Staging of The Fischer Tropsch Reactor W

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Reactor design 1

Ahmad Rafiee Research Article


Magne Hillestad

Norwegian University of Science


Staging of the Fischer-Tropsch Reactor
and Technology (NTNU), with a Cobalt-Based Catalyst
Department of Chemical
Engineering, Trondheim,
Norway. A method for systematic reactor design, described by Hillestad [1], is applied to
the Fischer-Tropsch synthesis. The reactor path is sectioned into stages and design
functions are optimized to maximize an objective function. Two different objec-
tive functions are considered: the yield of wax and a measure of the profitability.
With the chosen kinetic model [2] and the path temperature constrained by
240 °C, staging of the Fischer-Tropsch synthesis based on the first criteria will
increase the yield of wax. By introducing the cost of heat transfer area in the
objective function, the total heat transfer area requirement of a two-stage reactor
is significantly less than of a single-stage reactor.

Keywords: Fischer-Tropsch synthesis, Path optimization, Plug flow, Reactor design, Staging
Received: December 14, 2012; revised: May 29, 2013; accepted: June 07, 2013
DOI: 10.1002/ceat.201200700

1 Introduction The earliest catalysts used for FT synthesis were iron and
cobalt. Iron is a highly active catalyst and exhibits water-gas
In the last decades, the conversion of natural gas through a shift (WGS) activity whereas cobalt catalysts do not have this
gas-to-liquids (GTL) process proved to be an alternative for activity, leading to improved hydrocarbon yield. Cobalt cata-
the use of remote natural gas reserves to produce liquid trans- lysts yield mainly straight-chain hydrocarbon products and no
portation fuels. An increasing world-wide demand for clean- oxygenates as with iron catalysts. However, cobalt catalysts are
burning fuels has sparked a renewed interest in the study of 230 times more expensive than iron, but still a very good alter-
the Fischer-Tropsch (FT) synthesis. native to iron catalysts. The reason is that cobalt catalysts dem-
A GTL plant consists of three main sections: (i) synthesis gas onstrate activity at lower operating pressures [5].
(syngas) production: conversion of natural gas to a mixture Staging in chemical engineering is not a new idea and has
consisting mainly of H2 and CO, (ii) FT synthesis: conversion been applied on many processes. Androulakis and Reyes [6]
of H2 and CO to a wide range of hydrocarbons, (iii) upgrading studied oxidative coupling of methane (OCM) and the role of
of products. oxygen distribution and product removal on a staged plug-
There are different routes for syngas production: auto-ther- flow reactor. Maretto and Krishna [7] introduced staging of an
mal reforming (ATR), steam methane reforming, combined FT slurry-bubble column reactor and optimized the reactor
reforming, and heat exchange reforming involving series and conversion. The advantages of this reactor configuration com-
parallel arrangements [3]. Over the years, four types of reac- pared to a single-stage reactor are increased syngas conversion
tors have been utilized for FT synthesis: the fixed-bed tubular and reactor productivity. In their study, the multi-stage reactor
reactor known as the ARGE reactor, circulating fluidized-bed requires additional cooling tubes. Waku et al. [8] performed
reactors known as Synthol reactors, the Sasol Advanced studies on staged oxygen introduction and selective hydrogen
Synthol reactor, and the slurry-bubble column reactor [4, 5]. combustion during propane dehydrocyclodimerization reac-
The upgrading unit involves mainly separation and hydropro- tions on cation-exchanged zeolite. Hwang and Smith [9] inves-
cessing. The primary upgrading starts with the removal of light tigated the effect of catalyst dilution and distribution of feed to
ends and dissolved gases. The same basic technologies used in control the reactor temperature of two case studies involving
crude oil refineries have been adapted for FT product refining nitrobenzene hydrogenation and ethylene oxidation. Diakov
[3]. and Varma [10] investigated the effect of feed distribution in a
membrane reactor for methanol oxidative dehydrogenation.
Guillou et al. [11] examined the influence of hydrogen distri-
– bution between stages in a micro-channel FT reactor. Rafiee
Correspondence: Prof. M. Hillestad (magne.hillestad@chemeng.ntnu.
and Hillestad [12] evaluated staging of an FT reactor with an
no), Norwegian University of Science and Technology (NTNU), iron-based catalyst. The design functions, i.e., fluid mixing,
Department of Chemical Engineering, Sem Sælandsvei 4, 7491 Trond- hydrogen distribution, heat transfer area distribution, coolant
heim, Norway. temperature, and catalyst concentration are optimized to max-

