You are on page 1of 20

DE GRUYTER International Journal of Chemical Reactor Engineering.

2019; 20180122

Ali Asghar Jafari1 / Somayeh Tourani1 / Farhad Khorasheh2

Simulation of Methanol Carbonylation Reactor


in Acetic Acid Production Plant: Selection of
an Appropriate Correlation for Mass Transfer
Coefficients
1 Department of Chemical Engineering, Mahshahr branch, Islamic Azad University, Mahshahr, Iran, E-mail:

s.tourani@mhriau.ac.ir. https://orcid.org/0000-0003-3462-1545.
2 Department of Chemical and Petroleum Engineering, Sharif University of Technology, Tehran, Iran, E-mail:

khorashe@sharif.ir. https://orcid.org/0000-0002-8477-4753.

Abstract:
This paper deals with mathematical modeling and simulation of methanol carbonylation reactor in acetic acid
production plant that consisted of a continuous stirred tank reactor (CSTR), a flash drum, a Joule-Thomson
valve, and a condenser. The model was based on material and energy balances that considered liquid-gas mass
transfer, thermodynamics, and reactor hydrodynamics. The most important aspect of the model was the selec-
tion of an appropriate correlation for prediction of mass transfer coefficient. Several correlations were examined
and comparison of the model results with plant data indicated that the correlation reported by Lemoine was
most appropriate. The simulation results were found to be in good agreement with plant data and the effects
of various operating parameters on the performance of the model such as temperature of reactor from 186 °C
to 183.6 °C, CH3 I concentration from 3.11 mol% to 2.21 mol% and agitator rotation speed from 62 rpm to 112
rpm were investigated.
Keywords: Acetic acid plant, CSTR, Hydrodynamic, Mass transfer coefficient, Simulation
DOI: 10.1515/ijcre-2018-0122

1 Introduction
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Acetic acid (AA) is widely used as a raw material for the production of vinyl acetate monomer (VAM) and acetic
anhydride. It is also a solvent used in purified terephthalic acid (PTA) production. Among the various routes for
synthesis of AA, the most important route for large-scale production is homogeneous methanol carbonylation.
One of the most common industrial processes for synthesis of AA is the Monsanto process in which methanol
is carbonylated by carbon monoxide (CO) at reaction temperature of 150-200 °C and reactor pressure of 30-60
bar. The reaction takes place in a continuous stirred tank reactor (CSTR) using a soluble rhodium catalyst with
CH3 I and HI as promoter (Golhosseini Bidgoli et al. 2012). In this study, a plant wide model for the methanol
carbonylation process to produce acetic acid was developed using Visual Basic.NET 2008 software. Following
the selection of a correlation for prediction of mass transfer coefficient, the simulation of the AA unit was per-
formed using the predicted liquid-gas mass transfer coefficient, thermodynamics, reactor hydrodynamics, and
the appropriate material and energy balance equations. The most important aspect for a successful simulation
of the AA plant is the accurate prediction of CO concentrations in the liquid phase using the appropriate val-
ues of the mass transfer coefficient (kL ). The value of kL is affected by many factors including the liquid phase
composition and properties, as well as the geometrical and operating characteristics of the reactor including
the type of impeller, the geometry of the reactor, and the agitation speed (Moutafchieva et al. 2013). A number
of algorithms for calculation of kL and modeling of agitated reactors have been previously reported in the lit-
erature (Dake, Kolhe, and Chaudhari 1984; Hjortkjaer and Jørgensen 1978; Tekie et al. 1997; Lu, Wu, and Chou
1999; Lemoine and Morsi 2005; Lemoine 2005; Soriano 2005; Kiełbus-Rąpała and Karcz 2011). Selected corre-
lations for prediction of kL a are presented in Table 1. In the present investigation, the method proposed by
Lemoine (Lemoine 2005) was found to be most appropriate for calculation of kL a and is subsequently used for

Somayeh Tourani is the corresponding author.


© 2019 Walter de Gruyter GmbH, Berlin/Boston.

Brought to you by | University of Western Ontario


Authenticated 1
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

the modeling and simulation of the AA plant to investigate the effects of various operating parameters on the
model performance.

Table 1: Selected correlations for prediction of kL a in GS-CSTR.


No Reference Correlation
𝑛
1 [23] 𝑘𝐿 𝑎 = 𝑐(𝑁 3 𝑑2𝑇 ) 𝑈𝐺
𝑚

2 [24] 𝑘𝐿 𝑎 =
𝑃∗ 0.74
0.36
3.89 × 10−3 ( 𝑉𝐺 ) 𝑈𝐺
𝐿
𝑛
𝑁 3.15 𝑑5.35 𝑚
3 [25] 𝑘𝐿 𝑎 = 𝑐( 𝑇
1.41 ) 𝑈𝐺
𝑉𝐿
4 [26] 𝑘𝐿 𝑎 =
𝑃∗ 1/3 0.8
𝑃∗
3.92 × 10−6 ( 𝜌𝐺𝐿 ) ( 𝑉 𝐺𝜌 )
𝐿 𝐿
∗ 0.67
0.56
5 [27] 𝑘𝐿 𝑎 = 1.06( 𝑉
𝑃
) 𝑈𝐺
𝐿
3
6 [4] 𝑘𝐿 𝑎 = 2.564 × 10 ×
𝐷0.5 0.575
𝐴𝐵 𝜀𝐺 𝑃 ∗ 𝛿 𝜂
𝜌0.06 𝑑0.402
(𝑉 𝐿
) 𝑈𝐺𝐺𝑆𝑅 exp (−2.402𝑋𝑊 )
𝐺 𝑆
7 [5] 𝑘𝐿 𝑎 =
0.374
𝑃∗
0.081 ( 𝑉 ) 𝑈𝐺 0.636
𝐿

2 Modeling of AA plant
The schematic of the AA plant is illustrated in Figure 1. CO is injected to the reactor through a sparger and
is transferred to the liquid phase containing methanol where AA is produced by the carbonylation reaction.
The liquid products are transferred to a flash drum via a JT-valve where two phases are separated; a gas phase
is transferred to the distillation section and a liquid phase containing the Rh catalyst that is returned to the
reactor. The mathematical model to describe the AA plant consisted of material and energy balances that in-
cluded the following considerations: gas-liquid mass transfer, reaction kinetics, reactor hydrodynamics, and
thermodynamic calculations for the JT-valve and flash drum.
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Figure 1: Schematic of the acetic acid plant.

2.1 Gas-liquid mass transfer

Gas sparged continuous stirred tank reactors (GS-CSTR) are commonly used in chemical processing industries
as they provide adequate mass transfer between gas and liquid phases (Furukawa et al. 2012). Injection of gas
through a sparger creates bubbles that collide with the agitator blades. Some bubbles will be broken, while oth-
ers are coalesced or dissolved in the liquid resulting in a distribution of bubbles with various diameters. Figure
2 provides a schematic diagram for the algorithm used for calculation of volumetric mass transfer coefficient,
kL a, in GS-CSTR using reactor geometry, physicochemical properties of the reacting media, and operating vari-

Brought to you by | University of Western Ontario


2 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

ables including the flow of injected gas, QGI , input power per unit volume of liquid, P*/VL , gas holdup, εG , and
Sauter mean bubble diameter, dS (Lemoine 2005).

