You are on page 1of 8

* Unknown * | ACSJCA | JCA11.2.5208/W Library-x64 | manuscript.3f (R5.0.i3:5004 | 2.

1) 2020/02/05 13:43:00 | PROD-WS-121 | rq_1126724 | 11/30/2020 01:22:43 | 8 | JCA-DEFAULT

pubs.acs.org/IECR Article

1 Scaling Up the Performance of a Reactor Model for Hydrotreating


2 Vegetable Oil from Bench-Scale to Pilot-Scale Reactors

3 Alexis Tirado and Jorge Ancheyta*

Cite This: https://dx.doi.org/10.1021/acs.iecr.0c04538 Read Online

ACCESS Metrics & More Article Recommendations

4 ABSTRACT: Scaling up and validation of a trickle-bed reactor model were


5 performed with the experimental information obtained during the hydro-
6 treatment of vegetable oil in a pilot-scale reactor. The scale-up is carried out
7 with a previously developed bench-scale reactor model that demonstrated
8 good agreement with the experimental data. The differences in the
9 characteristics of both reactor scales affected catalyst wetting efficiency,
10 diffusion limitations inside the catalytic particle, and wall effects. These
11 alterations were attributed mainly to the use of larger catalytic particles and
12 higher surface velocities of the feedstock used in the pilot-scale reactor
13 compared with the bench-scale reactor. In addition, the catalyst used in the
14 pilot-scale reactor showed more effectiveness than that in the bench-scale
15 reactor, thus presenting more triglyceride conversion. Simulations with the rector model also indicated that the higher feed flow rate
16 and the large amount of catalyst used in the pilot-scale reactor generate a greater release of heat that makes it difficult to stabilize the
17 reactor to reach steady state.

18

19
■ INTRODUCTION
Biofuels obtained by catalytic hydrotreating (HDT) are
system parameters, where kinetic and reactor models play an
important role for more accurate predictions.4,5
46
47
There are no reports in the literature devoted to reactor 48
20 valuable alternatives to meet the increasing levels of energy
scale-up studies for hydrotreating vegetable oil. The only 49
21 demand as well as overcome the high levels of greenhouse
available works reported experiments carried out in different 50
22 gases present in the world.1,2 The HDT process can be used to reactor scales, using different catalysts and reactor character- 51
23 treat different feedstocks such as petroleum distillates, istics in such a way that various hydrodynamic parameters have 52
24 vegetable oils, and lignin extracts. In the laboratory scale, it not been studied, making it more difficult to evaluate the effect 53
25 is common to use HDT reactors in studies of catalyst of reactor size on reaction behavior.5−7 The objective of this 54
26 screening, effect of feed properties, and collection of kinetic work is to evaluate the performance of a mathematical model 55
27 and hydrodynamic data for further scale-up purposes. During of a fixed-bed reactor for the catalytic hydrotreatment of 56
28 laboratory experiments, depending on the reaction conditions vegetable oil using experimental data obtained in bench-scale 57
29 and reactor characteristics, sometimes accurate data cannot be and pilot-scale reactor studies. 58


30 obtained if the influence of certain phenomena that alter the
31 behavior of the reactive system is not properly studied, such as REACTOR MODEL 59
32 mass and temperature macrogradients (deviation of the idea
33 pattern flow, efficiency of catalytic particle wetting and wall Equations and Assumptions. The mathematical model 60

34 effects) and resistance to mass and energy transfer at the consists of a series of partial differential equations (PDEs), 61

35 interfaces and interior of the catalytic particle. If these effects which are detailed in Table 1. They represent the mass and 62 t1

36 cannot be minimized, at least they have to be quantified to heat balances in the three-phase system (gas, liquid, and solid) 63

37 obtain reliable and repeatable data for the scale-up and design along the trickle-bed reactor over time-on-stream and spatial 64

38 of HDT reactors.3 variable z since the reactive species present a cocurrent 65

39 For the scale-up of fixed-bed reactors, it is necessary to take


40 care of the evaluation in different reactor scales of parameters Received: September 14, 2020
41 such as gas surface velocity (uG), liquid surface velocity (uL), Revised: October 27, 2020
42 and equivalent particle diameter (dpe) to mitigate the risks Accepted: November 20, 2020
43 associated with hydrodynamic changes during the escalation of
44 the process. Inclusion of these hydrodynamic parameters into
45 mathematical models is of great help for the calculation of

© XXXX American Chemical Society https://dx.doi.org/10.1021/acs.iecr.0c04538


A Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Table 1. Equations of the Fixed-Bed Reactor Model

i pG y
Mass Balance Equations

− K iLaLjjjj Hi − CiL zzzz


k i {
G
εG ∂pi u ∂piG
gaseous compounds = − RTG

i piG yz
RTG ∂t G ∂z

L j
+ K i aLjjj H − Ci zzz − fw K iSas(CiL − CiS)
k i {
∂CiL ∂CiL L
gaseous compounds in liquid phase εL ∂t
= − uL ∂z

