Вы находитесь на странице: 1из 20

Fuel 90 (2011) 3492–3508

Contents lists available at ScienceDirect

Fuel

j o u r n a l h o m e p a g e : w w w . e l s ev i e r . c o m / l o c a t e / f u e l

Detailed description of kinetic and reactor modeling for naphtha catalytic reforming

Miguel A. Rodríguez, Jorge Ancheyta
Instituto Mexicano del Petróleo, Eje Central Lázaro Cárdenas Norte 152, Col. San Bartolo Atepehuacan, Mexico City 07730, Mexico

article info abstract

Article history: Kinetic and reactor modeling of catalytic reforming of naphtha is described in the present work. The development of a kinetic
Available online 22 June 2011 reforming model is reported with detail. The validation of the developed kinetic model with bench-scale isothermal reactor
experiments is also carried out. The kinetic and reactor mod-els are applied for the simulation of commercial semi-
Keywords: regenerative reforming unit. The effect of benzene precursors in the feed in both laboratory and commercial reactors is also
Catalytic reforming simulated, and the use of the reactor model to predict other process parameters is highlighted.
Reactor modeling
Kinetic modeling 2011 Published by Elsevier Ltd.

1. Introduction without changing the boiling point range of the feed. During this
transformation catalytic reforming produces significant amounts of hydrogen,
Catalytic reforming is a chemical process used to convert petro-leum which is used in other processes such as hydrotreat-ing and hydrocracking, as
naphtha, particularly low-octane number straight-run naph-tha into high- well as small amounts of methane, eth-ane, propane and butanes.
octane gasoline called reformate. In addition to producing reformate catalytic
reforming is also a primary source of aromatics used in the petrochemical Catalytic reforming processes are commonly classified accord-ing to the
industry (BTX: benzene, toluene, and xylenes). frequency and mode of catalyst regeneration, into (1) semi-regenerative, (2)
cyclic regeneration, and (3) continuous regeneration. The main difference
The typical feed to catalytic reforming is a mixture of straight-run among the three types of pro-cesses is the need of unit shut-down for catalyst
naphthas: 30–90 LC light naphtha (C5 and C6), 90–150 LC med-ium naphtha regeneration in the case of semi-regenerative process, the use of an additional
(C7 and C9), and 150–200 LC heavy naphtha (C9 and swing or spare reactor for catalyst regeneration for the cyclic pro-cess, and the
C12). catalyst replacement during normal operation for the continuous regeneration
In commercial practice, the most-preferred feed to catalytic reforming is type. The most-used process worldwide is the semi-regenerative type
naphtha with boiling range of 85–165 LC, since the light fraction (85 LC ) is followed by the continuous regenera-tion and by the less common cyclic
not good feedstock due to its composition (low molecular weight paraffins regeneration. Currently, cata-lytic reformers are mostly designed with
tending to crack to C4– and to be benzene precursor formation, which is continuous regeneration and the former semi-regenerative plants are being
undesirable because of environmental regulations) and the heavy fraction revamped to operate as continuous regeneration.
(180 LC+) hydrocracks to excessive carbon lay-down on the reformer
catalyst.
A large number of reactions occur in catalytic reforming over bifunctional
Prior to catalytic reforming, the naphtha feed needs to be hydrotreated to catalysts, such as dehydrogenation and dehydroiso-merization of naphthenes
reduce the impurities content (sulfur, nitrogen, and oxygen compounds) to to aromatics, dehydrogenation of paraf-fins to olefins, dehydrocyclization of
acceptable levels, which if not removed will poison the reforming catalysts. paraffins and olefins to aromatics, isomerization or hydroisomerization to
This pretreatment is manda-tory since the catalyst is gradually poisoned isoparaffins, isomerization of alkylcyclopentanes and substituted aromatics
leading to excessive coking and rapid deactivation. and hydrocracking of paraffins and naphthenes to lower hydrocar-bons. All
reactions are desirable except hydrocracking, which oc-curs to a greater
To convert the low-quality naphthas, the catalytic reforming process re- extent at high temperature and converts valuable C 5+ molecules (reformate)
arranges the hydrocarbon molecules to form more com-plex molecular shape into light gases.
hydrocarbons with improved octane values. Although certain degree of
cracking occurs the conversion is done Various kinetic models for catalytic reforming reactions have been
reported in the literature [1–28],
which have been subject of several
reviews. The level of sophistication of these models var-ies
⇑ Corresponding author.
E-mail address: jancheyt@imp.mx (J. Ancheyta). from a few lumps to detailed kinetic models, and is related to
0016-2361/$ - see front matter 2011 Published by Elsevier Ltd.
the development of high-speed hardware and large-capacity
doi:10.1016/j.fuel.2011.05.022
M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508 3493

Nomenclature
a, b, c parameters of the hydrogen reaction rate equation N naphthenes with 10 atoms of carbon
10
þ
A aromatics N10 naphthenes with 10 + 11 atoms of carbon
N
A, B, C, D constant for calculating Cp 11 naphthenes with 11 atoms of carbon
A
10 aromatis with 10 atoms of carbon pi partial pressure of component i
þ
A10 aromatis with 10 + 11 atoms of carbon P reaction pressure; paraffins
A
11 aromatis with 11 atoms of carbon Po base reaction pressure
P
Cp molar specific heat 10 paraffins with 10 atoms of carbon
C þ
10 hydrocarbons with 10 atoms of carbon P10 paraffins with 10 + 11 atoms of carbon
þ P
C10 hydrocarbons with 10 + 11 atoms of carbon 11 paraffins with 11 atoms of carbon
C
11 hydrocarbons with 11 atoms of carbon ri rate of reaction of component i
dp particle diameter R ratio of hydrocarbon compositions
E
A activation energy Rg universal constant of gases
F molar flow Rep Reynolds number based on particle diameter
2
gc force to mass conversion factor, 9.8066 (kgm m)/(kgf s ) S cross sectional area
G superficial mass velocity SV space-velocity
o
DG reaction standard Gibbs energy T reaction temperature
T
DH heat of reaction o base reaction temperature
ki kinetic constant at T x conversion
o
k kinetic constant at T o
WABT weighted average bed temperature
k
i

10 kinetic constant for hydrocarbons with 10 atoms of car- WAIT weighted average inlet temperature
þ bon Wci weight fraction of catalyst in each reactor bed respect to
k10
kinetic constant for hydrocarbons with 10 + 11 atoms of the total
carbon WHSV weight hourly space velocity
k
11 kinetic constant for hydrocarbons with 11 atoms of car- yi molar composition of component i
bon z reactor length
K ratio of kinetic constants
Ke equilibrium constant Greek letters
LHSV liquid hourly space velocity e void fraction of the catalyst bed
MW molecular weight q density of the gas mixture
n reaction order qc density of the catalyst
N number of reactors; naphthenes l viscosity of the gas mixture
NC number of components

computers. A chronological evolution of the catalytic reforming ki-netic anticipate the effect of minor changes of process parameters on product yields
modeling is presented in Fig. 1. and quality. If a model with only few lumps is chosen, the predictive
Fig. 2 shows the reaction schemes used in the development of some capability surely is not sufficient to rep-resent the desired situation. But, if a
kinetic models. The rate equations, values of kinetic param-eters, properties detailed mechanistic model is selected, it may be too complex to implement,
of catalyst and feedstock used during experiments and other details can be not because of the solution of the model which with the modern computers
found in the respective references. and algorithms has become a relatively easy task, but due to the cost and
Most of the published kinetic models based on the lumping approach amount of experimental information needed to determine the model
report the rate constants to be dependent on feed and catalyst properties. parameters. Thus, why not use an intermediate ap-proach that maintains the
Some models are neither capable to pre-dict the composition of simplicity of the lumping approach and is detailed enough to correctly predict
alkylcyclopentanes nor the composition of n-paraffins and i-paraffins, nor the the behavior of com-mercial catalytic reforming plant? By intermediate
detailed composition of hydrocracking reaction products, not the whole range approach one must understand that the number of lumps is such that the
of hydro-carbons present in naphtha composition. The level of sophistica-tion composition of the product is predicted with all the desired components. The
varies from just a few lumps to a very detailed kinetic model. On the other answer to this question is the reason why lumped kinetic models are still
hand, sophisticated models based on funda-mental approaches, e.g., single commonly used to characterize reactive groups and describe the reaction
event kinetic model, although overcoming some drawbacks of the lumping kinetics of complex processes in a tractable manner. It is the general
models, still have to be validated under conditions, e.g., other feedstocks, conclusion of all the published scientific papers that the lumping approach is
different from those they were derived, and provide a more convincing sufficiently reliable for describing the relationships between pro-cess
comparison with industrial reality. Also, lumping models involve reduced variables and reaction rates. The comparisons of simulated results by using
number of kinetic parameters and require relatively small amount of lumping kinetic models with data obtained at different scales included those
experimental data for their estimation. While fundamental detailed models are at commercial scale support this conclusion.
quite complicated, with large number of parameters, and frequently need
larger number of experiments. Therefore a kinetic modeler faces to a big
dilemma, either one uses a lumped-kinetic model or a more fundamental
approach. The decision is not easy to make. However, there are some
important points that affect this decision. Most of the times, the model is The aim of this work is to propose a kinetic model for naphtha catalytic
needed to simulate a commercial unit and reforming and to describe in detail how it was developed and used together
with the reactor model to validate experimental information obtained in a
bench-scale isothermal reactor and to simulate commercial catalytic
reformers.
3494 M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508

Stijepovic et al. (2009)


Wei et al. (2008)
Sotelo and Froment (2008)
Shanyingu and Zhu (2004)
Rahimpour et al. (2003) Hou et al.
Taskar and Riggs (1997) Hu et al. (2003) (2006, 2007)

Model
Padmavathi and Chaudhuri (1997) Joshi et al. (1999)
Coppens and Froment (1996) Szczygiel (1999)
Taskar (1996)
Kinetic

Van Trimpont et al. (1988) Turpin (1992)


Marin and Froment (1982) Ramage et al. (1987)

Kmak and Stuckey (1973) Ramage et al. (1980)

Jenkins and Stepehns (1980)


Kmak (1972) Zhorov et al. (1970)
Krane et al. (1959)
Henningsen and Bundgarrd-Nielson (1970)
Smith (1959) Burnett et al. (1965)
Zhorov et al. (1965)

1955 1960 1965 1970 1975 1980 1985 1990 1995 2000 2005 2010
Year

Fig. 1. Evolution of kinetic modeling for catalytic reforming.

