Sei sulla pagina 1di 31

Industrial & Engineering Chemistry Research

This document is confidential and is proprietary to the American Chemical Society and its authors. Do not
copy or disclose without written permission. If you have received this item in error, notify the sender and
delete all copies.

Model-based optimization of an Acetylene Hydrogenation


reactor to improve overall ethylene plant economics

Journal: Industrial & Engineering Chemistry Research

Manuscript ID ie-2017-05234d.R2

Manuscript Type: Article

Date Submitted by the Author: 29-May-2018

Complete List of Authors: Aeowjaroenlap, Hattachai; SCG Chemicals Co Ltd, Process Technology
Center
Chotiwiriyakun, Kritsada; SCG Chemicals Co Ltd, Process Technology
Center
Tiensai, Nattawat; SCG Chemicals Co Ltd, Process Technology Center
Tanthapanichakoon, Wiwut; SCG Chemicals Co Ltd, Process Technology
Center
Spatenka, Stepan; Process Systems Enterprise Ltd
Cano, Alejandro; Process Systems Enterprise Ltd

ACS Paragon Plus Environment


Page 1 of 30 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
Model-based optimization of an Acetylene
9
10
11
12
Hydrogenation reactor to improve overall ethylene
13
14
15
16
plant economics
17
18
19
20
Hattachai Aeowjaroenlap*a, Kritsada Chotiwiriyakuna, Nattawat Tiensaia, Wiwut
21
22
23 Tanthapanichakoona, Stepan Spatenkab, Alejandro Canob
24
25
a
26 SCG Chemicals Co., Ltd, 271 Sukhumvit Rd., Map Ta Phut, Muang Rayong, Thailand 21150
27
28 b
Process Systems Enterprise Ltd, 26-28 Hammersmith Grove, London, W6 7HA, United
29
30
31 Kingdom
32
33
34 Keywords: acetylene hydrogenation, kinetics, modeling, optimization
35
36
37 ABSTRACT
38
39
40 The steam cracking process is a common technology for producing ethylene from
41
42
43
naphtha. However, one of the major contaminants in the ethylene product stream is acetylene,
44
45 which poisons catalysts used in downstream polymerization processes and needs to be converted
46
47 to ethylene by selective catalytic hydrogenation in order to upgrade product quality and increase
48
49
overall ethylene yield. The selective hydrogenation process is usually carried out in a multi-stage
50
51
52 fixed-bed catalytic reactor with internal cooling between stages to ensure that the outlet
53
54 concentration of acetylene does not exceed 1 ppm. Nevertheless, side reactions also take place,
55
56
57
58
59
60 ACS Paragon Plus Environment
1
Industrial & Engineering Chemistry Research Page 2 of 30

1
2
3 including total hydrogenation to ethane which increases energy consumption at the recycle
4
5
6 furnace and decreases overall plant productivity. Moreover, in a tail-end hydrogenation process,
7
8 the catalyst surface often becomes covered by so-called Green Oil generated from acetylene
9
10 oligomerization, which causes catalyst deactivation and lowers the selectivity to ethylene. A key
11
12
13
challenge of the tail-end acetylene hydrogenation process is to maximize the selectivity to
14
15 ethylene while maintaining a full conversion of acetylene and maximizing the run-length
16
17 between catalyst regenerations. In this work, a model of the tail-end 3-stage fixed-bed catalytic
18
19
selective hydrogenation reactor was developed and validated to accurately predict the reactor
20
21
22 outlet composition and other important variables. To achieve the optimum operating policy, the
23
24 model-based dynamic optimization was applied to maximize overall process economics through
25
26 enhancement of selectivity to production of ethylene. Implementation of the optimal operating
27
28
29 policy on a commercial acetylene hydrogenation reactor resulted in a 13% improvement of
30
31 ethylene selectivity and 10% increase of overall process economics while simultaneously
32
33 decreasing the rate of catalyst deactivation. This modelling and optimization approach should be
34
35
36
applicable to other fixed-bed hydrogenation processes, such as hydrogenation of methyl
37
38 acetylene and propadiene (MAPD) in propylene product stream.
39
40
41 INTRODUCTION
42
43
44 Olefins plants produce ethylene and propylene from naphtha and LPG feedstock in steam
45
46
47
crackers. The steam cracking process also produces a certain amount of acetylene impurity in the
48
49 ethylene product that is subsequently sent to a downstream polymerization process. A small
50
51 amount of acetylene can be a major cause of deactivation in the polymerization catalyst, and also
52
53
leads to undesired properties in the polymer products. To avoid such problems, the concentration
54
55
56
57
58
59
60 ACS Paragon Plus Environment
2
Page 3 of 30 Industrial & Engineering Chemistry Research

1
2
3 of acetylene impurity in the polymerization feed stream needs to be maintained at less than 1
4
5
6 ppm.
7
8
9 Acetylene is conventionally removed using selective hydrogenation to ethylene, which
10
11 also helps to increase overall plant ethylene yield. An acetylene hydrogenation unit is typically
12
13
14
located at either of two major locations in ethylene plant unit: front-end or tail-end (or back-end).
15
16 The front-end units are upstream from the demethanizer unit and the feed to these reactors
17
18 therefore includes all hydrogen produced in the cracking furnaces. The tail-end units require
19
20
hydrogen injection into the reactor feed. While both configurations have advantages and
21
22
23 disadvantages, the tail-end configuration is generally suitable, less demanding on the equipment,
24
25 catalyst and operation, and results in higher ethylene yields over the long term (Kaiser et al.,
26
27 1999).
28
29
30
31 The tail-end reactor is typically a multi-stage fixed-bed catalytic reactor consisting of
32
33 multiple stages with inter-stage cooling and inter-stage hydrogen injection. In a three-stage setup,
34
35 essentially all of the acetylene conversion takes place in the first two beds, with the last bed
36
37
38
providing protection to ensure that the acetylene is converted to the levels required for product
39
40 purity (a finishing bed). The conventionally used modern catalyst for acetylene hydrogenation is
41
42 a palladium catalyst promoted by silver on alumina support (Pachulski et al. 2012).
43
44
45
Numerous side reactions take place and result in loss of selectivity to ethylene. Over-
46
47
48 hydrogenation leads to ethane formation, which requires separation and recycle to the gas
49
50 cracker, thereby decreasing overall plant capacity and increasing energy consumption. In tail-end
51
52 acetylene hydrogenation units, the oligomerization of acetylene produces so-called Green Oil
53
54
55 (GO) (Vignola et al., 2018) which deposits on the catalyst surface, reducing the overall catalytic
56
57
58
59
60 ACS Paragon Plus Environment
3
Industrial & Engineering Chemistry Research Page 4 of 30

