Novel Energy Efficient Process For Acetic Acid

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

Novel energy efficient process for acetic acid


production by methanol carbonylation

Alexandre C. Dimian a,∗ , Anton A. Kiss b


aUniversity P̈olitehnicaöf Bucharest, Polizu 1-7, 011061 Bucharest, Romania
bDepartment of Chemical Engineering and Analytical Science, The University of Manchester, Sackville Street,
Manchester M13 9PL, United Kingdom

a r t i c l e i n f o a b s t r a c t

Article history: Acetic acid is an essential chemical product. The need for ‘green’ acetic acid has spurred
Received 16 January 2020 new research in developing more sustainable processes. This paper presents an original
Received in revised form 26 conceptual design of a low-energy and low-emissions sustainable process for acetic acid
February 2020 manufacturing by methanol carbonylation. The emphasis is set on energy efficiency, as the
Accepted 14 April 2020 exothermic reaction releases a large amount of energy that may be more effectively used.
Available online 27 April 2020 Two processes are investigated by rigorous simulation, based on homo- and heterogeneous
catalysis. The latter reveals as innovative feature the full valorization of the energy released
Keywords: in the reaction section, along with reducing the separation sequence to only two distillation
Acetic acid columns. The first dewatering column is driven by low-pressure steam generated by the
Process design reactor cooling. The second purification column uses heat pumping, in which the energy
Process simulation rejected in condenser is upgraded for reboiler heating by vapor compression (VC), using
Sustainable process water as working fluid. The electricity for the vapor compression is ensured by applying
Energy efficiency an Organic Rankine Cycle (ORC) for waste heat recovery by the reactor cooling. In this way,
the new green process needs very little energy (below 1 MJ/kg product). This eco-efficient
process shows also superior sustainability metrics (e.g. low emissions of 40 gCO2 e/kg) as
compared to the homogeneous catalyst process.
© 2020 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction Amidst growing environmental awareness and more strin-


gent regulations worldwide, there is an imminent shift of
Acetic acid is a major essential chemical product, being an acetic acid production strategy in favour of greener processes
important chemical reagent used primarily in the produc- (Pal and Nayak, 2017). The development and advances in acetic
tion of cellulose acetate, vinyl acetate monomer and polyvinyl acid processes and catalysts aim primarily to reduce the raw
acetate, acetic anhydride, monochloroacetic acid, acetate material consumption, energy requirements, and investment
esters, terephthalic acid, and synthetic fibres and fabrics (Le costs (Yoneda et al., 2001). The most important processes
Berre et al., 2014). It is often used in descaling agents, or as for acetic acid production are the carbonylation of methanol,
food additive. The global market size was 16.3 million tonnes vapor phase-oxidation (ethane and ethene), liquid-phase
in 2018, and it is further expected to reach a volume of about oxidation (n-butane and naphta), acetaldehyde oxidation,
20.3 million tonnes by 2024 (Marketwatch, 2019). Asia-Pacific oxidative and anaerobic fermentation (Yoneda et al., 2001;
has currently the highest market share, followed by North Le Berre et al., 2014). Among them, the manufacturing of
America, Europe, Middle East and Africa. acetic acid by methanol carbonylation takes the largest market
share (about 75%), in continuous growth. Research strategies
to improve the catalyst performance and process economics
of methanol carbonylation resulted in both rhodium- and
∗ iridium-catalyzed systems that operate with high activity at
Corresponding author.
E-mail address: ac dimian@chim.upb.ro (A.C. Dimian). reduced water concentration (Haynes, 2010). More sustainable
https://doi.org/10.1016/j.cherd.2020.04.013
0263-8762/© 2020 Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
2 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12

catalysts have been explored, such as immobilized rhodium 2.1. Catalysis and technology issues
and iodide species on cross-linked copolymers (Li et al., 2013),
heterogeneous rhodium-incorporated graphitic carbon nitride Methanol carbonylation may involve rhodium catalyst (Mon-
(Nam et al., 2017), a halide-free and noble metal-free cata- santo process) or iridium-base catalyst (CativaTM process) as
lyst based on pyridine-modified H-mordenite zeolite (Ni et al., homogeneous catalysts (Jones, 2000; Le Berre et al., 2014).
2017). The trend of using methanol carbonylation is favourable The reaction takes place in liquid phase at 150–200 ◦ C and
from a sustainability viewpoint, as both raw materials can be 30–60 bar. Methyl acetate is used as solvent for the catalytic
obtained from any feedstock leading to syngas, such as biogas complex. The elementary reactions are (Moulijn et al., 2013):
and biomass. For this reason, acetic acid can be a 100% green
product, and on this basis promotes the synthesis of other
CH3 -OH + HI ↔ CH3 I + H2 O Hr,298 = −53.1 kJ/mol (2)
bio-based chemicals. For example, the vinyl acetate monomer
(VAM) manufactured from acetic acid and ethylene (Dimian
and Bildea, 2008) becomes a bio-based product, as well as the CH3 I + CO → CH3 COI Hr,298 = −66.9 kJ/mol (3)
valuable vinyl paints. In this way traditional petrochemicals
can be replaced by green products obtained from a syngas
bio-refinery platform (Dimian, 2007; Dimian et al., 2019).
Other processes based on renewable materials include
CH3 COI + H2 O → CH3 COOH + HI Hr,298
hydrothermal processing of biomass waste (Jin et al., 2007),
integrated production in a pyrolysis oil-based biorefinery = −15.6 kJ/mol (4)
(Vitasari et al., 2015), and the fermentation of sugar residue
solutions driven by suitable bacteria (Vidra and Németh, 2018). Details about the reaction mechanism and implications on
Despite the ecological interest, the economic efficiency is kinetics may be found in Jones (2000). The key stage reac-
penalized by high energy use in the separation and purifi- tion (3) involves a metallic complex, with Rh or Ir. In the
cation operations (Kiss et al., 2016). The result is that the first case, the metallic complex [Rh(CO)2 I2 ] reacts initially
fermentation process is limited nowadays to only 10% of the with CH3 I giving [Rh(CO)2 MeI3 ]. This intermediate under-
market. goes a molecular rearrangement leading to the formation of
A novel route reported recently involves the hydro- an acyl group [Rh(CO)(CO-Me)I3 ]. Further, the addition of a
carboxylation of methanol with CO2 and H2 (Qian et al., 2016). CO molecule leads to an unstable intermediate [Rh(CO)2 (CO-
The reaction can be efficiently catalyzed by Ru–Rh bimetal- Me)I3 ]. It follows the expulsion of the acyl iodide CH3 COI and
lic catalyst using imidazole as ligand and LiI as promoter in the hydrolysis reaction (4) that produces the acetic acid and
1,3-dimethyl-2-imidazolidinone (DMI) solvent with outstand- HI. The initial catalytic complex is rebuilt, as well as HI that
ing reaction results. This method could open a road to fix CO2 re-joins the cycle. CH3 I, HI and H2 O implied in the catalytic
into bulk chemicals, but it is questionable whether the hydro- cycle are recovered and recycled.
gen manufacturing price could be competitive enough. This Controlling the amount of water is a key factor in process
might be the case if hydrogen is produced by electrolysis using design. Water is formed directly by the esterification of the
renewable energy, while methanol is used for energy storage. acetic acid with methanol, but it comes mainly by recycling
This study focuses on the conceptual design of a novel from the separation section. When the water content is higher
low-energy sustainable process for acetic acid manufactur- than 8 wt% the rate determining step in is the addition of CH3 I
ing by methanol carbonylation, using heterogeneous catalysis. to the Rh centre:
This key feature is obtained by full valorization of the energy
released by the exothermal reaction. The energy efficiency is
RCH3COOH = k × [Rh] × [CH3 I] (5)
improved by steam generation (by reactor cooling), as well as
using vapour compression (VC) coupled with an Organic Rank-
ine Cycle (ORC) that recovers the reactor’s effluent enthalpy. The reaction rate is essentially first order in both cata-
The plant capacity considered in this work is 200 ktpy acetic lyst and methyl iodide concentrations, but independent of
acid with a target purity of 99.7 wt%. The analysis and design is the reactants concentrations, CH3 OH and CO. Therefore, high
performed by employing the rigorous process simulator Aspen conversions can be obtained even in a CSTR of small volume
Plus v9.0. (Moulijn et al., 2013). If the water content is less than 8 wt%,
the rate determining step is the reductive elimination of the
acyl species.
2. Production processes and technologies
Keeping the integrity of the catalyst is essential. The
main cause of catalyst loss by precipitation is the reaction of
The methanol carbonylation can be described by the follow-
rhodium-acyl species with HI that leads to acetaldehyde and
ing stoichiometric equation, with a theoretical carbon yield of
the complex [Rh(CO)I4 ]− . The latter may cause a catalyst loss
100%:
by precipitation of the inactive and insoluble RhI3 , when the
CO concentration is insufficient.
CH3 -OH + CO → CH3 -COOH Hr,298 = −135.6 kJ/mol (1)