+ - P N5 A -
C8 Lumps: C5 N6 C5
C7 Lumps: C5 - P A -
Naphthene + H 2 Aromatics + 3H 2 N5 N6 C5

n Gas Paraffin

3 C6 Lumps: C5
- P N5 N6 A C5
-

Smith (1959) Ramage et al. (1980)


+ +
P9 N9 PX ↔ MX ↔ OX A9+

P8 N8
EB
N7
Alkyl- Alkyl- P7 A7
n-Paraffins
ciclohexanes Benzene P6 N6 A6
(NP)
(ACH) (ACH)
Cracked P5
Products
(0) P4
i-Paraffins Alkyl- P3
ciclopentanes
(IP)
(ACP) P2
Henningsen and Bundgaard-Nielson (1970) P1
Hou et al. (2007)

Fig. 2. Examples of some reaction schemes used to develop catalytic reforming kinetic models.

2. Development of the kinetic model To account for more reactions and have better naphtha compo-sition
predictability, the following modifications to the original model were done.
The kinetic model used for simulation of the catalytic reforming reactors
is an extension of that reported in [1], which utilizes lum-ped mathematical
representation of the reactions that take place. These representations are 2.1. Kinetic parameters for hydrocarbons with 11 atoms of carbon
written in terms of isomers of the same nature (paraffins, naphthenes or
aromatics). These groups range from 1 to 10 carbon atoms for paraffins, and Typical naphthas used as feed to catalytic reforming include hydrocarbons
from 6 to 10 carbon atoms for naphthenes and aromatics. The original with up to 11 atoms of carbon as can be seen in Table 2. The original model
reported model [1] considers reactions only for hydro- carbons with 10 atoms of
[1] includes 53 chemical reactions, which are summarized in Table 1.
carbon. To maintain the original values of
M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508 3495

kinetic parameters it was assumed that the reported hydrocarbons with 10 and 11 atoms of carbon: Cþ10 = C10 + C11. In such a way, the different
atoms of carbon are in fact a lump of hydrocarbons with 10 hydrocarbon species can be de-lumped as:
P10þ ¼ P10 þ P11 ð1Þ
Table 1
Chemical reactions considered in the original kinetic model reported by Krane et al. N10þ ¼ N10 þ N11 ð2Þ
[1], activation energies and factors for pressure effect for each reforming reaction. A10þ ¼ A10 þ A11 ð3Þ
Reaction Number of reactions EAj (kcal/mol)
The originally reported reaction rate equation for each
Paraffins
Pn ? Nn 5 45 hydro-carbon can be then expressed as:
Pn ? Pn i + Pi 21 55
dC10þ
þ þ þ
ð4Þ
Subtotal 26 –
Naphthenes dð1=SVÞ ¼ k C ¼ k10 ðC10 þ C11Þ
10 10
Nn ? A n 5 30
N ?N +P This equation can also be written as function of C 10 and C11 and their
n n i i 6 55
5 45
individual kinetic parameters (k10 and k11) as follow:
Nn ? Pn
Subtotal 16 –
Aromatics
dC10þ ð5Þ

An ? An i + Pi 6 40 dð1=SVÞ ¼ k10C10 þ k11C11


An ? Pn 4 45 By combining Eqs. (4) and (5) the following relationship can be derived:
An ? N n 1 30
Subtotal 11 –
k k10þðR þ 1Þ ð 6
Reaction ak 10 ¼
ðR þ KÞ Þ
Isomerization 0.370 where
Dehydrocyclization 0.700
R ¼ C10 ð 7
Hydrocracking 0.433
Hydrodealkylation 0.500 C11 Þ
Dehydrogenation 0.000
K ¼ k11 ð8Þ
n: number of atoms of carbon (1 6 i 6 5).
k10

Table 2
Typical compositions of various feeds to catalytic reforming process.

Naphthas Average value


1 2 3 4 5 6
n-Paraffins
C4 0 1.568 0 0 0 0 0.261
C5 1.818 11.368 9.818 10.362 1.983 1.392 6.124
C6 9.633 8.034 8.356 8.412 9.467 9.477 8.897
C7 8.116 6.778 7.114 7.148 8.386 8.402 7.657
C
8 6.464 5.326 5.602 5.616 6.640 6.683 6.055
C9 4.454 3.514 3.858 3.809 4.625 4.68 4.157
C
10 1.640 1.403 1.707 1.635 1.948 2.066 1.733
C
11 0.297 0.266 0.321 0.292 0.318 0.370 0.311
Subtotal 32.422 38.257 36.776 37.274 33.367 33.07 35.194
i-Paraffins

C4 0 0.076 0 0 0 0 0.013
C5 0.565 6.459 3.191 3.771 0.794 0.453 2.539
C6 8.868 7.289 7.548 7.495 5.373 5.413 6.998
C7 6.779 5.676 5.973 5.965 6.943 6.963 6.383
C8 7.070 5.897 6.310 6.187 7.289 7.344 6.683
C
9 6.241 5.066 5.499 5.311 6.448 6.509 5.846
C
10 3.526 2.84 3.384 3.221 3.899 4.402 3.545
C
11 0.212 0.203 0.281 0.254 0.289 0.374 0.269
Subtotal 33.261 33.506 32.186 32.204 31.035 31.458 32.275
Naphthenes

C5 0.897 0.977 0.973 0.978 0.333 0.286 0.741


C6 5.069 4.345 4.435 4.434 5.226 5.166 4.779
C7 6.934 6.038 6.071 6.065 7.179 7.157 6.574
C8 5.112 4.307 4.593 4.565 5.320 5.461 4.893
C9 1.842 1.535 1.655 1.578 1.938 1.970 1.753
C
10 0.495 0.398 0.558 0.492 0.561 0.641 0.524
C
11 0.096 0.085 0.106 0.099 0.105 0.125 0.103
Subtotal 20.445 17.685 18.391 18.211 20.662 20.806 19.367
Aromatics

C6 1.393 1.074 1.200 1.199 1.380 1.351 1.266


C7 3.506 2.676 3.024 3.038 3.634 3.576 3.242
C
8 5.326 4.015 4.529 4.542 5.507 5.428 4.891
C
9 2.908 2.186 2.956 2.671 3.488 3.218 2.905
C
10 0.707 0.569 0.903 0.830 0.891 1.056 0.826
C
11 0.032 0.032 0.035 0.031 0.036 0.037 0.034
Subtotal 13.872 10.552 12.647 12.311 14.936 14.666 13.164
3496 M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508

Eqs. (6)–(8) can be used for calculating the values of the individ-ual procedure in a graphic way for two reactions of hydrocracking of paraffins to
kinetic parameters for hydrocarbons with 10 and 11 atoms of carbon, k 10 and paraffins with less number of carbon atoms. The final set of individual kinetic
k11 respectively. For this calculation, the relation-ships defined by Eqs. (7) parameters for the reactions of hydrocar-bons with 10 and 11 atoms of carbon
and (8) are needed. are detailed in Table 3.
To determine the value of the constant R for each hydrocarbon type (Eq.
(7)), typical feed of a commercial catalytic reforming unit was analyzed in 2.2. Reactions for the formation of benzene
different periods of time, whose compositions are reported in Table 2. From
this table, the corresponding values of R for this specific feed are: The original model does not take into account neither the for-mation of
cyclohexane (N6) via methylcyclopentane (MCP) isomer-ization (MCP M N 6)
RP ¼ P10 ¼ 9:109 ð9Þ nor the production of MCP from P 6 (P6 M MCP). The Krane model (1) only
P11 considers the following path reaction: P6 M N6 M A6.
RN ¼ N10 ¼ 5:106 ð10Þ
Due to the importance for accurate prediction of benzene con-tent in
reformate, it is necessary to add those reactions in which benzene is taking
N11
part. Thus, the reaction network shown in Fig. 4, and the corresponding
RA ¼ A10 ¼ 24:414 ð11Þ
contribution to the reaction rate equations were added to the kinetic model.
Also, it was assumed that all the benzene is produced via cyclohexane
A11
dehydrogenation.
For calculation of RP, the total amount of paraffins was used, that is the
sum of n-paraffins plus i-paraffins. 2.3. Isomerization of paraffins
The values of K (Eq. (8)) were obtained for each reaction by extrapolation
of the relationships calculated with the original re-ported kinetic parameters The reactions of isomerization of n-paraffins to i-paraffins are highly
[1] as function of the number of atoms of carbon (e.g., k 7/k6, k8/k7, k9/k8 and desired during catalytic reforming of naphtha, since the pro-duced i-paraffins
k10/k9). Fig. 3 illustrates this contribute to the increase of octane number of

1.8
1.6678 P11 P9+P2
1.7 k11/k 10
1.6 P k10/k9
1.6499 11 P +P
10 11

1.5 k9/k8

1.4
k8/k7
Constant K 1.3
k /k
1.2 6 5

1.1

1.0
k7/k6
0.9

0.8
6 7 8 9 10 11 12 13
Number of atoms of carbon

Fig. 3. Example of the extrapolation procedure to calculate the constant K.

Table 3
Individual kinetic constants for hydrocarbons with 10 and 11 atoms of carbon.
Reaction C6/C5 C7/C6 C8/C7 C9/C8 C /C C11/C10
a
k10
c
k11
c
10 9 kþ b
10
P?N – – 2.2931 1.3609 1.4033 1.4645 0.0254 0.0243 0.0356
Pn ? Pn 1 + P1 1.1667 1.0000 1.3571 1.5789 1.6333 1.6678 0.0049 0.0046 0.0077
Pn ? Pn 2 + P2 1.2000 1.0000 1.3888 1.5600 1.6154 1.6499 0.0063 0.0059 0.0097
Pn ? Pn 3 + P3 – 1.1852 1.3438 1.5814 1.6029 1.6170 0.0109 0.0103 0.0166
Pn ? Pn 4 + P4 – – – 1.5714 1.6182 1.6212 0.0089 0.0084 0.0135
Pn ? Pn 5 + P5 – – – – – – 0.0124 0.0117 0.0191
N?P – 0.1351 2.3500 1.1489 1.0000 1.0000 0.0054 0.0054 0.0054
N?A – 2.2587 2.3678 1.1395 1.0000 1.0000 0.2450 0.2450 0.2450
Nn ? Nn 1 + P1 – – – 14.111 1.0551 1.0000 0.0134 0.0134 0.0134
Nn ? Nn 2 + P2 – – – – 1.0551 1.0000 0.0134 0.0134 0.0134
Nn ? Nn 3 + P3 – – – – – 1.0000 0.0080 0.0080 0.0080
An ? An 1 + P1 – – – 5.0000 1.2000 1.0000 0.0006 0.0006 0.0006
An ? An 2 + P2 – – – – 1.2000 1.0000 0.0006 0.0006 0.0008
A?P – – 1.0000 1.0000 1.0000 1.0000 0.0016 0.0016 0.0016
a
Extrapolated values.
b Original kinetic parameters reported by Krane et al. [1]. c
Calculated values of kinetic parameters.
M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508 3497

the reformate. Isomerization is a fast reaction catalyzed by acid sites and it The effect of temperature on equilibrium constant is given by
reaches the equilibrium at catalytic reforming condi-tions. Hence, paraffins [29]:
distribution can be estimated by thermody-namic equilibrium calculation. DGo
i
lnðKe Þ ¼ ð14Þ
i
For the following general isomerization reaction: Rg T
nPi $ iPi ð12Þ where DGo is the reaction standard Gibbs energy. DGo can be deter-
mined with þ Z Rg
Z Rg T ð Þ
the equilibrium constant (Ke) is Rg T ¼ Rg T0R T þ g T
y y ð13Þ DGo DG0o DH0o D H 0o 1 T
DCp dT T
DCp dT 15
i pi i pi