Figure 2: Algorithm for prediction of volumetric mass transfer coefficient (kL ) in GS-CSTR.

In GS-CSTR, the concentration of disolved CO in the liquid phase is calculated by eq. (1) where the calcula-
tion of kL a involves a methodology outlined by eqs. (2) to (34) (Lemoine 2005; Kiełbus-Rąpała and Karcz 2011;
Furukawa et al. 2012):

𝐶∗ − 𝐶𝐿,0
ln ( ) = 𝑘𝐿 𝑎 𝑡 (1)
𝐶∗ − 𝐶𝐿,𝑡

Where, C* is equilibrium gas solubility, t is time, CL,0 is gas concentration in the liquid at t=0 and CL,t is gas
concentration at any time. The following equation is used for calculation of gas holdup in GS-CSTR:

∗ 𝛼
𝑃𝐺𝑆𝑅 𝛽
𝜀𝐺 = 9.620 × 10−3 × ( ) 𝑈𝐺 × exp (−0.216) (2)
𝑉𝐿

where α and β are defined as follows:

𝛼 = 0.190 ∗ 𝑑𝑇−0.179 𝑁 0.043 𝜇−0.228 𝜎𝐿0.261 𝜌−0.011 (3)


Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

𝐿 𝐺

𝛽 = 1.86 × 1012 ∗ 𝑑𝑖𝑚𝑝


−0.087 𝐻 −0.279 𝑁 0.063 𝜌−4.27 𝜇−0.464 𝜎 1.38
𝐿 𝐿 𝐿 𝐿 (4)

where N is the mixing speed, dT is the reactor diameter, HL is the height of liquid, μL is liquid viscosity, ρL is
liquid density, ρG is gas density, σL is surface tension, dimp is the impeller diameter, and UG is superficial gas
velocity given by (Perry and Green 1997):

𝑄𝐺𝐼
𝑈𝐺 = 4 2
(5)
𝜋(𝑑𝑇 )

In GS-CSTRs, the power input per unit volume is calculated by the following equation (Lemoine 2005 and
Loiseau, Midoux, and Charpenntier 1977):

8 4 𝐵

𝑃𝐺𝑆𝑅 𝐴 ⎛ 𝑁𝑃 𝑑𝑖𝑚𝑝 𝜌𝐿 𝑁 ⎞
= .⎜ ⎟ + 𝑈𝐺 𝜌𝐿 𝑔 (6)
𝑉𝐿 𝑉𝐿 ⎝ 𝑄0.56
𝐺 ⎠
where A and B for foaming and non-foaming system respectively are equal to: A=0.83, B=0.45 and A=0.69,
B=0.45 (Loiseau, Midoux, and Charpenntier 1977).
The GS-CSTR investigated in this study was equipped with a dual impeller agitator. The lower impeller was a
Rushton turbine and the upper one was propeller type. The properties of these impellers are given in Table 2. In
order to calculate the power number, NP , eqs. (7) to (26) were used for the two types of impeller, Rushton turbine
and Propeller, in the following three-steps procedure (Furukawa et al. 2012). It should be noted that the power

Brought to you by | University of Western Ontario


Authenticated 3
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

consumption of the impeller in a mixing vessel is increased by insertion of baffle plates and reaches a maximum
value at a certain level of baffle conditions where the tangential flow is suppressed and the cylindrically rotating
zone disappears. The fully baffled condition is defined as that at which the power consumption of an impeller
reaches a maximum value by insertion of baffle plates on the mixing vessel wall (Sano and Usui 1987).

Table 2: Properties of agitator impellers.


Properties Lower Upper
impeller impeller
Type Rushton propeller
turbine
Diameter (m) 1.143 1.702
Height (m) 0.229 0.336
Distance to 1.057 3.538
bottom of
reactor (m)
Angle (Deg) 90 45
Numbers of 6 3
blades
Rotation speed 87 87
(rpm)

Step 1: Unbaffled condition

1.2𝜋 4 𝛽2 ⎤
𝑁𝑃,𝑢𝑛𝑏𝑎𝑓 𝑓 𝑙𝑒 = ⎡
⎢ 3 2 ⎥𝑓 (7)
⎣ 8𝑑𝑖𝑚𝑝 /𝑑𝑇 𝐻𝐿 ⎦

−1 1/𝑚 𝑚
𝐶 𝐶𝑡𝑟 𝑓 ∼
𝑓 = 𝐿 + 𝐶𝑡 ⎡
⎢( + 𝑅𝑒𝐺 ) +( ) ⎤
⎥ (8)
𝑅𝑒𝐺 ⎣ 𝑅𝑒 𝐺 𝐶𝑡 ⎦

𝜋𝜂 𝑙𝑛 (𝑑𝑇 /𝑑𝑖𝑚𝑝 ) ⎞
𝑅𝑒𝐺 = ⎛

⎜ ⎟
⎟ 𝑅𝑒𝑑 (9)
4𝑑/𝛽𝑑𝑇
⎝ ⎠
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

2 .𝜌
𝑁. 𝑑𝑖𝑚𝑝
𝑅𝑒𝑑 = (10)
𝜇

2
𝑑𝑖𝑚𝑝 ⎛ 𝑑𝑖𝑚𝑝 ⎞ 1/3
𝐶𝐿 = 0.215 𝜂𝑛𝑝 ⎛
⎜ ⎞
⎟⎜ ⎜ 2 ⎞
⎜1 − ⎛ ⎟ + 1.83 (𝑏𝑠𝑖𝑛 𝜃/𝐻) (𝑛𝑝 /2𝑠𝑖𝑛 𝜃)
⎟⎟ (11)
⎝ 𝐻𝐿 ⎠ ⎝ ⎝ 𝑑𝑇 ⎠ ⎠

−7.8 −1/7.8
𝐶𝑡 = [(1.96 𝑋 1.19 ) + 0.25−7.8 ] (for Rushton turbine) (12)

−7.8 −1/7.8
𝐶𝑡 = [(3 𝑋 1.5 ) + 0.25−7.8 ] (for propeller) (13)

−7.8 −1/7.8
𝑚 = [(0.71𝑋 0.373 ) + 0.333−7.8 ] (for Rushton turbine) (14)

−7.8 −1/7.8
𝑚 = [(0.8𝑋 0.373 ) + 0.333−7.8 ] (for propeller) (15)

−3.24
𝑑𝑖𝑚𝑝 −1.18
𝐶𝑡𝑟 = 23.8⎛
⎜ ⎞
⎟ (𝑏𝑠𝑖𝑛 𝜃/𝑑𝑇 ) 𝑋 −0.74 (16)
⎝ 𝑑𝑇 ⎠
Brought to you by | University of Western Ontario
4 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

𝑓 ∼= 0.0151 (𝑑𝑖𝑚𝑝 /𝑑𝑇 ) 𝐶𝑡 0.308 (17)

