Вы находитесь на странице: 1из 9

Fuel 216 (2018) 852–860

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Modeling of CSTR and SPR small-scale isothermal reactors for heavy oil T
hydrocracking and hydrotreating

Cristian J. Calderóna, Jorge Ancheytab,
a
Facultad de Química, UNAM, Ciudad Universitaria, Ciudad de México 04510, Mexico
b
Instituto Mexicano del Petróleo, Eje Central Lázaro Cárdenas Norte152, Col. San Bartolo Atepehuacan, Ciudad de México 07730, Mexico

G RA P H I C A L AB S T R A C T

A R T I C L E I N F O A B S T R A C T

Keywords: The dynamic modeling and simulation of a continuous slurry phase reactor for catalytic hydrocracking and
Modeling hydrotreating of an atmospheric residue (312 °C+) are reported. The reactor model is based on an axial dis-
Simulation persion. The hydrocracking kinetic model takes into account a five-lump model previously reported in the lit-
Slurry-phase reactor erature. The hydrotreating reactions simulated are: hydrodesulfurization (described by Langmuir–Hinshelwood
Heavy oil
kinetics), hydrodenitrogenation (for basic and non-basic nitrogen), hydrodeasphaltenization and hydro
Hydrocrackig
Hydrotreating
Conradson carbon removal (modeled with power-law approach). All the intrinsic kinetic parameters and cor-
relations were taken from the literature. The performance for the slurry-phase reactor was compared with a
continuous stirred tank reactor. Dynamic simulations and steady-state predictions agreed with the expected
behavior of the heavy fractions and impurities hydroprocesing.

1. Introduction alumina or silica-alumina, the gas phase is mainly composed of hy-


drogen and the liquid phase is the hydrocarbon. There are different
Hydrocracking (HDC) is one of the most important technologies in types of reactors, such as fixed, moving, and ebullated bed, however
oil refining. It consists of disintegrating heavy cuts with high molecular these reactors may face some complications due to high impurities
weight into lighter with low molecular weight fractions. This operation content of residue feeds, and excessive hydrocracking of heavy fractions
is carried out in multiphase reactors where the solid phase is the cat- that leads to coke formation and metal deposition, which eventually
alyst, usually sulfide of cobalt, molybdenum, or nickel supported on deposit on the catalyst surface and result in catalyst deactivation [1].


Corresponding author.
E-mail address: jancheyt@imp.mx (J. Ancheyta).

https://doi.org/10.1016/j.fuel.2017.11.089
Received 11 September 2017; Received in revised form 19 November 2017; Accepted 22 November 2017
0016-2361/ © 2017 Elsevier Ltd. All rights reserved.
C.J. Calderón, J. Ancheyta Fuel 216 (2018) 852–860

Nomenclature gases, gTn gR−1 gCat


−1
hr−1
k5 intrinsic kinetic parameter for vacuum gas oil hydro-
−1
Symbols gracking to distillates, gTn gCat hr−1
k6 intrinsic kinetic parameter for vacuum gas oil hydro-
−1
API API gravity gracking to naphtha, gTn gCat hr−1
Asph asphaltenes k7 intrinsic kinetic parameter for vacuum gas oil hydro-
−1
BN basic nitrogen gracking to gases, gTn gCat hr−1
k8 intrinsic kinetic parameter for distillates hydrogracking to
D distillates −1
naphtha, gTn gCat hr−1
Da axial dispersion coefficient, m2 / hr
k9 intrinsic kinetic parameter for distillates hydrogracking to
G gases −1
gases, gTn gCat hr−1
HDC hydrocracking reaction k10 intrinsic kinetic parameter for naphtha hydrogracking to
HDT hydrotreating reaction −1
gases gTn gCat hr−1
i Refering to i component. L reactor length, m
kCAsph catalytic kinetic parameter for hydrodesasphaltenization mT total mass composition, gT
reaction, wt%−0.503hr−1 Naph naphthenes
kCBN catalytic kinetic parameter for hydrodenitrogenation re- NBN non basic nitrogen
action of basic nitrogen, ppm−0.792hr−1 P pressure, MPa
kCNBN catalytic kinetic parameter for hydrodenitrogenation re- R residue
action of non-basic nitrogen, ppm−1.154hr−1 S sulfur
kCS catalytic kinetic parameter for hydrodesulfurization re- SG specific gravity at 15.6 °C
action, wt%−0.503hr−1 t time, hr
kD deactivation parameter, hr−1 tf Final time, hr
kTAsph thermal kinetic parameter for hydrodesasphaltenization T temperature, K
reaction, wt%0.795hr−1 u feed velocity, m/hr
kTBN thermal kinetic parameter for hydrodenitrogenation re- V volume, m3
action of non-basic nitrogen, ppm0.137hr−1 VGO vacuum gas oil
kTNBN thermal kinetic parameter for hydrodenitrogenation re- WP amount of catalyst in the feed, g
action of non-basic nitrogen, ppm0.137hr−1 xi content of i component in the feed, gi / gT , ppm
kTS thermal kinetic parameter for hydrodesulfurization re- x i0 initial content of i component, gi / gT , ppm
action, wt%0.062hr−1 Z reactor position, m
k1 intrinsic kinetic parameter for residue hydrogracking to
vacuum gas oil, gTn gR−1 gCat −1
hr−1 Greek symbols
k2 intrinsic kinetic parameter for residue hydrogracking to
distillates, gTn gR−1 gCat
−1
hr−1 η Effectiveness factor
k3 intrinsic kinetic parameter for residue hydrogracking to νi Stechiometric coefficient of i component
naphtha, gTn gR−1 gCat−1
hr−1
φ Function of catalyst deactivation
k4 intrinsic kinetic parameter for residue hydrogracking to

