order kinetic
model was developed based on a lumped reaction scheme that
of Co(III), leading to the recovery of bromide ions. includes a parallel step describing the formation of 4-
Simultaneously, Co(III) is reduced to Co(II) by hydrogen (bromomethyl)benzoic acid (BPTA). Kinetic rate constants
bromide (faster) and also by benzyl bromide, albeit at a much regressed from the experimental data are shown to be at least
slower rate. Partenheimer35 reported that the competing an order of magnitude greater than those reported in the
mechanisms of oxidation and solvolysis for the disappearance literature. This is attributed to the fact that the conversion data
of benzylic bromide occur at approximately the same rates were obtained under homogeneous conditions in this study
under the studied experimental conditions. (unlike previous literature data that were obtained in gas−
4.5. Effect of Substrate Concentration. Experiments liquid−solid systems) avoiding solid formation in the liquid
with different pX concentrations (runs 6, 11, and 12 of Table 1) phase, and thereby eliminating gas−solid and liquid−solid mass
were carried out to study the effect of substrate concentration transfer limitations. Our investigations concur with previous
on the regressed rate constants. Solubility studies show that reports that the slowest step is the oxidation of p-toluic acid to
approximately 1.6 g of TPA is completely dissolved in 100 g of 4-carboxybenzaldehyde. Systematic investigations of the effects
acetic acid at 200 °C and 15 bar,12 suggesting a maximum pX of stirrer speed reveal that it may not be possible to completely
concentration of 0.1 M to maintain a homogeneous liquid eliminate the gas−liquid mass transfer limitations in even a
phase. Table 5 shows the estimated rate constants. It was found laboratory scale stirred reactor. This further confirms that other
that the rate constants at the different concentrations are reactor configurations (such as a spray reactor in which the

Table 5. Estimated Reaction Rate Constants at Different pX Concentrations (P = 15 bar; T = 200 (°C; Stirring Rate =1000 rpm;
All Other Experimental Conditions Are As Shown in Table 1)a
run no. CpX,0 (mM) k1 (min−1) k2 (min−1) k3 (min−1) k4 (min−1) k5 (min−1) k6 (min−1)
a
12 14 11.8 (2.56) 40.1 (42.0) 5.20 (1.27) 18.73 (6.60) 5.64 (1.37) 0.225 (0.0255)
6 25 11.8 (1.60) 39.6 (25.1) 7.75 (1.45) 29.3 (10.3) 7.69 (1.42) 0.307 (0.0227)
11 35 11.0 (2.10) 38.9 (29.9) 8.81 (2.36) 30.3 (16.9) 5.87 (1.61) 0.302 (0.0445)

a
95% confidence level shown in parentheses.