Chem. Eng. Technol. 2013, 36, No. 00, 1–11 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
2 A. Rafiee, M. Hillestad

imize the concentration of C11+ at the end of the reactor path. liquid phase and the other species are in the gas phase. There
Jess et al. [13, 14] described the effect of particle size and sin- are small traces of these components in the gas phase which
gle-tube diameter on the thermal behavior of FT reactors. are neglected here. The average heat capacities of lump 1–5
The potential of sectioning the FT synthesis reactor into are: 2.86, 2.56, 2.52, 3.15, and 3.00 kJ kg–1K–1, respectively. The
stages based on the method proposed by Hillestad [1] is dem- price of products (lump 3–5) are 0.8, 0.9, and 1.1 kg–1, respec-
onstrated. Each reactor stage is represented by design func- tively [16].
tions. Staging of the reactor path will provide more degrees of The mass production rates of lump 2–5 are according to the
freedom for optimization. This study focuses on the FT unit ASF distribution model:
only and investigates the possibilities of cost reduction of the
FT reactor. ~ 2 ˆ 2aASF 1
R aASF †2 B (5)

~ 3 ˆ 3a2ASF 1
R aASF †2 ‡4a2ASF 1 2
aASF † B (6)
2 Fischer-Tropsch Kinetic Model
~ 4 ˆ 5a4ASF 1 aASF †2 ‡6a5ASF 1 aASF †2 ‡ . . . ‡10a9ASF 1 aASF †2 B

R
In an FT reactor, the synthesis gas (syngas) reacts to form a (7)
mixture of hydrocarbons.
~ 5 ˆ 11a10 aASF †2 ‡ . . . B

CO + UH2H2 → (–CH2–) + H2O (1) R ASF 1 (8)

The kinetic model applied here is the one given by Iglesia The conservation of mass requires that the sum of all com-
et al. [2] for a cobalt catalyst, where methane production and ponent reaction rates add up to zero.
CO consumption are defined by Eqs. (2) and (3), respectively.
~2 ‡ R
~1 ‡ R ~3 ‡ R
~4 ‡ R ~ CO ‡ R
~5 ‡ R ~H ‡ R
~H O ˆ 0

This kinetic model is chosen because it takes into account a R 2 2
(9)
higher selectivity of methane. However, other available kinetic
models could be applied [15]. The first five reaction rates in Eq. (9) are the rates of lump
1–5 on mass basis. The value of B is determined from Eq. (9).
k1 PH2 PCO 0:05 The consumption rate of H2 is calculated according to the stoi-
rCH4 ˆ (2)
1 ‡ K1 PCO chiometry of the reactions and the distribution of paraffins
and olefins.
k2 PH2 0:6 PCO 0:65
rCO ˆ (3)
1 ‡ K1 PCO
3 Problem Formulation
The kinetic and adsorption parameters are given in Tab. 1.
The process considered in this study is a once-through FT 3.1 Staging of the Reactor Path
reactor (s) and the pressure of the reactor is 20 bar.
The weight fractions of FT products heavier than methane A path is a line of production on which basic operations take
(Wn) are assumed to follow the ideal Anderson-Schulz-Flory place which are represented by design functions. The path is
(ASF) distribution. sectioned into a number of stages and the design functions
(decision variables) are optimized so as to maximize an objec-
Wn ˆ n 1 aASF †2 aASF n 1
n ˆ 2; 3; :::; ∞ (4)
tive function. The flow model given by Eq. (10) is a concise
formulation, representing the change of state variables (mass
In order to reduce the number of components, a lumping
fractions and temperature) along the path. The derivation of
technique is used for hydrocarbon species: (1) CH4, (2) C2, (3)
Eq. (10) is described in detail by Hillestad [1].
LPG (C3 and C4), (4) middle distillate (C5–C10), and (5) wax
(C11+). dx
At 20 bar and a temperature range of 210–240 °C, vapor- ‰cI uM r~JŠ ˆ ruA R
~ x† ‡ uF K  xF x† uH E  x xw †
xn
liquid equilibrium calculations using the UNISIM DESIGN (10)
process simulator indicate that lump 4 and 5 are mainly in the