Also, these reactors experience high pressure drop of the bed which Recent investigations concerning to kinetic models for hydro-
makes difficult to maintain a normal operation, therefore shut-down is cracking in slurry-phase have been reviewed [4]. Due to heavy oil
more frequent due to short catalyst life and unstable operation of the consists of a large amount of components, different approximations are
reactor [2]. used to represent hydrocracking reactions, being the most common the
To address the disadvantages in HDC in conventional technologies, lumping techniques. The method consists of lumping various com-
slurry-phase hydrocracking processes have been developed. This tech- pounds in a few pseudocompounds that differ by boiling temperature
nology consists of mixing the oil feed, hydrogen, and dispersed catalysts range. This approximation is easy to implement in a reactor model
together going through the reactor. It has the same processing as because it reduces the number of kinetic equations and parameters to
thermocracking but also reduces coking due to the presence of hy- be estimated. On the other hand it is known that as long as the reaction
drogen and catalyst that promote hydrogenation reactions [3]. The rate coefficients have been obtained under kinetic regime, the kinetic
catalyst also acts as a support for the low amount of coke that could be model can be implemented in the reactor model independently if those
formed, and due to the catalyst leaves the reactor continuously this parameters were obtained in a different type of reactor.
effect has no greater relevance in the operation. Depending on the oil to Moreover, the literature concerning mathematical modeling of hy-
be treated, the catalysts are designed for the removal of impurities such drocracking in SPR’s is scarce, as shown in a previous review work [5].
as sulfur, nitrogen, metals and asphaltenes, even for the saturation of Most of the reports are based on computational fluid dynamics models
aromatics and olefins. All these reactions occur simultaneously and are (CFD) formulated in steady-state and focused on the performance of the
known as hydrotreating reactions (HDT). hydrodynamic variables inside the reactor and not on residue conver-
While slurry-phase reactors (SPR) for hydrocracking of heavy oil are sion [6–10]. It should be pointed out that due to the difficulty for ob-
a promising technology, plants at a commercial scale have not been taining proper information, all those reactor models were validated
developed or are still in the design stage due to high catalyst cost and with air–water systems. On the other hand, there are recent modeling
elevated operating conditions necessary to achieve the conversion of the reports dealing with hydrocracking in industrial and pilot plant units
heaviest fractions of the feed [4]. In addition, SPR technology has some [11,12]. However, such models are based on ideal plug-flow patterns in
major disadvantages compared with fixed, moving, or ebullated bed steady-state.
reactors, such as the difficult separation of catalyst and the liquid pro- Recent models for SPR’s have been published. Those reactor models
duct, uncertain scale-up, catalyst sedimentation and agglomeration, as were formulated for different reacting systems, such as Fischer–Tropsch
well as the hard understanding of reaction kinetics and flow patterns [5]. synthesis, methanol synthesis, dimethyl ether synthesis, and diesel