[Rh(CO)(CO-CH3 )I3 ]− + HI → CH3 -CHO


The reaction is moderately exothermic and thermodynam-
ically favoured by lower temperature. The industrial operation + [Rh(CO)I4 ]− → RhI3 + I− + CO (6)
range is usually 15–200 ◦ C and pressures of 30–60 bar (Le Berre
et al., 2014). Although the overall stoichiometry is simple, the
chemistry is complex considering the catalytic reaction mech- Although rhodium-catalyzed carbonylation of methanol is
anism. highly selective, it suffers from some disadvantageous side
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12 3

reactions, as for example the water-gas shift reaction. Further ated by the density difference caused by the CO consumption.
methane appears by reaction of hydrogen with methanol: The reaction takes place in the riser, while the downcomer
ensures catalyst recycle and heat transfer through an external
CO + H2 O → CO2 + H2 (7) exchanger.

CH3 OH + H2 → CH4 + H2 O (8) 2.3. Process design

These side reactions may reduce the amount CO avail- A generic flowsheet is available regarding the technology
able for reaction down to a point where the catalyst integrity employed by the Monsanto process (Le Berre et al., 2014). This
is threatened. Another unwanted reaction is the formation employs a three columns separation sequence: the first two
of acetaldehyde. The reaction with hydrogen gives ethanol, handle the removal of light-ends and water, while the third the
which by carbonylation forms propionic acid as the main acetic acid purification. By this approach the species involved
heavy impurity. High boilers can be formed by aldol conden- in the catalyst complex, as described before, are fully recycled
sation reactions. to reactor. The same conceptual flowsheet is used by Cativa
H2 CO
and CT-ACETICA processes. However, no details about the pro-
CH3 -CHO→CH3 -CH2 -OH→CH3 -CH2 -COOH (9) cess design were given in terms of design and performance of
reactor and separators, as well as energy integration of units.
The alternative of using iridium as coordinating metal is This paper presents the first approach of analysis and
justified by its price, 17 times lower than that for rhodium design of such process, which is based on computer simu-
(Jones, 2000). Moreover, this brings a significant modification lation. The energy released by reaction, considering kinetic
in the reaction mechanism with strong impact on the process industrial data, reveals substantial capabilities for sustaining
performance. The oxidative addition of methyl iodide to the the energetic needs of the separation section. Two flowsheet
iridium centre is about 150 times faster compared to rhodium, alternatives are proposed, based on homogeneous and het-
and thus does not determine the reaction rate. The slowest erogeneous catalysis. The first is close to the actual industrial
step becomes the methyl migration to form the acyl complex. flowsheet, but proposes a quantitative assessment of the reac-
This step involves substitution of iodide centre with CO. Thus, tion section, the evaluation of the separations, as well as a first
the reaction rate dependence is totally different from that for energy integration level. This is also the reference case in the
the Rh catalyzed process: sustainability analysis. The second alternative employs only
two distillation columns and incorporates advanced energy
RCH3COOH = k × [Ir][CO]/[I− ] (10) integration methods, as heat pumping and on-site electricity
production by Organic Ranking Cycle (ORC). The comparison
The inverse dependence on the ionic iodide concentra- of alternatives by sustainability analysis reveals that the sec-
tion implies that removing it increases the reaction rate. This ond one demonstrates very low energy consumption, CO2
operation is done by promoters, such as iodide complexes emissions and process water usage. To put it in a nutshell, this
of zinc, cadmium, mercury, gallium and indium, as well as work reports the best process design performance to date, to
carbonyl-iodide complexes of tungsten, rhenium, ruthenium the best of our knowledge.
and osmium. The use of promoters can increase considerably
the reaction rate, up to 20 times by a molar ratio Ir/Ru of 5.
3. Process design and simulation based on
The Japanese company Chiyoda (www.chiyodacorp.com)
homogeneous catalyst
developed the CT-ACETICATM process that uses a heteroge-
neous catalyst in which the rhodium complex is immobilized
on a vinyl-pyridine resin. In this way the catalyst remains in The plant capacity corresponds to a production rate of
the reaction space and its integrity is preserved. Since the 25,000 kg/h or 416.3 kmol/h. The stoichiometric amounts of
amount of water drops under 2%, the energy for dehydrating raw materials are both 416.3 kmol/h methanol and CO, or
the acetic acid is drastically reduced. The same type of reac- 13,339 and 11,661 kg/h, respectively. A preliminary plant mate-
tion system is used in the new integrated process proposed rial balance based on industrial data for Rhodium catalyst (Le
here, which has the additional advantages of improved overall Berre et al., 2014), the yield estimates the yield as 98% for
efficiency due to effective energy integration and waste heat methanol, and 95% for CO. The lower value for CO may be
recovery by vapor compression and ORC. explained by secondary reactions leading to impurities and
gases as CO2 , H2 and CH4 .
2.2. Reactor technology The analysis and simulation work was done by employing
the process simulator Aspen Plus v9.0. The thermodynamic
Fig. 1 depicts the type of industrial reactors employed by modelling involves highly non-ideal mixtures, namely acetic
methanol carbonylation. Processes based on homogeneous acid/water, and polar supercritical gases, for which suitable
catalyst makes use of CSTR-type reactor with mechani- models are given in Appendix A.
cally stirring or gas injection device, working at pressures
of 30–40 bar and temperatures of 180–250 ◦ C (Jones, 2000), as 3.1. Reaction section
depicted in Fig. 1 left-hand. The reactor is provided with exter-
nal heat exchanger for cooling by steam generation (SG). A Since the reaction is fast it achieves practically full CO conver-
slurry gas–liquid tower reactor was developed for using het- sion with excellent yield. In the workable regime, the reaction
erogeneous catalyst (Yoneda et al., 2001), as illustrated by rate can be mastered by keeping constant the composition of
Fig. 1 right-hand. The gas is injected at the bottom through the catalytic complex, including co-catalyst and promoters,
an efficient distribution device to saturate the liquid phase since the reaction is zero-order with respect to the concen-
in CO. An external loop for liquid-phase circulation is cre- tration of reactive species. For the purpose of this paper a
4 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12