Table 1. Kinetic and adsorption parameters in Eqs. (2) and (3) [2]. The state vector x is the vector of mass fractions
augmented with the temperature (Rafiee and Hil-
Parameter Arrhenius expression Unit lestad [12]).

x ˆ ‰x 1 ; x2 ; x3 ; x4 ; x 5 ; xH2 ; xCO ; xH2 O ; hŠT


 
37 326
k1 8:8 × 10 6 exp kmolCH4 Pa 1:05 3
mreactor s 1
11†
RT
  hˆ T Tref †Tref1
68 401:5
K1 1:096 × 10 12 exp Pa 1
RT The first five mass fractions in Eq. (11) are the
mass fractions of lump 1–5, respectively.
 
37 326
k2 1:6 × 10 5 exp kmolCO Pa 1:25 3
mreactor s 1
~ x†
RT The dimension of the reaction rate vector R
–3 –1
is kg m s .

www.cet-journal.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2013, 36, No. 00, 1–11
Reactor design 3

~ x† ˆ R
R

~1; R
~2; R
~3; R
~4; R~5; R
~H ; R
~ CO ; R ~h T
~H O; R
 4.2 Objective Functions
2 2
(12)
~ h ˆ D; H †cp 1 Tref1 rCO
R The concentration of wax at the end of the path is a possible
objective function, J1 = xwax [1]. The design functions repre-
Heat of reaction (–DrH) at 25  C is 172 kJ per mole of CO sent costs, such as heat transfer area, extra feed points, and
consumed in the FT reactor. Simultaneously, the total mass is utility streams. A measure of the profit is another possible
conserved. objective function (J2).
dc Here, it is focused only on the FT synthesis, where the syngas
ˆ uF (13) is imported. With a syngas capacity of around 67.45 kg s–1, the
dn
price of syngas is 0.23 kg–1. The pressure of the reactor is
The total mass flow rate W is independent of pressure 20 bar.
change, temperature change, phase change, or mole number The costs associated with this study are categorized as:
change along the reactor path [1]. 1. Capital cost: Cost of FT reactor including reactor shell
Mixing: The function uM(n) is a design function represent- [19], tubes [20], and the price of cobalt catalyst which for
ing the fluid mixing. When uM is zero, Eq. (10) represents a the initial load amounts to ∼ 54 kg–1. The investment
plug-flow model and as the  functionincreases proportionally required for the power production unit is assumed to be
duM
with the length of path ˆ 1 , Eq. (10) becomes the 300 kW–1. It is presupposed that process water is free.
dn The selling price of the produced electricity in the power
completely mixed volume. For an intermediate slope
  plant is 0.06 kW–1h–1 [19].
duM
0< < 1 , Eq. (10) represents the plug-flow model with 2. Equivalent Annual Operating Cost (EAOC): This cost is
dn
the annual operating cost plus the annual capital charge
recycle [1].
[19]. The annual capital charge is the amortized capital
Distributed feed: The design function uM = ar defines extra
cost over the operating life of the plant to establish an
feed distribution along the path volume. The distribution of
annual cost. For an operating life time of 20 years and an
extra feed can be continuous or pointwise. In this study, a
interest rate of 8 %, the annual capital charge is 10 % of
pointwise distributed feed is applied [1].
the capital cost [19].
Heat transfer: The design function uH = br represents the
The annual profit is equal to the income minus the EAOC.
heat transfer area distribution along the path. Heat transfer
The amount of steam produced in the FT reactor is calcu-
area distribution is parameterized by a piecewise constant
lated as the total heat transfer divided by the latent heat of va-
function [1].
porization (assuming 100 % efficient heat transfer):
Chemical reactions: The design function uA is the catalyst R
activity which can vary between 0 and 1. A value of 1 is consid- UA T TW †dn
ered to be the maximum catalyst concentration. Catalyst activ- m_ steam ˆ (16)
k
ity is parameterized by a piecewise constant function.
Parameterization of design functions leads to a set of in- The latent heat of vaporization (k) is 1889 kJ kg–1.
equality constraints [1]. The model applied here is a pseudo- Finally, J2 is formulated as below (tax is neglected):
heterogeneous model. The gas and liquid phases flow in the
same direction and have the same degree of dispersion. J2 ˆ Income EAOC†
EAOC ˆ annual operating cost ‡ 0:1 × capital cost
nS nS
4 Optimization
X X
ˆ 4:792 × 108 ‡ 8:660 × 105 uA;i Di ‡ 1:027 × 104 b 2i Di C2;i
iˆ1 iˆ1
4.1 Optimal Control Analogy (17)