Kei ¼ y ¼ 1y T
0
T
0
n p i pi
o
To calculate DG with Eq. (15), the following dependence of heat capacity
P6 MCP on temperature can be used:
(C6 Paraffins) (Methyl Cyclopentane)
Cpi ¼ Ai þ BiT þ CiT2 þ DiT3 ð16Þ

By sustituting Eq. (16) in Eq. (15), the integrals can evaluated with the
following final forms:
A6 Z T DB 2 2 DC 3 3
N6 DCpdT ¼ DAðs 1ÞT0 þ ðs 1ÞT þ ðs 1ÞT
0 0
(Benzene) T02 3
(Cyclohexane) þ DD ðs4 1ÞT04 ð17Þ
Fig. 4. Reaction network for benzene formation.
4

Table 4
Thermodynamic data of various paraffins [30].
Name A B C D o o
H G
n-Butane 2.266 7.91E 02 2.65E 05 6.74E 10 30.15 4.10
i-Butane 0.332 9.19E 02 4.41E 05 6.92E 09 32.15 4.99
n-Pentane 0.866 1.16E 01 6.16E 05 1.27E 08 35.00 2.00
2 Methyl butane 2.275 1.21E 01 6.52E 05 1.37E 08 36.92 3.54
2,2 Dimethyl propane 3.963 1.33E 01 7.90E 05 1.82E 08 39.67 3.64
n-Hexane 1.054 1.39E 01 7.45E 05 1.55E 08 39.96 0.06
2 Methyl pentane 2.524 1.48E 01 8.53E 05 1.93E 08 41.66 1.20
3 Methyl pentane 0.570 1.36E 01 6.85E 05 1.20E 08 41.02 0.51
2,2 Dimethyl butane 3.973 1.50E 01 8.31E 05 1.64E 08 44.35 2.30
2,3 Dimethyl butane 3.489 1.47E 01 8.06E 05 1.63E 08 42.49 0.98
n-Heptane 1.229 1.62E 01 8.72E 05 1.83E 08 44.88 1.91
2 Methyl hexane 9.408 2.06E 01 1.50E 04 4.39E 08 46.59 0.77
3 Methyl hexane 1.683 1.63E 01 8.92E 05 1.87E 08 45.96 1.10
2,2 Dimethyl pentane 11.966 2.14E 01 1.52E 04 4.15E 08 49.27 0.02
2,3 Dimethyl pentane 1.683 1.63E 01 8.92E 05 1.87E 08 47.62 0.16
2,4 Dimethyl pentane 1.683 1.63E 01 8.92E 05 1.87E 08 48.28 0.74
3,3 Dimethyl pentane 1.683 1.63E 01 8.92E 05 1.87E 08 48.17 0.63
3 Ethyl pentane 1.683 1.63E 01 8.92E 05 1.87E 08 45.33 2.63
n-Octane 1.456 1.84E 01 1.00E 04 2.12E 08 49.82 3.92
2 Methyl heptane 21.435 2.97E 01 2.81E 04 1.10E 07 51.50 3.05
3 Methyl heptane 2.201 1.88E 01 1.05E 04 2.32E 08 50.82 3.28
4 Methyl heptane 2.201 1.88E 01 1.05E 04 2.32E 08 50.69 4.00
2,2 Dimethyl hexane 2.201 1.88E 01 1.05E 04 2.32E 08 53.71 2.56
2,3 Dimethyl hexane 2.201 1.88E 01 1.05E 04 2.32E 08 51.13 4.23
2,4 Dimethyl hexane 2.201 1.88E 01 1.05E 04 2.32E 08 52.44 2.80
2,5 Dimethyl hexane 2.201 1.88E 01 1.05E 04 2.32E 08 53.21 2.50
3,3 Dimethyl hexane 2.201 1.88E 01 1.05E 04 2.32E 08 52.61 3.17
3,4 Dimethyl hexane 2.201 1.88E 01 1.05E 04 2.32E 08 50.91 4.14
3 Ethyl hexane 2.201 1.88E 01 1.05E 04 2.32E 08 50.40 3.95
2,2,3 Trimethyl pentane 2.201 1.88E 01 1.05E 04 2.32E 08 52.61 4.09
2,2,4 Trimethyl pentane 1.782 1.86E 01 1.02E 04 2.19E 08 53.57 3.27
2,3,3 Trimethyl pentane 2.201 1.88E 01 1.05E 04 2.32E 08 51.73 4.52
2,3,4 Trimethyl pentane 2.201 1.88E 01 1.05E 04 2.32E 08 51.97 4.52
2 Methyl 3 ethyl pentane 2.201 1.88E 01 1.05E 04 2.32E 08 50.48 5.08
3 Methyl 3 ethyl pentane 2.201 1.88E 01 1.05E 04 2.32E 08 51.38 4.76
n-Nonane 0.751 1.62E 01 4.61E 05 7.12E 09 54.74 5.93
2,2,3 Trimethyl hexane 10.899 2.52E 01 1.71E 04 4.75E 08 57.65 5.86
2,2,4 Trimethyl hexane 14.405 2.64E 01 1.84E 04 5.23E 08 58.13 5.38
2,2,5 Trimethyl hexane 12.923 2.62E 01 1.85E 04 5.39E 08 60.71 3.21
3,3 Diethyl pentane 16.067 2.69E 01 1.91E 04 5.51E 08 55.44 8.38
2,2,3,3 Tetramethyl pentane 13.037 2.60E 01 1.81E 04 5.12E 08 56.70 8.20
2,2,3,4 Tetramethyl pentane 13.037 2.60E 01 1.81E 04 5.12E 08 56.64 7.80
2,2,4,4 Tetramethyl pentane 16.099 2.79E 01 2.06E 04 6.15E 08 57.83 8.13
2,3,3,4 Tetramethyl pentane 13.117 2.61E 01 1.82E 04 5.15E 08 56.46 8.15
n-Decane 1.890 2.30E 01 1.26E 04 2.70E 08 59.67 7.94
3,3,5 Trimethyl heptane 16.808 2.94E 01 2.07E 04 5.86E 08 61.80 8.02
2,2,3,3 Tetramethyl hexane 14.052 2.94E 01 2.11E 04 6.17E 08 61.66 11.28
2,2,5,5 Tetramethyl hexane 14.890 2.97E 01 2.14E 04 6.25E 08 68.32
4.66
3498 M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508
Table 5

Example of calculation of Ke for the isomerization of n-hexane (T = 400 K).


DA DB DC DD o o DG
o
Ke
DH0 DG0 RT
n-Hexane M 2 methyl pentane 1.1E 05 3.8E 09 1.7 1.14 1.200
1.47 0.0087 3.32
n-Hexane M 3 methyl pentane 5.95E 06 3.5E 09 1.06 0.45 0.303
0.484 0.0031 1.35
n-Hexane M 2,2 dimethyl butane 8.6E 06 8.5E 10 4.39 2.24 1.890
2.919 0.0113 6.62
n-Hexane M 2,3 dimethyl butane 6.1E 06 7.8E 10 2.53 0.92 0.455
2.435 0.0079 1.58

T D Cp D C 2 2 7 paraffins lumps. That is, 41 new paraffins lumps are incorporated to the
Z T dT ¼ DA lnðsÞ þ DBðs 1ÞT0 þ 2 ðs 1ÞT 0T
model.
0

þ DD ðs3 1ÞT03 ð18Þ


2.4. Effects of pressure and temperature on kinetic parameters
3
where
The model reported by [1] does not include the influence of
s¼ T 19
temperature and pressure on the kinetic parameters. To
T0 ð Þ overcome these limitations, an Arrhenius-type variation of
To apply the above described procedure for equilibrium calcula-tion of the the rate constants with activation energy values for each type
paraffins isomerization reactions that occur in catalytic reforming, some of reaction, and a factor that accounts for the pressure effect
thermodynamic data are required, which are de-picted in Table 4. For
example, for the isomerization of n-hexane, four isomers are obtained: 2 can be used. The equation for the combined effect of
Methyl pentane, 3 Methyl pentane, 2,2 Dimethyl butane, and 2,3 Dimethyl temperature and pressure on kinetic param-eters can be
butane. Each isomerization reaction needs to be considered separately. The
o
calculated values of all parameters required to evaluate DG are reported in expressed as follows:
o ¼ T ð21Þ
Table 5. With DG , Ke is then computed with Eq. (14). Once the values of all ki kiR g T0 P0
0
Ke have been determined, the formulae to calculate the composi-tion (y i) of EAj 1 1 P ak
all isomers is deduced from Eq. (13), which is

The values of activation energies (E Aj) for each reaction j, and pressure
yi ¼ Kei ð 20 Þ effect factors (ak) are given in Table 1. Krane et al. [1] re-ported values of
n