Fig. 1 – Chemical reactors for methanol carbonylation (left: homogeneous process, right: heterogeneous process).

practical kinetic approach is assumed by considering indus-


trial reaction rate data. As indicated in Table A-1 (see Appendix (12) Esterification of acetic acid with methanol:
A) by experiments 5–7, a reaction rate R of about 20 kmol/m3 /h
is assumed as achievable for a mixture composition with
methyl iodide as 6–10 wt%, methyl acetate 16–20 wt%, and CH3 -COOH + CH3 -OH = CH3 -COOCH3
water 4–5 wt%. This requirement is solved by a separate input
+ H2 O (molarextentofreaction2 kmol/h) (12)
stream that considers the above species recovered by recycles.
In addition, this approach offer a realistic description of the
(13) Formation of propionic acid:
reaction mixture involved further in separations (Sunley and
Watson, 2000).
A first design issue is recovery and recycling all the com-
CH3 -COOH + CH3 -OH = CH3 -CH2 -COOH
ponents of the catalyst recipe carried out to the separation
section, namely the volatile methyl iodide and methyl acetate, + H2 O (molarextentofreaction4 kmol/h) (13)
as well as controlling the amount of water in reactor. The use
of on-line FTIR spectrometry allows monitoring the amounts
of species mentioned before, as well as quantifying the pro- The reactant inputs are methanol (424.8 kmol/h
duction rate (Keit Company, 2019). A second issue is keeping = 13,612 kg/h) and CO (438.2 kmol/h = 12,274 kg/h), accord-
constant the temperature in reactor and managing the large ing to the preliminary material balance formulated before.
amount of heat released by reaction. This requirement is Methanol is preheated at 100 ◦ C, while the CO stream is
achieved by building efficient cooling and energy recovery pro- compressed at 35 bar in a four-stage compressor with inter-
cedures, as presented in this paper. mediate cooling at 50 ◦ C except the last adiabatic stage
Fig. 2 presents the flowsheet of the reaction section. This where the temperature rises at 152 ◦ C. In addition, a recycle
comprises the reactor RSTOIC modelling the reaction, fol- stream from separations is considered containing acetic
lowed by a flash unit FL-1 introduced for describing the acid (1000 kg/h), methyl acetate (2000 kg/h), methyl iodide
vapour–liquid equilibrium. The gas stream is condensed at (2000 kg/h), and water (1500 kg/h). The metal catalyst is not
5 ◦ C, the liquid being recycled to the reactor, while the gas introduced explicitly in simulation due to obvious modelling
leaves as vent stream. The liquid reaction mixture is split constraints.
further in a process stream and a liquid recycle stream that Effective reactor cooling is an important feature of the pro-
enters the cooling loop. The reactor effluent is depressurized cess design since a significant amount of reaction heat can
and vaporized in the FLASHER unit in view of separating the be recovered as useful energy for separations. Fig. 2 presents
catalyst, recycled as bottom stream to the reactor. a flowsheet that includes steam generator and cooler. This
The reactor modelling is based on stoichiometry. The main solution is used in industry, as described in a recent patent
reaction has a fractional conversion of 0.98. Secondary reac- (Zinobile, 2014). In view of maximizing the steam amount, the
tions describe the formation of by-products and impurities: reaction temperature is set at 220 ◦ C, as indicated in the same
patent. The reactor pressure should be over 35 bar to ensure
(11) Formation of gaseous impurities (relations (7) and (8)), liquid phase reaction.
lumped in the reaction: The goal in designing the reaction system is achieving
constant temperature inside the reactor by manipulating the
cooling capabilities. The parameters of the cooling loop were
tuned by an iterative procedure, split fraction of the SPLIT unit
CO + CH3 OH = CO2 + CH4
and return temperature of the COOLER unit, such to bring
(molarextentofreaction2 kmol/h) (11) the duty of the RSTOIC reactor close to zero. In addition, a
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12 5

Fig. 2 – Simulation flowsheet of the reaction section.