1
2
3 activity and increasing overall operating costs (Zhang et al., 2018). A decrease in activity
4
5
6 towards acetylene hydrogenation can be compensated by operating the reactor under more
7
8 aggressive conditions – for example, higher temperature and higher hydrogen feed rates – but
9
10 this in turn can cause catalyst deactivation to accelerate, leading to an increase in undesired
11
12
13
ethane production. Eventually, the degree of catalyst deactivation is such that continued
14
15 operation is uneconomical and it becomes necessary to take the reactor out of service in order to
16
17 regenerate the catalyst. Optimal determination of operating policy throughout the hydrogenation
18
19
cycle can improve process economics by prolonging catalyst activity and selectivity towards the
20
21
22 main reaction while lowering catalyst deactivation rates. The challenge in operating the
23
24 acetylene hydrogenation reactor is thus to maximize the overall economics by enhancing
25
26 acetylene conversion and selectivity to ethylene while maintaining the product quality and run-
27
28
29 length. In order to maintain the optimized state of the reactor over time, the operating conditions
30
31 need to be carefully adjusted hourly, or at least daily, in order to avoid loss of selectivity and
32
33 conversion.
34
35
36
37
This paper describes the application of an advanced process modelling and dynamic
38
39 optimization approach to a typical tail-end acetylene hydrogenation process in order to determine
40
41 optimal inter-bed conditions that maximize the selectivity of the desired reaction over an entire
42
43
operation cycle. A pseudo-stationary reactor model was constructed in the gPROMS
44
45
46 ProcessBuilder® simulation and optimization software in order to simulate the reactor behavior
47
48 and determine the optimal operating policy for maximizing process economics. The analysis
49
50 demonstrated that, by altering on a daily basis the inlet conditions and thus the conversion in
51
52
53 each of the three beds, it was possible to realize a 13% increase in ethylene gain from the process
54
55 and a 10% improvement in process economics.
56
57
58
59
60 ACS Paragon Plus Environment
4
Page 5 of 30 Industrial & Engineering Chemistry Research

1
2
3 MODEL-BASED OPTIMIZATION APPROACH
4
5
6 The approach described here involves creating a pseudo-stationary model of the reactor
7
8 catalyst beds capable of predicting to a high degree of accuracy how reaction rates and catalyst
9
10 deactivation rates are affected by composition, temperatures and other important attributes. The
11
12
13
kinetic model for the main hydrogenation reactions and for catalyst deactivation were proposed
14
15 and validated against historical plant data to ensure the model accuracy across the range of
16
17 conditions to be investigated. The validated model is then used within gPROMS dynamic
18
19
optimization framework to determine optimal operating conditions over the operation cycle.
20
21
22 Multiple decision variables are varied to determine the combination of values that maximize the
23
24 overall economics of the process. In this case, the decision variables are the time-varying inlet
25
26 temperature to each bed, and the time-varying hydrogen feed rate to each bed, with constraints
27
28
29 including the maximum bed temperatures and outlet C2H2 concentration.
30
31 FIXED-BED REACTOR MODEL
32
33 The commercial acetylene hydrogenation unit considered in this work is a multi-stage
34
35
36
fixed-bed reactor with solid catalyst operated under adiabatic conditions. The reactor model
37
38 therefore needs to consider variations in reaction conditions along the length of each bed, while
39
40 variations of reaction conditions in radial direction can occur only due to mal-distribution of the
41
42
feed or non-uniform properties of the bed, i.e. phenomena that are difficult to predict based on
43
44
45 first-principles. As such a one-dimensional model of the bed is adequate, assuming no variations
46
47 of temperature, composition and pressure over the bed cross sectional area. The catalyst bed
48
49 model is assumed to be pseudo-stationary, with a slow deactivation process being the only
50
51
52 dynamic phenomenon considered in the model to be dependent on time.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
5
Industrial & Engineering Chemistry Research Page 6 of 30

1
2
3 The commercial catalyst loaded in the reactor studied in this work was in a form of
4
5
6 spherical alumina pellets coated with a thin catalytically active layer (i.e. egg-shell type catalyst).
7
8 Initial test simulations with a distributed pellet model confirmed negligible mass transfer
9
10 limitations in such thin layers. This enables simplification of the catalyst model to consider a
11
12
13
lumped catalyst layer, neglecting concentration and temperature distribution across the layer of
14
15 active catalyst.
16
17
18 Since axial dispersion and heat conduction in the bed have negligible effect in industrial
19
20
scale catalytic bed reactors, the required model can be reduced to a pseudo-stationary plug flow
21
22
23 adiabatic heterogeneous catalytic bed model. Such model is described, for example, in (Froment
24
25 and Bischoff, 1990) and can be represented by the following equations:
26
27
28 For component i in the gas phase and gas temperature inside each bed:
29


30
− + 
 −   = 0
31

(1)
32
33

−   + ℎ
 −  = 0
34

35 (2)
36
37
38 with boundary conditions at inlet (z=0):
39

 = , ;  = 
40
41
42
43 For solid phase:
44
45
46   , = 
 −   (3)
47 
48

 −∆,    = ℎ
 − 
49
50 (4)
51 
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
6
Page 7 of 30 Industrial & Engineering Chemistry Research