∂CiL ∂CiL
organic compounds in liquid phase εL ∂t
= − uL ∂z
− fw K iSas(CiL − CiS)
∂CiS N
compounds in solid phase ∈S(1 − ∈B ) ∂t
= − fw K iSas(CiL − CiS) + ∑ j =RL1 ηυ r
j i ,j i ,j
Heat Balance Equations
∂TG ∂T
gas phase εGρG CpG ∂t
= − uGρG CpG ∂zG − fw hGLaL(TG − TL)
∂T ∂T
liquid phase εLρL CpL ∂tL = − uLρL CpL ∂zL − fw hGLaL(TG − TL) − fw hLSas(TL − TS)
∂T N
solid phase εSρS CpS ∂tS = fw hLSas(TL − TS) + ∑ j =RL1 ηjvi , jri , j(−ΔHR )

Figure 1. Kinetic model for hydrotreatment of vegetable oils.

66 downflow. Correlations reported in the literature were used for equations) is only considered in the mass and energy balance 92
67 the evaluation of properties of the compounds at operating equations corresponding to the solid phase, fulfilling the 93
68 conditions as well as mass and energy transfer coefficients. The second assumption. In addition, this term is also influenced by 94
69 dynamic reactor model has been developed and evaluated the effectiveness factor, as well as the mass and energy transfer 95
70 previously with the experimental data obtained during terms are affected by hydrodynamic effects. 96
71 hydrotreatment of Jatropha oil in a bench-scale reactor Reaction Kinetic Model. The kinetic model was 97
72 operating in isothermal mode showing good agreement developed in a previous work with experimental data reported 98
73 between experimental and calculated yields.8 for the hydrotreatment of Jatropha oil carried out in a bench- 99
74 To solve the mass and energy balance equations, the scale reactor under the following reaction conditions: 340−420 100
75 following assumptions were considered in the model: °C reaction temperature, 80 bar hydrogen pressure, and 0.5− 101
76 • Operation in dynamic regime 12 h−1 weight hourly space velocity (WHSV). All of the 102
77 • Chemical reactions take place only on the catalyst experiments were conducted with a NiW/SiO2-Al2O3 catalyst. 103
78 surface The values of the determined kinetic parameters were 104
79 • Evaluation of plug-flow deviation and hydrodynamic calculated by minimization of an objective function based on 105
80 effects (backmixing, catalyst wetting and wall flow) by the differences between experimental and calculated yields. A 106
81 typical theoretical criteria sensitivity analysis on each parameter was also performed to 107
82 • Internal diffusion limitations calculated with effective- guarantee the appropriate set of kinetic parameters.11 108
83 ness factor All of the chemical compounds involved during the 109
84 • Insignificant catalyst deactivation due to short time-on- vegetable oil hydrotreatment were grouped into five lumps, 110 f1
85 stream experiments (100 h) as shown in Figure 1. The feedstock composed of 95% 111 f1
86 The mass and heat balance equations are based on a triglycerides (Tg) is grouped in a single lump. Heavy 112
87 mathematical model reported by Korsten and Hoffmann,9 hydrocarbons (Hv, C15−C18) are produced by hydrodeoxyge- 113
88 which was modified by Mederos and Ancheyta10 by adding the nation and decarbonylation/decarboxylation reactions, middle 114
89 term of accumulation (first term of each equation) based on (Md, C9−C14) and light (Lh, C5−C8) hydrocarbons are 115
90 the fact that the reactor operates under a dynamic regime. The generated by hydrocracking reactions, and oligomerized 116
91 consumption/generation term (last term in solid-phase compounds (Ol, > C18) are produced by triglycerides. 117

B https://dx.doi.org/10.1021/acs.iecr.0c04538
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

118 The kinetic model considers 10 reaction rate coefficients, Table 3. Properties of Catalyst and Operating Conditions
119 and the values of preexponential factors (A) and activation for Different Reactor Scales
t2 120 energies (EA) are reported in Table 2. Reaction heats were
scale bench scale pilot scale
121 calculated for each representative reaction pathway of the
mode isothermal isothermal
122 kinetic model and are also included in Table 2.12
catalyst mass, g 2 100
catalyst shape crushed extrudates
Table 2. Kinetic Parameters and Heat of Reactions for
equivalent particle diameter (dpe), cm 0.023 0.254
Vegetable Oil Hydrotreatment
bed length (LB), cm 3.5 45
reaction j A0j (h−1) EAj (kJ/mol) ΔHLR,j (MJ/mol) reactor diameter (dR), cm 1.3 3
3
bed volume, cm 4.64 318
Tg-Ol (1) 9.8723 × 10 9
128.4918 −1.0235
inlet temperature, °C 420 420
Tg-Hv (2) 157.5590 12.7951 −1.2620a
pressure, MPa 8 8
−0.8039b −1
WHSV, h 1 1
−0.9284c
H2/feed ratio, Nl/l 1500 2500
Tg-Md (3) 1.1004 × 1021 277.9029 −1.0574 −04 −02