temperature of 150 ◦ C is assumed for generating LP steam of acetic acid (117.9 ◦ C) and propionic acid (141.2 ◦ C). A direct
3.5 bar. The results are split fraction of 0.242 and return CW sequence of three columns is appropriate, as shown in Fig. 3.
temperature of 56 ◦ C that delivers 6.2 MW LP steam for the The first column C-1 removes lights (methyl iodide, methyl
separation section. In addition, 6.2 MW duty of the COOLER acetate, methanol) and gaseous species as top distillate
remains available for reactant preheating and/or for energy (vapor and liquid), while the bottom recovers completely the
saving by other means, as electricity generation by Organic entrained catalyst as acetic acid solution. The acetic acid is
Rankine Cycle (ORC) developed in this study. taken over as vapor side-stream and fed to the column C-2.
The key stream leaving the reaction section is PROCESS, Here the dehydration of acetic acid takes place, the distil-
obtained as vapor from the unit FLASHER after pressure reduc- late recovering completely the remaining light species and
tion from 35 to 1.5 bar. This operation allows recovery most of most of the water amount. The top vapor from C-1 and C-
the Rh/Ir catalyst as bottom stream and recycling it to reactor. 2 enters the column C-4 where the lights are recovered by
The stream PROCESS sent further to the separation section absorption in methanol. A final cooling at 5 ◦ C recovers the
has a flowrate of 32,046 kg/h (25,622 acetic acid, 2296 methyl entrained methanol too. As mentioned, the condensate and
acetate, 1608 water, 1999 methyl iodide, 148 propionic acid, the top distillates are recycled to the chemical reactor to pre-
235 CO, 78 CO2 and 22 CH4 ). Supercritical components as CO, serve the composition of the catalyst complex. A makeup of
CO2 and CH4 , leave in the stream VENT, while the light species methyl iodide might be necessary to compensate losses in exit
are recycled to the reactor, after absorption in methanol and gaseous streams. The water inventory should also be balanced
deep cooling. such to ensure the prescribed water concentration in the reac-
The reactor effluent flowrate from simulation tor by means of on-line process monitoring. Finally, the bottom
is 132,760 kg/h (density of 655 kg/m3 ) with a volu- stream of C-2 is sent to purification in the column C-3, where
metric flow Qv = 202 m3 /h. From the mass balance high purity acetic acid is obtained as top product, while the
the amount of acetic acid formed by reaction is: heavy impurities are removed as bottom stream.
NA = NMeOH × conversion = 424.8 × 0.99 = 420.55 kmol/h. The The sizing of internals for all columns has been performed
reaction volume is VR = NA /R = 21.03 m3 . Hence the reaction interactively in Aspen Plus by selecting different internals,
time is tR = VR /Qv = 373 s. Taking a volume factor of 0.75 gives trays or packing. Column C-1 has 17 theoretical stages with the
a reactor of 28 m3 . For a cylindrical vessel with ellipsoidal vapor feed stream close to the reboiler, equipped with FLEX-
heads, the reactor volume is VR = /4 × D3 × (H/D + 0.166), S trays. The diameter is varied in three steps (1.3/1.8/2.2 m)
leading to a vessel of 2.3 m diameter and 3.2 m height. to accommodate large variations in vapor and liquid flows,
while aiming an efficient hydraulic regime between 70% and
80% from flooding. The top condenser works at 35 ◦ C and may
3.2. Separation section
be cooled by air or water or a combination.
The column C-2 deals with water removal. This is a dif-
The strategy of separations starts by examining the state and
ficult separation as the equilibrium curve shows a tangent
composition of the reactor-outlet mixture (Dimian et al., 2014).
pinch in the region of water distillation. The column needs 35
The stream PROCESS (which leaves the flasher as vapor at
theoretical stages to ensure water recovery in top over 97.5%.
1.6 bar and 125 ◦ C) contains acetic acid in large quantity, and
Simple sieve trays are suitable. Changing the tower diameter
much lower amounts of methyl acetate, methyl iodide and
is necessary, this time larger in the middle (1.9 m for the stages
water (there are also small amounts of dissolved gases and
2–20, 2.3 m for 21–26 and 1.9 m for 27–34), with the maximum
entrained Rh/Ir catalyst). Vapor pressure is the characteristic
flooding approach of 75%, 79%, and 78% respectively. The feed
property for determining the separation sequence of compo-
is located above the stage 27. The column partition in three
nents. Methyl iodide is the most volatile (nbp 42.4 ◦ C), followed
regions corresponds to different operation characteristics. The
by methyl acetate (56.9 ◦ C), methanol (64.7 ◦ C), water (100.0 ◦ C),
6 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12

Fig. 3 – Flowsheet of the acetic separation section by homogeneous catalyst.

water removal increases continuously from bottom to the top, 4. Eco-efficient process based on
while opposite happens for the acetic acid. In the middle sec- heterogeneous catalyst
tion a larger vapor flowrate due to the feed needs a larger cross
section to avoid flooding. If heterogeneous catalyst and advanced process integration
Finally, in the column C-3 the acetic acid is obtained as top methods are applied, an eco-efficient process with superior
product of purity over 99.7 wt%. The bottom stream removes sustainability features can be designed, as shown in Fig. 4.
the heavies, simulated here as propionic acid. The separa-
tion is difficult (30 trays), the relative volatility of acetic acid
to propionic acid being around 1.20. Accordingly, for acetic 4.1. Reaction system
acid recovery of 98.5% the separation needs a very large reflux
flowrate, about the same as the top product flowrate. Higher The CSTR reactor is replaced by a gas-liquid tower reactor
recovery can be obtained by stripping the bottom product, not (Fig. 1, right hand) in which the heterogeneous catalyst is recy-
considered here. cled. The FLASHER unit is no longer necessary. The PRODUCT
stream is cooled from 210 to 120 ◦ C in the exchanger COOL-P
and sent to the distillation column D-1 after pressure reduc-
3.3. Integration of reaction and separation sections tion. The parameters of the recycle loop are slightly modified
in view of optimal heat integration. Thus, the exit temperature
The exothermic reaction develops an important amount of from ST-GEN unit increases to 160 ◦ C. The effect is reducing the
energy, equivalent to 15.7 MW. The reactor cooling creates duty of ST-GEN to 5.4 MW (correlated with the reboiler of D-
opportunities for supplying energy to the other units. The unit 1) but increasing the duty of the unit COOL-R to 7 MW. Thus,
ST-GEN can ensure 6.2 MW as steam at 3.6 bar and 140 ◦ C. In the units COOL-P and COOL-R cumulate a considerable duty of
this way the heating requirements of the separation section 9 MW, which raises the question of valorization. The tempera-
(13.5 MW) can be reduced by 45%. Pinch Analysis indicates ture level well above 100 ◦ C suggests that an Organic Rankine
that supplementary energy saving is possible for methanol Cycle (ORC) can be applied for producing electricity from the
and recycle preheating (units H-1 and H-2), as well as for residual heat. Then the process/process heat exchange in the
sustaining the vaporization in the unit FLASHER, with a cumu- reaction section is limited only to reactants’ preheating in the
lated duty of 2.69 MW. A H-curve plot of the unit COOLER units H-1 and H-2, with a cumulated duty of 1.23 MW.
(duty 6.24 MW) shows that 2.75 MW is available on the inter-
val 150–110 ◦ C, enough for driving the above units ensuring 4.2. Separation system
a Tmin of at least 10 ◦ C. Accordingly, the duty of COOLER
diminishes with 2.69 MW from 6.24 to 3.5 MW. The saving As the heterogeneous catalyst remains in the reactor (hence
with respect to enthalpy of the liquid recycle is 72%. This not removed with the effluent as in the case for homogeneous
is substantial, but can be further increased to full recov- catalysts), the separation system presented in Fig. 4 consist of
ery. only two columns. The first one D-1 handles the removal in
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12 7

Fig. 4 – Eco-efficient process for acetic acid manufacturing.

stage, so that the total column’s height is 17.5 m. A notable


Table 1 – Sizing characteristic of the distillation columns
for acetic acid separation. effect is the pressure drop reduction from 0.32 bar with trays
to less than 0.1 bar with packing. Lower column diameter gives
Item Unit D-1 D-2
lower capital cost too.
Feed kg/h 29,800 25,400 The column D-2 performs the final purification of the raw
Distillate V/L kg/h 400/4000 25,000 acetic acid to 99.7 wt%. It has 35 plates (feed on 17) and works
Bottom 25,400 400 at high boilup ratio (over 100). Mellapack structured pack-
Stages/feed/ss – 35/17 35/17
ing is employed, which leads to a diameter of 1.5 m. A very
Top pressure bar 1.3 1.1
Pressure drop bar <0.1 <0.1
low pressure drop (below 0.1 bar) is convenient for heat pump
Top temp. ◦
C 35 121 implementation, as discussed next.