1
2
3 The pressure profile in each bed is calculated using Ergun’s equation for pressure drop
4
5
6 over bed of catalyst pellets (Froment and Bischoff, 1990).
7
8
9 This model was configured in gPROMS ProcessBuilder using the Advanced Model
10
11 Library for fixed bed reactors (AML: FBCR) by choosing the appropriate simplifying options.
12
13
14
Figure 1 shows an example of specifying catalyst geometry and properties settings in a one-
15
16 dimensional model of a tube uniformly packed with catalyst particles. The model setup in a
17
18 flowsheeting environment of gPROMS ProcessBuilder is shown in Figure 2.
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41 Figure 1. Catalyst geometry and properties specification
42
43
44
45 The inputs to the reactor model included feed flowrate, feed composition, hydrogen
46
47 injection rate (or H2/C2H2 ratio) to each bed and inlet temperature and pressure for each bed. The
48
49 key outputs calculated by the model included product composition, acetylene conversion,
50
51
52 ethylene gain, temperature profile in the beds and economic gain from the process.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
7
Industrial & Engineering Chemistry Research Page 8 of 30

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31 Figure 2. Reactor flowsheet model setup in gPROMS ProcessBuilder.
32
33
34
35 REACTION KINETICS
36
37
38 A key challenge in building a predictive catalytic reactor model is to determine the
39
40 reaction kinetics, i.e. the rate laws and kinetic parameters. Several researchers have reported the
41
42
43
kinetics and mechanism of acetylene selective hydrogenation over the typical Pd/Al2O3 catalyst
44
45 (Menshchikov et al. 1975); Bos et al. 1993; Schbib et al, 1996; Molero et al. 1999; Borodzinski
46
47 and Cybulski, 2000; Mostoufi et al. 2005; Borodzinski and Bond, 2006; Borodzinski and Bond,
48
49
2008) and to much lower extent the Ag-promoted Pd/Al2O3 catalyst (Khan et al. 2006; Huang et
50
51
52 al. 2007; Pachulski et al. 2012). However, the reaction pathways and kinetic parameters reported
53
54 in literature are not only different but also not directly applicable to modeling of the present
55
56
57
58
59
60 ACS Paragon Plus Environment
8
Page 9 of 30 Industrial & Engineering Chemistry Research

1
2
3 industrial reactor since they were obtained at laboratory conditions which often differ
4
5
6 significantly from plant operating, and values of some model parameters are often missing. Most
7
8 importantly, the models were not validated against data obtained at industrial scale.
9
10 The approach for kinetic model development adopted in this work involved testing
11
12
13
various expressions identified in the literature by tuning the key kinetic parameters against
14
15 historical plant data. The reaction scheme was formulated such that it at least includes all
16
17 components that were measured in the plant, which included C2H2, C2H4, C2H6, C4H8 and C6H12.
18
19
The reaction scheme as adopted is shown in Figure 3. Two different active sites are considered:
20
21
22 A-site for reactions directly involving acetylene and E-site for hydrogenation of ethylene to
23
24 ethane (Borodzinski and Cybulski, 2000). Butadiene is generally accepted as an intermediate
25
26 product in oligomerization reactions (Ahn et al., 2007; Egorova et al., 2009), but since butadiene
27
28
29 was not measured in the plant, overall reactions are adopted for production of C4s and C6s from
30
31 acetylene (Sheth et al., 2003). Although the amount of Green Oil (“GO”) was not directly
32
33 measured, its formation is included in the model and it is considered as a precursor for formation
34
35
36
of carbonaceous deposits deactivating the catalyst (Yang et al, 2014; Zhivonitko et al., 2016).
37
38 However, since the amount of GO formed in the bed is negligible compared to the amount of
39
40 process gas, it is assumed that the GO formation reaction does not affect the mass and energy
41
42
balances in the bed (Gislason et al., 2002).
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
9
Industrial & Engineering Chemistry Research Page 10 of 30

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Figure 3. The adopted reaction scheme for acetylene hydrogenation in industrial reactor.
17
18
19 The results of testing various reaction rate expressions from literature did not show
20
21
22 significant differences in model predictions, which is in agreement with conclusions by Bos et al.
23
24 (Bos et al., 1993) where the authors concluded that the rate expressions should be regarded as
25
26 empirical correlations to be used for the prediction of the rates of reaction within the range of
27
28
29
experimental conditions. In addition, the plant data on its own do not provide enough
30
31 information for a model discrimination or for estimation of a large number of parameters.
32
33 The reaction rate equation for acetylene hydrogenation to ethylene and ethylene
34
35
hydrogenation to ethane, as shown in the equation (5) and (6), were adapted from Bos et al. (Bos
36
37
38 et al., 1993). To account for the combined effect of catalyst deactivation and imperfect
39
40 regeneration, activity parameters
and
! were incorporated in the rate expressions. The
41
42 formation of C4s and C6s was considered to take place on active site A and the rate equations
43
44
45 assume a similar form as the acetylene hydrogenation reaction (5) (Pachulski et al., 2012). The
46
47 adapted rate expressions for reactions from the scheme depicted in Figure 3 are summarized in
48
49 equations (5) to (9).
50
$% &' ()*) *)
" =
51
"+,% %
()*) ()*)  "+,*) *) 
52 Acetylene Hydrogenation; (5)
53
$. &) ()*/ *)
- =
54
.  12
Ethylene Hydrogenation; . .
(6)
55 0"+,()*) ()*) +,()*/ ()*/ +,*) *)
56
57
58
59
60 ACS Paragon Plus Environment
10
Page 11 of 30 Industrial & Engineering Chemistry Research