853
C.J. Calderón, J. Ancheyta Fuel 216 (2018) 852–860

hydrodesulfurization [13–18]. Such models are based on axial disper- t = 0, 0 ⩽ Z ⩽ L x i = x iin (2)
sion (ADM), despite of the well-known uncertainty in the parameters
especially the axial dispersion coefficient. In a previous investigation, ∂x i
0 ⩽ t ⩽ tf , Z = 0 D = u (x i−x i0)
modeling of a SPR for hydrotreating of gas oil was reported, which uses ∂Z (3)
a kinetic model for mild hydrocracking reactions for the light fractions
∂x i
[19]. The model was successfully validated versus the experimental 0 ⩽ t ⩽ tf , Z = L =0
∂Z (4)
data of a hydrodesulfurization system and was compared with the
predictions of the reactor model of Khadem-Hamedani et al. [18]. In the case of the CSTR, the well known consideration of perfect
The objectives of this investigation are to adapt this model for hy- mixing is assumed, then the reactor model approximates to
drocracking and hydrotreating of heavy oil in a slurry phase reactor, as C R
well as to present a comparison between the dynamic behavior and ∂x i W
= x iin−x i + P ∑∑ νi ri,k
steady-state predictions of a SPR and a continuous stirred tank reactor ∂t mT i k (5)
(CSTR) based on intrinsic reaction rate coefficients and activation en-
with the initial condition
ergies taken from the literature with the aim of finding the best oper-
ating conditions and to explore the influence of catalyst deactivation x i (0) = x iin (6)
and effectiveness factors in the reactor performance.
x i stands for the total mass composition of the component i in feed
and product (gases + hydrocrabon), the subscript i represents hy-
2. Model formulation
drogen, hydrogen sulfide, sulfur, nitrogen, residue, vacuum gas oil,
distillates, naphtha and light gases in feed and product, mT represents
2.1. Reactor models
the total mass (gases + hydrocarbon) in the reactor, u the feed velocity,
WP the amount of catalyst in the feed and D the axial dispersion coef-
A schematic representation of the small-scale reactors to be modeled
ficient, which was estimated through the correlation proposed by
is presented in Fig. 1. In order to compare the performance of both C R
Deckwer et al. [13]. The therm ∑i ∑k νi ri,k represents the rate of re-
reactors, the same operating conditions were assumed: An isothermal
action of each component i in reaction k. Being νi the stoichiometric
operation, 891 std m3/ m3 of H2−hydrocarbon ratio, 100 ml of catalyst
coefficient and ri,k the intrinsic reaction rate.
loaded in the reactors, and a constant pressure of 9.8 MPa. The di-
mensions of the SPR are 38 cm of length and 7.7 cm of diameter while
2.2. Reaction kinetics
the CSTR total volume is 1.7 Lt. It is also assumed that in both reactors
the heavy oil feed is preheated and mixed with hydrogen and then
Heavy oils are a complex mixture of large number of components,
passed through the reactor.
therefore the study of all the reactions involved in hydrocracking and
Axial dispersion model was selected to approximate the SPR per-
hydrotreating is quite complicated. For the particular case of hydro-
formance. The model is based on previous considerations reported in
cracking, there are different types of kinetic models that have been
the literature, such as dynamic regime, isothermal operation, and a
reported to approximate all the reactions [20]. Despite the dependence
pseudo–homogeneous slurry-phase [13–19]. It is also assumed that the
of the kinetic parameters with properties of the feed and the use of an
gas phase is present in excess, then the mathematical model can be
invariant distillation range of products, a pseudo-component based
approximated with an homogeneous model. Based on these con-
model will be used to represent the hydrocracking kinetic due to its
siderations, the dynamic equations of the ADM are as follows:
simplicity and utility in simulation and control applications. The kinetic
C R
∂x i ∂x ∂ 2x W model is presented in Fig. 2, this model was proposed by Sanchez et al.
= −u i + Da 2i + P ∑∑ νi ri,k [21] and consists of five pseudo-components: residue (538 °C+), va-
∂t ∂Z ∂Z mT i k (1)
cuum gas oil (343–538 °C), middle distillates (204–343 °C), naphtha
With the initial and boundary conditions: (IBP-204 °C) and gases, and 10 reaction paths for hydrocracking of all

Fig. 1. Schematic representation of the experi-


mental reactors.

854
C.J. Calderón, J. Ancheyta Fuel 216 (2018) 852–860

n nC
xS CS x H2 H2 n
rHDS = φS ηS kCS + kTS xS TS
(1 + K H2 S x H2 S )2 (12)