𝑋 = 𝛾𝑛𝑝 0.7 𝑏 𝑠𝑖𝑛1. 6 𝜃/𝐻𝐿 (18)

2 ln (𝑑𝑇 /𝑑𝑖𝑚𝑝 )
𝛽= (19)
𝑑𝑇 /𝑑𝑖𝑚𝑝 − 𝑑𝑖𝑚𝑝 /𝑑𝑇

1/3
⎡ 𝜂𝑙𝑛 (𝑑𝑇 /𝑑𝑖𝑚𝑝 ) ⎤
𝛾=⎢
⎢ ⎥
5 ⎥
(20)
⎣ (𝛽𝑑 𝑇 /𝑑𝑖𝑚𝑝 ) ⎦

0.611
0.157 + [𝑛𝑝 𝑙𝑛 (𝑑𝑇 /𝑑𝑖𝑚𝑝 )]
𝜂 = 0.711 2
(21)
𝑛0.52
𝑝 [1 − (𝑑𝑖𝑚𝑝 /𝑑𝑇 ) ]

where 𝛽, b, m, ReG , Red , f, θ, np , N, Ct are pitch of the impeller blade, height of impeller blade, turbulent term,
modified Reynolds number, impeller Reynolds number, friction factor, angle of impeller blade, number of im-
peller blades, mixing speed, and turbulent term, respectively.
Step 2: Fully baffled condition
For Rushton turbine:

1.3
𝑁𝑃 𝑚𝑎𝑥 = 10(𝑛𝑝 0.7 𝑏/𝑑𝑖𝑚𝑝 ) (𝑛𝑝 0.7 𝑏/𝑑𝑖𝑚𝑝 ) ≤ 0.54
𝑁𝑃 𝑚𝑎𝑥 = 8.3 (𝑛𝑝 0.7 𝑏/𝑑𝑖𝑚𝑝 ) 0.54 < (𝑛𝑝 0.7 𝑏/𝑑𝑖𝑚𝑝 ) ≤ 1.6 (22)
0.6
𝑁𝑃 𝑚𝑎𝑥 = 10(𝑛𝑝 0.7 𝑏/𝑑𝑖𝑚𝑝 ) 1.6 < (𝑛𝑝 0.7 𝑏/𝑑𝑖𝑚𝑝 )

For propeller:

1.7
𝑁𝑃𝑚𝑎𝑥 = 6.5(𝑛0.7 1.6
𝑝 𝑏 𝑠𝑖𝑛 𝜃/𝑑𝑖𝑚𝑝 ) (23)

Step 3: Baffled condition


Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

−1/3
𝑁𝑃 = [(1 + 𝑥−3 ) ] 𝑁𝑃𝑚𝑎𝑥 (24)

(𝐵𝑤 /𝑑𝑇 ) 𝑛0.8


𝐵
𝑁𝑃, 𝑢𝑛𝑏𝑎𝑓 𝑓 𝑙𝑒
𝑥 = 4.5 0.2
+ (for Rushton turbine) (25)
𝑁𝑃𝑚𝑎𝑥 𝑁𝑃𝑚𝑎𝑥

(𝐵𝑤 /𝑑𝑇 ) 𝑛0.8


𝐵
𝑁𝑃, 𝑢𝑛𝑏𝑎𝑓 𝑓 𝑙𝑒
𝑥 = 4.5 0.72
+ (for propeller) (26)
(2𝜃/𝜋) 0.2
𝑁𝑃𝑚𝑎𝑥 𝑁𝑃𝑚𝑎𝑥

Where Bw and nB are the width of baffle plate and the number of baffles, respectively.
Sauter mean bubble diameter, dS , in GS-CSTR can be calculated by eqs. (27) to (29) (Lemoine 2005):

𝛾 𝜆
𝑑𝑆 = 9.380 × 10−3 × 𝑈𝐺 𝜀𝐺 (27)

−0.878 𝑑0.351 𝑁 0.563 𝐻 0.185


𝛾 = 1.380 × 10−2 × 𝑑𝑖𝑚𝑝 (28)
𝑇 𝐿

𝜆 = 1.3 × 10−20 × 𝜌7.49


𝐿 𝜎𝐿
−0.24 𝜌−0.196 exp (−8.470𝑋 )
𝐺 𝑊 (29)

There are a number of correlations to estimate the volumetric mass transfer coefficient, kL a, in a GS-CSTR. Two
methods are used in this investigation. Method 1 (Lemoine 2005) utilizes eqs. (30) to (33) and method 2 is given
by eq. (34). (Kiełbus-Rąpała and Karcz 2011)

Brought to you by | University of Western Ontario


Authenticated 5
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

𝐷0.5 0.575
𝐴𝐵 𝜀𝐺 𝑃∗ 𝜂
𝛿
𝑘𝐿 𝑎 = 2.564 × 103 × ( ) 𝑈𝐺𝐺𝑆𝑅 exp (−2.402) (30)
𝜌0.06
𝐺 𝑑𝑆
0.402 𝑉𝐿

𝛿 = 4.664 × 10−4 × 𝑑𝑇0.124 𝑁 0.593 𝜇−0.769


𝐿 (31)

0.363 𝑁 0.967 𝜌−0.47 𝜇−0.884 𝐻 −1.44


𝜂 = 9.475 × 10−5 × 𝑑𝑖𝑚𝑝 (32)
𝐿 𝐿 𝐿

Where DAB is diffusivity of gas in liquid and is calculated as follows (Poling et al. 2001):

𝑀𝐵0.5 𝑇
𝐷𝐴𝐵 = 1.1728 (10−16 ) 0.6
(33)
𝜇 𝐵 𝑉𝐴

In eq. (33), VA is the molar volume of the diffusing gas at its normal boiling point and MB is the molecular
weight of the liquid.

0.374
𝑃∗
𝑘𝐿 𝑎 = 0.081 ( ) 𝑈𝐺 0.636 (34)
𝑉𝐿

The calculated mass transfer coefficient and the rate of mass transfer in the agitated reactor using the above two
methods are compared with actual plant data in Table 3 indicating that the first method is superior as it takes
into account a number of system parameters including gas holdup, Sauter mean diameter, and diffusivity.