1
2
$% &2 ()*) *)
3 =
3
"+,% %
()*) ()*)  "+,*) *) 
4 Butene Formation; (7)
5
$% &/ ()*) *)
6 4 = % %  
"+,()*) ()*)  "+,*)
Hexene Formation; (8)
7 *)
8
9 All rate constants and adsorption constants in equations (5) to (8) where considered to
10
11 have an Arrhenius-type temperature dependency:
12
13
14 .:,; ' '
89 = 8 BC
 =  7
56
15 < > >?@A
16 for reaction i = 1 – 4 and GO (9)
17
18 ∆*:EF,G ' '
89 = 8 BC
D = D 7
19 56 < > >?@A
for all adsorbing components (10)
20
21
22
23
24
25 CATALYST DEACTIVATION MODEL
26
27
28 Another modeling challenge represents the fact that none of the published literature
29
30
describe a mechanistic model for the combined effect of catalyst deactivation and imperfect
31
32
33 regeneration (Kuhn et al. 2015). There are basically 5 intrinsic mechanisms of catalyst decay:
34
35 poisoning; fouling; thermal degradation (including sintering); loss of catalytic phases by vapor-
36
37 solid and/or solid-solid reactions (including formation of volatile compounds); and attrition
38
39
40 (Butt, 1972; Bartholomew et al. 2006). In addition, a catalyst activity drop was observed after
41
42 each regeneration cycle due to inevitable imperfect regeneration or irreversible changes in
43
44 catalyst properties (Pachulski et al., 2011). Here the main deactivation mechanism is considered
45
46
47 to be fouling by carbon deposition as well as incomplete regeneration (Mosafer et al., 2016). For
48
49 practical reason, the approach taken in this work involved application of a semi-empirical model
50
51 similar to that of Smith et al. (Smith et al. 2003) and its tuning together with the reaction kinetics
52
53
using historical plant data from multiple operation cycles.
54
55
56
57
58
59
60 ACS Paragon Plus Environment
11
Industrial & Engineering Chemistry Research Page 12 of 30

1
2
3 The model assumes that the catalyst activity depends on amount of Green Oil
4
5
6 accummulated on the catalyst, as shown in equations (11) to (14). However, since there was no
7
8 suitable reaction rate equation for GO formation reported in literature, the model development
9
10 approach involved testing various expressions against the historical plant data. The best results in
11
12
13
terms of fit to data and confidence in model parameters were obtained when assuming the GO
14
15 formation to be a first order with respect to acetylene and C6s, and a zero order with respect to
16
17 hydrogen. The drop in activity for both active sites is accounted for by empirical terms that
18
19
consider the number of operation cycles already performed on the current catalyst (HI ) and the
20
21
22 deactivation effect represented by CGO, the cumulative Green Oil concentration in the catalyst
23
24 pellets.
25
26
27 Rate of Green Oil formation: JK =
JK LI-M- LINM"- (11)
28
OIPQ
= JK
29
OR
30 Green Oil accumulation: (12)
31
32 Activity for sites A:
= 7 8&( ( 8" 7 8&% IPQ (13)
33
34 Activity for sites E:
! = 7 8&( ( 8" 7 8&. IPQ S
!T + 1 −
!T  1 − 7 8&. IPQ V (14)
35
36
37
38
39 PARAMETER ESTIMATION AND KINETIC MODEL VALIDATION
40
41
42 Kinetic parameters in selected reaction rate expressions need to be fixed based on
43
44 evidence from the literature or established from available experimental data. Ideally the kinetic
45
46
47 model development is supported by a large number of laboratory experiments with a small
48
49 sample of catalyst pellets, under well-defined and reproducible conditions. However, in this
50
51 work it was not realistic to support the model development with small-scale experimentation. For
52
53
54
operational reasons it was also not practical or safe to conduct a series of experiments with the
55
56 target industrial-scale reactor. The challenge was therefore to develop a useful model based only
57
58
59
60 ACS Paragon Plus Environment
12
Page 13 of 30 Industrial & Engineering Chemistry Research

1
2
3 on evidence from literature and on a detailed model-based analysis of historical plant data from a
4
5
6 number of historical cycles from the existing two reactors.
7
8
9
The controlled variables in the plant data, used as model inputs, included trajectories in
10
11
12 time for the feed rate, feed composition and pressure, as well as the inlet temperature and H2
13
14 flow to each bed. The measured variables, used for model validation, included trajectories in
15
16 time for C2H2 and H2 concentrations at the outlet from each bed, temperature measurements at
17
18
19 the outlet of and inside each bed, and the C4s and C6s product flowrates.
20
21
22
The reaction kinetic parameters together with deactivation parameters were estimated
23
24
25 simultaneously in gPROMS ProcessBuilder using plant data from several cycles. The attempt
26
27 was originally made to estimate a single set of parameters for all three reactor beds. However, it
28
29 was not possible to satisfactorily reproduce the behavior of all three beds with a single model,
30
31
32 and different rate expressions and parameters had to be applied in the third bed. This suggests
33
34 that the rate expressions are not representing all important aspects of the reaction mechanism.
35
36 Moreover, the evidence in the literature suggests that there could be a significant shift in reaction
37
38
39
mechanism at around 2000 ppm of acetylene (Pachulski et al. 2012). The inlet concentration of
40
41 acetylene to Bed 1 ranged from 13,000 to 16,000 ppm, whereas its outlet concentration from Bed
42
43 2 ranged from 500 to 1,000 ppm. Obviously, Bed 3 mainly served as the finishing bed to ensure
44
45
that the final acetylene concentration is less than 1 ppm, which had negligible impact on the
46
47
48 overall process economics. In addition, optimization of this unit would require an extremely
49
50 accurate, predictive model (to <1 ppm levels) and even if such model is available, changes in
51
52 operating policy of the third bed represents a high risk of acetylene leakage. For the ultimate
53
54
55 purpose of hydrogenation rector optimization, it is therefore reasonable to take account of only
56
57
58
59
60 ACS Paragon Plus Environment
13
Industrial & Engineering Chemistry Research Page 14 of 30