Asph n Asph nT
rHDAsph = φAsph ηAsph kCAsph xAsph + kTAsph xAsph (13)
nCCR nTCCR
rHDCCR = φCCR ηCCR kCCCR x CCR + kTCCR x CCR (14)
nNBN nTNBN
rHDNNBN = φNBN ηNBN kCNBN xNBN + kTNBN xNBN (15)
nBN n
rHDNBN = φBN ηBN kCBN xBN + kTBN xBNTBN (16)
nNBN TNBN n
−φNBN ηNBN kCNBN xNBN −kTNBN xNBN (17)
Catalyst deactivation is originated by coke deposition, however the
estimation of this phenomenon is based on deactivation functions. For
hydrocracking reactions a time-dependant non-selective expression was
used, which means that the selectivity of the hydrocracking reactions is
not affected by coke deposition, then the deactivation constant k DHDC is
the same for all the reactions. On the contrary for hydrotreating reac-
tions a time-dependant selective model was used, in this case it is ne-
cessary to take into account the effect of coke deposition on each hy-
drotreating reaction, then there would be a deactivation constant for
each reaction, (k DS,k DAsph,k DCCR,k DNBN ,k DBN ). In both cases the deactiva-
tion function is presented in Eq. 18 where mi is the deactivation order
for reaction i, (HDC ,HDS,HDAsph,HDCCR,HDNBN ,HDBN ).
1
φ=
(1 + k Di t )mi (18)
In the case of the CSTR Eq. 18 can be used directly in the reactor
Fig. 2. Five-lump kinetic model for hydrocracking reported by Sanchez et al. [21]. model due to the catalyst does not leave the reactor during the opera-
tion, however for the SPR it is considered that the catalyst does not
lumps, as well as 10 reaction rate coefficients to be calculated. Recently remain in the reactor, on the contrary it enters and leaves the reactor
Martínez and Ancheyta [22] conducted an experimental study based on along with the feed, then the deactivation function changes to Eq. 19.
the kinetic model presented in Fig. 2 to obtain the 10 intrinsic reaction 1
rate coefficients in a CSTR. The kinetic expressions for hydrocracking φ=
reactions are presented in Eqs. (7)–(11). Although most literature re-
(1 + k Di ( ))
Z
u
mi
(19)
ports propose that hydrocracking reactions of heavy oils follow a first- Z
Where represents the variation of time that the catalyst remains in
u
order reaction, in this work the hydrocracking reaction of vacuum re- the reactor. Also for the SPR, due to the catalyst particles are very small,
sidue was considered to be of second-order while the rest of the hy- the effectiveness factor is considered to be the unity for all the reactions
drocracking reactions were assumed to be of first-order [23,22]. The (ηi = 1).
second-order reaction is attributed to the broad distribution of reaction
rate values of compounds present in the residue fraction, which gen- 2.3. Model equations
erates that the reaction rate declines faster than that of a simple first-
order reaction due to the most-reactive compounds disappear first and The generalized SPR model is presented in Eq. 1, along with the
fast, leaving the most-refractory compounds in the residue [23]. initial and boundary conditions given by Eqs. (2)–(4). These equations
All the reaction rate expressions involve the calculation of effec- need to be formulated for each one of the components involved in the
tiveness factors and catalyst deactivation function, due to the large hydrocracking and hydrotreating system considering the reaction rate
particle size and the deposition of coke on the catalyst surface. expressions presented in Eqs. (7)–(17), then the set of partial differ-
rR = −φHDC η (k1 + k2 + k3 + k 4 ) yR2 (7) ential equations (PDE) is obtained as shown in Eqs. (20)–(31). That
system of PDE needs to be reduced into an ordinary differential equa-
rVGO = φHDC η [k1 yR2 −(k5 + k6 + k 7 ) yVGO ] (8) tion (ODE) system in order to find the numerical solution with a MA-
TLAB subroutine, particularly the ode23s. The orthogonal collocation
rD = φHDC η [k2 yR2 + k5 yVGO −(k8 + k 9 ) yD ] (9) method was selected to perform the calculations.

rN = φHDC η [k3 yR2 + k6 yVGO + k8 yD −k10 yN ] ∂xR ∂x ∂ 2x W


(10) = −u R + D 2R − P φHDC [k1 + k2 + k3 + k 4 ] xR2
∂t ∂Z ∂Z mT (20)
rG = φHDC η [k 4 yR2 + k 7 yVGO + k 9 yD + k10 yN ] (11)
∂xVGO ∂x ∂ 2x WP
= −u VGO + D VGO + φ [k1 xR2−(k5 + k6 + k 7 ) xVGO]
On the other hand, as hydrogen is present in the system, hydro- ∂t ∂Z ∂Z 2 mT HDC
treating reactions also occur simultaneously, then kinetic expressions (21)
for all these reactions are needed. It is known that thermal and catalytic
∂xD ∂x ∂ 2x WP
reactions act sequentially via thermal scission and hydrogenation. Thus = −u D + D D2 + φ [k2 xR2 + k5 xVGO−(k8 + k 9 ) xD]
in order to take into account catalytic and thermal reactions Martínez ∂t ∂Z ∂Z mT HDC (22)
and Ancheyta [24] proposed the expressions presented in Eqs. ∂xNaph ∂xNaph ∂2xNaph WP
(12)–(17). As well as for hydrocracking reactions, all the expressions for = −u +D + φ [k3 xR2 + k6 xVGO + k8 xD
∂t ∂Z ∂Z 2 mT HDC
hydrotreating reactions require the knowledge of catalyst effectiveness
factor and deactivation rates. −k10 xN ] (23)

855
C.J. Calderón, J. Ancheyta Fuel 216 (2018) 852–860

∂x G ∂x ∂ 2x W
= −u G + D G2 + P φHDC [k 4 xR2 + k 7 xVGO + k 9 xD + k10 xN ]
∂t ∂Z ∂Z mT
(24)
n nx
∂xS ∂x ∂ 2x W ⎛ xS xS x H2 H2
n ⎞
= −u S + D 2S − P ⎜φS k xS + kTS xS TS ⎟
∂t ∂Z ∂Z mT (1 + K H2 S x H2 S )2 (25)
⎝ ⎠