Table 3: Comparison of two methods for calculating the rate of mass transfer.
Plant MeOH CO flow kLa kLa mass Deviation mass Deviation mass
capacity flow (Nm3/h) (method (method transfer (%) transfer (%) transfer
(%) (kg/h) 1) 2) (method (method (Ac-
1)(kg/h) 2)(kg/h) tual)(kg/h)
68.6 7000 6206 0.062349 0.043726 7099 2.335 5007 −27.822 6937
70.6 7200 6335 0.061923 0.044299 7223 1.561 5177 −27.208 7112
72.5 7400 6468 0.061865 0.044888 7432 1.990 5438 −25.374 7287
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

74.5 7600 6600 0.061648 0.045468 7558 1.287 5620 −24.685 7462
76.5 7800 6732 0.0614201 0.046042 7726 1.165 5836 −23.583 7637
78.4 8000 6864 0.061135 0.046612 7894 1.050 6062 −22.401 7812
80.4 8200 6996 0.060938 0.047182 8062 0.939 6285 −21.310 7987
82.4 8400 7128 0.060449 0.047741 8230 0.833 6541 −19.860 8162
84.3 8600 7261 0.060194 0.048305 8398 0.732 6780 −18.676 8337
86.3 8800 7392 0.059697 0.048854 8566 0.634 7048 −17.199 8512
88.2 9000 7524 0.059435 0.049409 8735 0.553 7298 −15.989 8687

% deviation = (calculated mass transfer – actual mass transfer)/actual mass transfer ×100

2.2 Thermodynamics

An equation of state (EOS) must be used to determine the properties of different streams in the AA unit that
contain up to eight components including AA, water, CH3 I, HI, CO, CO2 , H2 and N2 . The Peng-Robinson-
Stryjek-Vera (PRSV) EOS (Li et al. 2005; Šerbanovic, Djordjevic, and Grozdanic 1994; Stryjek and Vera 1986) was
used in this investigation for (i) to calculate C* in the reactor, (ii) to perform flash calculations of JT-valve, and
(iii) to determine gas and liquid composition at the condenser output. The PRSV EOS is as follows:

𝑅𝑇 𝑎𝛼
𝑃= − (35)
𝑉 − 𝑏 𝑉 2 + 2𝑏𝑉 − 𝑏2
Where

Brought to you by | University of Western Ontario


6 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

0.457235𝑅2 𝑇𝑐 2
𝑎= (36)
𝑃𝑐

0.077796𝑅𝑇𝑐
𝑏= (37)
𝑃𝑐

2
𝛼 = [1 + 𝑘 ∗ (1 − 𝑇𝑟 0.5 )] (38)

𝑘 = 𝑘0 + 𝑘1 (1 + 𝑇𝑟 0.5 ) (0.7 − 𝑇𝑟 ) (39)

𝑘0 = 0.378893 + 1.44897153𝜔 − 0.17131848 𝜔2 + 0.0196554 𝜔3 (40)

where, k1 is an adjustable parameter specific for each component. Residual enthalpy is calculated by eq. (41)
(Gmehling and Kleiber 2012; Neto and Barbosa JR 2012)

𝐻𝑅 𝐴 𝑇 𝑑𝑎 𝑍 + (20.5 + 1) 𝐵
= 𝑍 − 1 − 1.5 [1 − ] 𝑙𝑛 [ ] (41)
𝑅𝑇 2 𝐵 𝑎 𝑑𝑇 𝑍 − (20.5 − 1) 𝐵

where A and B are defined as follows:

P𝛼 𝑃𝑏
𝐴= 2
, 𝐵= (42)
(RT) 𝑅𝑇

2.2.1 Vapour Liquid Equilibrium (VLE)

VLE calculation to obtain gas solubility in the liquid, C* in eq. (1), reactor temperature, pressure, and composi-
tion was performed with PRSV EOS using the algorithm given in Figure 3 (Chang 1991).
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Brought to you by | University of Western Ontario


Authenticated 7
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

Figure 3: VLE calculation algorithm for prediction of gas solubility in liquid.

2.2.2 Flash calculations


Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

There is a JT-valve at the reactor outlet before the flash drum. This valve drops the pressure from 28 barg to
1.3 barg. This expansion results in a temperature decrease of approximately 125 °C along with the separation
of liquid and vapor phases. By specifying the pressure downstream from the valve, a flash calculation would
determine the temperature as well as the amount and composition of the vapor phase during the isenthalpic
expansion in the JT-valve. Figure 4 shows the adiabatic flash calculation algorithm (Gmehling and Kleiber 2012)
where HIN is the enthalpy of the feed, q0 is an initial estimate of molvapour /molfeed ratio in the outlet stream,
and T0 is an initial estimate of the outlet temperature.

Brought to you by | University of Western Ontario


8 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

Figure 4: Flash calculation algorithm.

2.2.3 Condenser calculation

The vapor phase inside the reactor contains N2 , CO2 , H2, and vapors of AA, water, CH3 I, and HI. The presence of
these components would decrease the partial pressure of CO in the vapor phase. To stabilize the partial pressure
of CO, other gases are removed from the vapor phase by passing the vapor stream through a condenser where
AA, water, CH3 I and HI are condensed and returned to the reactor. The outlet temperature of the condenser is
controlled by the flow of cooling water. The algorithm for condenser calculations (Gmehling and Kleiber 2012)
to determine the condenser cooling power requirement as well as return condensate flow and compositions is
presented in Figure 5.
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Figure 5: Algorithm for condenser calculations.

Brought to you by | University of Western Ontario


Authenticated 9
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

2.3 Kinetics

Two reactions proceed in the reactor include carbonylation of methanol to AA and water-gas shift reaction.
Both of these reactions take place in the presence of a rhodium complex as catalyst (Khimtekhnologiya 2004).

CH3 OH + CO → CH3 COOH (43)

H2 O + CO → H2 + CO2 (44)

The soluble catalyst system comprises of the rhodium complex and methyl iodide-hydrogen iodide as promoter
allowing the reactions to proceed at a reasonable rate (Dake, Kolhe, and Chaudhari 1984; Khimtekhnologiya
2004). This reaction is exothermic and the heat released at reactor conditions is about 2342 kJ/kg of AA pro-
duced (Khimtekhnologiya 2004; Hjortkjaer and Jensen 1976).
In this study the following rate is used for the reaction (43) (Khimtekhnologiya 2004):

𝑅1 = 1795 exp (−7830 (1/𝑇 − 1/443)) [I]1.05 [Rh]0.99 (45)

where, R1 is the theoretical reaction rate of the carbonylation reaction expressed in gram-moles of methanol re-
acted per hour per dm3 liquid volume, [I] is total iodine concentration used as promoter, and [Rh] is the concen-
tration of the rhodium catalyst. According to the above rate expression, the rate of carbonylation reaction is inde-
pendent of the concentrations of CH3 OH and CO as long as both are available for reaction (Khimtekhnologiya
2004; Hjortkjaer and Jensen 1976). CO is introduced as a gas stream and for the reaction to proceed in the liquid
phase, CO must be transferred to the liquid phase. Mass transfer should be sufficient so that the rate of CO
transfer to the liquid phase is at least equal to the rate at which it is removed by reaction. This is achieved by ag-
itation and maintaining a sufficient partial pressure of CO in the gas phase. With increasing agitation speed, the
interfacial surface area between dispersed gas bubbles and the liquid phase increases allowing for enhanced
mass transfer. CO partial pressure in the gas phase also affects the driving force for CO to transfer in to the
liquid. These two factors acting together satisfy the requirements in the AA plant design (Khimtekhnologiya
2004).
It is worth noting that according to the eq. (45), the theoretical reaction rate can be estimated for a given
temperature, rhodium concentration, and iodide concentration. Actual reaction rate is approximately 67 % of
the theoretical reaction rate. The actual reaction rate is determined by CH3 OH feed rate (Khimtekhnologiya
2004).
The rate for the water-gas shift reaction is given by eq. (46):
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

𝑅2 = (0.03933r2 − 0.1143r + 0.08516) 𝑇𝑅 (46)

Where, R2 is the experimental rate of water-gas shift reaction (mol/h), and TR is the theoretical rate of car-
bonylation reaction (mol/h) and r is the actual injected methanol/theoretical rate. eq. (46) is the experimental
rate and as derived from different actual operating conditions. At the reactor conditions, with no methanol
injected, only reaction (44) takes place. By injecting methanol, reaction (43) takes place and the remaining CO
reacts with water. In fact, CO reaction with methanol is preferred over its reaction with water. Therefore, the
water-gas shift reaction depends on reaction (43) and eq. (46) includes a term for the actual rate of the main
reaction. Some actual data at various conditions are presented in the Table 4. eq. (46) was derived using data
from rows 12 and 15 of Table 4 by curve-fitting.