9024 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026


Industrial & Engineering Chemistry Research Article

liquid is the dispersed phase) are needed to completely (3) Landau, R.; Heights, R.; Saffer, A. Process for the preparation of
overcome gas−liquid mass transfer limitations. The effects of terephthalic acid. U.S. Patent 2833818, 1958.
bromide and initial pX concentrations on the reaction rate were (4) Tomás, R. A. F.; Bordado, J. C. M.; Gomes, J. F. P. p-Xylene
also investigated. The reaction rate shows mixed order oxidation to terephthalic acid: A literature review oriented toward
dependence on bromide concentration; positive order at process optimization and development. Chem. Rev. 2013,
relatively low values and zero order at higher bromide DOI: 10.1021/cr300298j.
(5) Raghavendrachar, P.; Ramachandran, S. Liquid-phase catalytic
concentrations where the intermediate BPTA elimination by oxidation of p-xylene. Ind. Eng. Chem. Res. 1992, 31, 453−462.
oxidation is favored. Decreases in the reaction rate constants (6) Weissermel, K.; Arpe, H.-J. Industrial Organic Chemistry; VCH
with increasing pX concentration, reported in traditional three- Publishers, Inc.: New York, 1997.
phase reaction systems, were not observed under the (7) Ure, A. M.; Parker, D. Methods, processes, and system for
homogeneous liquid phase conditions of our study. This treating and purifying crude terephthalic acid and associated process
further confirms that the presence of the solid TPA phase streams. WO 2010/122304 A1, 2010.
hinders the reaction rate and the kinetic rate constants reported (8) Hashmi, S. M. A.; Al-Luhaidan, S. Process for preparing purified
in the literature are influenced by mass transfer limitations. terephthalic acid. EP 1671938 A1, 2006.
(9) Meyer, D. H. Fiber-grade terephthalic acid by catalytic hydrogen
6. TRIBUTE TO PROFESSOR MASSIMO MORBIDELLI treatment of dissolved impure terephthalic acid. U.S. Patent 3584039,
1971.
On the occasion of Professor Morbidelli’s 60th birthday, we (10) BP. Annual Report and Form 20-F 2011. bp.com/annualreport
wish him good health and cheer. We hope that this (accessed Jan. 2013).
contribution is a fitting tribute to Professor Morbidelli’s (11) Subramaniam, B.; Busch, D. H.; Niu, F. Spray process for
outstanding contributions to the field of chemical engineering selective oxidation. WO 2010/111288 A2, 2010.
and to his profession. The diversity of topics in which Professor (12) Li, M. A Spray reactor concept for catalytic oxidation of p-xylene
Morbidelli has made impactful contributions, ranging from to produce high-purity terephthalic acid. Ph.D. Dissertation, University
parametric sensitivity in chemical reactors to adsorption-based of Kansas, Lawrence, KS, 2013.
separations by simulated moving beds to polymer reaction (13) Li, M.; Niu, F.; Zuo, X.; Metelski, P. D.; Busch, D. H.;
engineering, is truly impressive. It should come as no surprise Subramaniam, B. A spray reactor concept for catalytic oxidation of p-
that he has also published in the area of p-xylene oxidation to xylene to produce high-purity terephthalic acid. Chem. Eng. Sci. 2013,
TPA, the subject of this dedication. It is especially a pleasure for 104, 93−102.
one of the authors (B.S.) to have enjoyed a special friendship (14) D’oro, P. C.; Danoczy, E.; Roffia, P. Low temperature oxidation
of p-xylene. Oxid. Commun. 1980, 1 (2), 153−162.
with Professor Morbidelli, sharing memorable times during (15) Morbidelli, M.; Paludetto, R.; Carra, S. Gas−liquid autoxidation
their graduate student days at the University of Notre Dame, a reactors. Chem. Eng. Sci. 1986, 41, 2299−2307.
sabbatical stay at ETH, Zürich, and a visit to Politecnico di (16) Jacobi, R.; Baerns, M. The effect of oxygen transfer limitation at
Milano. It is indeed a distinct privilege to be invited to the gas-liquid interphase. Kinetics and product distribution of the p-
contribute to this special issue honoring him.


xylene oxidation. Erdoel Kohle Erdgas P. 1983, 36 (7), 322−326.
(17) Sun, W.; Pan, Y.; Zhao, L.; Zhou, X. Simplified free-radical
ASSOCIATED CONTENT reaction kinetics for p-xylene oxidation to terephthalic acid. Chem. Eng.
*
S Supporting Information Technol. 2008, 31 (10), 1402−1409.
Details of the estimation of the gas−liquid mass transfer (18) Cao, G.; Pisu, M.; Morbidelli, M. A lumped kinetic model for
liquid-phase catalytic oxidation of p-xylene to terephthalic acid. Chem.
coefficient at the reactor operating conditions; transient
Eng. Sci. 1994, 49 (24B), 5775−5783.
concentration data used in Figures 6 and 8. This material is (19) Cao, G.; Servida, A.; Pisu, M. Kinetics of p-xylene liquid-phase
available free of charge via the Internet at http://pubs.acs.org.


catalytic oxidation. AIChE J. 1994, 40 (7), 1156−1166.
(20) Cincotti, A.; Orrù, A.; Cao, G. Effect of catalyst concentration
AUTHOR INFORMATION and simulation of precipitation processes on liquid-phase catalytic
Corresponding Author oxidation of p-xylene to terephthalic acid. Chem. Eng. Sci. 1997, 52
*E-mail: bsubramaniam@ku.edu. Tel.: 785-864-2903. Fax: 785- (21/22), 4205−4213.
864-6051. (21) Cincotti, A.; Orrù, A.; Cao, G. Kinetics and related engineering
aspects of catalytic liquid-phase oxidation of p-xylene to terephthalic
Notes acid. Catal. Today 1999, 52, 331−347.
The authors declare no competing financial interest.