Tg-Lh (4) 3.4066 × 109 127.7768 −1.0574 liquid mass flow, g/s 5.55 × 10 2.77 × 10
−03 −03

Ol-Hv (5) 127.1923 20.7014 −0.0795a uL, cmL3/(cmR2 s) 1.01 × 10 7.27 × 10


−05 −03

−0.0732b KSTg, s−1 8.24 × 10 1.51 × 10


0.0317c aS, cmS2/cmR3 126.20 11.86
Ol-Md (6) εL, cmL3/cmR3 0.430 0.288
Hv-Ol (7) 6.1409 × 1012 173.1821 0.0299
Hv-Md (8) 1.3894 × 1019 260.8500 −0.0424
Hv-Lh (9) 0.0424 2.1775 −0.0430
must be clear. One of the main errors during the scale-up and 146
Md-Lh (10) 0.2817 6.5759 −0.0202
a b
modeling of reactors is the anticipated assumption of a plug- 147
HDO: Hydrodeoxygenation. HCO: Hydrodecarbonylation. flow behavior since deviation from this pattern would mask the 148
c
HCO2: Hydrodecarboxylation. results giving a misinterpretation of the experimental data. 149
Different analytical criteria based on reactor parameters are 150
123 Model Solution. For the reactor model solution, the used to delimit the influence of this phenomenon. Mederos et 151
124 method of lines was selected, where the PDEs were reduced al.3 presented a review of criteria to evaluate the effects of 152
125 into ordinary differential equations (ODEs) by discretization radial and axial mass dispersion in packed beds based on the 153
126 in the axial directions using backward finite difference method. LB/dpe ratio. The results of the evaluation of these effects under 154
127 In the solution of the mathematical model, only the length of the operating conditions of each reactor are listed in Table 4. It 155 t4
128 the catalytic bed is considered for the discretization of the can be observed that both experimental reactors operated 156
129 spatial derivatives leaving the independent variable time (t) under ideal behavior since they fulfill the criteria for axial and 157
130 without discretizing. Thus, the ODEs were solved using a radial dispersion. Although the catalyst bed length in the pilot- 158
131 fourth-order Runge−Kutta method and the following initial scale reactor is 12.8 higher than that in the bench-scale reactor, 159
132 conditions to predict the concentration profiles of the lumps as the LB/dpe ratios for both cases have similar magnitudes due to 160
133 well as temperature profiles of the phases by working under the smaller equivalent particle diameter used in the bench-scale 161
134 nonisothermal conditions: reactor. 162
Catalyst Wetting Efficiency. The distribution of the liquid 163
At z = 0 through the catalytic bed influences the performance of the 164

piG = (piG )0 , i = H2 process. As the liquid flow is subjected to a frictional force 165
greater than the gravity force, it will flow more evenly through 166

CiL = (CiL)0 , i = Tg, H 2 the cross section spreading through every interstitial channel 167
available, thus improving contact with the catalytic surface. 168

CiL = 0, i = Ol, Hv, Md, Lh Otherwise, a partial wetting of the catalyst will be obtained, 169
reducing the performance of the catalytic system. Table 4 170
CiS = 0, i = H 2 , Tg, Ol, Hv, Md, Lh shows the ratios of forces of friction and gravity for both 171
reactor scales. It is observed that in both cases, the influence of 172
TG = TL = TS = T0 the force of gravity is greater, which indicates an incomplete 173


wetting of the catalyst.3 It was then decided to estimate the 174

135 RESULTS AND DISCUSSION wetting efficiency as an essential scale-up parameter that 175
determines the extent of catalyst wetting. The wetting 176
136 A three-phase reactor model was used to simulate the efficiency was calculated using the correlation reported by El- 177
137 performance of a pilot-scale reactor hydrotreating vegetable Hisnawi et al.13 178
138 oil. The model results were compared with those obtained in a
139 bench-scale reactor. In both cases, the simulated results were fw = 1.617ReL0.146GaL−0.071 (1) 179
140 compared with the available experimental literature data. The
141 operating parameters used for both reactor scales are shown in where ReL and GaL are Reynolds and Galileo numbers, 180

t3 142 Table 3. respectively, calculated by the following equations: 181

143 Review of Criteria to Ensure Ideal Behavior. Ideal Plug d peuLρL


144 Flow. To carry out the scale-up and design of reactors, ReL =
145 knowledge of the phenomena that occur within the reactor μL (2) 182

C https://dx.doi.org/10.1021/acs.iecr.0c04538
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Table 4. Analytical Criteria to Evaluate the Hydrodynamic Effects in Packed-Bed Reactor


bench-scale pilot-scale
LB uLdR 152.17 > 4.74 × 10−5 177.16 > 3.26 × 10−2
radial dispersion > 0.04
d pe εLDrL fulfilled fulfilled
LB 152.17 > 100 177.16 > 100
axial dispersion d pe
> 100
fulfilled fulfilled
0.0298 < 1 0.0933 < 1
(dP / dz)flow 180μL (1 − εL)2 uL
catalyst wetting = 2 > 1 fw not fulfilled not fulfilled
(dP / dz)gravity ρL d pegεL 4
0.435 0.539
dR 56.52 > 20 11.81 > 20
wall effects d pe
> 20
fulfilled not fulfilled