1 P Kei kinetic parameters ðkoiÞ at temperatures ranging from 471 to 515 LC and P o
isomers

þ i¼1 1
= 300 psig, H2-to-oil ratio of 2–8 mol/ mol, and WHSV of 0.7–5.0 h .
The values of Ke and equilibrium composition calculated by using this Because for each reaction, only one value of the corresponding kinetic
procedure for the isomers of paraffins ranging from 6 to 11 atoms of carbon at parameter for each reaction rate was reported in this range of temperatures,
different temperatures are summarized in Table 6. the base temper-ature was considered to be the average of both values (T o =
493 LC).
Splitting the total paraffins lumps, as reported by (1), into n-paraffins and
i-paraffins, allows then for extending the model to consider 48 paraffins plus 2.5. The extended proposed kinetic model
i-paraffins lumps instead of the original
Based on all the previous considerations, a kinetic model with 33 lumps
plus 41 paraffins isomers was developed. The proposed reaction scheme is
Table 6
Equilibrium constants and molar composition for hexane isomerization. shown in Fig. 5. All reactions are assumed to be pseudo-first order with
respect to the hydrocarbon. The reaction rate of each lump is given by the
Reaction Equilibrium constants following equations, which com-prise the kinetic model. Each rate reaction
n-Hexane M 300 K 400 K 500 K 600 K 700 K 800 K
equation is function of the kinetic constants (k i) and the concentration of each
2 Methyl pentane 6.73 3.32 2.20 1.69 1.41 1.24
pseudo-component (Pi, Ni, Ai). The values of all rate constants are reported in
3 Methyl pentane 2.11 1.35 1.04 0.87 0.76 0.69
2,2 Dimethyl 41.83 6.62 2.20 1.07 0.65 0.45 Table 7.
butane
2,3 Dimethyl 4.60 1.58 0.82 0.54 0.39 0.32 Paraffins:
butane
Equilibrium molar composition dP11 ¼ k34N11 þ k58A11 ðk1 þ k2 þ k3 þ k4 þ k5 þ k6ÞP11 ð22Þ
dð1=SVÞ
Isomers 300 K 400 K 500 K 600 K 700 K 800 K
2 Methyl pentane 11.96 23.94 30.30 32.76 33.51 33.59 dP10 ¼ k2P11 þ k39N10 þ k61A10
3 Methyl pentane 3.76 9.76 14.26 16.78 18.09 18.76
2,2 Dimethyl 74.33 47.73 30.34 20.73 15.33 12.08 dð1=SVÞ ð23Þ
butane ðk7 þ k8 þ k9 þ k10 þ k11 þ k12ÞP10
2,3 Dimethyl 8.17 11.36 11.34 10.37 9.36 8.52
butane dP9 ¼ k3P11 þ k8P10 þ k44N9 þ k65A9
n-Hexane 1.78 7.21 13.76 19.36 23.71 27.05
Total 100.00 100.00 100.00 100.00 100.00 100.00 dð1=SVÞ ð24Þ
ðk13 þ k14 þ k15 þ k16 þ k17ÞP9
dP8 ¼ k4P11 þ k9P10 þ k14P9 þ k48N8 þ k68A8
dð1=SVÞ ð25Þ
ðk18 þ k19 þ k20 þ k21 þ k22ÞP8
M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508 3499
A11

(nP11 iP11) N11


A10
(nP10 iP10) A9 N10

(nP9 iP9) A8 N9

(nP8 iP8) N8
A7
(nP7 iP7) N7
MCP
(nP6 iP6) N6 A6

(nP5 iP5)

(nP4 iP4)

P3

P2

P1

Fig. 5. Proposed reaction scheme for catalytic reforming of naphtha.

Table 7
Kinetic parameters of the proposed model (T = 490 LC).

Reaction step 1 Reaction step 1 Reaction step 1


k (h ) k (h ) k (h )
0.0356 0.0018 k 0.2150 k
P11 ? N11 + H2 k1 P7 + H2 ? P5 + P2 25 N8?A8 + 3H2 49
k k
P11 + H2 ? P10 + P1 0.0075 k2 P7 + H2 ? P4 + P3 0.0043 26 N8 + H 2 ? N 7 + P 1 0.0007 50
k k k
P11 + H2 ? P9 + P2 0.0100 3 P6 ? N6 + H2 0.0000 27 N7 + H2 ? P7 0.0019 51
k k
P11 + H2 ? P8 + P3 0.0135 k4 P6 ? MCP + H2 0.0042 28 N7 ? A7 + 3H2 0.0788 52
k k
P11 + H2 ? P7 + P4 0.0135 k5 P6 + H2 ? P5 + P1 0.0018 29 N6 + H2 ? P6 0.0204 53
k k
P11 + H2 ? P6 + P5 0.0191 k6 P6 + H2 ? P4 + P2 0.0016 30 N6 ? A6 + 3H2 0.1368 54
P ?N +H k k k
l0 l0 2 0.0243 7 P6 + H2 ? 2P3 0.0025 31 N6 ? MCP 0.0040 55
k k
Pl0 + H2 ? P9 + P1 0.0015 k8 P5 + H2 ? P4 + P1 0.0018 32 MCP + H2 ? P6 0.0008 56
k k
Pl0 + H2 ? P8 + P2 0.0054 k9 P5 + H2 ? P3 + P2 0.0022 33 MCP ? N6 0.0238 57
k k k
Pl0 + H2 ? P7 + P3 0.0160 10 N11 + H2 ? P11 0.0050 34 A11 + 4H2 ? P11 0.0016 58
k k k
Pl0 + H2 ? P6 + P4 0.0095 11 N11 ? A11 + 3H2 0.6738 35 A11 + H2 ? A10 + P1 0.0006 59
k k k
Pl0 + H2 ? 2P5 0.0095 12 N11 + H2 ? N10 + P1 0.0134 36 A11 + H2 ? A9 + P2 0.0006 60
k k k
P9 ? N9 + H2 0.0500 13 N11 + H2 ? N9 + P2 0.0134 37 A10 + 4H2 ? P10 0.0016 61
k k k
P9 + H2 ? P8 + P1 0.0030 14 N11 + H2 ? N8 + P3 0.0080 38 A10 + H2 ? A9 + P1 0.0006 62
k N + H ?P k k
P9 + H2 ? P7 + P2 0.0039 15 10 2 10 0.0054 39 A10 + H2 ? A8 + P2 0.0006 63
k k k
P9 + H2 ? P6 + P3 0.0068 16 N10 ? A10 + 3H2 0.3198 40 A10 + H2 ? A7 + P3 0.0000 64
k k k
P9 + H2 ? P5 + P4 0.0058 17 N10 + H2 ? N9 + P1 0.0134 41 A9 + 4H2 ? P9 0.0016 65
k k k
P8 ? N8 + H2 0.0266 18 N10 + H2 ? N8 + P2 0.0134 42 A9 + H 2 ? A8 + P 1 0.0005 66
k k k
P8 + H2 ? P7 + P1 0.0019 19 N10 + H2 ? N7 + P3 0.0080 43 A9 + H 2 ? A7 + P 2 0.0005 67
k k k
P8 + H2 ? P6 + P2 0.0056 20 N9 + H2 ? P9 0.0054 44 A8 + 4H2 ? P8 0.0011 68
k k k
P8 + H2 ? P5 + P3 0.0034 21 N9 ? A9 + 3H2 0.2205 45 A8 + H 2 ? A7 + P 1 0.0001 69
k k k
P8 + H2 ? 2P4 0.0070 22 N9 + H 2 ? N 8 + P 1 0.0127 46 A7 + 4H2 ? P7 0.0016 70
k k k
P7 ? N7 + H2 0.0076 23 N9 + H 2 ? N 7 + P 2 0.0127 47 A6 + 3H2 ? N6 0.0015 71

P7 + H2 ? P6 + P1 0.0027 k N8 + H2 ? P8 0.0025 k
24 48

dP7 ¼ k5P11 þ k10P10 þ k15P9 þ k19P8 þ k51N7 þ k70A7 dP5 ¼ k6P11 þ 2k12P10 þ k17P9 þ k21P8 þ k25P7 þ k29P6
dð1=SVÞ dð1=SVÞ
ðk23 þ k24 þ k25 þ k26ÞP7 ð26Þ ðk32 þ k33ÞP5 ð28Þ

dP6 ¼ k6P11 þ k11P10 þ k16P9 þ k20P8 þ k24P7 þ k53N6 dP4 ¼ k5P11 þ k11P10 þ k17P9 þ 2k22P8 þ k26P7 þ k30P6 þ k32P5

dð1=SVÞ dð1=SVÞ
þ k56MCP ðk27 þ k28 þ k29 þ k30 þ k31ÞP6 ð27Þ ð29Þ
3500 M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508

dP3 ¼ k4P11 þ k10P10 þ k16P9 þ k21P8 þ k26P7 þ 2k31P6 dA7 ¼ k52N7 þ k64A10 þ k67A9 þ k69A8 k70A7 ð44Þ
dð1=SVÞ ð30Þ dð1=SVÞ
þ k33P5 þ k38N11 þ k43N10 þ k64A10 dA6 ¼ k54N6 k71A6 ð45Þ
dP2 ¼ k3P11 þ k9P10 þ k15P9 þ k20P8 þ k25P7 þ k30P6 dð1=SVÞ
dð1=SVÞ The hydrogen balance can be derived from the stoichometry of the
þ k33P5 þ k37N11 þ k42N10 þ k47N9 þ k60A11 ð31Þ reactions in which it is involved, either those that consume hydrogen or those
that produce hydrogen. The resulting reaction rate for hydrogen is then:
þ k63A10 þ k67A9
7 6 6
dP1 ¼ k2P11 þ k8P10 þ k14P9 þ k19P8 þ k24P7 þ k29P6 dH2
¼X aP þ X bN X
dð1=SVÞ dð1=SVÞ i¼1 i 12 i i¼1 i 12 i i¼1ciA12 i k56MCP
þ k32P5 þ k36N11 þ k41N10 þ k46N9 þ k50N8 ð32Þ ð46Þ

þ k59A11 þ k62A10 þ k66A9 þ k69A8 The values of ai, bi, and ci, are reported in Table 8.
Naphthenes:
ð33Þ 3. Results and discussions
dN11 ¼ k1P11 ðk34 þ k35 þ k36 þ k37 þ k38ÞN11
dð1=SVÞ 3.1. Validation of the kinetic model with bench-scale reactor
dN10 ð34Þ experiments
¼ k7P10 þ k36N11 ðk39 þ k40 þ k41 þ k42 þ k43ÞN10
dð1=SVÞ 3.1.1. Experiments in an isothermal bench-scale reactor
To validate the modified kinetic model various experiments were
dN9 ¼ k13P9 þ k37N11 þ k41N10 ðk44 þ k45 þ k46 þ k47ÞN9
conducted with a hydrodesulfurized straight-run naphtha recovered from a
dð1=SVÞ ð35Þ commercial naphtha HDS unit (82–168 LC distil-lation range, 0.74 g/mL
density at 20 LC, and <0.5 wppm sulfur). The detailed composition of this
feed for catalytic reforming exper-iments determined by GC analysis is
dN8 ¼ k18P8 þ k38N11 þ k42N10 þ k46N9 ðk48 þ k49 þ k50ÞN8 presented in the first column of Table 9.
dð1=SVÞ
ð36Þ

dN7 ¼ k23P7 þ k43N10 þ k47N9 þ k50N8 ðk51 þ k52ÞN7 ð37Þ Table 9


Molar composition of the feed for bench-scale reactor and of different reformates obtained at
dð1=SVÞ 2 1
10.5 kg/cm pressure, 6.5 H2/oil molar ratio, and 3.54 h WHSV.