Bottom temp. C 128 127
Reflux kg/h 15,000 2500 4.3. Heat pump assisted distillation
Internals Mellapak Mellapak
Stages/diameter –/m 2–8/1.4 2–34/1.5
9–17/1.6
The reboiler duty of 5.4 MW of the column D-1 is covered by
Total height m 18–34/1.8 17.5 the steam generated in the reaction section. A simple criterion
17.5 based on Carnot efficiency can be effectively used to decide
Condenser duty MW −5.5 −4.9 whether a heat pump is worth exploring (Kiss and Infante
Reboiler duty MW 5.10 4.8 Ferreira, 2016): Q/W = Tc /(Tr − Tc ) > 10, where Q is the reboiler
duty, W the work provided, Tr the reboiler temperature and
top of water together with light species, while in the second Tc the condenser temperature. When the Q/W ratio exceeds
D-2 the purification of acetic acid takes place. Table 1 presents 10 then a heat pump should be considered. Referring to the
the sizing characteristics of the columns. purification column D-2, after implementing structured pack-
The design of D-1 aims high recovery of water to ensure ing, the temperature difference top-bottom is about 12 ◦ C and
delivering high-purity acetic acid from D-2. The condenser of consequently Q/W is above 30, which is clearly favourable.
D-1 is partial vapor-liquid. The liquid distillate contains the The condenser temperature (121 ◦ C at 1.1 bar) suggests that
water and most of the methyl acetate and methyl iodide, while water may be used as thermal agent for vapor compression.
the vapor distillate takes over the supercritical gases together Compressing LP steam is a widespread technique in industry.
with some methyl acetate and methyl iodide. A mass reflux In this case a heat pump can be installed around the col-
rate of 15,000 kg/h, combined with a tall column of 35 stages umn D-2 driving a steam Rankine cycle (Fig. 6, left-hand).
(feed on 17) results in recovery of 99% of water and 100% of Steam is generated at top by taking over the condensation
lighter components. Fig. 5 shows the composition profiles. The heat. Considering the vapor pressures of acetic water and
dewatering process is particularly difficult in the lower part of water, a mean temperature difference (MTD) of 20 ◦ C is cre-
the stripping region. ated between the hot and cold sides, while high heat transfer
In view of heat integration, the column pressure drop, and coefficients are reached. Then the steam is compressed to a
as consequence the bottom temperature should be at the low- pressure level enough for driving the reboiler. Implementing
est. This goal can be achieved by switching from trays to structured packing leads to a very low pressure drop of maxi-
structured packing. By implementing Mellapak from Sulzer mum 0.1 bar, and a bottom temperature of 126 ◦ C. Considering
the tower diameter drops significantly. The hydraulic calcu- a MTD of 24 ◦ C gives a target steam temperature of 150 ◦ C,
lations in Aspen Plus indicate that a shape of three diameters which corresponds to a pressure of 4.8 bar. The thermal bal-
is suitable: 1.4 m stages 1–8, 1.6 m 9–17, and 1.8 m 18–34, at ance indicates that a steam amount of 9500 kg/h is necessary
flooding of about 80%. A HETP of 0.5 m is selected for a real for running the heat pumping.
8 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12

Fig. 5 – Concentration profiles (as vapor and liquid molar fractions) along the acetic acid dehydration column.

Fig. 6 – Heat pumping by vapor compression (left) and power generation using an Organic Rankine Cycle (right).

Table 2 – Organic Rankine Cycles for waste heat recovery in the reaction section.
Fluid R-ORC recycle stream P-ORC product stream Expandera

Flowrate (G) Duty (Q) Power (W) Efficiency (Eff) G Q W Eff T-in T-out
◦ ◦
kg/s MW kW – kg/s MW kW – C C

R-600a 17 7.0 1360 0.194 4.8 2.0 386 0.193 125 35


R-1233ze 18.5 4.92 786 0.160 7.5 2.0 314 0.157 142 44
R-365fa 18.5 4.76 704 0.148 8 2.0 301 0.151 145 60

a
Note: Inlet pressure 30 bar, outlet pressure 2 bar.

The pressure ratio of 4.8 may be achieved by a three-stage the column working at a very large reflux rate. The bottom
compressor each with a compression ratio of 1.667, which lim- stream is very small compared with the distillate. As con-
its the outlet temperature below 200 ◦ C. For the first two stages sequence, the effect of disturbances is reduced. A practical
intermediate vapor de-superheating is necessary, at 120 and solution is using the heat pumping for covering 90–95% of
130 ◦ C respectively. This may be obtained by air cooling, or bet- the boiling duty by implementing an external heat exchanger,
ter by spraying liquid water. Assuming an overall efficiency of while a small reboiler handles disturbances (Kiss and Infante
0.72 the simulation in Aspen Plus gives a compressor power of Ferreira, 2016). Direct injection of LP steam from grid can be
1080 kW. Hence, a good coefficient of performance is obtained, used for start-up.
COP = Qr /Wc = 5.93/1.078 = 5.5. The design of the heat exchangers was done rigorously
The practical implementation of the heat pump might raise by using the EDR tool in Aspen Plus. The evaporator is of
controllability problems because of coupling the dynamics of horizontal BEM-TEMA type with organic vapor condensation
two units. A significant simplification comes from the fact that in tubes (2 passes) and water evaporation in shell. The heat
the duties of the hot and cold sources are practically identical, transfer coefficient (HTC) reaches a high value of 1800 W/m2 K
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12 9