The optimal path configuration can be found by solving a C2,i is the cost of tubes per meter ($ m–1) [20].
problem similar to that of an optimal control problem.
C2,i = 2.5 + 98.958(di – 0.02) (18)
max J
‰r; uŠ ∈ U All performed approximations have the same effect on all
(14)
dz the cases.
s:t: ˆ f z; u†; z 0† ˆ z0
dn

The objective function (J) is formulated as a function of 4.3 Mathematical Programming


state variables z ˆ ‰xT ; cŠT at the end of the path. Here, it is re-
quired that the temperature along the path does not exceed The decision variables (design functions) and state vector
240 °C, i.e., T(n) ≤ 240 °C [7, 17, 18]. z = [xT, c]T are discretized. The design functions are discretized
The path constraints are represented by nonlinear inequality by piecewise linear or piecewise constant functions [1]. The
constraints [1]. system of ordinary differential equations given by Eq. (14) is
h(z,u) ≤ 0 (15) discretized by orthogonal collocation. On each reactor stage

Chem. Eng. Technol. 2013, 36, No. 00, 1–11 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
4 A. Rafiee, M. Hillestad

and each collocation point, Eq. (14) is formulated as nonlinear optimizing the path, no extra feed is beneficial, because it will
equality constraints [1]. dilute the wax concentration. The mixing structure for this
case is a PFR and there is no catalyst dilution. The reactor tem-
In this study, the optimization algorithm based on SQP, fmin- perature hits the maximum to obtain the maximum conver-
con in Matlab is applied. The objective functions are maximi- sion and then declines. In this case, the concentration of water
zation of wax production at the end of the path and a measure is 41.87 wt %.
of profitability. The inequality constraints, h(z,u), are path Case 2: Two stages are selected. The concentration of wax
constraints and bounds on decision variables. The equality and the average heat transfer area density are higher than in
constraints are the discretized state variables at each colloca- case 1. There is no cost on heat transfer area. All the design
tion point on each reactor stage. functions are free and there are 13 degrees of freedom, i.e., two
design parameters for each of heat transfer area density, cool-
ant temperature, catalyst activity, and extra feed and five for
mixing structure. The results are indicated in Fig. 2.
5 Results Removing liquid wax from the outlet stream of the first
stage has no effect on the objective function, because the reac-
In all cases (if otherwise stated), the residence time (s) and tion rates are calculated based on the partial pressure of H2
chain growth probability are kept constant at 2.97 m3s kg–1 and and CO in the gas phase. On the other hand, removing water
0.9 [7, 17, 18], respectively. The maximum operating tempera- from the first stage and reoptimizing the path will increase the
ture of the reactor is limited by catalyst deactivation. Catalyst concentration of wax to 23.98 %. In this case, CO conversion
deactivation is not considered here. and average heat transfer area are 81.90 % and 163.43 m2m–3,
Tab. 2 presents the optimization results. In each case, the respectively.
input parameters are: objective function criteria, number of Case 2a: In this case, the heat transfer density is required to
stages, number of design parameters for coolant temperature, be uniform along the path. The reason may be that similar
and number of design parameters for heat transfer area den- reactors are wanted. The mass fraction of wax and CO conver-
sity. In Tab. 2, the value of 1 for nuT (or nuH) means that all the sion are increased compared to case 1. The first stage is plug
stages have the same coolant temperature or heat transfer area flow with recycle (rP) due to less degrees of freedom compared
density. The optimization outputs are: value of objective func- to case 2, and rP is there to level out the temperature peak. In
tion, mixing structure (PFR, CSTR, or PFR with recycle), aver- this case, there are 12 degrees of freedom.
age heat transfer area density, coolant temperature, catalyst Case 2b: The same criteria as above are applied and both