Reaction temperature
dN6 ¼ k27P6 þ k57MCP þ k71A6 ðk53 þ k54 þ k55ÞN6 ð38Þ Feed 490 LC 500 LC 510 LC
dð1=SVÞ n-Paraffins
P 2.20 0.01 0.01 0
dMCP ¼ k28P6 þ k55N6 ðk56 þ k57ÞMCP ð39Þ P
11

10 2.97 0.09 0.00 0


dð1=SVÞ P9 5.09 0.40 0.28 0.18
P
8 6.36 1.22 0.91 0.63
Aromatics: P7 3.20 2.92 2.45 1.97
dA11 ¼ k35N11 ðk58 þ k59 þ k60ÞA11 ð40Þ P6 4.40 5.51 5.22 4.40
P
5 3.80 5.26 4.97 4.85
dð1=SVÞ Total 28.02 15.41 13.84 12.03
i-Paraffins
dA10 ¼ þ ð þ þ þ Þ ð41Þ iP 6.22 0.28 0.17 0.85
k40N10 k59A11 k61 k62 k63 k64 A10 10

dð1=SVÞ iP9 8.31 1.50 1.24 0.67


iP8 6.51 3.76 2.75 2.01
ð42Þ iP7 6.20 8.01 7.29 6.07
dA9 ¼ k45N9 þ k60A11 þ k62A10 ðk65 þ k66 þ k67ÞA9 6.70 9.41 9.50 9.83
iP6
dð1=SVÞ iP5 3.40 6.40 6.19 5.53
Total 37.34 29.36 27.14 24.96
dA8 ¼ k49N8 þ k63A10 þ k66A9 ðk68 þ k69ÞA8 ð43Þ Naphthenes
N
dð1=SVÞ 11 0.40 0.00 0.00 0
N
10 0.60 0.00 0.00 0
Table 8 N
9 3.56 0.01 0.02 0.01
N8 4.71 0.63 0.66 0.38
N7 5.80 0.33 0.31 0.27
N
6 3.21 0.01 0.01 0.01
Coefficients of the balance equation of hydrogen.
MCP 0.42 1.35 1.23 1.15
Total 18.70 2.33 2.23 1.82
i ai bi ci
Aromatics
1 k1 (k2 + k3 + k4 + k5 + k6) 3k35 (k34 + k36 + k37 + k38) 4k58 + k59 + k60 A
11 0.30 0.97 1.10 1.25
2 k7 (k8 + k9 + k10 + k11 + k12) 3k40 (k39 + k41 + k42 + k43) 4k61 + k62 + k63 + k64 A
10 2.70 5.61 5.76 5.99
3 k13 (k14 + k15 + k16 + k17) 3k45 (k44 + k46 + k47) 4k65 + k66 + k67
4k + k A9 4.21 12.53 13.20 14.17
4 k18 (k19 + k20 + k21 + k22) 3k49 (k48 + k50) 68 69
A8 4.71 15.66 16.90 18.2
5 k23 (k24 + k25 + k26) 3k52 k51 4k70
A7 3.22 12.82 13.91 15.02
6 k27 + k28 (k29 + k30 + k31) 3k54 k53 3k71
– – A6 0.80 5.30 5.91 6.56
7 (k32 + k33)
Total 15.94 52.90 56.78 61.19
M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508 3501

PIC
PI

Reactor Separator
On-line
gas
TI chromatography

TI PI
PI
TI

Feedstock

Separator
Feed H
Feedstock 2

vessel dryer dryer Reformate

Compressor

Fig. 6. Experimental bench-scale unit for catalytic reforming experiments.

The tests were carried out in a fixed-bed bench-scale stainless-steel tively. The most important increase is observed for lighter aro-matics,
reactor (2.5 cm internal diameter and 25 cm length) with hydrogen recycle. particularly A6, A7 and A8.
The reactor is operated in isothermal mode by independent temperature Naphthenes react relatively easily and are highly selective to aromatics
control of a three-zone electric furnace. A general diagram of the compounds via dehydrogenation. This reaction pro-ceeds essentially to
experimental setup is depicted in Fig. 6. All the experiments were done at completion. For instance N6, N9, N10 and N11 disappear completely and
2
pressure of 10.5 kg/cm , H2/oil molar ratio of 6.5 and temperatures of 490, the conversions of N7 and N8 are higher than 87%. It is also confirmed
500 and 510 LC. To sim-ulate a series of three reforming reactors, the bench- that naphthenes dehy-drogenation is favored by high reaction temperature
scale reactor was loaded with different amounts of catalyst, i.e. 6, 15 and 30 as they
mL, keeping the same naphtha flowrate at constant value of were almost completely converted at temperatures higher than 490 LC
(>87% conversion of total amount of naphthenes). This is the main reason
102 mL/h in order to have different space-velocities (WHSV): 17.72, 7.09 and why naphthenes are the most desirable com-ponents in reforming
1
3.54 h , respectively. These amounts of catalyst and WHSV were selected in feedstocks.
order to have 20% of the total mass of catalyst in the first reactor, 30 % in the Paraffins isomerization reaction is important because naphtha contains
second reactor and 50% in the third reactor, which are typical percentages of high percentage of n-paraffins, which after isomeriza-tion, yield products
catalyst load-ing in commercial catalytic reforming reactors. with higher octane number. The naphtha used in the present experiments
has high paraffin content: 28.02 mol% n-paraffins and 37.34 mol% i-
The catalyst used in all experiments was a commercial available Pt-Re paraffins. This reaction occurs rapidly at commercial operating
reforming sample (0.29 wt.% Pt, 0.29 wt.% Re), having specific surface area temperatures and is limited by the thermodynamic equilibrium. The
2
of 221 m /g, pore volume of 0.36 mL/g and particle diameter of 1.6 mm. The temperature has little influence on it because the heat of reaction is low.
catalyst beds were diluted with an inert material both with the same particle
size in order to have a better distribution of heat losses over the reactor, so The most difficult reaction to promote is the dehydrocyclization of
that uniformity of the temperature can take place more easily. The degree of paraffins, which consists of molecular arrangements of paraf-fins to
dilution was varied depending on the amount of catalyst loaded to the reac- naphthenes. Heavy paraffins (P9, P10 and P11) have con-versions higher
tor. The highest dilution was used for experiments with 20% of the total mass than 92% and lighter paraffins showed lower values. This is because the
of catalyst. The temperature drop, measured with an ax-ial thermocouple, was increase in the probability of ring for-mation is high as the molecular
less than 5 LC thus the operation can be as-sumed as isothermal. weight of the paraffin increases. Similarly to naphthenes dehydrogenation
reaction, paraffins dehydrocyclization is favored at high reaction
temperatures.
Reformate samples were collected in a high pressure product receiver. The
remaining C4 (butane and lighters) cracking prod-ucts were removed by
distillation afterwards (stabilization). The stabilized reformate was analyzed 3.1.2.2. Validation of the kinetic model. The kinetic model described
on paraffins, i-paraffins, naphth-enes and aromatics by gas chromatography. previously was incorporated into an isothermal pseudo-homoge-neous one-
dimensional reactor model. The use of diluent and prop-er particle size of the
catalytic bed ensures that the data were collected under kinetic regime and
3.1.2. Results of the simulation transport effects can be ne-glected. To evaluate the product composition as
3.1.2.1. Reforming experiments. Table 9 shows the detailed molar function of the reactor length, the mass balance equations were solved with a
composition of the reformate as function of reaction temperature. From this Runge Kutta Method.
table, the following observations can be made:
Fig. 7 presents a comparison of the experimental and predicted molar
The total amount of aromatics increases from 15.94 mol% to 52.90, 56.78 reformate composition for some selected hydrocarbons: nP 5, iP5, nP6, nP7,
and 61.19 mol% at 490, 500 and 510 LC, respec- MCP, N6, N7 and A6) at 510 LC as function of catalyst
3502 M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508

8 catalyst, i.e., catalytic reforming is conducted as a gas–solid reac-tion. Radial


7 and axial dispersion effects can also be neglected since the reactor diameter
and length are much larger than the diameter of the catalyst particle. In
6 addition, at steady-state conditions the catalyst activity can be assumed to be
constant. Thus, the commer-cial semi-regenerative reforming reactors can be
5 A6 represented with the one-dimensional pseudo-homogeneous adiabatic model.
The following ordinary differential equations constitute the reactor model,
mol %

4 which are integrated through each reactor bed to describe the reformate
composition, temperature and pressure profiles along the length of the
3 N7 reactors. The Ergun equation was used to predict the total pressure drop of the
reactors.
2 N6
1 dyi ¼ MWi ðriÞ ð47Þ
dz zWHSV
MCP
0
dT S NR r H
7 ð iÞð D Þ
dz ¼ P i¼1 NC Ri ð48Þ
6
P
i-P5 i¼1FiCpi
dP ¼ 1:75 10 5ð1 eÞ G2 1:5
10 3ð1 eÞ
2
Gl
5 n-P5 þ e3
dz e 3
qdP gc qdP2gc
%

4 P6 ð49Þ
mol

3 P Eqs. (47)–(49) are solved simultaneously with the kinetic model rate
7 equation for each component (ri) described previously using a fourth-order
Runge–Kutta method. To solve the energy balance equation, heats of reaction
2 (DHRi) are necessary which are deter-mined with the following equations:
1
X X ð50Þ
0 D HR ¼ mpHfpmr Hfr
ZT
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Hfi ¼ Hfi0 þ
298K
Fractional catalyst weigth CpdT ð51Þ
Fig. 7. Experimental (d) and predicted (–) composition of the reformate obtained at 510 LC in 2 3 ð52Þ
Cp ¼ A þ BT þ CT þ DT
the bench-scale reactor.