∂xAsph ∂xAsph ∂2xAsph WP n nTAsph


= −u +D + (φ k x x Asph + kTAsph xAsph )
∂t ∂Z ∂Z 2 mT Asph Asph Asph
(26)

∂x CCR ∂x ∂2x CCR W nTCCR


= −u CCR + D + P (φCCR k xCCR x CCR
nCCR
+ kTCCR x CCR )
∂t ∂Z ∂Z 2 mT
(27)

∂xBN ∂x ∂ 2x W n
= −u BN + D BN + P (φBN kCBN xBN
nBN
+ kTBN xBNTBN
∂t ∂Z ∂Z 2 mT
nNBN TNBN n
−φNBN k xNBN xNBN −kTNBN xNBN ) (28)

∂xNBN ∂x ∂ 2x W nTNBN
= −u NBN + D NBN + P (φNBN k xNBN xNBN
nNBN
+ kTNBN xNBN )
∂t ∂Z ∂Z 2 mT
(29)

∂x H2 ∂x H2 ∂2x H2 W
= −u +D + P rH2
∂t ∂Z ∂Z 2 mT (30)

n nx
∂x H2 S ∂x H2 S ∂2x H2 S W ⎛ x H2xSS x H2 H2 n ⎞
= −u +D + P φS k xS + kTS xS TS
∂t ∂Z ∂Z 2 mT ⎜ (1 + K H2 S x H2 S )2 ⎟
⎝ ⎠
Fig. 3. Comparison of reactor performance of heavy oil hydrocracking. (–) SPR, (- -)
(31) CSTR, (Symbols) Experiment. (o) Residue, (*) Vacuum gas oil, (+) Middle distillates, (Δ )
For the CSTR, the general model is based on Eq. 2, then including Naphtha, (∇) Gases.
the reaction rate expression (7)–(17) the CSTR model is obtained as
shown in Eqs. (32)–(43). These equations form an ordinary differential dxNaph WP
equation system which can be solve directly as well in MATLAB with
in
= xNaph −xNaph + φ [k3 xR2 + k6 xVGO + k8 xD−k10 xN ]
dt mT HDC (35)
the subroutine ode23s.
dx G
dxR W = xGin−x G + WP φHDC [k 4 xR2 + k 7 xVGO + k 9 xD + k10 xN ]
= xRin−xR− P φHDC [k1 + k2 + k3 + k 4 ] xR2 dt (36)
dt mT (32)
n nx
dxS W ⎛ xS xS x H2 H2 n ⎞
dxVGO in
= xVGO
W
−xVGO + P φHDC [k1 xR2−(k5 + k6 + k 7 ) xVGO] = xSin−xS − P ⎜φS k xS + kTS xS TS ⎟
dt mT (33) dt mT (1 + K H2 S x H2 S )2 (37)
⎝ ⎠

dxD W dxAsph WP n nTAsph


= xDin−xD + P φHDC [k2 xR2 + k5 xVGO−(k8 + k 9 ) xD] in
= xAsph −xAsph + (φ k x x Asph + kTAsph xAsph )
dt mT (34) dt mT Asph Asph Asph (38)

Table 1
Reaction rate coefficients and feed properties reported by Martínez and Ancheyta [22,24] for hydrocracking and hydrotreating reactions at 380 °C.

Hydrocracking Value Hydrotreating Value Feed property Value

k1 (gTn gR−1 gCat


−1
hr−1) 7.598 kCS (wt%−0.503hr−1) 28.8071 sg 60/60 °F 1.028
n −1 −1 4.037 kCAsph (wt%−0.503hr−1) 0.3819 API gravity 6.15
k2 (gT gR gCat hr−1)
k3 (gTn gR−1 gCat
−1
hr−1) 1.571 kCBN (ppm−0.792hr−1) 0.0062 Carbon (wt%) 82.67
n −1 −1
k 4 (gT gR gCat hr−1) 0.783 kCNBN (ppm−1.154hr−1) 5.0341E −4 Hydrogen (wt%) 10.07
−1 0.031 kTS (wt%0.062hr−1) 0.1179 6.17
k5 (gTn gCat hr−1) Sulfur (wt%)
−1
k 6 (gTn gCat hr−1) 0.004 kTAsph (wt%0.795hr−1) 1.1998 BN (ppm) 5014
−1
k7 (gTn gCat hr−1) 0.003 kTBN (ppm0.137hr−1) 0 NBN (ppm) 5886
−1 0.051 kTNBN (ppm0.137hr−1) 0 30.6
k8 (gTn gCat hr−1) Asph (wt%)
n −1
k 9 (gT gCat hr−1) 0.0009 kDS (hr−1) 0.0235 – –
−1
k10 (gTn gCat hr−1) 0.054 kDAsph (hr−1) 0.1199 – –
kD (hr ) −1 3.799 kDBN (hr−1) 0.0125 – –
m 0.222 kDNBN (hr−1) 0.0083 – –
– – mS 0.228 – –
– – mAsph 0.211 – –
– – mBN 0.293 – –
– – mNBN 0.533 – –