dilution, and mass fraction of each lump of components. The the coolant temperature and the heat transfer density are re-
fresh syngas and extra feed contain only H2 and CO and H2/ quired to be the same along the path. The reason of having
CO to the reactor is 2.1. equal coolant temperatures is that it is intended to use the
Case 1: One stage is selected and the target is to maximize same pressure level of boiling water as coolant. The mixing
the mass fraction of wax. The profile is illustrated in Fig. 1. By structure for this case is PFR with recycle for the first stage
and PFR for the second stage.
The number of degrees of free-
Table 2. Cases with different optimization criteria, number of stages (ns), number of design pa-
dom for this case is 11.
rameters for coolant temperature (nuT), and heat transfer area density (nuH).
Case 2c: Chain growth probabil-
Input Results ity is varying along the path as a
function of temperature and par-
Criteria Case ns nuT a nuH b J uM c ad xWax CO
tial pressures of H2 and CO based
conversion [%]
on the correlation proposed by
1 1 1 1 21.92 P 123.5 21.92 74.91 Song et al. [21]. The chain growth
probability at the end of the path is
2 2 2 2 22.81 P-P 156.5 22.81 77.97
0.78 and the concentration of wax
2a 2 2 1 22.66 rP-P 151.7 22.66 77.5 is 8.05 wt %. The concentration of
J1 lump 4 is increased compared to
2b 2 1 1 22.64 rP-P 151.0 22.64 77.43
case 2 due to a low chain growth
2c 2 2 2 8.05 rP-P 157.2 8.05 75.55 probability.
3 3 3 3 23.02 rP-P-P 155.1 23.02 78.72 Case 3: Three stages are applied
here. The optimal mixing structure
4 1 1 1 100.00 % P 117.2 21.79 74.47
is rP-P-P (Fig. 3) and the concen-
J2 4a 2 2 2 122.00 % C-P 97.6 22.21 75.97 tration of wax is increased com-
pared to case 1. All design func-
4b 2 2 2 125.00 % C-P 96.1 22.31 76.28
tions are free and 20 degrees of
a
Number of design parameters for coolant temperature; 1 means the same coolant temperature. freedom are given.
b
Number of design parameters for heat transfer area density; 1 means the same heat transfer Case 4: One stage is selected and
area density. cP is PFR, C is CSTR, and rP is a PFR with recycle. dAverage heat transfer area den- a measure of the annual profit is
sity along the path. maximized. The mixing structure

www.cet-journal.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2013, 36, No. 00, 1–11
Reactor design 5

Figure 1. Case 1 and the cri-


teria defined by J1.

is a single PFR and the wax concentration of wax is 6 Discussion


21.79 wt %.
Case 4a: Two stages are applied and there is no distributed 6.1 Model Verification
feed. The mixing structure is CSTR for the first stage and
plug flow for the second stage. The value of objective func- The model presented in this study is verified by: (i) UNISIM
tion is increased by 22 % which is due to the lower heat DESIGN process simulator, and (ii) performing two-phase cal-
transfer area requirements compared to the single-stage reac- culations based on the correlations proposed by Marano and
tor (case 4). Comparing the heat transfer area density of this Holder [22] for FT systems. The values of b, coolant tempera-
case with case 2 demonstrates that by introducing the cost ture, and mixing structure are the same as for case 1. The
of heat transfer area, the total area is reduced. The results results presented in Tab. 3 are close to the results given by
are presented in Fig. 4. If the CSTR in Fig. 4 is set to a PFR, UNISIM DESIGN and two-phase calculations. Consequently,
the peak temperature exceeds the maximum temperature.
The CSTR serves to level out the temperature peak with less Table 3. Verification of model. I: this study, II: UNISIM DESIGN,
heat transfer area. III: two-phase calculations.
Case 4b: The same criteria as above are applied but the origi-
nal feed is distributed between the stages. In this case, 97.48 % I II III
of the original syngas is fed to the first reactor and the objec- xWax [wt %] 21.92 21.70 21.63
tive function increases by 25 % compared to case 4. The con-
centration of wax at the end of the path is 22.31 wt %. CO conversion [%] 74.91 74.04 74.47