The values of heats of formation ðH0fÞ, and constants (a, b, c, and d) for
bed length. The following conclusions can be pointed out from this calculating the specific heat (Cpi) are reported in Table 10.
comparison:
3.2.2. Reaction conditions of the commercial reformer
The predicted product molar composition agrees very well with The commercial semi-regenerative reforming unit has four reactors with
experimental data. The average deviation between experimen- interstage heaters, and operates at the following reaction conditions: 495 LC
tal and predicted values is less than 3%. 2
inlet temperature, 10.5 kg/cm reactor pressure, 6.3 mol/mol hydrogen-to-oil
As the naphtha passes through the catalyst bed, A6 concentra-tion ratio, and feedstock flow rate of 30,000 BPD. The hydrodesulfurized straight-
increases. The same behavior is found for all aromatics run naphtha used as feedstock has the following main properties: 0.7406 spe-
compounds.
cific gravity 60/60 LF, 104.8 g/mol molecular weight, 88–180 LC distillation
The concentrations of N6 and N7 and heavy paraffins (P7–P11, only P7 is range, 59.11 mol% total paraffins, 20.01 mol% total naphthenes, and 20.88
shown in the figure) decrease as they undergo conver-sion. High rate of mol% total aromatics. The catalyst used in the commercial reforming unit is
conversion of naphthenes is found in the first 30 % of the catalyst bed. the same than that employed in the bench-scale experiments described
After 60 % catalyst bed, naphthenes concentration approaches a very low previously.
steady-state value.
The relative reaction rates of naphthenes and paraffins are dif-ferent in the The main characteristics of each of the four reactors of the semi-
first 20–30 % of the catalyst bed. While N6 and N7 are almost totally regenerative reforming unit are detailed in Table 11. The inverse of weight
converted in this section, MCP and paraf-fins have low conversion. This hourly space velocity (100/WHSV) is reported in this table since this
means that MCP is much less reactive than N6 and N7. parameter is more common to indicate the increased reaction severity and
reactants reactor positions. As can be seen the first reactor is always smaller
The calculated A6 (benzene) composition matches very well with than the other reactors and the last reactor is the largest. This difference in
experimental data with a maximum deviation of 2%, which is the result of reactor sizes is because some of the reactions occurring in the first reactors
proper description of benzene reaction mechanism. are very fast, and those taking place in the last reactors are slow.

3.2. Simulation of commercial semi-regenerative reforming reactors 3.2.3. Results of the simulation
3.2.3.1. Composition of the reformate. Fig. 8 presents the results of
3.2.1. The reactor model simulated reformate molar composition and the commercial
At the typical catalytic reforming conditions the feed (naphtha) enters the
reactor in gas phase, and is put in contact with the solid
re-ported values at the exit of the fourth reactor for different
hydro-carbons. From this table, the following observations
can be drawn:
M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508 3503
Table 10

Properties of pure compounds contained in reforming feed and products.


o
PM sg Tb (K) PVR (bar) RON MON A B C D H (J/mol)
f
Gases 9.0E 5 2 3 1.381E 5 7.645E 9
Cp = A + BT + CT + DT
Hydrogen 2.016 20.3 – – – 2.714E+1 9.274E 3 0.0
Methane 16.043 0.3000 111.6 – – – 1.925E+1 5.213E 2 1.197E 5 1.132E 8 7.490E+4
Ethane 30.070 0.3564 184.6 – – – 5.409E+0 1.781E 1 6.938E 5 8.713E 9 8.474E+4
Propane 44.094 0.5077 231.1 13.100 98.90 97.10 4.224E+0 3.063E 1 1.586E 4 3.215E 8 1.039E+5
Butane 58.124 0.5844 272.7 3.558 93.80 89.60 9.487E+0 3.313E 1 1.108E 4 2.822E 9 1.262E+5
n-Paraffins 3.626E+0 4.873E 1 2.580E 4 5.305E 8 1.465E+5
Pentane 72.151 0.6310 309.2 1.073 61.70 62.60
Hexane 86.178 0.6640 341.9 0.342 24.80 26.00 4.413E+0 5.820E 1 3.119E 4 6.494E 8 1.673E+5
Heptane 100.205 0.6882 371.6 0.112 0.00 0.00 5.146E+0 6.762E 1 3.651E 4 7.658E 8 1.879E+5
Octane 114.232 0.7068 398.8 0.037 19.00 15.00 6.096E+0 7.712E 1 4.195E 4 8.855E 8 2.086E+5
Nonane 128.259 0.7217 424.0 0.012 17.00 20.00 8.374E+0 8.729E 1 4.823E 4 1.031E 7 2.292E+5
Decane 142.286 0.7342 447.3 0.004 16.01 21.10 7.913E+0 9.609E 1 5.288E 4 1.131E 7 2.498E+5
Undecane 156.313 0.7439 469.1 0.001 14.10 23.00 8.395E+0 1.054E+0 5.799E 4 1.237E 4 2.705E+5
i-Paraffins 1.390E+0 3.847E 1 1.846E 4 2.895E 8 1.346E+5
Isobutane 58.124 0.5631 261.4 4.978 93.00 92.09
2 Methylbutane 72.151 0.6247 301.0 1.409 92.30 90.30 9.525E+0 5.066E 1 2.729E 4 5.723E 8 1.546E+5
2,2 Dimethylbutane 86.178 0.6540 322.8 0.679 95.60 93.40 1.663E+1 6.293E 1 3.481E 4 6.850E 8 1.857E+5
2,2 Dimethylpentane 100.205 0.6782 352.4 0.241 92.80 95.60 5.010E+1 8.956E 1 6.360E 4 1.736E 7 2.063E+5
2,2,4 Trimethylpentane 114.232 0.6962 372.4 0.118 100.00 100.00 7.461E+0 7.779E 1 4.287E 4 9.173E 8 2.243E+5
2,2 Dimethylheptane 128.242 0.7146 405.9 0.029 101.10 100.30 2.089E+1 9.668E 1 6.120E 4 1.570E 7 2.470E+5
3,3,5 Trimethylheptane 142.286 0.7469 428.9 0.012 94.20 90.55 7.037E+1 1.232E+0 8.646E 4 2.455E 7 2.587E+5
Naphthenes 5.011E+1 6.381E 1 3.642E 4 8.014E 8 1.068E+5
Methylcyclopentane 84.162 0.7536 345.0 0.310 91.3 80.0
Cyclohexane 84.162 0.7834 353.8 0.225 83.0 77.2 5.454E+1 6.113E 1 2.523E 4 1.321E 8 1.232E+5
Methylcyclohexane 98.189 0.7740 374.1 0.111 74.8 71.1 6.192E+1 7.842E 1 4.438E 4 9.366E 8 1.549E+5
Ethylcyclohexane 112.216 0.7922 404.9 0.033 45.6 40.8 6.389E+1 8.893E 1 5.108E 4 1.103E 7 1.719E+5
Propylcyclohexane 126.243 0.7977 429.9 0.012 17.8 14.0 6.252E+1 9.889E 1 5.795E 4 1.291E 7 1.934E+5
Butylcyclohexane 140.260 0.8031 454.1 0.004 70.31 68.50 6.296E+1 1.081E+0 6.305E 4 1.400E 7 2.133E+5
Hexylcyclopentane 154.297 0.8006 476.3 0.001 70.00 68.00 5.832E+1 1.128E+0 6.536E 4 1.473E 7 2.096E+5
Aromatics 3.392E+1 4.739E 1 3.017E 4 7.130E 8
Benzene 78.114 0.8844 353.2 0.222 108.00 98.00 8.298E+4
Toluene 92.141 0.8718 383.8 0.071 120.10 105.00 2.435E+1 5.125E 1 2.765E 4 4.911E 8 5.003E+4
Ethylbenzene 106.168 0.8718 409.3 0.025 107.90 97.90 4.310E+1 7.072E 1 4.811E 4 1.301E 7 2.981E+4
Propylbenzene 120.195 0.8665 432.4 0.010 101.50 98.70 3.129E+1 7.486E 1 4.601E 4 1.081E 7 7.830E+3
Butylbenzene 134.222 0.8646 456.5 0.003 100.40 95.50 2.299E+1 7.934E 1 4.396E 4 8.570E 8 1.382E+4
Pentylbenzene 148.250 0.8624 478.6 0.000 110.00 92.00 4.218E+1 9.772E 1 6.242E 4 1.570E 7 3.380E+4
Table 11

Main characteristics of the commercial catalytic reforming reactors.

Reactor 1 2 3 4 Total
Length, m 4.902 5.410 6.452 8.208 24.972
% of reactor length 19.6 21.7 25.8 32.9 100.0
Accumulated % of reactor length 19.6 41.3 67.1 100.0 –
Diameter, m 2.438 2.819 2.971 3.505 –
Catalyst weight, ton 9.13 13.82 22.82 42.58 88.35
wt.% of catalyst bed 10.3 15.7 25.8 48.2 100
Accumulated wt.% of catalyst bed 10.3 26.0 51.8 100 –
1
WHSV, h 16.0 10.6 6.4 3.4 36.4
1 16.0 26.6 33 36.4 –
Accumulated WHSV, h
100/WHSV 6.25 9.43 15.63 29.41 60.72
Accumulated 100/WHSV 6.25 15.68 31.31 60.72 –

Good agreement is obtained between simulated and reported commercial light paraffins is increased because they are produced by hydro-cracking
values. The content of hydrocarbons with ten and or hydrogenolysis.
eleven atoms of carbon and those involved in benzene forma-tion (MCP, The contents of N6–N10 and heavy paraffins (P8–P11) decrease as they
A6, P6 and N6) as well as iso-paraffins are well calcu-lated. Thus the undergo conversion. High conversion of naphthenes is found in the first
reactor and kinetic models predict with good accuracy the reformate and second reactors. After the third reactor, naphthenes content approaches
composition. a very low steady-state value.
As the feedstock passes through the reactor in series the content of all The relative rates of these naphthenes and paraffins conversions are very
aromatic hydrocarbons increases. The most important change is observed different in the first two reactors. While N 6 and N7 are almost totally
for A6, A7, A8 and A9, while for A10 the increase is low and A11 remains converted in this section, paraffins have a very low conversion.
essentially unchanged.
Naphthenes with six and more atoms of carbon react relatively easy and There is an insignificant change in iso-paraffins content in the first and
their reactions proceed essentially to completion. Sim-ilar behavior is second reactors, and in the last two reactors the change is higher. Similar
found for heavy paraffins, i.e. P10 and P11. P8 and P9 also exhibited high behavior is found with light paraffins. This means that the increase in
conversions, while the content of some aromatics content in the third and
3504 M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508

60 A 6 5
50 5 4

40 4 3
mol %

30 i-P 3
2
20 2
n-P MCP 1
10 1 N8
N7
N N N9
0 6
0 N
10
0
6 12 20 A8

5 nP6 10 iP6 16

iP7 A7
4 8 12 A
9
mol %

3 nP5 6

nP7 4 8
2 iP8 A6

1 2 4 A
nP8 iP5 10

nP9 iP9 A
11
nP10
0 0 0
nP11
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
100/WHSV 100/WHSV 100/WHSV