856
C.J. Calderón, J. Ancheyta Fuel 216 (2018) 852–860

Fig. 6. Dynamic axial profile in the SPR for hydrocracking reactions.

dx CCR W nTCCR
Fig. 4. Comparison of reactor performance of heavy oil hydrotreating. (–) SPR, (- -) CSTR, in
= x CCR −x CCR + P (φCCR k xCCR x CCR
nCCR
+ kTCCR x CCR )
(Symbols) Experiment. (o) Sulfur, (*) Asphaltenes, (+) CCR, (Δ ) Basic Nitrogen, (∇) Non dt mT (39)
Basic Nitrogen.
dxBN W n
in
= xBN −xBN + P (φBN kCBN xBN
nBN nNBN
+ kTBN xBNTBN −ϕNBN k xNBN xNBN
dt mT
n
TNBN
−kTNBN xNBN ) (40)

dxNBN W nTNBN
in
= xNBN −xNBN + P (φNBN k xNBN xNBN
nNBN
+ kTNBN xNBN )
dt mT (41)

dx H2 W
= x Hin2 −x H2 + P rH2
dt mT (42)
n nx
dxS W ⎛ xS xS x H2 H2 n ⎞
= xSin−xS + P ⎜φS k xS + kTS xS TS ⎟
dt mT (1 + K H2 S x H2 S )2 (43)
⎝ ⎠

3. Results and discussion

3.1. Dynamic simulations of CSTR and SPR

Martínez and Ancheyta [22] performed an experimental study in


order to obtain the intrinsic reaction rate coefficients for hydrocracking
reactions of heavy oil involving catalyst deactivation and the catalyst
effectiveness factor in a continuous stirred tank reactor. In their re-
search the five-lump kinetic model shown in Fig. 2 was used. All the
reaction rate coefficients and feed properties were taken from that in-
vestigation in order to simulate and compare the performance of both
CSTR and SPR reactors. A WHSV of 0.98 hr−1 and a temperature of
38 °C were considered, the rest of the data are presented in Table 1.
The dynamic behavior for the small-scale reactors for hydrocracking
reactions is presented in Fig. 3. The continuous line stands for the si-
mulated composition obtained at the exit of the SPR, and the dotted line
the simulated composition obtained in the CSTR. In both simulations
the same volume and weight hourly space velocity were considered.
Fig. 5. Influence of catalyst deactivation in hydrodesulfurization and hydrocracking re-
actions. (–) SPR, (- -) CSTR.
The symbols in each plot represent the experimental data obtained in
the CSTR. As can be seen, the simulation of the CSTR is in good
agreement with the experimental points. It is also observed that the

857
C.J. Calderón, J. Ancheyta Fuel 216 (2018) 852–860

Fig. 8. Dynamic axial profile of gases in the SPR.

Fig. 7. Dynamic axial profile in the SPR for hydrotreating reactions.

catalyst starts to deactivate, from the beginning of the reactions, so that


the rate of the different hydrocracking reactions reduces until it reaches
the steady-state near the 150 h. This situation does not occur in the SPR
due to the catalyst leaves the reactor along with the hydrocarbon,
therefore a major residue conversion is reached in the SPR in a shorter
period of time, around 5 h.
For thermal and catalytic hydrotreating reactions, the effects of
temperature and space-velocity were reported in the literature [22,24].
All reaction rate coefficients were obtained in kinetic regime, therefore
they are considered to be intrinsic. Since the catalyst used in the ex-
periments is of commercial size, the effectiveness factor is included in
the kinetic model for the CSTR. All the parameters needed for the re-
actor simulations are presented in Table 1. The dynamic simulation of
the small-scale reactors for hydrotreating reactions is presented in
Fig. 4. Similar to hydrocracking reactions, in catalytic hydrotreating the
conversion is favored in the SPR due to the small catalyst size and that
the catalyst does not remain inside the reactor which disfavors catalyst Fig. 9. Yield of hydrocracked lumps as a function of residue conversion in the SPR.
deactivation. On the other hand, the effects of catalyst deactivation and
internal diffusion are less severe in hydrotreating reactions compared
with hydrocracking reactions, since hydrocracking reactions are

858
C.J. Calderón, J. Ancheyta Fuel 216 (2018) 852–860

Fig. 10. Yield of hydrocracked lumps as a function of residue conversion in the CSTR.