1
2
3 Bed 1 and Bed 2. As a result, the operating policy for Bed 3 remained unchanged and all
4
5
6 limitation and constraints, including outlet concentration of acetylene, were considered at outlet
7
8 Bed 2 instead. In conclusion, the following model parameters were estimated for the first and
9
10 second bed:
T! , W , , ! , JK
T
, X$,JK , ",T , -,T , 3,T , 4,T , DI-M- , DM- , DI-M-
!
, DM-
!
. The published
11
12
13 values were used for the remaining parameters as shown in Table 1.
14
15
16
17 Table 1: Kinetic parameters used in parameter estimation. (Reference temperature (Tref) for rate
18 constants and adsorption constants is 310K.)
19
20 Reference for
Parameters Unit Estimated value Initial value
21 initial value
22
23
T! - 0.02 – 1 - -
24
25 I - 0.3 – 15 - -
26
27 - 0.01 – 0.5 - -
28
29 ! - 0.1 – 1.3 - -
30
31 "T mol. kgcat-1.s-1.bar-2 0.015 – 0.025 15.8 Bos, 1993
32
33 -T mol. kgcat-1.s-1.bar-2 0.5 – 1.5 15.6x10-3 Bos, 1993
34
35 3T mol. kgcat-1.s-1.bar-2 8.3x10-4 - -
36
37
4T mol. kgcat-1.s-1.bar-2 8.5x10-4 - -
38
JK
T
39
40
mol. kgcat-1.s-1.bar-2 135 - -

DI-M-
41
42 bar-1 38 1800 Bos, 1993

DM-
43
44 bar-1 1.4 3.95 Bos, 1993

DI-M-
!
45
46 bar-1 1.1 28.7 Bos, 1993

DI-M4
47
!
48 bar-1 0.009 0.04 Bos, 1993

DM-
49
! bar-1 0.001 4.15 Bos, 1993
50
51
52 X$," kJ.mol-1 10.1 10.1 Bos, 1993
53
54 X$,- kJ.mol-1 26.9 26.9 Bos, 1993
55
56
57
58
59
60 ACS Paragon Plus Environment
14
Page 15 of 30 Industrial & Engineering Chemistry Research

1
2

X$,3
3 Pachulski,
4 kJ.mol-1 30 16.3
5
2012

X$,4 16.3 (same as X$,3 )


6
7 kJ.mol-1 30 -

X$,JK
8
9 kJ.mol-1 (-100) – 0 0 -

∆$O ,I-M-
10
11 kJ.mol-1 -14.4 -14.4 Bos, 1993
12
13 ∆$O ,M- kJ.mol-1 0 0 Bos, 1993
14
15 ∆$O ,I-M-
!
kJ.mol-1 -8.0 -8.0 Bos, 1993
16
17
∆$O ,I-M4
!
kJ.mol-1 -36.4 -36.4 Bos, 1993
18
19
20
21
22
In accordance with best practice, the parameter estimation was performed using a subset
23
24 of the available data, with other data reserved for model validation of predictions using the
25
26 estimated parameters. Comparisons with plant data showed that the model reproduced historical
27
28
cycles sufficiently well to be used in subsequent optimization. As an example of model
29
30
31 validation, Figure 4(a) shows the comparison between simulated results of the model against the
32
33 measurements from plant data, in a fitting stage, for concentration of C2H2, H2 and temperature
34
35 at the outlet of Beds 1 and 2 respectively as a function of days on stream (DOS) over a run. In
36
37
38 order to validate the model, figure 4(b) shows the prediction results of the model with tuned
39
40 kinetic parameters in comparison with another set of plant data.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
15
Industrial & Engineering Chemistry Research Page 16 of 30

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
16
Page 17 of 30 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 4. Comparison plots of the results from the model (solid blue line) versus historical data
39
40
41 (red dots) (a) in the fitting stage and (b) in the validating stage.
42
43
44
45
USING THE MODEL TO SIMULATE OPERATIONAL SCENARIOS
46
47
48
49 The developed model was initially used to explore various operating scenarios by
50
51 performing a series of dynamic simulations of the remainder of an ongoing operation cycle. The
52
53 objective was to understand an effect of changing key manipulated variables. For example, the
54
55
56 inlet temperature profile and feed H2 to C2H2 ratio were varied to see the effect on product stream
57
58
59
60 ACS Paragon Plus Environment
17
Industrial & Engineering Chemistry Research Page 18 of 30

1
2
3 compositions, the product outlet conditions, catalyst performance and in particular ethylene gain.
4
5
6 As an example, after operating the reactor for a period, the model was used to simulate the effect
7
8 of changing some operating conditions, i.e. H2 to C2H2 ratio, on Ethylene Gain. Figure 5 shows
9
10 the effects of maintaining the same temperature profile but decreasing the H2 to C2H2 ratio in the
11
12
13
feed. This resulted in an ethylene gain of 5% in the initial change of the run, compared to
14
15 operating with the existing condition.
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
Figure 5. Simulation-based analysis
41
42
43 The simulation results in Figure 5 also show that, although a lower H2 to C2H2 ratio can
44
45 initially result in a higher ethylene gain (higher selectivity), it also enhances catalyst deactivation
46
47 and therefore affects the selectivity later in the cycle. This emphasizes the fact that the whole
48
49
50 operation cycle need to be considered which requires the dynamic optimization runs (Bilimoria
51
52 and Bailey, 1978).
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
18
Page 19 of 30 Industrial & Engineering Chemistry Research

1
2
3 OPTIMIZING OPERATING POLICY
4
5
6 The model was applied to determine the optimal operating policy for an operation cycle
7
8 via dynamic optimization in gPROMS. As discussed above, only Bed 1 and Bed 2 were
9
10 considered in the optimization. Any optimization problem is defined by an objective function,
11
12
13
decision variables and constraints (Shin et al. 20017; Bäumer et al. 2007).
14
15 The objective function to be maximized was the overall economics of the process
16
17 (Economic Gain, ECG) over the whole operation cycle, defined by Equation 15 as an average
18
19
profit from operation cycle minus the fixed costs calculated per day.
20
$ 1 f
21
22 XY Z ] = _ S `I-M4 × Xbℎ\c7d7 Le7
23 
\ ^ T
24
25 + g × `I-M4 × Xbℎ
d7 Le7