because condensation and boiling, respectively, and combined


Table 3 – Sustainability metrics for acetic acid
with MTD of 18 ◦ C allows the exchange of 6 MW for an area of manufacturing.
195 m2 . The condenser that plays the role of column’s reboiler,
Metrics Material Energy Water CO2e
is BKM-TEMA kettle type. HTC of 1500 W/m2 K and MTD of
kg/kg MJ/kg kg/kg kg/kg
22 ◦ C gives an exchange area of 268 m2 .
A quick economic assessment can be done in term of credit Homogeneous 0.0354 2.16 0.82 0.17
of saving thermal energy against compressor investment and Eco-efficient process 0.0248 0.78 0.38 0.038
Improvement, % 29.94 63.90 53.85 77.6
operation costs. The evaluation depends on local conditions.
Considering a nominal reboiler duty is 6 MW and assuming 10
$/GJ for LP steam and 8000 h annual operation time gives the 1360 kW (for G = 16.7 kg/s), while 2 MW in the Product-ORC in
steam cost of (10 × 6 × 8000 × 3600/1000) = 1.73 M$. 386 kW (for G = 4.8 kg/s). Hence, 9 MW waste heat may be con-
On the other hand, an estimation of the compressor cost verted in 1746 kW electricity with an overall thermal efficiency
can be obtained by means of the correlation C = a + bSn , (Towler W/Q of 19.4%.
and Sinnott, 2012) in which a = 580,000, b = 20,000, n = 0.6, S The simulation shows that using safer hydrofluorocarbon
being the power in kW and the resulting cost in USD. The (HFC) fluids gives lower thermal efficiency, the operational
power of 1080 kW gives a compressor purchased cost of duty of the R-ORC dropping to about 5 MW. The reason is that
1.92 M$. An installation factor of 2 leads to a total installed cost a temperature-crossing of TQ profiles takes place due to a
of 3.84 M$. By assuming an electricity price of 0.07 $/kwh gives steeper slope of the liquid saturation (bubble points) curve
a compressor operating cost of (1080 × 8000 × 0.07) = 0.605 M$, in the T-S plot. However, a total heat of about 7 MW is con-
only 35% from the cost of thermal energy. verted in 1100 kW by R-1233ze (16% efficiency) and 1005 kW by
The estimation of profitability can be done based on Total R-365fa (15% efficiency).
Annual Cost (TAC). This considers that the capital of installed The conclusion of this section is that by steam genera-
compressor is paid back within 3 years, while the operating tion, heat pumping and waste heat recovery by ORC the large
cost each year. As result the TAC is (3.84/3 + 0.605) = 1.89 M$, amount of heat available from the exothermic reaction can be
to be compared with the annual cost of thermal energy of fully used for driving the separation section. The reboiler of the
1.73 M$. This alternative has the minimum utility require- dewatering column C-1 is covered by steam generation, while
ments. The only energy necessary for separations is required the purification column C-2 driven by heat pumping with
only by the VC system. The implementation is profitable if an electricity generated by applying ORC on wasted heat, oth-
advantageous price is obtained for the compression system, erwise rejected in the environment. The reduction in energy
which is the case since steam compression is widespread in requirements has a strong impact on eco-efficiency that will
process industries. In addition, electricity can be generated in be quantified by evaluating the sustainability measures.
situ, as shown in the next section.
5. Sustainability analysis
4.4. Organic Rankin Cycle (ORC) for waste heat
recovery A simple but efficient approach capturing the key feature of
sustainability in process design proposed by industrial experts
Heat curve calculation in Aspen Plus reveals that 7.5 MW waste (Schwartz et al., 2002) is based on six metrics: material inten-
heat is available over the interval 160/49 ◦ C from COOL-R, as sity, energy intensity, water consumption, toxic emissions,
well as 2 MW over 210/120 ◦ C from COOL-P. Waste heat can be pollutant emissions, and greenhouse gas (GHG) emissions.
advantageously recovered as electrical power by employing These reference output may be the unit of product, the rev-
an ORC system (Dimian et al., 2014). The principle is shown enues (sales), or the value-added (VA). Note that lower values
in Fig. 6 right-hand. The selection of working fluid may fol- mean better performance. Table 3 presents a comparative per-
low the recommendations of international agencies (IAE, 2014, formance of homogeneous catalyst process vs. eco-efficient
www.iea.org) or experts (Quoilin et al., 2013). heterogeneous process.
Since the hot source is available as sensible heat spread
over a temperature interval, the feasibility of the ORC method (A) Material intensity expresses the mass of wasted materials
require the top source temperature above the Tc of the work- per unit of output.
ing fluid. In addition, the TQ profiles inside the evaporator For the base-case process (homogeneous catalyst) the raw
must avoid temperature crossing and respect a Tmin of 5 ◦ C. materials input is methanol 13,612 kg/h + CO 12,273 kg/h,
The fluid pressure in evaporator and at expander inlet is set for an acetic acid production of 25,000 kg/h. The mate-
at 30 bar, located inside the optimal range of 0.8–0.9 from Pc rial intensity is: wasted materials = 0.0354 kg/kg. In the
(Castelli et al., 2019). The outlet expander pressure is set at eco-efficient process (heterogeneous catalyst) the yield of
2 bar. The feasibility was checked by simulation in Aspen Plus. methanol rises from 0.98 to 0.99 and the yield of CO from
Isentropic expansion is 0.85 and mechanical efficiencies 0.9. 0.95 to 0.96. Therefore, the material intensity drops to
The evaporator specification is supplying superheated vapor 0.0248 kg/kg, which is 30% reduction.
with 2 ◦ C. The objective of ORC is getting the highest work (B) Energy intensity represents the primary energy consumed
delivered. per unit of output.
Table 2 presents the results: fluid flowrate (G), source duty A difference should be observed between the energy devel-
(Q), generated power (W) and thermal efficiency (Eff), as well oped by utility and the primary energy spent for produce
as the inlet/outlet temperatures of expander for inlet/outlet it. An efficiency of 80% can be assumed for steam genera-
pressures of 30 and 2 bar. When using hydrocarbon fluids, tion by fuel combustion and 40% for electricity production
i-butane gives the best results. The i-pentane fluid failed in R- by modern power plants. Here is assumed that an electri-
ORC since it has a higher Tc than the source inlet-temperature. cal kWh needs three times primary (thermal) energy. Key
The whole duty of 7 MW in the Recycle-ORC is converted in elements of the energy balance are given in Figs. 3 and 4.
10 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12