considered to occur only in the catalyst surface, and hydrotreating in-


volves thermal and catalytic conversions. In comparison both reactors
have a high conversion at the beginning of the operation due to the
catalyst has its major activity as seen in Figs. 3 and 4 for hydrocracking
and hydrotreating reactions respectively. In the case of the CSTR, it is
observed that after two hours of operation the highest conversion is
reached, and then the conversion decrease due to catalyst deactivation.
Catalyst deactivation is also observed in Fig. 5 for reactions of hydro-
desulfurization and hydrocracking of residue, in which a major effect is
presented in hydrocracking reactions due to the heavy molecules pre-
sent in the liquid phase are easily clogged in the catalyst due to the Fig. 11. Steady-state simulations of hydrocracking reactions.
presence of coke. Evidently as the reaction time increases, the catalyst
deactivation will also increase generating a negative effect in reactions
catalyst deactivation since in the CSTR, the catalyst remains inside the
conversion. In the SPR the reacting system needs only around two hours
reactor during the operation then there is a great deactivation due to
to reach the steady-state, after that the conversion remains constant
coke deposition.
while fresh catalyst continuously enters the reactor as the deactivated
catalyst leaves it. This behavior is also observed in Figs. 6–8 in which
the dynamic axial profile for hydrocracking and hydrotreating reactions 3.2. Steady-state simulations in a SPR
as well as for hydrogen and sulfhydric acid is reported is presented. As
can be seen another major conclusion is that in the SPR a constant axial Once the SPR model is validated with the CSTR experimental data,
profile is presented thus SPR can be modeled as a CSTR. it was used to simulate all the hydrocracking and hydrotreating reac-
Figs. 9 and 10 compare the yield of each lump for the conversion of tions simultaneously. The feed velocity (WHSV) and temperature vary
feedstock as a function of the yield of each lump vs conversion of re- from 0.98 to 2.56 hr−1 and from 380 to 420 °C respectively as shown in
sidue in the SPR and CSTR respectively. As can be observed in the SPR a previous experimental reports [22,24]. Figs. 11 and 12 show the si-
high feed conversion is obtained. The residue can reach a constant 60 mulation results of hydrocarbon and impurities compositions at the exit
percent of conversion producing lighter fractions after 200 h of opera- of the reactor once the steady-state has been reached. Fig. 12 shows the
tion, producing mainly vacuum gas oil and distillates which changes steady-state lumps composition. Since residue is only present as a re-
from 32 and 9 weight percent to 50 and 18 weight percent in the actant in the kinetic model, a lineal effect between the WHSV and
product respectively. A low production of naphtha and gases is also temperature variation is observed. In this case, increasing both the re-
observed. On the other hand the CSTR present a similar selectivity, the sidence time of the feed and temperature in the reactor increases the
main difference is observed after a long time of operation, at which the residue conversion, therefore a low composition of residue would be
conversion of residue decreases after it reaches its higher value, around present in the hydrocarbon product. On the contrary the vacuum gas oil
58 percent of residue conversion. This phenomenon is originated due to appears in the kinetic model as both, reactant and product, therefore
VGO conversion depends on the selectivity of residue and lighter

859
C.J. Calderón, J. Ancheyta Fuel 216 (2018) 852–860

simulations have shown similar behavior in the CSTR and SPR. As ex-
pected higher conversion of residue is reached in the SPR (effectiveness
factor equals one) and deactivation of catalyst is not taken into account.
The results of steady-state simulation have shown that it is possible to
approximate a slurry-phase reactor for hydrocracking with an axial
dispersion model with appropriate reaction rate coefficients.
As expected, for experimental proposes the CSTR has some ad-
vantages compared with the SPR. It promises lower cost in the ex-
perimental framework due to the amount of catalyst used. It was also
found that for modeling purposes, the experimental SPR can be con-
veniently approximated by means of an ADM, however to reduce its
complexity, the SPR can be approximated to an ideal CSTR model. On
the other hand for industrial applications the SPR promises to obtain
better results in the conversion of residues to light fractions.

Acknowledgement

The authors thank Instituto Mexicano del Petróleo for its financial
support. C.J.C also thanks CONACyT for his Ph.D. scholarship.

References

[1] Speight JG. New approaches to hydroprocessing. Catal Today 2004;98(1–2):55–60.