Table 4: Actual plant data for calculating the water-gas shift reaction.
1 CH3I % 9.5 11.9 11.8 11.9 12.2 12.8 11.8 12.8 11.1
2 HI% 3.3 3 3.1 3.2 3 2.9 3.1 3.1 4
3 Rh (ppm) 290 290 280 304 292 288 280 284 280
4 Reactor Temp. (C) 186.4 186.6 187.7 187.7 186 187.7 187.7 187.7 187.7
5 Reactor Level (cm) 165 170 165 160 165 160 165 160 160
6 MeOH Flow (ton/h) 7.7 8 9.2 9.6 8 9.9 9.3 10.2 6.4
7 Total purge gas flow 1350 1220 1200 1350 1250 1550 1250 1580 1400
(nM3/h)
8 H2% 13.1 13.2 12.8 10.6 12.7 9.4 10.4 9.9 14
9 CO2% 13.6 13.1 10.6 10 13 9.5 11.1 9.5 12.8

Brought to you by | University of Western Ontario


10 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

10 Theoretical Rate of 14.94 14.11 14.02 15.32 14.1 15.24 14.02 16.86 14.956
carbonylation
reaction (ton/h)
11 Theoretical Rate of TR 466.292440.387437.578478.152440.075475.655437.578526.217466.790
carbonylation
reaction (mol/h)
12 Actual injected r 0.515 0.567 0.656 0.627 0.567 0.650 0.663 0.605 0.428
methanol/theoreti-
cal rate
(MeOH/TR)

13 Flow of H2 in the 176.85 161.04 153.60 143.10 158.75 145.70 130.00 156.42 196.00
purge gas (Nm3 /h)
14 Flow of H2 in the R2 15.790 14.379 13.714 12.777 14.174 13.009 11.607 13.966 17.500
purge gas (mol/h)
15 R2 /TR 0.034 0.033 0.031 0.027 0.032 0.027 0.027 0.027 0.037

Details of the carbonylation reaction are illustrated in Figure 6 (Khimtekhnologiya 2004) where injected
methanol reacts with AA, reaction (1), to generate methyl acetate(MA) and water. Consequently, the generated
MA reacts with HI and generates AA and CH3 I according to reaction (2). CH3 I then reacts with the Rh com-
plex and generates a larger Rh complex via reaction (3). Finally in reaction(4), the injected CO reacts with the
generated larger Rh complex to produce AA and water and complete the cycle for regeneration of HI and the
Rh complex catalyst (Khimtekhnologiya 2004).
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Figure 6: Details of the carbonylayion reaction.

2.4 Material and energy balance

Mass and energy balances were obtained by the following procedure. Assuming that the reactor conditions are
at steady-state (step 1), after a small time interval Δt, incremental changes would occur and another steady-state
condition is reached (step 2). Since the total mass of liquid and vapor inside the reactor was controlled and kept
constant during time interval Δt, the temperature of the fluid and the concentration of different components
for step 2 are calculated by the following equations:

QΔt (Ci,IN −Ci,1 ) + MCi,1 ±Ri


Ci,2 = (47)
M

M cp,1 T1 +QΔt (cp,IN TIN - cP,1 T1 ) + Hr


𝑇2 = (48)
M cp,2

Brought to you by | University of Western Ontario


Authenticated 11
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

The solution procedure is outlined in Figure 7. Q is the total mass flow rates of inlets or outlets of the reactor, M is
the total mass of liquid and vapor inside the reactor, Ri is the mass of component i produced or consumed during
time Δt, Ci,1 is the mass concentration of component i at time t, Ci,2 is the mass concentration of component i
at time t+Δt, Ci,IN is inlet mass concentration of component i, Hr is the heat generated or consumed by the
reaction during time Δt, and cP is the specific heat. Once the operating parameters including the methanol flow,
flashing flow, flash drum pressure, rhodium and CH3 I concenterations are set to the desired vlaues, steady state
conditions are achieved.

Figure 7: Performance of unsteady material and energy balances in CSTR.

2.5 Physical properties calculations


2.5.1 Density

Rackett recommended the following equation to claculate the density of pure compounds (Spencer and Danner
1972):

2
⎡ 7 ⎤
𝑅𝑇𝑐 ⎢⎣1+(1− 𝑇𝑐 )
𝑇
1 ⎥

𝑉𝑠 = = 𝑍 (49)
𝜌 𝑃𝑐 𝑐
where VS is saturated liquid molar volume. A modification of Rackett equation is presented by Gmehling et al.
(Gmehling and Kleiber 2012):
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

A
Density = T D
(50)
(1 + ( 1 - ) )
C
B
where A, B, C and D are constants for pure compounds. The following equation was used to calculate the
density of a mixture:

1 𝑥
= ∑( 𝑖) (51)
𝜌 𝜌𝑖
where xi is the mass fraction and ρi is density of pure component i (Perry and Green 1997).

2.5.2 Viscosity

The following equation has been proposed to calculate the viscosity of pure compounds (Viswanath et al. 2007):

𝐵
ln 𝜇𝐿 = 𝐴 + (52)
𝑇+𝐶
A modified version of the above equation (Gmehling and Kleiber 2012) is as follows:

B
𝜇L = exp [A+ +C.Ln (T) +DTE ] (53)
T
Brought to you by | University of Western Ontario
12 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

Where T is temperature (K), and A, B, C, and D are constants for pure compounds. The following equation can
be used to calclate the viscosity of a mixture:

𝐿𝑛 𝜇𝑚 = 𝑥1 𝐿𝑛 𝜇1 + 𝑥2 𝐿𝑛 𝜇2 + … (54)

Where xi is the mole fraction of each compound in the mixture (Poling et al. 2001).