Chem. Eng. Technol. 2013, 36, No. 00, 1–11 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
6 A. Rafiee, M. Hillestad

Figure 2. Case 2 and the cri-


teria defined by J1.

assuming an ideal split between gas and liquid phases is a good where all the design functions are free, the number of degrees
approximation. of freedom is 13. In case 2a, the heat transfer area density is
forced to be the same for both stages, and the first stage is a
plug flow with recycle to level out the temperature peak.
6.2 Interpretation of the Results By introducing the cost of heat transfer area in the objective
function, the total heat transfer area requirement of a two-
This study focuses on the FT unit only and investigates the stage reactor is 16.7 % less than for a single-stage reactor. The
possibilities of cost reduction of an FT reactor. The kinetic optimal mixing structure of a two-stage FT reactor is com-
model is the one given by Iglesia et al. [2] which takes into pletely mixed (CSTR) for the first stage and plug flow (PFR)
account a higher selectivity of methane. The pressure drop is for the second stage. The CSTR is to level out the temperature
neglected here. The model applied is a pseudo-heterogeneous peak with less heat transfer area requirement.
model. A case study is conducted to investigate the effect of distri-
With the chosen kinetic model, staging of the FT synthesis buting the original syngas feed between the stages. In this case,
based on the first criteria will increase the mass fraction of wax 97.48 % of the original syngas is fed to the first reactor and the
and CO conversion. The concentration of wax for a single-, concentration of wax at the end of the path is 22.31 wt %.
two-, and three-stage FT reactor of the same total volume is For a single-stage FT reactor with iron-based catalyst and
21.92, 22.81, and 23.02 wt %, respectively. By optimizing the syngas flow rate of 67.45 kg s–1, the conversion of CO and con-
path, no extra feed is beneficial because it will dilute the wax centration of wax at the end of the path are 75.23 % and
concentration. The reactor temperature hits the maximum to 13.70 %, respectively, [12]. In this case, the water-gas shift
obtain the maximum conversion. For a two-stage reactor reaction is active and part of CO is converted to CO2, i.e., the

www.cet-journal.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2013, 36, No. 00, 1–11
Reactor design 7

Figure 3. Case 3 and the cri-


teria defined by J1.

concentration of CO2 at the end of the path is 34.81 %. This case. Systematic reactor staging will provide more degrees of
will increase H2/CO at the end of the reactor path to 5.33. The freedom for reactor optimization.
optimal mixing structure is plug flow [12]. The mixing structure of a two-stage reactor based on the
Maretto and Krishna [7] studied the staging of an FT slurry- profitability measure (second objective function) is CSTR for
bubble column reactor and optimized the reactor conversion. the first stage and plug flow for the second stage. The geometry
Staging of the reactor increases syngas conversion and reactor of the slurry reactor, gas and liquid velocities, and the arrange-
productivity compared to a single-stage reactor. In their study, ment of the cooling tubes will affect the flow pattern inside the
the catalyst activity is equal to 1 and a temperature difference reactor [7]. With the same value of b for fixed-bed and slurry-
of 10 °C is considered between the reactor and the coolant. bubble column reactors, more area is needed for a fixed-bed
Furthermore, there is no cost on heat transfer area and the reactor due to a lower heat transfer coefficient.
multi-stage reactor requires additional cooling tubes. Here,
catalyst activity, reactor temperature, and heat transfer area are
optimized. 7 Conclusions
The method developed by Hillestad [1] was applied to FT syn-
6.3 Realization of the Optimal Design thesis by means of a kinetic model given by Iglesia et al. [2].
The design functions such as fluid mixing, heat transfer area
In an industrial GTL plant, the FT reactor will require large density, catalyst dilution, and distribution of extra feed are
volumes and sectioning of the reactor may be necessary in any optimized along the path to maximize an objective function.

Chem. Eng. Technol. 2013, 36, No. 00, 1–11 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
8 A. Rafiee, M. Hillestad

Figure 4. Case 4a and the


criteria defined by J2.