3
d pegρL2 reactor has around 10% more catalyst wetting than that in the 212
GaL = bench-scale reactor. 213
183
μL2 (3) Wall Effects. In laboratory-scale reactors, maldistributions 214

184 Equation 1 was developed with experimental data in the range commonly occur in the liquid phase, which tend to flow 215

185 of (0.55 × 10−3) ≤ uL ≤ 7.5 × 10−3. The values of superficial toward the reactor walls increasing the dispersion. This is due 216

186 velocity of liquid phase (uL) for bench-scale and pilot-scale to the fact that near the wall, the fluid velocities are higher and 217

187 reactors in this work are 1.01 × 10−3 and 7.27 × 10−3, the conversions are lower as a result of the low catalytic activity 218

188 respectively, which are in between the applicable range of this in it, which severely affects hydrodynamic and kinetic 219

189 correlation. The wetting efficiency for the bench-scale reactor phenomena. Different researchers have related these effects 220

190 is calculated to be 0.435, while that for the pilot-scale reactor is to the dR/dpe ratio; however, it is not easy to know the 221

191 0.539, which agree with the range of 0.12−0.6 for experimental optimum value of this ratio to guarantee negligible wall effects 222

192 reactors reported in the literature.14,15 Although larger particle because the orientation of the catalytic particles influences the 223

193 size is used in the pilot-scale reactor, it is not subjected to a results. Therefore, different criteria have been developed based 224

194 superficial velocity of the liquid high enough to achieve on this relationship.3,17 Based on the criteria given in Table 4, 225

195 complete wetting efficiency as has been reported in different it can be observed that the pilot-scale reactor suffers from wall 226

196 studies.14,15 When working with superficial velocities similar to effects. 227

197 those used in industrial reactors, the wetting efficiency With the results of wetting efficiency and wall effects, it can 228

198 approaches unity. be concluded that under the operating conditions and 229

f2 199 Figure 2 depicts some typical results reported in the characteristics of the catalytic particle and bed, the catalyst 230

200 literature of catalyst wetting efficiency as a function of wetting in the pilot-scale reactor is better than that of the 231
bench-scale reactor; however, it is more influenced by wall 232
effects. 233
Catalyst Effectiveness Factor. The quantification of the 234
internal diffusion effects of the catalyst particle was carried out 235
by the effectiveness factor. The following Thiele modulus 236
equation for nth-order irreversible reaction was used.17

1 ijjj Vp yzzz
237

ij n + 1 yz ρS k in(Ci )
j z jj zz
ϕS jj Sp zz
S n−1

k { k 2 {
Φj =
DeiL (4) 238

where 239

ϵS 1
DeLi = [ L L
]
τ (1/DM ) + (1/ D K ) (5)
i i 240

ϵS = ρS Vg (6) 241
Figure 2. Catalyst wetting efficiency as a function of GmL: (Δ) Alvarez ρB
and Ancheyta,16 (●) Korsten and Hoffmannn,9 and (◊) this work. ρS =
(1 − ϵB) (7) 242
201 superficial mass flow velocity of liquid phase for different The effectiveness factor for each reaction pathway is calculated 243
202 cases: hydrotreatment of heavy crude oil and hydrotreatment using the following equations10,17 244
203 of gas oil.9,16 The data correspond to experiments carried out for Φj < 3 245
204 in pilot-scale reactor with LB = 147.3/dR = 2.94 cm and LB =
66.5 cm/dR = 3 cm, respectively. The results obtained in this tan h(Φj)
205
ηjL =
206 work are also included in the figure. It is clearly seen that when Φj (8) 246
207 the liquid flow rate increases, so does the catalyst wetting
208 efficiency in such a way that large commercial reactors operate for Φj ≥ 3 247
209 with complete wetting of catalyst and that our results follow 1
210 this tendency and are within the reported values. From these ηjL =
Φj (9)
211 values, it is anticipated that the catalytic bed in the pilot-scale 248

D https://dx.doi.org/10.1021/acs.iecr.0c04538
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

249 For the crushed catalyst used in the bench-scale reactor,


250 effectiveness factor was found to be 1, while for the commercial
251 size catalyst used in the pilot-scale reactor, different values
252 were obtained for each reaction pathway, which are presented
t5 253 in Table 5. These values indicate that reaction pathways 2 and

Table 5. Effectiveness Factor for Each Reaction Pathway in


the Pilot-Scale Reactor
pathway value
1 0.9724
2 0.8139
3 0.9829
4 0.9890
5 0.9537
7 0.9926
8 0.9959
9 0.9996
10 0.9989