6 W$R
26 (15)
27 − `M- × \eh7d edij7V b − −
28 ^ ^HRkR
29
 l
30 −
31 ^
32
33 The Economic Gain considers the profit from produced ethylene, partially profit from
34
35 produced ethane which is recycled back to cracking furnaces, cost of hydrogen injection and the
36
37 fixed costs for catalyst regeneration, for new catalyst and for switching between the reactors.
38
39
40 The decision variables (controls) were the inlet temperature and inlet H2/C2H2 ratio for
41
42 each bed varied over the whole cycle. The time horizon was divided into a number of intervals
43
44 i.e. daily, weekly or biweekly basis, and the optimizer would optimize values for the decision
45
46
47 variables in each interval. The process constraints over the whole run were maximum
48
49 temperatures in the beds and maximum C2H2 concentration at Bed 2 outlet.
50
51 OPTIMIZATION APPROACH
52
53
Initially, the approach was first demonstrated on historical cycles. The model inputs
54
55
56 (flowrate, feed composition, pressure) were read from historical data, then the manipulated
57
58
59
60 ACS Paragon Plus Environment
19
Industrial & Engineering Chemistry Research Page 20 of 30

1
2
3 parameters i.e. inlet temperature and inlet H2/C2H2 ratio were optimized retroactively for the
4
5
6 whole run length. Figure 6(a) shows the example results of the control parameters i.e. outlet
7
8 concentration of C2H2 and H2 at Bed 1 which were still in the operating window during the
9
10 optimization. This approach also demonstrated the expected increase of Economic Gain if the
11
12
13
selected historical cycle had been operated optimally.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49 Figure 6. Demonstration of optimization results of the control parameters (not to scale) based on
50
51
52 (a) historical plant data and (b) initial portion of plant data
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
20
Page 21 of 30 Industrial & Engineering Chemistry Research

1
2
3 In order to perform offline optimization of a current (ongoing) cycle, inputs are read from
4
5
6 historical data until the current time. From the current time onwards, the model inputs are
7
8 assumed fixed (feed rate, feed composition) and the manipulated parameters are optimized. The
9
10 simulated results of the optimized control parameters were shown in Figure 6 (b). In this way an
11
12
13
advice can be provided on the operation of the remainder of the run, including optimal run length
14
15 prediction. For the optimization of a new cycle, the time profile of key inputs, such as flowrate
16
17 and feed composition, need to be assumed and the manipulated parameters are optimized during
18
19
the whole run.
20
21
22 For the current operating cycle described above, the optimization was performed for the
23
24 remainder of the operation cycle after 20 days since start of run. The optimization was repeated
25
26 on a daily basis based on actual values for model inputs – flowrate, feed composition and
27
28
29 pressure. Because of possible long-term effect of some changing operating conditions, the
30
31 optimized results from the model needed to be carefully discussed with plant operators before
32
33 implementation, as it might result in accelerated deactivation of the catalyst in long term.
34
35
36
Therefore, during the optimization runs, a weekly meeting with plant operators was set up in
37
38 order to review the optimization results before implementation.
39
40 OPTIMIZATION RESULTS
41
42
After careful discussion with the plant operators, the model and its recommended optimal
43
44
45 operation policy were tested and validated successfully and then put to real use in the industrial
46
47 reactor. Figure 7 shows optimized trajectories for the manipulated variables after implementing
48
49 the identified optimal strategy in a new cycle. The optimized new cycle is compared to the
50
51
52 immediately preceding (base case) cycle run without any optimization. The results show that the
53
54 optimal trajectories are to operate at a lower temperature than the usual operating policy for an
55
56
57
58
59
60 ACS Paragon Plus Environment
21
Industrial & Engineering Chemistry Research Page 22 of 30

1
2
3 initial period before increasing the temperature when conversion and selectivity in the bed
4
5
6 decrease. The lower temperature increases selectivity in the first bed, thus reducing production of
7
8 unwanted ethane and Green Oil. It also reduces conversion in the first bed, which is taken up by
9
10 spare capacity in the second bed. The lower temperature and corresponding lower rates of
11
12
13
reaction in the first bed have the effect of reducing catalyst deactivation, hence extending
14
15 catalyst lifetime.
16
17 The optimization results suggested to initially decrease the inlet temperature of Bed 1
18
19
while increasing temperature in Bed 2 and appropriately adjusting the H2/C2H2 inlet ratios for
20
21
22 both beds. This resulted in a shift of some conversion of acetylene from Bed 1 to Bed 2 due to a
23
24 significantly higher selectivity of the acetylene hydrogenation to ethylene in Bed 2 when
25
26 compared to Bed 1. The insights obtained in the optimization allowed development of a new
27
28
29 operating policy that generally reduced loss of ethylene by reducing acetylene conversion in the
30
31 first bed and increasing the conversion in the second bed, thereby achieving a better balance in
32
33 their conversion duty.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
22
Page 23 of 30 Industrial & Engineering Chemistry Research

1
2
3
4
5
6 (a)
7
8
9
10
11
12 (b)
13
14
15
16
17
18
19
20
21 (c)
22
23
24
25
26
27
28 Figure 7. The optimized operating policy and results compared to baseline after implementation.
29
30
31 (a) Bed 1 (b) Bed 2 and (c) Total Ethylene Gain
32
33 As shown in Figure 7 (c), by implementing the optimized policy on the new operation,
34
35 total ethylene gain was estimated to improve by 13%, which corresponded to a 10% increase in
36
37
38 economic gain. The total benefit realized was estimated to be approximately 1.0 MUSD/year.
39
40 CONCLUSION
41
42 In this work, a realistic model of the industrial acetylene hydrogenation unit was
43
44
45 developed in gPROMS ProcessBuilder, validated against historical data from several operation
46
47 cycles, and then used to optimize the operating policy. The optimized operating policy included a
48
49 predicted benefit of 1 MUSD/year when compared to a preceding cycle operation. An important
50
51
additional benefit realized from the optimization is a prolonged duration of the operation cycles.
52
53
54 The major benefit of developing and optimizing the model is that it identifies significant
55
56 improvements in economic gain without the need for capital expenditure. The knowledge gained,
57
58
59
60 ACS Paragon Plus Environment
23
Industrial & Engineering Chemistry Research Page 24 of 30