2.5.3 Surface tension

Poling et al. (Poling et al. 2001)recommended the following equation to calculate the surface tension of pure
compounds:

𝑛
𝑇
𝜎 = 𝐴(1 − ) (55)
𝑇𝑐

The above equation modified by DIPPR (Design Institute for Physical Properties) as follows (Gmehling and
Kleiber 2012):

𝐵+ 𝐶𝑇𝑟 +𝐷𝑇𝑟2 + 𝐸𝑇𝑟3


𝜎 = 𝐴(1 − 𝑇𝑟 ) (56)

Where A, B, C and D are constants for each pure compound. The following equation can be used to calclate the
surface tension of a mixture (Winterfeld, Scriven, and Davis 1978):

1/2
𝑛 𝑛
1
𝜎𝑀𝑖𝑥 = ∑ ∑ 2
(𝑥𝑖 𝑣𝐿𝑖 ) (𝑥𝑗 𝑣𝐿𝑗 ) (𝜎𝐿𝑖 𝜎𝐿𝑗 ) (57)
𝑛
𝑖=1 𝑗=1 (∑𝑘=1 𝑥𝑘 𝑣𝐿𝑘 )

Where xi is mole fraction, vi is molar volume, and σLi is surface tension of each compound in the mixture.

2.5.4 Enthalpy
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

The following equation can be used to calculate the enthalpy of a mixture:

𝑖𝑔
𝐻 = 𝐻 𝑅 + 𝐻 𝑖𝑔 = (𝐻 − 𝐻 𝑖𝑔 ) + 𝐻 𝑖𝑔 = (𝐻 − 𝐻 𝑖𝑔 ) + ∑ 𝑥𝑖 𝐻𝑖 (58)

Where HR is residual enthalpy obtained from eq. (41) and Hig is the ideal gas enthalpy of each compound (Perry
and Green 1997; Chou and Prausnitz 1986):

𝑇
𝑖𝑔 𝑖𝑔
𝐻𝑖 = ∫ 𝐶𝑝 𝑑𝑇, 𝐶𝑝 = 𝐴 + 𝐵𝑇 + 𝐶𝑇 2 + 𝐷𝑇 3 (59)
𝑇𝑅

Where A, B, C, and D are constants for each pure compound.

2.6 Simulation of AA unit

The reactor specifications and the composition of feed are presented in Table 5. A computer code was developed
using Visual Basic.NET 2008 software to perform the simulation of the AA unit using the unsteady material and
energy balances that included appropriate calculations of mass transfer coefficients, kinetics, thermodynamics,
hydrodynamics, and physical properties including density, viscosity, enthalpy, and surface tension as oulined
above. A comparison of simulation results with plant data is presented in Table 6 indicating a good agreement.

Table 5: Physical specifications of the reactor and feed composition.

Brought to you by | University of Western Ontario


Authenticated 13
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

Properties Value
Height (m) 6.858
width (m) 3.658
volume (m3 ) 59
Numbers of baffles 4
Pressure (bar) 28
Temperature ( °C) 185
Agitator type Vertical dual impeller
Acetic acid concentration (mol%) 55.86
Water concentration (mol%) 38.31
CH3 I concentration (mol%) 3.51
HI concentration (mol%) 1.2
CO concentration (mol%) 0.22
CO2 concentration (mol%) 0.08
N2 concentration (mol%) 0.08
H2 concentration (mol%) 0.03

Table 6: Comparation of simlation results and actual plant data.


Conditions MeOH CO Flashing Flash Flash Return Reactor MeOH CO Purge
Temp. Temp. Flow drum drum pump Temp. Flow Flow Flow
(T) (T) (ton/h) Temp. Pres- Flow (T) (ton/h) (Nm3 /h) (Nm3 /h)
(T) sure (ton/h)
(bar)
Actual 92 35 172 127.1 1.37 132 187.7 9.2 7238 710
1 Simulation 92 35 171.7 125.45 1.37 133.5 187.7 9.2 7472 710
Deviation(%) 0.00 0.00 −0.17 −1.30 0.00 1.14 0.00 0.00 3.23 0.00
Actual 90.3 34.7 152 127.46 1.34 118 186 8 6330 678
2 Simulation 90.3 35 153 125 1.34 120 186 8 6540 672
Deviation(%) 0.00 0.86 0.66 −1.93 0.00 1.69 0.00 0.00 3.32 −0.88
Actual 98 33 182.5 126.73 1.37 140 187.7 9.9 7756 810
3 Simulation 98 35 184.7 125.6 1.37 145.8 187.7 9.8 7950 790
Deviation(%) 0.00 6.06 1.21 −0.89 0.00 4.14 0.00 −1.01 2.50 −2.47
Actual 90 35.4 188 127.1 1.38 145 187.7 10.2 7960 810
4 Simulation 90 35 191.2 125.8 1.38 151.6 187.7 10.2 8230 810
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Deviation(%) 0.00 −1.13 1.70 −1.02 0.00 4.55 0.00 0.00 3.39 0.00
Actual 49 37 135 130 1.4 105 186.7 7.3 6270 1050
5 Simulation 49 35 133 126.1 1.4 106 186.7 7.3 6154 1050
Deviation(%) 0.00 −5.41 −1.48 −3.00 0.00 0.95 0.00 0.00 −1.85 0.00

3 Simulation results
3.1 Effect of flash drum pressure

The pressure of the flash drum which is 2.3 bar under normal operation conditions can be increased under
certain situations. To investigate the effects of flash drum pressure on the performance of the AA plant two
scenarios were considered: i) reactor tempeperature of 175.2 °C and flashing flow of 80,000 kg/h, ii) Reactor
temperature of 185.36 °C and flashing flow of 179,000 kg/h. For both scenarios, the pressure of the flash drum
was varied in the range of 2.3 to 3.35 bar. Figure 8 illustrates the effect of flash drum pressure on temperature
and molar ratio of vapor phase to the total fluid flow after the JT-valve. Increasing flash drum pressure led to a
decrease in the molar ratio of vapor phase to total fluid flow that would result in an increase in the liquid level
in the flash drum. To control the liquid level, more liquid is therefore required to be pumped and returned to
the reactor thus increasing the energy requirements for pumping of excess liquid.

Brought to you by | University of Western Ontario


14 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

Figure 8: Effect of flash drum pressure on temperature and molar ratio of vapor phase to feed after JT-valve.

3.2 The effect of reactor temperature on the reaction rate

Reactor temperature is one of the most important operating parameters affecting the performance of the AA
plant. Small fluctuations in reactor temperature can cause significant variations in plant output. Figure 9 il-
lustrates the effect of temperature on the rate of methanol and CO consumption and product flows when the
reactor temperarure is reduced from 186 to 183.6 °C. A reduction in the reactor temperature would cause a de-
crease in the reaction rate leaving some methanol unreacted and consequently resulting in a reduction in both
the reactor liquid level as well as the rate of CO injection leading to lower AA production.
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Figure 9: The effect of reactor temperature on methanol reaction rate, CO flow, and product flow.

3.3 Effect of CH3 I concentration

Another important parameter is the concentration of CH3 I and HI as promoters. In order to investige the effect
of CH3 I concentration on the performance of the AA plant, CH3 I concentration in the reactor was varied in the
range of 3.11 mol% to 2.21 mol% with reactor temperature set at 186.8 °C, methanol feed rate at 10,200 kg/h,
and HI molarity at 1.1 mol%. Figure 10 shows the reaction rate for methanol and the product flow rates illus-
trating that with decreasing concentration of CH3 I, the reaction rate for methanol consumption is decreased
consequently resulting in a decrease in the flow of inejcted CO, the liquid level in the reactor, the reactor tem-
perature, and the rate of AA production.