Fig. 8. Profiles of predicted (–) and commercial (s) molar composition of the reformate.

fourth reactors appears to be solely because of the disappear-ance of takes place, the remaining naphthenes are dehydrogenated and a moderate
paraffins. temperature drop is observed (DT 2 = 30 LC). From the second reactor the
effluent is again reheated to 495 LC before entering the third reactor, and
3.2.3.2. Delta-T of reactors. Fig. 9 shows the temperature profiles in the four finally, the same sequence is fol-lowed in the fourth reactor. The temperature
reactors in series. In the first reactor, the major reactions are endothermic and drops across the third and fourth reactors are lower (17 and 14 LC,
fast (e.g., dehydrogenation of naphthenes to aromatics), causing a very sharp respectively), which are due to the exothermic hydrocracking of paraffins
temperature drop from 495 to 443 LC. This remarkable decrease in reaction. Dehydrogenation and cracking reactions take place in the last two
temperature (DT1 = 52 LC) al-most quenches all reactions and more catalyst reactors. The average temperature and the amount of catalyst in the last
reactors are higher that those of the first ones. This al-lows the transformation
in the first reactor would not provide additional conversion.
of paraffins into aromatics by dehydrocyc-lization and into light paraffins by
hydrocracking.
The outlet stream of the first reactor is reheated to the same temperature as
that of the inlet of the first reactor (495 LC) and fed to the second reactor.
The comparison of simulated and actual temperature drops indicates that
Here, mostly the isomerization reaction
the reactor model predictions match well with the information reported in the
commercial reforming unit. The accu-mulated absolute difference between
500 predicted and actual reactor delta-T is smaller than 3 LC.
R-1 R-2 R-3 R-4
490 To =495°C

3.3. Simulation of the effect of benzene precursors in the feed

3.3.1. Preparation of the feed


Reactor temperature,°C

480
470 The feed used in this case is slightly different from that em-ployed in
previous sections. The new feed (Feed A) has the follow-ing main properties:
0.730 specific gravity 60/60 LF, 110 g/mol molecular weight, and 60–190 LC
460 distillation range. Having an ini-tial boiling point (IBP) of 60 LC indicates
that the precursors of ben-zene formation are present in this feed, since the
hydrocarbons involved in the benzene formation network have the following
450 normal boiling points: 80.1 LC benzene, 68.7 LC hexane, 71.8 LC methyl
cyclopentane, and 80.7 LC cyclohexane.

440
To evaluate the effect of benzene formation precursors Feed A was
fractionated to adjust its initial boiling point to 88 LC in order to prepare a
0 10 20 30 40 50 60 70 free-benzene precursors feed. This other feed (Feed B) resulted with the
following main properties: 0.734 specific gravity 60/60 LF, 112 g/mol
100/WHSV
molecular weight, and 88–190 LC distillation range.
Fig. 9. Profiles of predicted reactor temperature (–) and commercial value (s).
The compositions of Feed A and Feed B determined by gas chro-
matography are presented in Table 12. It is clearly observed that Feed A
possesses certain amount of the precursors of benzene for-mation and benzene
itself (1.12 mol% A6, 7.69 mol% P6, 0.64 mol% MCP, and 4.23 mol% N6),
while in Feed B their contents are substan-tially reduced (0.26 mol% A 6, 0.44
mol% P6, 0 mol% MCP, and 3.45 mol% N6).

Because of separation of light compounds, Feed B became a little bit


heavier than Feed A, which is observed as an increase in specific gravity and
molecular weight. i-P5, i-P6 and MCP were totally removed by adjusting IBP of
Feed A, and n-P5 and n-P6 were reduced considerably (>94%). Cyclohexane
(N6) exhibited a reduction of 18.4%, while reduction in benzene (A 6) was of
68.3%.

The main benzene formation precursors are N 6 and MCP, and in less
importance n-P6. By adjusting the IBP of Feed A from 60 to 88 LC, MCP and
n-P6 are almost totally eliminated, but N 6 is still present in high amount in
Feed B. It is expected that at the beginning of the reaction N 6 will not be
produced via MCP isomerization, since Feed B does not have MCP in its
composi-tion. Therefore, benzene will be obtained only by dehydrogena-tion of
the N6 originally present in Feed B. As the reaction proceeds, MCP will be
formed from P6, and N6 and hence A6 pro-duction will increase.

3.3.2. Results of the simulations


For this case, the simulations were carried out first in isother-mal mode of
operation to compare the predicted reformate compo-sition with experimental
values obtained in the bench-scale reactor, and then in adiabatic mode to
predict the behavior of a commercial reforming unit.

Table 12
Molar composition of feeds with and without benzene formation precursors.

Feed A

0.77
2.72
4.05
5.52
6.77
7.69
6.80
34.32

4.10
4.52
6.50
5.64
6.72
2.46
29.94


0.87
3.56
4.04
5.95
4.23
0.64
19.29

0.96
1.34
4.24
5.77
3.02
1.12
16.45

nP9 nP8 nP7 nP6 nP5


n-Paraffins Total
nP11 nP10
i-Paraffins M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508 3505
iP10
iP9 8
iP8 7
iP7
iP6
iP5 6
Total
Naphthenes 5
N N7
11
N

mol %
10

N9 A6
N8 4
N
7
N
MCP
6
N
3 6
Total
Aromatics 2
A
11
A
10 1
A9
A
8
MCP
A7
A 0
6 0 10 20 30 40 50 60 70 80 90 100
Total
Fractional catalyst weight
Fig. 10. Comparison of experimental (d) and predicted (–) molar composition of

the reformate obtained at 510 LC in the bench-scale reactor.


3.3.2.1. Isothermal model predictions versus experimental data. Fig. 10

shows a comparison of experimental and predicted reformate mo-


lar composition of some selected hydrocarbon types (MCP, N6, N7
and A6) as function of the position in the catalyst bed. It is observed
that the calculated compositions agree well with experimental
bench-scale reactor information with average deviation less than
3%. Particularly, the calculated A6 (benzene) composition matches
with high accuracy the experimental data with maximum devia-
tion of 2%.
It is worthy to mention that rate constants of lumped kinetic
models usually depend on feedstock and catalyst properties, and
it may be inappropriate to use them for simulating reforming reac-
Feed B tors for feed conditions different to those from which the parame-
ters were determined. However, if the kinetic model is sufficiently
0.79 detailed, rate constants can be considered to be independent of ini-
3.55 tial feedstock composition, and thus be used to simulate the reac-
5.82 tor for other feed conditions. A problem when using more detailed
7.88 models is that the simplicity of kinetic representations used in
9.75
0.44 models with small number of lumps is partially lost, since the
0.09 use of kinetic models with large number of lumps, where the num-
28.32 ber of parameters is significantly increased, means that greater
amounts of experimental data are also required. Despite this, the
5.53 results shown in Fig. 10 clearly indicate that the developed kinetic
5.20 model is sufficiently detailed to consider the kinetic parameters to
8.39
6.68 be independent of feed composition.

–3.3.2.2. Predictions with the adiabatic model. Fig. 11 presents the ef-
25.80 fect of feed composition on benzene and its formation precursors
hydrocarbon as well as total aromatics content in the reformate

and commercial reactor delta-T, as function of the inverse of
1.15 space-velocity (100/WHSV) and temperature. The following effects
4.72
7.00 are observed:
8.43
3.45 Feed B produces less benzene in the reformate than Feed A (3.4

versus 6.1 mol% at the exit of the fourth reactor).
24.75

The benzene production rate in the first two reactors using Feed A is
higher than that of Feed B. In these two reactors, benzene is
0.90
1.25
mainly produced by N6 dehydrogenation. After reactor 2 there is
5.75 no considerable increase in benzene content when Feed B is
8.45 used (from 2.95 mol% at the exit of the second reactor to
4.52 3.41 mol% at the exit of fourth reactor), while this increase
0.26 was indeed higher with Feed A (from 4.5 to 6.1 mol% at the exit
21.13
of the same reactors).
3506 M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508
7 150

R-1 R-2 R-3 R-4


6 140
Benzene, mol % .

-T, °C
4 130

3 120

Total delta
2
1 110
T=490°C
0 100
0 10 20 30 40 50 60
100/WHSV
8 90
A6
6 nP6
80

Total aromatics, mol %


4 A6
mol % . 70

nP6 60
2 MCP

50
MCP
0 N6 40
470 480 490 500 510 520 470 480 490 500 510 520
Temperature,°C Temperature,°C

Fig. 11. Simulation of the operation of commercial reforming reactors with feeds with benzene formation precursors (–, Feed A) and without benzene formation precursor (- - -, Feed B).

The complete separation of MCP as in Feed B aids to reduce cyclohexane model can be used to analyze other aspects of the catalytic reform-ing
formation, and thus benzene production. process. Turpin [14] proposed a procedure to validate the per-formance of a
N6 conversion is almost 100% at all reaction temperatures for the two catalytic reforming unit. This procedure considers the calculation of the
feeds. As the reaction temperature is increased MCP content in reformate global and hydrogen material balances, com-parison of experimental and
obtained from both feeds also increases calculated feed and product densities, analysis of the gas composition
with respect to feed initial values, which is produced mainly from P 6. The (calculation of some quotients, iso-butene to n-butane molar ratio),
increase in MCP contributes to N 6 formation, and hence, benzene content probability plots of feed and refor-mate components, and ring balance. All
in reformate is also increased. this information and more can be generated with the developed kinetic and
The effect of temperature on benzene formation is lesser in Feed A reactor models.
compared with Feed B. A6 increases from 1.12 mol% to 8 mol% at 510 As examples, Tables 13 and 14 show the global and hydrogen mass
LC for Feed A, while this increase for Feed B is from 0.26 mol% to 4 mol balances for a case of study simulated at the following condi-tions: bench-
2
%. This indicates that the separation of ben- scale isothermal reactor, 750 LC inlet temperature, 10.5 kg/cm reactor
zene formation precursors from the reforming feed also helps to decrease pressure, and 6.3 mol/mol hydrogen-to-oil ratio.
A6 formation as the reaction temperature is increased.
From the data of these tables, the balance errors are calculated with:
Reactor delta-T is less when Feed B is used, which means that this feed
has lower content of those reacting compounds that contribute in higher Feed mass flowrate Product mass flowrate
extent to reaction exothermality than Feed A. Mass balance error ¼
Feed mass flowrate
Another reason for the differences of reactor delta-T with both feeds is the 100
different heats of reforming reaction, and the place and extent where they ð53Þ
take place. For instance, in the first reac-tor the major reactions are 2
endothermic and very fast, such as dehydrogenation of naphthenes to H2 balance error ¼ H mass flowrate H2 mass flowrate 100
aromatics, in the second reactor, mostly the isomerization takes place and
H2 mass flowrate
the remaining naphthenes are dehydrogenated, in the third and fourth reac- ð54Þ
tors exothermic hydrocracking of paraffins and dehydrogena-tion
reactions occur. Table 13
Global mass and molar balances.