(22) Wang, Q.; Li, X.; Wang, L.; Cheng, Y.; Xie, G. Kinetics of p-
xylene liquid-phase catalytic oxidation to terephthalic acid. Ind. Eng.
ACKNOWLEDGMENTS Chem. Res. 2005, 44, 261−266.
This research was supported in part by the National Science (23) Yan, X. Data mining macrokinetic approach based on ANN and
Foundation (EEC-0310689) and in part by the U.S. Depart- its application to model industrial oxidation of p-xylene to terephthalic
ment of Agriculture (2011-10006-30362). We gratefully acid. Chem. Eng. Sci. 2007, 62, 2641−2651.
(24) Partenheimer, W. Methodology and scope of metal/bromide
acknowledge valuable discussions with Dr. Peter Metelski autoxidation of hydrocarbons. Catal. Today 1995, 23, 69−158.
(BP Products North America Inc.) regarding the effects of (25) Wang, Q.; Cheng, Y.; Wang, L.; Li, X. Semicontinuous studies
bromide concentration on the reaction kinetics.


on the reaction mechanism and kinetics for the liquid-phase oxidation
of p-xylene to terephthalic acid. Ind. Eng. Chem. Res. 2007, 46, 8980−
REFERENCES 8992.
(1) ICIS. ICIS pX, PTA, MEG Market Outlook Webinar: Asia and (26) Dong, Y.; Yan, X. Hybrid model of industrial p-xylene oxidation
Middle East Markets. Personal contact (communicated May 2013). incorporated fractional kinetic model with intelligent models. Ind. Eng.
(2) ICIS. http://www.icis.com/Articles/2007/11/06/9076461/ Chem. Res. 2013, 52, 2537−2547.
purified-terephthalic-acid-pta-uses-and-market-data.html (accessed (27) Qian, F.; Tao, L.; Sun, W.; Du, W. Development of a free radical
Feb. 2013). kinetic model for industrial oxidation of p-xylene based on artificial

9025 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026


Industrial & Engineering Chemistry Research Article

neural network and adaptive immune genetic algorithm. Ind. Eng.


Chem. Res. 2012, 51, 3229−3237.
(28) Abrams, K. J. Process for preparing aromatic carboxylic acids
with efficient energy recovery by the oxidation of aromatic
hydrocarbon feedstocks. WO 96/11899 A1, 1996.
(29) Viola, A.; Cao, G. A rapid direct analysis of p-xylene oxidation
products by reverse phase HPLC. J. Chromatogr. Sci. 1996, 34, 27−33.
(30) Athena Visual Studio, v 14.2; Athena Visual Software, Inc.:
Naperville, IL, 2012.
(31) Sawada, K.; Tanaka, M. Formation of bromo complexes of
cobalt(II) in acetic acid. J. Inorg. Nucl. Chem. 1977, 39 (2), 339−344.
(32) Partenheimer, W. A Chemical Model for the Amoco “MC”
Oxygenation Process to Produce Terephthalic Acid. Catalysis of
Organic Reactions; Marcel Dekker: New York, 1990; pp 321−346.
(33) Kamiya, Y.; Nakajima, T.; Sakoda, K. Autoxidation of p-xylene
catalyzed with cobalt monobromide in acetic acid. Bull. Chem. Soc. Jpn.
1966, 39 (10), 2211−2215.
(34) Metelski, P. D.; Adamian, V. A.; Espenson, J. H. Mechanistic
role of benzylic bromides in the catalytic autoxidation of methylarenes.
Inorg. Chem. 2000, 39, 2434−2439.
(35) Partenheimer, W. The complex synergy of water in the metal/
bromide autoxidation of hydrocarbons caused by benzylic bromide
formation. Adv. Synth. Catal. 2004, 346, 297−306.
(36) Patenheimer, W. The complex synergy of water in the metal/
bromide autoxidations. Part II. Effect of water and catalyst on the
aerobic oxidation of benaldehydes and the effect of water on the
elementary catalytic pathways. Adv. Synth. Catal. 2005, 347, 580−590.
(37) Saha, B.; Espenson, J. H. Low-bromide containing MC catalyst
for the autoxidation of para-xylene. J. Mol. Catal. A 2007, 271, 1−5.

■ NOTE ADDED AFTER ASAP PUBLICATION


This paper was published ASAP on December 26, 2013. Minor
text corrections were made to the Tribute section, and the
corrected version was reposted on January 2, 2014.

9026 dx.doi.org/10.1021/ie403446b | Ind. Eng. Chem. Res. 2014, 53, 9017−9026

Вам также может понравиться