In the base case, the separation section requires 13.5 MW per GJ. With respect to water make-up the EPA GHG cal-
steam from which 6.2 MW available as LP generated steam, culator (www.epa.gov/p2) estimates 3300 kWh consumed
hence only 7.3 MW must be supplied as LP and MP steam. for 1 million gallons for treatment and distribution, or
The electricity consumption regards 1450 kW power for 0.951 kWh/m3 , which corresponds to 0.672 kg CO2 e/m3 .
the CO compression, plus 150 MW for pumping, mixing For the base-case process the contributions are electricity
and process control, hence a total of 1600 kW or 4.8 MW (16,000 × 0.707 = 1311), steam (12.6 × 3.6 × 65 = 2948), and
thermal energy. To this one may add 0.26 MW required by CW (20.5 × 0.672 = 13.8), leading to 4273 kg/h CO2 e in total
cooling-water conditioning via cooling towers (≈2% from or 0.17 kg CO2 e/kg. For the eco-efficient process, the GHG
13 MW of C-1 and C-2 condensers) and 0.2 MW for refriger- contributions are: electricity (1014 × 0.707 = 717), steam
ation. The total used energy is then 12.6 MW. Taking an (1 × 3.6 × 65 = 234), water (9.4 × 0.672 = 6.4), hence 958 kg/h
uncertainty factor of 20%, the total load is 15 MW. The CO2 e in total or 0.038 kg CO2 e/kg (78% reduction in GHG
energy intensity is then 2.16 MJ/kg. emissions mainly due to less steam).
For the eco-efficient process, the reaction section gener- (E) Toxic emissions include toxics and hazardous materials for
ates 5.4 MW as LP steam, which covers the reboiler of D-1, the operating personnel.
while 9 MW of waste heat is converted into 1746 KW elec- The toxicity metrics characterize the threats on health
tricity. Column D-2 is fully driven by a heat pump that of species leaving in the process outputs. The main exit
needs 1160 kW electricity. In addition, 1450 kW is used points of in gas streams in the process are VENT (Fig. 2)
for CO compression and 150 kW for pumping and pro- and GAS (Fig. 3). Methyl iodide is the most toxic species,
cess control. Hence the electricity consumption is 2760 kW, very volatile and difficult to separate from large amount
but only 2760–1746 = 1014 kW should be imported, which of gas components, such as CO and CH4 . The absorption
is equivalent to 3.1 MW of primary energy. The process in methanol can ensure over 99.8% recovery. However, the
needs also 1 MW for feed preheating (units H-1 and H-2). exit gas stream may carry some methanol, which should
The cooling water conditioning referring to D-1 con- be recovered by deep cooling.
denser is 0.02 × 10 = 0.2 MW, while for refrigeration 0.2 MW (F) Pollutant emissions express the amount of pollutants pro-
is assumed. The total load for the reaction and separation duced per unit of output.
sections is 4.5 MW, which rises to 5.4 MW after apply- The process requires paying attention to Health, Safety
ing 20% uncertainty. It results that energy intensity is and Environment aspects. These are well known in the
0.78 MJ/kg, or 64% reduction compared with the base-case. industrial practice and not treated here. According to
(C) Water consumption expresses the amount of water used material balance got by simulation, the gaseous effluents
per unit of output. contain small amounts of unreacted CO, as well as CO2
Water consumption should include process water that and CH4 with traces of CH3 I, methyl acetate methanol and
should be treated because chemical contamination, losses other light organics. These are submitted to VOC treat-
in vent and purge streams, as well as make-up water for ment by combustion. About 1000 kg/h CO2 leaves as plant
the cooling system. In this case we consider negligible the emissions. For the eco-efficient process utility this amount
first two sources. For the cooling water (CW) conditioning is comparable with the CO2 emissions due to utilities.
via cooling tower is assumed. Turton et al. (2013) estimates Thus, the overall CO2e emissions may be estimated to
the water loss by evaporation as 1.83% from inflow for 2000 kg/h or 0.08 kg/kg product, which is a very low value.
10 ◦ C temperature range. For the base-case water cooling
is needed for the condensers of C-1 and C-2 since top tem-
Table 3 shows a summary of the sustainability metrics.
peratures of 35 ◦ C. The cumulated duty of 13 MW results
The eco-efficient heterogeneous process shows superior sus-
in 1118 m3 /h CW, and water make-up of 20.5 m3 /h. Then
tainability metrics compared to the homogeneous catalyst
the water consumption is 0.82 kg/kg product.
process. A better yield leads to notable reduction in material
In the eco-efficient process, water cooling is used only for
intensity. The complete use of the reaction heat for steam gen-
the condenser D-1 with a duty of 6 kW. Similarly, one gets
eration, heat pumping and ORC power generation, gives very
CW flow rate of 516 m3 /h and water make-up 9.4 m3 /h.
low energy intensity below 1 MJ/kg. The new process shows
The freshwater consumption drops to 0.38 kg/kg, or 54%
low CO2 emissions as well (below 50 g/kg), and the consump-
reduction.
tion of cooling water is also very low.
However, the freshwater consumption may be brought to
zero if air-cooling is adopted for taking out most of the
condenser duty (temperature range 94.7–35 ◦ C) but limit- 6. Conclusions
ing water-cooling in closed circuit (dry cooling) only at the
cold end. Acetic acid can be efficiently produced by a sustainable pro-
(D) Greenhouse gas (GHG) emissions expresses the total GHG cess based on methanol carbonylation using a heterogeneous
emitted per unit of output. rhodium/iridium catalyst. As the raw materials are available
GHG emissions cumulate the amounts of CO2 and equiv- from renewable feedstock (biogas or biomass) the acetic acid
alent GHG gases released, as well as involved in the produced by this method is a bio-based product.
production of heat and power, electricity, and cooling This study proposed innovative process designs in which
water. The GHG estimations depend on the conversion the reaction and separation sections are highly integrated.
factors, in turn depending on the mix used for energy Since the highly exothermic reaction delivers a large amount
production. Here we adopt the USA average value of of energy, the focus is achieving low-energy consumption, as
0.707 kg/h CO2 e per kwh (US Energy Information Agency, well as superior sustainability metrics for material and energy
2019). The CO2 conversion factor for steam is also variable. intensity, water consumption and GHG emissions.
A reasonable assumption is considering a primary energy A first innovative feature in this study is getting zero steam
ratio electricity/steam of 3 that leads to 65 kg/h CO2 e consumption for driving the separation section. In the het-
Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12 11