Feed B produces more aromatics than Feed A, which is due to the higher Stream Inlet (g/h) Outlet (g/h)
initial contents of aromatics and naphthenes in Feed B (21.12 and 24.75 Naphtha (C5–C12) 119.9916 110.4971
mol% respectively) compared with Feed A (16.45 and 19.29 mol% Hydrogen 19.6989 20.8477
respectively). Gases (C1–C4) 0.0000 8.3458
Total 139.6905 139.6905
Inlet (mol/h) Outlet (mol/h)

3.4. Use of the model to predict other process parameters Naphtha (C5–C12) 1.0857 1.0653
Hydrogen 9.7713 10.3411
Apart from calculating the reformate composition, temperature and Gases (C1–C4) 0.0000 0.2064
Total 10.8570 11.6125
pressure profiles along the reactors system, the developed
M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508 3507
Hydrogen balance.
Table 14 Hydrogen content in hydrocarbons Inlet (g/h) Outlet (g/h)
n-Paraffins
nP11 0.0000 0.0000 rules, such as density and molecular weight. Other properties, e.g. RON,
nP10 2.4727 0.7595 MON, RVP, depend non-linearly on the mixture composi-tion thus requiring
nP9 2.5915 1.5639
non-linear mixing rules or other approaches, e.g. structural group
nP8 2.9667 2.2383
nP7 2.2658 2.3005 contribution.
nP6 1.8799 2.2477
nP5 0.4216 1.1713
4. Concluding remarks
nP4 0.0000 0.6009
nP3 0.0000 0.5552
nP2 0.0000 0.2581 The complexity of the various kinetic models reported in the lit-erature
nP1 0.0000 0.1411 varies from a few lumps to detailed reaction schemes. Sim-ple old-fashion
Sub-total 12.5981 11.8365 models, based on only a few lumps, should not be used for rigorous
Naphthenes
simulation of catalytic reforming reactors. The most detailed kinetic models,
N 0.0000 0.0000
N
11 although their authors claim that they represent the catalytic reforming
0.0000 0.0412
10
process in a rigorous man-ner, have still to be validated under conditions
N9 1.2115 0.0615
N8 0.7757 0.0827 different from those from which they were derived, e.g. comparisons of the
N7 0.8411 0.1792 predicted reactor behavior against commercial data of catalytic reforming
N6 0.4163 0.1305 units are mandatory for such complex models in order to provide a more
MCP 0.0000 0.0000 convincing applicability. The kinetic model proposed in this work combines
Sub-total 3.2446 0.4951
Aromatics
the simplicity of the lumping-based models with the complexity of the most
A advanced model, since it can predict the detailed composition of the reformate
0.0000 0.0000
A
11
and takes into consider-ation the most important process variables that affect
10 0.3309 0.5416
A9 0.6146 1.4930 the behavior of the reacting system. The validation and application of the
A8 0.5921 1.2787 model are indeed carried out with real data obtained from controlled iso-
A7 0.1996 0.6850 thermal bench-scale reactor experiments and with data recovered from
A6 0.0420 0.1432 commercial plants.
Sub-total 1.7793 4.1416
Total H2 in hydrocarbons 17.6220 16.4732
Hydrogen gas 19.7003 20.8491
Total 37.2323 37.2323
Acknowledgments

The authors thank Instituto Mexicano del Petróleo for its finan-cial
Table 15 support. M.A. Rodríguez also thanks CONACyT for financial support (PDF).
Amounts of hydrogen determined from chemical reaction calculations.

mol/mol g/g SCF/Bbl 3


m /Bbl
6.3000 0.1157 5600.5 158.6
References
Inlet H2
Produced H2 1.8094 0.0329 1594.8 45.2
[1] Krane HG, Groh AB, Shulman BD, Sinfeit JH. Reactions in catalytic reforming of
Consumed H2 0.7681 0.0140 677.0 19.2
naphthas. In: Proceedings of the fifth world petroleum congress; May 30, 1959. p. 39–51
Net H2 (produced-consumed) 1.0413 0.0189 917.8 26.0 [Sec. III].
Oulet H2 7.3413 0.1346 6518.3 184.6 [2] Smith RB. Kinetic analysis of naphtha reforming with platinum catalyst. Chem Eng Prog
1959;55(6):76–80.
[3] Burnett RL, Steinmetz HL, Blue EM, Noble EM. An analog computer model of conversion
in a catalytic reformer. Division of petroleum chemistry, American chemical society,
Detroit meeting; April 1965. p. 17–24.
For both balances the error was zero, while the maximum acceptable [4] Zhorov YM, Panchenkov GM, Zel’tser SP, Tirakyan YA. Mathematical description of
errors in the global balance and in the hydrogen balance are ±1% and ±0.5% platforming for optimization of a process (I). Kineetika i Kataliz 1965;6(6):1092–8.
respectively. Values outside these ranges indi-cate errors when measuring the
[5] Zhorov YM, Panchenkov GM, Shapiro IY. Mathematical description of Platforming
flowrates or in the analysis of the gas and liquid streams. carried out under severe conditions. Khim Technol Topl Masel 1970;15(11):37–40.

Given that it is possible to determine the amount of hydrogen entering the [6] Henningsen J, Bundgaard-Nielson M. Catalytic reforming. Brit Chem Eng 1970;15:1433–
6.
reactor and also that which has been produced or con-sumed by chemical [7] Kmak WS. A kinetic simulation model of the powerformimg process. Paper presented at
reactions, the amount of hydrogen leaving the reactor can be calculated with a AIChE national meeting. Houston, Texas; 1972.
mass balance as shown in Table 15. If the severity of the reforming reactor is [8] Kmak WS, Stuckey TW. Powerforming process studies with a kinetic simulation model.
AICHE national meeting, New Orleans, Paper No. 56a; 1973.
modified the determi-nation of the amount of hydrogen at the exit of the [9] Ramage MP, Graziani KR, Krambeck FJ. Development of Mobil’s kinetic
reforming reactor becomes important since it gives the potential quantity of reforming model. Chem Eng Sci 1980;35:41–8.
hydrogen supply to hydrotreating/hydrocracking units. [10] Jenkins JH, Stephens TW. Kinetics of cat reforming. Hydrocarb Process 1980;59:163–7.

[11] Marin GB, Froment GF. Reforming of C6 hydrocarbons on a platinum-alumina catalyst.


Once the composition of the reformate product has been calcu-lated with Chem Eng Sci 1982;37(5):759–73.
the model, some of its properties can be determined with mixing rules by [12] Ramage MP, Graziani KR, Schipper PH, Krambeck FJ, Choi BC. KINPTR (Mobil’s
using properties of the pure components, such as octane numbers (RON, kinetic reforming model): a review of Mobil’s industrial process modeling philosophy.
Adv Chem Eng 1987;13:193–266.
MON), Reid vapor pressure (RVP), density, molecular weight, average [13] Van Trimpont PA, Marin GB, Froment G. Reforming of C7 hydrocarbons on a sulfided
boiling points, among others. Data reported in the literature to calculate some commercial Pt/Al2O3 catalyst. Ind Eng Chem Res 1988;27:51–7.
of these properties are summarized in Table 10. In general, those properties [14] Turpin LE. Cut benzene out of reformate. Hydrocarb Process 1992:81–92.
[15] Coppens MO, Froment GF. Fractal aspects in the catalytic reforming of naphtha. Chem
that are additive, i.e. depend on the mass, are evaluated by linear mixing Eng Sci 1996;51:2283–92.
[16] Taskar U. Modeling and optimization of a catalytic naphtha reformer. PhD thesis. Texas
Tech University, USA; 1996.
[17] Taskar U, Riggs JB. Modeling and optimization of a semi-regenerative catalytic naphtha
reformer. AIChE J 1997;43(3):740–53.
[18] Padmavathi G, Chaudhuri KK. Modeling and simulation of commercial catalytic naphtha
reformers. Can J Chem Eng 1997;75(5):930–7.
3508 M.A. Rodríguez, J. Ancheyta / Fuel 90 (2011) 3492–3508

[19] Joshi PV, Klein MT, Huebner AL, Leyerle RW. Automated kinetic modeling of catalytic [25] Hou W, Su H, Mu S, Chu J. Multiobjective optimization of the industrial naphtha catalytic
reforming at the reaction pathways level. Rev Proc Chem Eng 1999;2(3):169–93. reforming process. Chin J Chem Eng 2007;15(1):75–80.
[26] Wei W, Bennett CA, Tanaka R, Hou G, Klein MT. Detailed kinetic models for catalytic
[20] Szczygiele Jerzy. On the kinetics of catalytic reforming with the use of various raw reforming. Fuel Process Technol 2008;89:344–9.
materials. Energy Fuels 1999;13:29–39. [27] Sotelo R, Froment GF. Fundamental kinetic modeling of catalytic reforming. Ind Eng
[21] Rahimpour MR, Esmaili S, Bagheri GNA. Kinetic and deactivation model for industrial Chem Res 2009;48:1107–19.
catalytic naphtha reforming. Iran J Sci Tech, Trans B: Tech 2003;27(B2):279–90. [28] Stijepovic MZ, Vojvodic-Ostojic A, Milenkovic I, Linke P. Development of a kinetic
model for catalytic reforming of naphtha and parameter estimation using industrial plant
[22] Hu Y, Su H, Mu S, Chu J. Modeling, simulation and optimization of commercial naphtha data. Energy Fuels 2009;23:979–83.
reforming process. In: Proceedings of the 42nd IEEE conference on decision and control, [29] Smith JM, Van Ness HC, Abbott MM. Introduction to Chemical Engineering
Maui, Hawaii, USA, December. p. 6206–11. Thermodynamics, EE UU, Mc-Graw Hill, 1996.
[23] Shanyinghu F, Zhu XX. Molecular modeling and optimization for catalytic reforming. [30] Reid RC, Prausnitz JM, Sherwood TK, The Properties of Gases and Liquids, 3rd Ed., Mc-
Chem Eng Commun 2004;191:500–12. Graw Hill, New York, 1977.
[24] Hou W, Su H, Hu Y, Chu J. Modeling, simulation and optimization of a whole industrial
catalytic naphtha reforming process on Aspen Plus platform. Chin J Chem Eng
2006;14(5):584–91.

Вам также может понравиться