254 5 present the greatest diffusion limitations in the catalytic


255 particle since they show the lowest values of effectiveness
256 factor (0.814 and 0.953, respectively), while the other
257 pathways have negligible internal diffusion effects (>0.96) at
258 420 °C, 80 bar, and WHSV of 1 h−1.
259 Validation of the Mathematical Model. The simulation Figure 3. Comparison of predicted (lines) and experimental
260 of the bench-scale and pilot-scale reactor models was (symbols) yields in isothermal operation: (a) bench-scale reactor
261 conducted first under isothermal conditions to compare the and (b) pilot-scale reactor: Tg (, ○), Ol ( , ⧫), Hv ( − ·· −,
Δ), Md (---, ◊), Lh ( − · −,*), and Hv + Ol (···, □).
262 predicted yields of the different lumps during catalytic
263 hydrotreatment of vegetable oil. Experimental data for
264 hydrotreatment of Jatropha oil at 420 °C, 8 MPa, and commercial size catalyst, which indicates that the 297

265 WHSV of 1 h−1 were used in both reactor scales for model reactions of formation of this lump (pathways 2 and 298

266 validation.12,18 The differences between the two reactors and 5) presented the lowest values (0.814 and 0.953, 299

267 catalysts are highlighted in Table 3. respectively) compared with those found in the bench- 300

268 For better presentation of the results in both reaction scales, scale reactor of 1 due to the use of a crushed catalyst. 301

269 the following figures include the same simulated results for • The enhanced formation of medium hydrocarbons in 302
270 bench-scale reactor at the top and pilot-scale reactor at the the pilot-scale reactor is due to the reduction of 303
f3 271 bottom. Figure 3 shows the calculated yield profiles of each formation of heavy hydrocarbons, since all reaction 304
272 lump along the catalytic bed under steady-state conditions for pathways involved with this lump do not suffer from 305
273 bench-scale and pilot-scale reactors. The X-axis was normalized limitations in intraparticle diffusion (effectiveness factor 306
274 to 100% of catalyst bed in both cases for proper comparison; greater than 0.98 in both scales). 307
275 however, the catalyst bed length is different, as indicated in The mean absolute error (MAE) was calculated to measure 308
276 Table 3. The product yields at the exit of the reactors are how model simulations fit with experimental data. The MAE 309
t6 277 summarized in Table 6. At a first glance, both simulation values for the bench-scale and pilot-scale reactors were 2.756 310
278 results seem to be quite similar; however, there are indeed and 1.068, respectively. This clearly indicates that the reactor 311
279 some important differences. At the end of the catalytic bed, the model accurately fits the experimental data. 312
280 calculated profiles fit quite well with experimental data, taking Conversion of Triglycerides. Figure 4 illustrates the 313 f4
281 into account that the experimental products were fractionated dynamic triglyceride concentration profiles along the catalytic 314
282 into naphtha (<C9 hydrocarbons), kerosene (C9−C14 hydro- bed for both reactor scales at 420 °C, 8 MPa, and WHSV of 3 315
283 carbons), and diesel (C15−C18 plus > C18) range cuts.18 Some h−1. A higher WHSV was used for the simulations to magnify 316
284 other differences in predictions are: the differences, since at a WHSV of 1 h−1, the predictions 317
285 • Triglycerides are converted a little faster in the pilot- practically overlap. This figure shows how the concentration of 318
286 scale reactor. For instance, at 10% of catalytic bed triglycerides increases as the feed flows through the catalyst 319
287 length, the yield of triglycerides is 31.82% in the bench- particles until the catalytic bed becomes as wet as possible 320
288 scale reactor, while for the pilot-scale reactor, it is under the established operating conditions. It is observed that 321
289 22.68%. This behavior indicates that the 10% lower the concentration of triglycerides in the reactor at pilot scale is 322
290 catalytic wetting efficiency in the bench-scale reactor lower than that at bench scale under the different times 323
291 compared with the pilot-scale reactor has the strongest evaluated. This is attributed to the smaller fraction of liquid 324
292 effect among the other parameters (wall effects, internal volume in the reactor presented at a pilot scale (εL = 0.288) 325
293 diffusion). compared with the bench-scale reactor (εL = 0.450), which 326
294 • The heavy hydrocarbon lump showed a lower yield directly affects the dynamic term in the reactor model. The 327
295 along the catalytic bed for the pilot-scale reactor. This liquid holdup is affected by the lower liquid−solid interfacial 328
296 can be attributed to the effectiveness factor of the area in the pilot-scale reactor (aS = 11.86) with respect to that 329

E https://dx.doi.org/10.1021/acs.iecr.0c04538
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Table 6. Comparison of Experimental Data with Calculated Yields


bench scale pilot scale
experimental calculated experimental calculated
Tg 4.6 × 10−02 2.2 × 10−03 <0.99 3.6 × 10−05
Ol (>C18) 11.952 9.937 8.539
Hv (C15−C18) 57.318 59.852 48.510
Hv + Ol (>C14) 58 57.05
Md (C9−C14) 18.767 23.33 30 31.530
Lh (<C9) 11.5 6.876 12 10.307