erogeneous process this reduces to two columns: dewatering Haynes, A., 2010. Catalytic methanol carbonylation. Adv. Catal.
and acetic acid purification. Since from VLE viewpoint the sep- 53, 1–45.
aration of acetic acid/water is difficult, both columns need Honeywell | Refrigerants, Take Advantage of Honeywell Ultra Low
GWP refrIgerants. www.honeywell-refrigerants.com/europe
considerable energy. The LP steam generated in the reaction
(Last accessed 26 November 2019).
section covers the reboiler duty of the dewatering column. Jin, F., Zhou, Z., Kishita, A., Enomoto, H., Kishida, H., Moriya, T.,
Furthermore, no steam is needed for the purification column 2007. A new hydrothermal process for producing acetic acid
by applying heat pumping by vapor compression using water from biomass waste. Chem. Eng. Res. Des. 85, 201–206.
as thermal agent. The principle is that the energy available Jones, J.H., 2000. The CativaTM process for the manufacture of
in condenser at about 120 ◦ C can be upgraded as steam for acetic acid. Platinum Met. Rev. 44, 94–105.
reboiler heating at 150 ◦ C and 4.8 bar by using vapor compres- Kiss, A.A., Infante Ferreira, C.A., 2016. Heat Pumps in Chemical
Process Industry. CRC-Press, Taylor & Francis Group, Boca
sion.
Raton, USA.
A second innovative feature is that the power for heat Kiss, A.A., Lange, J.P., Schuur, B., Brilman, D.W.F., van der Ham,
pumping is ensured by an Organic Rankine Cycle engine that A.G.J., Kersten, S.R.A., 2016. Separation technology – making a
upgrades waste heat available in the reaction section from difference in biorefineries. Biomass Bioenergy 95, 296–309.
the recycle and product coolers. Among several candidates Keit Company, UK. Solid-state FTIR for in-process monitoring of
the isobutane appears as the most suitable ORC fluid, as it acetic acid production.
www.keit.co.uk/Acetic-Acid-Keit-IRmadillo.pdf (Last accessed
gives full waste heat recovery. Additionally, a large part for
20 December 20 2019).
the energy for CO compression is covered too.
Le Berre, C., Serp, P., Kalck, P., Torrence, G.P., 2014. Acetic acid. In:
The overall result is a sustainable eco-efficient process with Ullmann’s Encyclopedia of Industrial Chemistry. Wiley-VCH,
very low energy intensity (below 1 MJ/kg) and low GHG emis- Germany.
sions (40 g CO2e/kg), as well as nearly zero water consumption. Li, F., Chen, B., Huang, Z., Lu, T., Yuanab, Y., Yuan, G., 2013.
Sustainable catalysts for methanol carbonylation. Green
Chem. 15, 1600–1607.
Conflict of interest
Marketwatch, Press Release: Global Acetic Acid Market to Reach
20.3 Million Tons by 2024. www.marketwatch.com (Last
None declared. accessed 05 December 2019).
Moulijn, J.A., Makkee, M., Van Diepen, A.E., 2013. Chemical
Acknowledgement Process Technology, 2nd ed. Wiley, Chechester, UK.
Nam, J.S., Kim, A.R., Kim, D.M., Chang, T.S., Kim, B.S., Bae, J.W.,
Financial support of the European Commission through the 2017. Novel heterogeneous Rh-incorporated graphitic-carbon
European Regional Development Fund and of the Roma- nitride for liquid-phase carbonylation of methanol to acetic
acid. Catal. Commun. 99, 141–145.
nian state budget, under the grant agreement POC P-37-449
Ni, Y., Shi, L., Liu, H., Zhang, W., Liu, Y., Zhu, W., Liu, Z., 2017. A
(acronym ASPiRE) is gratefully acknowledged. AAK thank- green route for methanol carbonylation. Catal. Sci. Technol. 7,
fully acknowledges the Royal Society Wolfson Research Merit 4818–4822.
Award (No. WM170003). Nowicki, L., Ledakowicz, S., Zarzycki, R., 1992. Kinetics of
rhodium-catalysed methanol carbonylation. Ind. Eng. Chem.
Appendix A. Supplementary data Res. 31, 2472–2475.
Pal, P., Nayak, J., 2017. Acetic acid production and purification:
critical review towards process intensification. Sep. Purif. Rev.
Supplementary material related to this article can be 46, 44–61.
found, in the online version, at https://doi.org/10.1016/ Quoilin, S., Van Den Broek, M., Declaye, S., Dewallef, P., Lemort, V.,
j.cherd.2020.04.013. 2013. Techno-economic survey of Organic Rankine Cycle
(ORC) systems. Renew. Sustain. Energy Rev. 22, 168–186.
Qian, Q., Zhang, J., Cui, M., Han, B., 2016. Synthesis of acetic acid
References via methanol hydrocarboxylation with CO2 and H2 . Nat.
Commun. 7, Article number: 11481.
Bidgoli, R.G., Naderifar, A., Mohammadrezaei, A.R., Nasr, M.R.J., Prausnitz, J.M., Anderson, T.F., Grens, E.A., Eckert, C.A., Hsieh, R.,
2012. Kinetic study, modeling and simulation of homogeneous O’Connell, J.P., 1980. Computer Calculation of Multicomponent
rhodium-catalyzed methanol carbonylation to acetic acid. Vapour–Liquid and Liquid–Liquid Equilibria. Prentice Hall.
Iran. J. Chem. Chem. Eng. 31, 57–73. Schwartz, J., Beloff, B., Beaver, E., 2002. Use sustainability metrics
Castelli, A.F., Elsido, C., Scaccabarozzi, R., Nord, L.O., Martelli, E., to guide decision-making. Chem. Eng. Prog. 98 (7),
2019. Optimization of Organic Rankine Cycles for waste heat 58–63.
recovery from aluminum production plants. Front. Energy Sunley, G.J., Watson, D.J., 2000. High productivity methanol
Res., http://dx.doi.org/10.3389/fenrg.2019.00044. carbonylation catalysis using iridium, CativaTM process for
Dimian, A.C., 2007. Renewable raw materials: chance and the manufacture of acetic acid. Catal. Today 58, 293–307.
challenge for computer-aided process engineering. Comput. Towler, G., Sinnott, R., 2012. Chemical Engineering Design:
Aided Chem. Eng. 24, 309–318. Principles, Practice and Economics of Plant and Process
Dimian, A.C., Bildea, C.S., 2008. Vinyl acetate monomer process. Design. Butterworth-Heinemann, USA.
In: Chemical Process Design: Computer-Aided Case Studies, Turton, R., Bailie, R.C., Whiting, W.B., Shaeiwitz, J.A.,
Wiley-VCH. Bhattacharyya, D., 2013. Analysis, Synthesis and Design of
Dimian, A.C., Bildea, C.S., Kiss, A.A., 2014. Integrated Design and Chemical Processes, 4th ed. Prentice Hall, USA.
Simulation of Chemical Processes, 2nd ed. Elsevier, US Energy Information Agency, Energy and environment: Green
Amsterdam. Gas Explained. www.eia.gov (Last accessed 26 November
Dimian, A.C., Bildea, C.S., Kiss, A.A., 2019. Applications in Design 2019).
and Simulation of Sustainable Chemical Processes. Elsevier, Vidra, A., Németh, Á., 2018. Bio-produced acetic acid: a review.
Amsterdam. Period. Polytechn. Chem. Eng. 62 (3), 245–256.
12 Chemical Engineering Research and Design 1 5 9 ( 2 0 2 0 ) 1–12

Vitasari, C.R., Meindersma, G.W., de Haan, A.B., 2015. Conceptual Yoneda, N., Takeshi, M.J., Spehlmann, B., 1999. The Chiyoda
process design of an integrated bio-based acetic acid, AceticaTM process: a novel acetic acid technologies. Surf. Sci.
glycolaldehyde, and acetol production in a pyrolysis oil-based Catal. 121, 93–98.
biorefinery. Chem. Eng. Res. Des. 95, 133–143. Zinobile, R.J., 2014. Production of acetic acid with an increased
Yoneda, N., Kusano, S., Yasui, M., Pujado, P., Wilcher, S., 2001. production rate. US Patent 8, 637,698 B2, Celanese Corp.
Recent advances in processes and catalysts for the production
of acetic acid. Appl. Catal. A: Gen. 221, 253–265.

You might also like