1
2
3 as well as the model itself, can be applied to other fixed bed reactors, such as hydrogenation of
4
5
6 methyl acetylene and propadiene (MAPD) in propylene product stream. There is also potential to
7
8 connect the model to the process historian for a real-time monitoring of catalyst activity.
9
10 NOMENCLATURE
11
12
13
a, a0 catalyst activity and initial (start-of-run) activity
14
15 av catalyst surface to bed volume ratio (m2/m3)
16
17 C molar concentration (mol/m3)
18
19
Ccat cost of catalyst replacement ($)
20
21
22 CGO moles of Green Oil per mass of catalyst (mol/kgcat)
23
24 cp specific heat of the process gas (J/kg.K)
25
26 Creg cost of regeneration cycle ($)
27
28
29 Csw switching cost (loss of ethylene) per cycle ($)
30
31 Ea activation energy (J/mol)
32
33 ECG economic gain from the reactor operation ($/day)
34
35
36
GO Green Oil
37
38 Hads heat of adsorption (J/mol.K)
39
40 hf gas-solid heat transfer coefficient (J/s.m2.K)
41
42
K adsorption constant (for active site A/E and component i) at Tref
43
44
45 k0 reaction rate constant at Tref
46
47 kA, kE deactivation parameters (m3/mol)
48
49 kC kinetic constant for a deactivation due to regeneration
50
51
52 kg gas-solid mass transfer coefficient (m/s)
53
54 NC number of operation cycles on the same catalyst (= number of regenerations - 1)
55
56
57
58
59
60 ACS Paragon Plus Environment
24
Page 25 of 30 Industrial & Engineering Chemistry Research

1
2
3 p partial pressure (bar)
4
5
6 P selling price of each product ($/tonne)
7
8 r rate of reaction (kmol/kg cat.s)
9
10 R gas constant (J/mol.K)
11
12
13
T temperature (K)
14
15 ug gas superficial velocity (m/s)
16
17 Y furnace yield of ethylene from cracking ethane
18
19
z distance from inlet to the reactor bed (m)
20
21
22 Greek letters
23
24  catalyst density (kg/m3)
25
26 τ run length of an operation cycle (days)
27
28
29 Subscripts and superscripts
30
31 A active site A
32
33 E active site E
34
35
36
g gas phase
37
38 i component i
39
40 S solid (catalyst) phase
41
42
43
44
45 REFERENCES
46
47 (1) Kaiser ,V.; Laugier, J. P.; DiCinto, R.; Picciotti, M. Acetylene Removal from Cracked
48
49 Gas: A Technology Overview, AIChE Ethylene Producers Conference 68c (1999).
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
25
Industrial & Engineering Chemistry Research Page 26 of 30

1
2
3 (2) Pachulski, A.; Schödel, R.; Claus, P. Kinetics and reactor modeling of a Pd-Ag/Al2O3
4
5
6 catalyst during selective hydrogenation of ethyne. Applied Catalysis A: General 445-446
7
8 (2012) 107-120.
9
10
11 (3) Menshchikov, V. A.; Falkovich, Y. G.; Aerov, M. E.; Hydrogenation kinetics of
12
13
14
acetylene on a palladium catalyst in the presence of ethylene. Kinet. Catal. 16 (1975)
15
16 1338.
17
18
19 (4) Molero, H.; Bartlett, B. F.; Tysoe, W. T. The Hydrogenation of Acetylene Catalyzed by
20
21
Palladium: Hydrogen Pressure Dependence. Journal of Catalysis 181 (1999) 49–56.
22
23
24
25 (5) Borodzinski, A.; Cybulski, A. The kinetic model of hydrogenation of acetylene–ethylene
26
27 mixtures over palladium surface covered by carbonaceous deposits. Applied Catalysis A:
28
29 General 198 (2000) 51–66.
30
31
32
33 (6) Borodzinski, A.; Bond, G. C. Selective Hydrogenation of Ethyne in Ethene‐Rich
34
35 Streams on Palladium Catalysts. Part 1. Effect of Changes to the Catalyst During
36
37 Reaction. Catalysis Reviews 48 (2006) 91-144.
38
39
40
41
(7) Borodzinski, A.; Bond, G. C. Selective Hydrogenation of Ethyne in Ethene‐Rich
42
43 Streams on Palladium Catalysts, Part 2: Steady‐State Kinetics and Effects of Palladium
44
45 Particle Size, Carbon Monoxide, and Promoters. Catalysis Reviews 50 (2008) 379-469.
46
47
48
(8) Bos, A. N. R.; Bootsma, E. S.; Foeth, F; Sleyster, W. J.; Westerterp, K. R. A kinetic
49
50
51 study of the hydrogenation of ethyne and ethene on a commercial Pd/Al2O3 catalyst;
52
53 Chemical Engineering and Processing 32 (1993) 53-63.
54
55
56
57
58
59
60 ACS Paragon Plus Environment
26
Page 27 of 30 Industrial & Engineering Chemistry Research