The objective function is here defined to be the yield of wax Staging with optimal distribution of heat transfer area, dis-
(J1) and a measure of profitability (J2). Staging of the FT reac- tribution of syngas, and mixing configuration will increase the
tor based on J1 will increase the mass fraction of wax and CO profit measure as demonstrated here. The results of this study
conversion. No distributed feed is beneficial because it will can serve as initial points for staging of the FT reactor in the
only dilute the wax concentration. overall GTL process with different syngas production config-
By introducing the cost of heat transfer area in the objective urations [23].
function (J2), the total heat transfer area requirement is re-
duced which will increase the annual profit. The optimal mix-
ing structure for a two-stage FT is completely mixed (CSTR) Acknowledgment
for the first stage and plug flow (PFR) for the second stage.
Having less heat transfer area, the purpose of the CSTR is to The authors gratefully acknowledge financial support from the
level out the temperature peak. If the CSTR is set to a PFR, the Research Council of Norway through the GASSMAKS pro-
peak temperature will exceed the maximum temperature. The gram.
syngas feed can be distributed between the stages. In this case,
97.48 % of the main syngas feed is fed to the first reactor and The authors have declared no conflict of interest.
the objective function will increase compared to a single-stage
reactor due to less heat transfer area requirements.

www.cet-journal.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2013, 36, No. 00, 1–11
Reactor design 9

Symbols used V [m3] residence volume along the path


VR [m3] total residence volume of the path
a [m2m–3] heat transfer area density W [kg s–1] total mass flow rate along the path
A [m2] heat transfer area W0 [kg s–1] total mass flow rate at the inlet
B [–] scaling parameter to ensure a Wn [–] weight fraction of hydrocarbons
consistent ASF distribution for all heavier than methane
components heavier than methane x [–] state vector of mass fractions and
cp,i –1 –1
[kJ kg K ] specific heat capacity of component i temperature, x = [x1,...,xn,h]T
cp,ref [kJ kg–1K–1] reference heat capacity xF [–] state vector of extra feed
C2,i [m–1] cost of tubes per meter xW [K] coolant temperature
di [m] tube diameter z [–] state vector augmented with mass
E [–] diagonal matrix, flow rate, z = [xT,c]
E = diag (0,0,...., cp,ref cp–1) z(0) [–] z at the inlet of the reactor path
h [–] path constraints
I [–] identity matrix Greek symbols
J1 [wt %] wax concentration a [kg m–3s–1] feed distribution
J2 [–] measure of profitability aASF [–] chain growth probability
~J [–] partial derivative of component b [kg m–3s–1] heat transfer area design function,
reactions with respect to x, b = Ua cp,ref–1
~
~J ˆ ∂R0 x† ‡ diag 0; 0; . . . ; 1† × c [–] mass flow rate relative to the inlet,
  ∂x   W W0–1
a 1 cp;f cp;01 bcp;ref cp;01 Di [–] volume fraction of stage i, Vi VR–1
K [–] diagonal matrix, k [kJ kg–1] latent heat of vaporization of water
diag (1,1,...., cp,F cp–1) h [–] dimensionless temperature,
k1 [kmolCH4Pa–1.05mreactor–3s–1] h = (T–Tref )Tref–1
kinetic parameter n [–] dimensionless volume of the path
k2 [kmolCOPa–1.25mreactor–3s–1] (independent variable), V VR–1
kinetic parameter r [m3s kg–1] space time or residence time,
K1 [Pa–1] adsorption parameter VR W0–1
ṁsteam [kg s–1] steam production rate xi [–] mass fraction of component i
n [–] carbon number
nc [–] number of components
ns [–] number of stages References
nuT [–] number of design parameters for
coolant temperature [1] M. Hillestad, Chem. Eng. Sci. 2010, 65, 3301–3312.
nuH [–] number of design parameters for [2] E. Iglesia, S. C. Reyes, S. L. Soled, in Computer-Aided Design
heat transfer area density of Catalysts (Eds.: E. R. Becker, C. J. Pereira), Exxon Research
PCO [Pa] partial pressure of CO and Engineering Co., New York 1993, 199–257.
PH2 [Pa] partial pressure of H2 [3] M. E. Dry, A. P. Steynberg, in Fischer-Tropsch Technology
rCO [kmol mreactor–3s–1] consumption rate of CO (Eds: M. E. Dry, A. P. Steynberg), Elsevier Science & Technol-
rCH4 [kmol mreactor–3s–1] production rate of CH4 ogy Books, Amsterdam 2004, 406–481.
Ri [kmol mreactor–3s–1] component reaction rate on molar [4] B. H. Davis, Catal. Today 2002, 71, 249–300.
basis [5] http://www.fischer-tropsch.org/DOE/DOE_reports/510/
~i
R [kg mreactor–3s–1] component reaction rate on mass 510-34929/510-34929.pdf
basis [6] I. P. Androulakis, S. C. Reyes, AIChE J. 1999, 45, 860–868.
T [K] temperature along the path [7] C. Maretto, R. Krishna, Catal. Today 2001, 66, 241–248.
Tref [K] reference temperature [8] T. Waku, S. Y. Yu, E. Iglesia, Ind. Eng. Chem. Res. 2003, 42,
Tw [K] temperature of the cooling medium 3680–3689.
along the path [9] S. Hwang, R. Smith, Chem. Eng. Sci. 2004, 59, 4229–4243.
uA [–] catalyst dilution [10] V. Diakov, A. Varma, Ind. Eng. Chem. Res. 2003, 43 (2), 309–
uF [–] feed distribution design function, 314.
uF = ra [11] L. Guillou, S. Paul, V. L. Courtis, Chem. Eng. J. 2008, 136,
uH [–] heat transfer area distribution 66–76.
design function, uH = rb [12] A. Rafiee, M. Hillestad, Comput. Chem. Eng. 2012, 39, 75–83.
uM [–] mixing design function [13] A. Jess, C. Kern, Chem. Eng. Technol. 2012, 35 (2), 369–378.
u [–] vector of decision variables [14] A. Jess, C. Kern, Chem. Eng. Technol. 2012, 35 (2), 379–386.
U [kWm–2K–1] overall heat transfer coefficient [15] I. C. Yates, C. N. Satterfield, Energy Fuels 1991, 5, 168–173.
UH2 [–] usage ratio of H2 [16] http://www.opec.org/opec_web/en/publications/338.htm