Brought to you by | University of Western Ontario


Authenticated 15
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

Figure 10: Effect of CH3 I concentration on methanol reaction rate and product flow rates.

3.4 Effect of liquid level in the reactor

A reduction in the liquid level inside the reactor would result in lower CO transfer from the gas phase to liquid
phase (lower amounts of liquid for mass transfer) as well as in lower rate of methanol consumption (lower
reaction bed). To illustrate the effect of reactor liquid levels on the performance of the AA unit, simulations
were performed with reactor liquid levels in the range of 81 % to 67.5 % under the same concentrations of AA,
water, CH3 I, HI, and Rh catalyst with the results presented in Figure 11. According to eq. (45), the value of the
rate should be multiplied by the volume of the the liquid in the reactor:

𝐹𝑀𝑒𝑂𝐻
𝑅𝑎𝑡𝑒 = (60)
𝑉𝑟𝑒𝑎𝑐𝑡𝑜𝑟

𝐹𝑀𝑒𝑂𝐻 = 𝑅𝑎𝑡𝑒. 𝑉𝑟𝑒𝑎𝑐𝑡𝑜𝑟 (61)


Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Figure 11: Effect of liquid level in reactor on methanol reaction rate and product flow rate.

As shown in Figure 11, reducing the liquid level inside the reactor would reduce the rate of methanol consump-
tion and the product flow rates.

3.5 Effect of agitator rotation speed

The most important parameter for the rate of gas to liquid mass transfer in the reactor is rotation speed of agita-
tor. To investigate the effect of agitation speed on the performance of the reactor, simulations were performed
with agitator rotation speed in the range of 62 to 112 rpm. An increase in the agitation speed would increase the

Brought to you by | University of Western Ontario


16 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

total power input according to eq. (6), increase the gas holdup according to eqs. (2) to (4), increase the Sauter
mean diameter according to eqs. (27) to (29), increase the volumetric mass transfer coeficient, kL , according to
eqs. (30) to (33), and increase the total mass transfer rate. The increases in kL and total mass transfer rate with
increasing agitation speed are presented in Figure 12. The most important criteria for desired reactor perfor-
mance is the presence of CO in the liquid phase inside the reactor. The proper agitation speed is an important
operating parameter since the rate of mass transfer of CO to the liquid phase is highly dependent on it.

Figure 12: Effect of agitation speed on mass transfer coefficient and CO mass transfer rate.

3.6 Effect of methanol feed rate to the reactor

Under identical reactor liquid level, temperature, and feed concentrations, the rate of the water-gas shift reac-
tion is decreased with increasing flow of injected methanol to the reactor (Figure 13) as the presence of excess
mathanol favors the main carbonylation reaction over the competing water-gas shift reaction.
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Figure 13: Effect of methanol flow on the rate of water-gas shift reaction.

3.7 Optimum value of flashing flow

The return pump flowrate is an important parameter that affects the reactor temperature. This flowrate depends
on the flashing flow. If the flashing flow is higher (or lower) than the optimum value, the reactor temperature
would increase (or decrease). To investigate the effect of flahing flow on the performance of the reactor, sim-
ulation were carried out with reactor temperature of 186.5 °C, methanol flow of 9000 kg/h, and injected CO
temperature of 25 °C. Under these conditions, the flashing flow was varied to obtain the optimum value that
would maintain the reactor temperature. According to Figure 14, when the flashing flow to methanol flow ratio
was 19.2, the reactor temperature had remained constant. The columns in Figure 14 show the change in reactor
temperature over time of 15 minutes when the flashing flow is not optimum.

Brought to you by | University of Western Ontario


Authenticated 17
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

Figure 14: Effect of the rate of flashing flow on the reactor temperature.

4 Conclusions
Mathematical modeling and simulation of the acetic acid production system was performed by considering
mass transfer, kinetics, thermodynamics, hydrodynamics, and material and energy balances simultaneously.
Two methods (Lemoine 2005; Kiełbus-Rąpała and Karcz 2011) were used for the calculation of mass transfer
coefficients in the agitated reactor. The results from Lemoine’s method were closer to actual plant values since
it considers a number of system parameters including gas holdup, Sauter mean diameter, and diffusivity. The
following conclusions could be drawn from the simulation results:

1. Increasing the pressure of flash drum, led to decrease in the molar ratio of vapor phase to the total inlet
flow and leading to increased energy requirements of steam or electricity consumption for pumping excess
liquid in the flash drum.
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

2. Reducing the reactor temperature led to a decrease in the value of the theoretical reaction rate and some
methanol remained unreacted. This led to a decrease in both the flow of injected CO and liquid level in
reactor and a lower rate of AA production.
3. Reducing the CH3I concentration in the reactor also decreased the value of the theoretical reaction rate
leaving some methanol to remain unreacted consequently leading to a decrease in both the flow of injected
CO and liquid level in reactor.

4. A reduction in the liquid level in the reactor led to a reduction in both the rate of methanol and CO con-
sumption and the product flows.
5. An increase in the rotation speed of the agitator resulted in an increase in total power input, an increase in
gas holdup, and an increase in the volumetric mass transfer coefficient, kL .

Nomenclature
a  gas-liquid interfacial area per unit liquid volume, m−1

b height of impeller blade, m


C clearance between bottom and impeller, m
C* equilibrium gas solubility in the liquid, kmol.m−3
Ci mass concentration of component i

Brought to you by | University of Western Ontario


18 Authenticated
Download Date | 5/6/19 3:19 PM
DE GRUYTER Jafari et al.

cp specific heat, J / kg°C


DAB diffusivity of the gases in toluene, m2 .s−1
dS Sauter mean bubble diameter, m or mm (when specified)
dT diameter of the tank, m
dimp impeller blade dimeter, m
f friction factor
H liquid height above the bottom of the reactor, m
Hr generated heat of reaction during the time Δt, J
hB baffle length, m
kL a volumetric liquid-side mass transfer coefficient, s−1
M mass, kg
N mixing speed, rpm or Hz (when specified)
NCR critical mixing speed, rpm or Hz (when specified)
NP power number
NPunbaffled power number in unbaffled condition
NPmax power number in fully baffled condition
nB number of baffle plates
np number of impeller blades
P* total power input, W
P pressure, bar
P power consumption, W
Q mass flow, kg/h
R universal gas constant, kJ.kmol−1 .K−1
Ri mass of component i that produced or consumed, kg
Red impeller Reynolds number
ReG modified Reynolds number
T temperature, K or °C (when specified)
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

TC critical temperature, K
UG superficial gas velocity, m.s−1
V volume, m3
v phase molar volume, m3 .kmol−1
Z compressibility factor
εG gas holdup, %
μ viscosity, kg.m−1 .s−1 or Pa.s
ρ density, kg.m−3
σ surface tension, N.m−1
ω accentric factor
φ fugacity, bar