1
2
3 (9) Huang, W.; McCormick, J. R.; Lobo, R. F.; Chen, J. G. Selective hydrogenation of
4
5
6 acetylene in the presence of ethylene on zeolite-supported bimetallic catalysts. J. Catal.
7
8 246 (2007) 40–51.
9
10 (10) Khan N. A., Shaikhutdinov S., Freund H. J. Acetylene and ethylene hydrogenation on
11
12
13
alumina supported Pd-Ag model catalyst. Catalysis Letters 108, 2006, 159-164.
14
15 (11) Smith C. M.; Akiyan N.; Bajer J. Measurement of the Green Oil Formation Kinetics and
16
17 Catalyst Deactivation During C2H2 Hydrogenation. AIChE Ethylene Producers
18
19
Conference, 2003, 36–47.
20
21
22 (12) Ahn, I. Y.; Lee, J. H.; Kum, S. S.; Moon S. H. Formation of C4 species in the
23
24 deactivation of a Pd/SiO2 catalyst during the selective hydrogenation of acetylene.
25
26 Catalysis Today 123, 2007.
27
28
29 (13) Froment G. F.; Bischoff K. B. Chemical Reactor Analysis and Design., 2nd ed., John
30
31 Wiley & Sons, 1990.
32
33
34 (14) Bäumler C.; Urban, Z.; Matzopoulos M. Enhanced methods optimize ownership costs for
35
36
37
catalysts. Hydrocarbon Processing 6, 2007, 71-78.
38
39
40 (15) Shin S. B.; Han S. P.; Lee W. J.; Chae H. J.; Lee D. I.; Lee W. H.; Urban Z. Optimize
41
42 terphthaldehyde reactor operations. Hydrocarbon Processing 4, 2007, 83-90.
43
44
45
(16) Mostoufi N.; Ghoorchian A.; Sotudeh-Gharebagh R. Hydrogenation of Acetylene:
46
47
48 Kinetic Studies and Reactor Modelling. International Journal of Chemical Reactor
49
50 Engineering 3 (2005) A14.
51
52
53 (17) Bartholomew C.H.; Farrauto R.J. Fundamentals of Industrial Catalytic Processes, 2nd Ed.,
54
55
56 Wiley (2006), Chapter 5.
57
58
59
60 ACS Paragon Plus Environment
27
Industrial & Engineering Chemistry Research Page 28 of 30

1
2
3 (18) Vignola E., Steinmann S. N., Al Farra A., Vandegehuchte B. D., Curulla D., and Sautet
4
5
6 P., Evaluating the Risk of C–C Bond Formation during Selective Hydrogenation of
7
8 Acetylene on Palladium, ACS Catalysis, 2018, 8 (3), 1662-1671
9
10
11 (19) Zhang X., Sun Z., Wang B., Tang Y., Nguyen L., Li Y., and Tao F. F., C–C Coupling on
12
13
14
Single-Atom-Based Heterogeneous Catalyst, Journal of the American Chemical Society,
15
16 2018, 140 (3), 954-962
17
18
19 (20) Sheth P. A., Neurock M., and, and Smith C. M., A First-Principles Analysis of Acetylene
20
21
Hydrogenation over Pd(111), The Journal of Physical Chemistry B, 2003, 107 (9), 2009-
22
23
24 2017
25
26
27 (21) Schbib N. S., García M. A., Gígola C. E. and Errazu A. F., Kinetics of Front-End
28
29 Acetylene Hydrogenation in Ethylene Production, Industrial & Engineering Chemistry
30
31
32 Research, 1996, 35 (5), 1496-1505
33
34
35 (22) Bilimoria M. R. and Bailey J. E., Dynamic Studies of Acetylene Hydrogenation on
36
37 Nickel Catalysts, ACS Symposium Series, 1978, 65(43), 526–536
38
39
40
41
(23) Yang B., Burch R., Hardacre C., Hu P., and Hughes P., Mechanistic Study of 1,3-
42
43 Butadiene Formation in Acetylene Hydrogenation over the Pd-Based Catalysts Using
44
45 Density Functional Calculations, The Journal of Physical Chemistry C, 2014, 118 (3),
46
47
1560-1567
48
49
50
51 (24) Gislason J., Xia W. and Sellers H., Selective Hydrogenation of Acetylene in an Ethylene
52
53 Rich Flow:  Results of Kinetic Simulations, The Journal of Physical Chemistry A, 2002,
54
55 106 (5), 767-774
56
57
58
59
60 ACS Paragon Plus Environment
28
Page 29 of 30 Industrial & Engineering Chemistry Research

1
2
3 (25) Zhivonitko V. V., Skovpin I. V., Crespo-Quesada M., Kiwi-Minsker L. and Koptyug I.
4
5
6 V., Acetylene Oligomerization over Pd Nanoparticles with Controlled Shape: A
7
8 Parahydrogen-Induced Polarization Study, The Journal of Physical Chemistry C, 2016,
9
10 120 (9), 4945-4953
11
12
13
14
(26) Kuhn M., Lucas M. and Claus P., Long-Time Stability vs Deactivation of Pd–Ag/Al2O3
15
16 Egg-Shell Catalysts in Selective Hydrogenation of Acetylene, Industrial & Engineering
17
18 Chemistry Research, 2015, 54 (26), 6683-6691
19
20
21
(27) Butt J. B., Catalyst Deactivation, Advances in Chemistry, 1972, 109(7), 259–518
22
23
24
25 (28) Mosafer M., Hafizi A., Rahimpour M.R. and Bolhasani A., Optimization of regeneration
26
27 protocol for Pd/Ag/a-Al2O3 catalyst of the acetylene hydrogenation process using
28
29 response surface methodology, Journal of Natural Gas Science and Engineering, 2016,
30
31
32 34, 1382-1391
33
34
35 (29) Egorova S. P., Lamberov A. A., Galimzyanova L. R., Nazarov M. V., Shatilov V. M. and
36
37 Gil’manov Kh. Kh., Deactivation of the Pd–Ag–Al2O3 Catalyst and Ag–Al2O3
38
39
40
Chemisorbent at Joint Operation in the Process of Selective Acetylene Hydrogenation,
41
42 Catalysis in Chemical and Petrochemical Industry, 2009, 1(2), 102–110
43
44
45 (30) Pachulski A., Schödel R. and Claus P., Performance and regeneration studies of Pd–
46
47
Ag/Al2O3 catalysts for the selective hydrogenation of acetylene, Applied Catalysis A:
48
49
50 General, 2011, 400, 14–24
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
29
Industrial & Engineering Chemistry Research Page 30 of 30

1
2
3 GRAPHICAL ABSTRACT
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
30

Potrebbero piacerti anche