Chem. Eng. Technol. 2013, 36, No. 00, 1–11 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
10 A. Rafiee, M. Hillestad

[17] R. Krishna, S. T. Sie, Fuel Process. Technol. 2000, 64, 73–105. [21] H. S. Song, D. Ramkrishna, S. Trinh, H. Wright, Korean
[18] C. Maretto, R. Krishna, Catal. Today 1999, 52, 279–289. J. Chem. Eng. 2004, 21 (2), 308–317.
[19] R. Turton, R. C. Bailie, W. B. Whiting, J. A. Shaeiwitz, Analy- [22] J. J. Marano, G. D. Holder, Fluid Phase Equilib. 1997, 138, 1–
sis, Synthesis, and Design of Chemical Processes, 3rd ed., Pre- 21.
ntice Hall, New York 2008. [23] A. Rafiee, M. Hillestad, Chem. Eng. Technol. 2012, 35 (5),
[20] M. S. Peters, K. Timmerhaus, R. West, Plant Design and Eco- 870–876.
nomics for Chemical Engineers, 3rd ed., McGraw-Hill, New
York 2003.

www.cet-journal.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2013, 36, No. 00, 1–11
Reactor design 11

Research Article: A cobalt-based catalyst Staging of the Fischer-Tropsch Reactor


Fischer-Tropsch (FT) reactor is with a Cobalt-Based Catalyst
sectioned into stages. Design functions
are optimized to maximize the two A. Rafiee, M. Hillestad*
selected objective functions, namely, the
wax concentration at the end of the Chem. Eng. Technol. 2013, 36 (䊏),
reactor path and a measure of XXX … XXX
profitability. Staging of the FT reactor
increases the production rate of wax DOI: 10.1002/ceat.201200700
based on the first objective function
criterion.

Chem. Eng. Technol. 2013, 36, No. 00, 1–11 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com

You might also like