Figure 4. Dynamic concentration profiles of triglycerides along the Figure 5. Dynamic gas temperature profiles at nonisothermal
catalytic bed at 420 °C, 8 MPa, and WHSV of 3 h−1: (a) bench scale conditions for (a) bench scale and (b) pilot scale: (···) 3 min, (- -
and (b) pilot scale: 60 s ( − −), 180 s ( ), 300 s (·· -) 10 min, ( ) 30 min, ( − · −) 60 min, ( − · · −) 120 min, ()
), 480 s (). 240 min.

hydrocracking of triglycerides in the pilot-scale reactor, the 348


330 in the bench-scale reactor (aS = 126.20), which is inversely
most exothermic reaction that takes place. Therefore, the heat 349
331 proportional to the diameter of the equivalent particle as
release increases as the feed flows through the catalyst bed. 350
332 observed below17,19
This affects the second half of the catalytic bed, which 351

333 εL = 0.185 ϵB aS0.333χ 0.22 (10) undergoes more pronounced stabilization under bench-scale 352
conditions. Moreover, the highest heat release in the reactor 353
6(1 − ϵB) affects the reactive system, which makes it difficult to reach the 354
aS = steady state. The calculated temperature difference between 355
d pe (11)
334 the solid and liquid phases for the pilot-scale reactor was 2 °C, 356
335 It has been reported that the decrease in particle size causes an unlike the profiles reported for the bench-scale reactor of ∼0 357
336 increase in liquid holdup due to a higher capillary pressure.20,21 °C since for the pilot-scale reactor, a commercial particle size is 358
337 Dynamic Simulation of a Nonisothermal Pilot-Scale used, causing a slightly higher difference at the solid−liquid 359
338 Reactor. To analyze the temperature profiles of each phase interface.22,23 360
339 along the catalytic bed under nonisothermal operating Final Remarks. The proposed reactor model was first 361
340 conditions, the heat balance equations were integrated to the validated with experimental data obtained under isothermal 362
f5f6 341 reactor model. Figures 5 and 6 illustrate the predicted conditions in a bench-scale unit.8 Reaction kinetics was 363
342 temperature profiles at an inlet temperature of 420 °C, 8 developed from a series of experiments during hydrotreatment 364
343 MPa, and WHSV of 1 h−1 for bench-scale and pilot-scale of Jatropha oil.11 The model was then used to simulate the 365
344 reactors for gas and liquid phases, respectively. It is observed same small reactor under nonisothermal conditions.23 Now, 366
345 that the pilot-scale reactor presents a greater increase in this model was utilized for predicting the performance of a 367
346 temperature, in both phases. This increase in the upper part of pilot-scale reactor working at isothermal and nonisothermal 368
347 the catalytic bed has been attributed to the accelerated modes. All of these steps constitute the scale-up of the reactor 369

F https://dx.doi.org/10.1021/acs.iecr.0c04538
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research


pubs.acs.org/IECR Article

AUTHOR INFORMATION 402

Corresponding Author 403


Jorge Ancheyta − Instituto Mexicano del Petróleo, Mexico City 404
07730, Mexico; orcid.org/0000-0001-9626-637X; 405
Email: jancheyt@imp.mx 406

Author 407
Alexis Tirado − Instituto Politécnico Nacional, Centro de 408
Investigación en Ciencia Aplicada y Tecnología Avanzada 409
Unidad Legaria, Mexico City 11500, Mexico; orcid.org/ 410
0000-0002-1259-1835 411

Complete contact information is available at: 412


https://pubs.acs.org/10.1021/acs.iecr.0c04538 413

Notes 414
The authors declare no competing financial interest. 415

■ ACKNOWLEDGMENTS
The authors thank the Mexican Institute of Petroleum for
416

417
supporting this research project. A.T. also thanks Consejo 418
Nacional de Ciencia y Tecnologiá (CONACYT) for the Ph.D. 419
scholarship grant. 420

Figure 6. Dynamic liquid temperature profiles at nonisothermal



A0j
NOMENCLATURE
frequency factor for reaction j, h−1
421

422
conditions for (a) bench-scale and (b) pilot scale: (···) 3 min, (- - -) aL gas−liquid interfacial area per unit reactor volume, 423
10 min, ( ) 30 min, ( − · −) 60 min, ( − · · −) 120 min, ()
240 min.
cmS2/cmR3
aS liquid−solid interfacial area per unit reactor volume, 424
cmS2/cmR3
370 model, which has been demonstrated to be successful in terms Ci concentration of component i, moli/cmL3 425
371 of accuracy of predictions when different phenomena related to Cpi specific heat capacity of compound i, J/(mol K) 426
372 plug-flow pattern, catalyst wetting efficiency, wall effects, and DLei effective diffusivity of compound i inside porous solid, 427
373 internal catalyst diffusion were taken into consideration. cmL3/(cmS s)
374 For the nonisothermal simulations, both reaction scales D Ki Knudsen diffusion coefficient of compound i, cm3/ 428
375 exhibited high ΔT along the catalytic bed, which is not a (cmS s)
376 suitable operation for a commercial reactor. Therefore, these DM i molecular diffusion coefficient of compound i, cm3/ 429
377 results evidence the need of using quenching streams in (cmS s)
378 between the reactor to control the exothermicity of the dpe equivalent particle diameter, cmS 430
379 reaction. This issue is currently under study and will be dR internal reactor diameter, cmR 431
reported in future works.