Brought to you by | University of Western Ontario


Authenticated 19
Download Date | 5/6/19 3:19 PM
Jafari et al. DE GRUYTER

References
Chang, M.-y. 1991. “Mass Transfer Characteristics of Gases in Aqueous and Organic Liquids at Elevated Pressures and Temperatures in Agi-
tated Reactors.” Chemical Engineering Science 46 (10): 2639–50.
Chou, G. F., and J. M. Prausnitz. 1986. “Adiabatic Flash Calculations for Continuous or Semicontinuous Mixtures Using an Equation of State.”
Fluid Phase Equilibria 30: 75–82.
Dake, S., D. Kolhe, and R. Chaudhari. 1984. “Carbonylation of Ethanol Using Homogeneous Rh Complex Catalyst: Kinetic Study.” Journal of
Molecular Catalysis 24 (1): 99–113.
Furukawa, H., Y. Kato, Y. Inoue, T. Kato, Y. Tada, and S. Hashimoto. 2012. “Correlation of Power Consumption for Several Kinds of Mixing
Impellers.” International Journal of Chemical Engineering 2012, 1-6.
Gmehling, J., and M. Kleiber. 2012. Chemical Thermodynamics for Process Simulation. Weinheim, Germany: John Wiley & Sons.
Golhosseini Bidgoli, R., A. Naderifar, A. R. Mohammadrezaei, and M. R. Jafari Nasr. 2012. “Kinetic Study, Modeling and Simulation of Homo-
geneous Rhodium-Catalyzed Methanol Arbonylation to Acetic Acid.” Iranian Journal of Chemistry and Chemical Engineering (IJCCE) 31 (1):
57–73.
Hjortkjaer, J., and V. W. Jensen. 1976. “Rhodium Complex Catalyzed Methanol Carbonylation.” Industrial & Engineering Chemistry Product
Research and Development 15 (1): 46–49.
Hjortkjaer, J., and J. A. Jørgensen. 1978. “Thermodynamic Activation Parameters for the Rhodium Complex Catalyzed Carbonylation of Pri-
mary Alcohols.” Journal of Molecular Catalysis 4 (3): 199–203.
Khimtekhnologiya. 2004. “Training Manual for Operation of Acetic Acid Plant.” In Training Manual for Operation of Acetic Acid Plant. Ukraine:
National Petrochemical Company- Fanavaran Petrochemical Company. 1–42.
Kiełbus-Rąpała, A., and J. Karcz. 2011. “Mass Transfer in Multiphase Mechanically Agitated Systems.” In Mass Transfer in Multiphase Systems
and Its Applications, edited by Anna Kiełbus-Rąpała and Joanna Karcz. Szczecin, Poland: West Pomeranian University of Technology. In-
Tech.
Lemoine, R. 2005. “Hydrodynamics, Mass Transfer and Modeling of the Toluene Oxidation Process.” vol. 69. Ph. D. Dissertation, Pittsburgh
university.
Lemoine, R., and B. I. Morsi. 2005. “An Algorithm for Predicting the Hydrodynamic and Mass Transfer Parameters in Agitated Reactors.”
Chemical Engineering Journal 114 (1-3): 9–31.
Li, J., M. Rodrigues, H. A. Matos, and E. Gomes de Azevedo. 2005. “VLE of Carbon Dioxide/Ethanol/Water: Applications to Volume Expansion
Evaluation and Water Removal Efficiency.” Industrial & Engineering Chemistry Research 44 (17): 6751–59.
Loiseau, B., N. Midoux, and J. C. Charpenntier. 1977. “Some Hydrodynamics and Power Input Data in Mechanically Agitated Gas‐Liquid Con-
tactors.” AIChE Journal 23 (6): 931–35.
Lu, W.-M., H.-Z. Wu, and C.-Y. Chou. 1999. “Effect of Impeller Blade Number on K L A in Mechanically Agitated Vessels.” Korean Journal of
Chemical Engineering 16 (5): 703–08.
Moutafchieva, D., D. Popova, M. Dimitrova, and S. Tchaoushev. 2013. “Experimental Determination of the Volumetric Mass Transfer Coeffi-
cient.” Journal of Chemical Technology and Metallurgy 48 (4): 351–56.
Neto, M. A. M., and J. R Barbosa JR. 2012. “A Departure Function-Based Method to Determine the Effect of the Oil Circulation Ratio on the
Performance of Vapor Compression Refrigeration Systems.” International Refrigeration and Air Conditioning Conference at Purdue. West
Lafayette:[sn].
Automatically generated rough PDF by ProofCheck from River Valley Technologies Ltd

Perry, R. H., and D. W. Green. 1997. Perry’s Chemical Engineers’ Handbook. Seventh, International ed. New York: McGraw-Hill, Inc.
Poling, B. E., J. M. Prausnitz, O. C. John Paul, and R. C. Reid. 2001. The Properties of Gases and Liquids, vol. 5, 5th ed. New York: McGraw-Hill.
Sano, Y., and H. Usui. 1987. “Effects of Paddle Dimensions and Baffle Conditions on the Interrelations among Discharge Flow Rate, Mixing
Power and Mixing Time in Mixing Vessels.” Journal of Chemical Engineering of Japan 20 (4): 399–404.
Šerbanovic, S. P., B. D. Djordjevic, and D. K. Grozdanic. 1994. “Correlation of Excess Molar Volumes of Mixtures by Means of PRSV Eos Coupled
with Mixing Rule Containing Regular Solution and Residual Excess Free Energy (KTK).” Journal of Chemical Engineering of Japan 27 (5): 671–
74.
Soriano, J. P. 2005. “Mass Transfer Characteristics in an Agitated Slurry Reactor Operating Under Fischer-Tropsch Conditions.” Master’s The-
sis, University of Pittsburgh.
Spencer, C. F., and R. P. Danner. 1972. “Improved Equation for Prediction of Saturated Liquid Density.” Journal of Chemical and Engineering Data
17 (2): 236–41.
Stryjek, R., and J. Vera. 1986. “PRSV—An Improved peng‐Robinson Equation of State with New Mixing Rules for Strongly Nonideal Mix-
tures.” The Canadian Journal of Chemical Engineering 64 (2): 334–40.
Tekie, Z., J. Li, B. I. Morsi, and M.-Y. Chang. 1997. “Gas-Liquid Mass Transfer in Cyclohexane Oxidation Process Using Gas-Inducing and
Surface-Aeration Agitated Reactors.” Chemical Engineering Science 52 (9): 1541–51.
Viswanath, D., T. Ghosh, D. Prasad, N. Dutt, and K. Rani. 2007. Viscosity of Liquids: Theory, Estimation, Experiment, and Data. Washington, United
States: Springer.
Winterfeld, P., L. Scriven, and H. Davis. 1978. “An Approximate Theory of Interfacial Tensions of Multicomponent Systems: Applications to
Binary Liquid‐Vapor Tensions.” AIChE Journal 24 (6): 1010–14.

Brought to you by | University of Western Ontario


20 Authenticated
Download Date | 5/6/19 3:19 PM

You might also like