380 EA j activation energy for j reaction, J/moli 432
fw catalyst wetting efficiency, cm2S, wett/cmS2 433
381 CONCLUSIONS GmL superficial mass flow velocity of liquid phase, gL/ 434
382 A robust mathematical model was used to simulate the (cmR2 s)
383 behavior of fixed-bed reactors, which was demonstrated to be Ga Galileo number, dimensionless 435
384 capable of predicting the yield profiles of catalytic hydrotreat- g gravitational constant, cm2/s 436
385 ment of vegetable oil products for the production of renewable Hi Henry’s law coefficient of component i, MPa cm3/mol 437
386 fuels. The reactor model was validated with experimental hI heat transfer coefficient for interface I, J/(cmS2 s K) 438
387 information obtained in bench-scale and pilot-scale reactors, kj reaction rate coefficient for reaction j, s−1 439
L
388 showing good agreement between real and predicted yields. Ki gas−liquid mass transfer coefficient of component i, 440
389 Different hydrodynamics and diffusional aspects that influence cmL3/(cmI2 s)
S
390 reactor performance were evaluated. It is concluded through Ki liquid−solid mass transfer coefficient of component i, 441
391 analytical correlations that both reactor scales operate under cmL3/(cmS2 s)
392 the plug-flow regime. The pilot-scale reactor presented a LB length of catalyst bed, cmR 442
393 higher wetting efficiency of the catalytic bed and was more NRL number of reactions in the liquid phase, dimensionless 443
394 influenced by wall effects under the operating conditions n reaction rate order, dimensionless 444
395 studied. The use of larger catalytic particles in the pilot-scale pGi partial pressure of component i in the gas phase, MPa 445
396 reactor affected diffusion limitations inside the catalytic R gas law constant, MPa cm3/(mol K) 446
397 particle. All of these phenomena altered the performance of Re Reynolds number, dimensionless 447
398 the hydrotreating reactions. It was determined that the higher r rate of reaction per unit of volume, mol/(cm3 s) 448
399 feed flow rate and the large amount of catalyst used in the Sp total geometric external surface area of catalyst 449
400 pilot-scale reactor generates a greater release of heat that makes particle, cmS2
401 it difficult to stabilize the reactor to reach steady state. T temperature, K 450

G https://dx.doi.org/10.1021/acs.iecr.0c04538
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

451 t time-on-stream, s (8) Tirado, A.; Ancheyta, J. Modeling of a bench-scale fixed-bed 514
452 uf superficial velocity of phase f, cmf3/(cmI2 s) reactor for catalytic hydrotreating of vegetable oil. Renewable Energy 515
453 Vg total pore volume of catalyst particle, cm(G + L)3/gS 2020, 148, 790−797. 516

454 Vp total geometric volume of catalyst particle, cmS3 (9) Korsten, H.; Hoffmann, U. Three-phase reactor model for 517
hydrotreating in pilot trickle-bed reactors. AIChE J. 1996, 42, 1350− 518
455 W wetting factor, dimensionless 1360. 519
456 WHSV weight hourly space velocity (10) Mederos, F. S.; Ancheyta, J.; Elizalde, I. Dynamic modeling and 520
457
458 z axial reactor coordinate, cmR2 simulation of hydrotreating of gas oil obtained from heavy crude oil.


521
Appl. Catal., A 2012, 425−426, 13−27. 522
459 GREEK SYMBOLS (11) Tirado, A.; Ancheyta, J. Defining appropriate reaction scheme 523
for hydrotreating of vegetable oil through proper calculation of kinetic 524
460 ΔHR heat of reaction, J/mol parameters. Fuel 2019, 242, 167−73. 525
461 ∈B catalyst bed void fraction, cm(G + L)3/cmr3 (12) Anand, M.; Farooqui, S. A.; Kumar, R.; Joshi, R.; Kumar, R.; 526
462 ∈S catalyst particle porosity, cm(G + L)3/cms3 Sibi, M. G.; Singh, H.; Sinha, A. K. Kinetics, thermodynamics and 527
463 εf external holdup of phase f, cmf3/cmR3 mechanisms for hydroprocessing of renewable oils. Appl. Catal., A 528
464 η catalyst effectiveness factor, dimensionless 2016, 516, 144−52. 529
465 μ dynamic viscosity, cP (13) El-Hisnawi, A. A.; Dudukovic, M. P.; Mills, P. L. Trickle-Bed 530
466 ρB catalyst bed density, gS/cm3cat Reactors: Dynamic Tracer Tests, Reaction Studies, and Modeling of 531
467 ρf density of phase f at process conditions, gf/cmf3 Reactor Performance. ACS Symp. Ser. 1982, 421−440. 532
468 τ tortuosity factor for catalyst particle, cmL/cmS (14) Bhaskar, M.; Valavarasu, G.; Sairam, B.; Balaraman, K. S.; Balu, 533

469 υ stoichiometric coefficient of reaction, dimensionless K. Three-Phase Reactor Model to Simulate the Performance of Pilot- 534
Plant and Industrial Trickle-Bed Reactors Sustaining Hydrotreating 535
470 ϕS shape factor, dimensionless Reactions. Ind. Eng. Chem. Res. 2004, 43, 6654−6669. 536
471 Φ Thiele modulus, dimensionless (15) Satterfield, C. N. Trickle-bed reactors. AIChE J. 1975, 21, 209− 537
472 χ correlation of the pressure drop of the gas and liquid 228. 538
473 phases, dimensionless (16) Alvarez, A.; Ancheyta, J. Transient behavior of residual oil 539

474

475

f
SUBSCRIPTS AND SUPERSCRIPTS
phase (gas, liquid or solid)
front-end hydrodemetallization in a trickle-bed reactor. Chem. Eng. J.
2012, 197, 204−14.
(17) Froment, Gilbert F.; Bischoff, K. B. Chemical ReactorAnalysis
and Design, 3rd ed.; John Wiley & Sons: Nwy York, 1990; Vol. 2.
540
541
542
543
476 H2 nolecular hydrogen (18) Anand, M.; Farooqui, S. A.; Kumar, R.; Joshi, R.; Kumar, R.; 544
477 Hv heavy compounds Sibi, M. G.; Sihng, H.; Sinha, A. K. Optimizing renewable oil 545
478 i component i hydrocracking conditions for aviation bio-kerosene production. Fuel 546
479 I gas−liquid interface Process. Technol. 2016, 151, 50−58. 547
480 j reaction j (19) Shah, Y. Gas-Fluid-Solid Reactor Design; McGraw-Hill Interna- 548
481 L liquid phase tional Book Co.: London;;New York, 1979. 549
482 Lh light compounds (20) Carruthers, J. D.; DiCamillo, D. J. Pilot plant testing of 550
hydrotreating catalysts. Influence of Catalyst Condition, Bed Loading 551
483 Md middle compounds
and Dilution. Appl. Catal. 1988, 43, 253−276. 552
484 Ol oligomerized compounds (21) Macías, M. J.; Ancheyta, J. Simulation of an isothermal 553
485 Tg triglycerides hydrodesulfurization small reactor with different catalyst particle 554
486 S solid phase shapes. Catal. Today 2004, 98, 243−252. 555
487
488 0 initial condition (22) Perego, C. Experimental methods in catalytic kinetics. Catal. 556


Today 1999, 52, 133−145. 557
489 REFERENCES (23) Tirado, A.; Trejo, F.; Ancheyta, J. Simulation of bench-scale 558
hydrotreating of vegetable oil reactor under non-isothermal 559
490 (1) Kubičková, I.; Kubička, D. Utilization of triglycerides and related conditions. Fuel 2020, 275, No. 117960. 560
491 feedstocks for production of clean hydrocarbon fuels and
492 petrochemicals: A review. Waste Biomass Valorization 2010, 1, 293−
493 308.
494 (2) Sotelo-Boyas, R.; Trejo-Zarraga, F.; Hernandez-Loyo, F. deJ.
495 Hydroconversion of Triglycerides into Green Liquid Fuels, In
496 Hydrogenation; InTech, 2012; Vol. 8, pp 135−152.
497 (3) Mederos, F. S.; Ancheyta, J.; Chen, J. Review on criteria to
498 ensure ideal behaviors in trickle-bed reactors. Appl. Catal., A 2009,
499 355, 1−19.
500 (4) Hickman, D. A.; Weidenbach, M.; Friedhoff, D. P. A comparison
501 of a batch recycle reactor and an integral reactor with fines for scale-
502 up of an industrial trickle bed reactor from laboratory data. Chem. Eng.
503 Sci. 2004, 59, 5425−5430.
504 (5) Tirado, A.; Ancheyta, J.; Trejo, F. Kinetic and reactor modeling
505 of catalytic hydrotreatment of vegetable oils. Energy Fuels 2018, 32,
506 7245−7261.
507 (6) Forghani, A. A.; Jafarian, M.; Pendleton, P.; Lewis, D. M.
508 Mathematical modelling of a hydrocracking reactor for triglyceride
509 conversion to biofuel: model establishment and validation. Int. J.
510 Energy Res. 2014, 38, 1624−34.
511 (7) Muharam, Y.; Adevia, R. T. Modelling and simulation of a slurry
512 bubble column reactor for green fuel production via hydrocracking of
513 vegetable oil. E3S Web Conf. 2018, 67, No. 02032.

H https://dx.doi.org/10.1021/acs.iecr.0c04538
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

You might also like