[2] Zhang S, Liu D, Deng W, Que G. A review of slurry-phase hydrocracking heavy oil
technology. Energy Fuels 2007;21(6):3057–62.
[3] Zhang S, Deng W, Luo H, Que G. Slurry-phase residue hydrocracking with dispersed
nickel catalyst. Energy Fuels 2008;22:3583–6.
[4] Quitian A, Ancheyta J. Experimental methods for developing kinetic models for
hydrocracking reactions with slurry-phase catalyst using batch reactors. Energy
Fuels 2017;30(6):4419–37.
[5] Calderon CJ, Ancheyta J. Modeling of slurry-phase reactors for hydrocracking of
heavy oils. Energy Fuels 2016;30(4):2525–43.
[6] Matos-Carbonell M, Guirardello R. Modelling of a slurry bubble column reactor
applied to the hydroconversion of heavy oils. Chem Eng Sci 1997;52(21):4179–85.
[7] Parulekar SJ, Shah YT. Steady-state behavior of gas-liquid-solid fluidized-bed re-
actors. Chem Eng J 1980;20(1):21–33.
[8] Matos EM, Guirardello R. Modeling and simulation of a pseudo-two-phase gas-li-
quid column reactor for thermal hydrocracking of petroleum heavy fractions. Braz J
Chem Eng 2002;19(3):319–34.
[9] Matos EM, Nunhez JR. The effect of different feed flow patterns on the conversion of
bubble column reactors. Chem Eng J 2006;116(3):163–72.
[10] Matos EM, Guirardello R, Mori M, Nunhez JR. Modeling and simulation of a
pseudo-three-phase slurry bubble column reactor applied to the process of petro-
leum hydrodesulfurization. Comput Chem Eng 2009;33(6):1115–22.
[11] Faraji Davood, Sadighi Sepehr, Mazaheri Hossein. Modeling and evaluating zeolite
and amorphus based catalyst in vacuum gas oil hydrocracking. Int J Chem Reactor
Eng 2017.
[12] Manek Eduard, Haydary Juma. Hydrocracking of vacuum residue with solid and
dispersed phase catalyst: Modeling of sediment formation and hydrodesulfuriza-
Fig. 12. Steady-state simulations of hydrotreating reactions. tion. Fuel Process. Technol. 2017;159:320–7.
[13] Deckwer WD, Serpemen Y, Ralek M, Schmidt B. Modeling the fischer-tropsch
synthesis in the slurry-phase. Ind Eng Chem Proc D.D. 1982;21(2):231–41.
fractions. For WHSV between 0.98 and 2.56 hr−1 and temperatures [14] Maretto C, Krishna R. Modelling of a bubble column slurry reactor for fischer-
below 410 °C a lineal function in the VGO composition is observed, tropsch synthesis. Catal Today 1999;52(2–3):279–89.
[15] de Toledo ECV, de Santana PL, Wolf-Maciel MR, Maciel-Filho R. Dynamic modelling
which indicates that the rate of hydrocracking of residue is faster than of a three-phase catalytic slurry reactor. Chem Eng Sci 2001;56(21–22):6055–61.
that of the hydrocracking of VGO. For temperatures higher than 410 °C [16] de Swart JKA, Krishna R. Simulation of the transient and steady-state behaviour of a
low composition of VGO is observed, which is sign of a high hydro- bubble column slurry reactor for fischer-tropsch synthesis. Chem Eng Process
2002;41(1):35–47.
cracking of this pseudo-component in order to produce lighter fractions, [17] Rados N, Al-Dahhan M, Dudokivic MP. Dynamic modeling of slurry bubble column
such as middle distillates, naphtha and gases. Finally the light fractions; reactors. Ind Eng Chem Res 2005;44(16):6086–94.
distillates, naphtha and gases show a lineal behavior at steady-state. In [18] Khadem-Hamedani B, Yaghmaei S, Fattahi M, Mashayekhan S, Hosseini-Ardali SM.
Mathematical modeling of a slurry bubble column reactor for hydrodesulfurization
those cases as well as residence time and temperature increases, the of diesel fuel: single and two bubble configurations. Chem Eng Res Des
conversion of residue and vacuum gas oil is higher and a major com- 2015;100:362–76.
position of these fractions is present in the product. A similar effect is [19] Calderon CJ, Ancheyta J. Dynamic modeling and simulation of a slurry-phase re-
actor for hydrotreating of oil fractions. Energy Fuels 2017;31(5):5691–700.
shown in Fig. 12 for impurities contents, in which a high removal of
[20] Ancheyta J, Sanchez S, Rodriguez M. Kinetic modeling of hydrocracking of heavy
impurities such as sulfur, asphaltenes, Conradson carbon, and nitrogen oil fractions: a review. Catal Today 2005;1(109):76–92.
is observed as temperature and residence time increase. [21] Sanchez S, Rodriguez M, Ancheyta J. Kinetic model for moderate hydrocracking of
heavy oils. Ind Eng Chem Res 2005;44(25):9409–13.
[22] J. Martínez, Ancheyta J. Kinetic model for hydrocracking of heavy oil in a cstr
4. Conclusions involving short term catalyst deactivation. Fuel 2012;100:193–9.
[23] Sanchez S, Ancheyta J. Effect of pressure on the kinetics of moderate hydrocracking
Hydrocracking as well as catalytic and thermal hydrotreating have of maya crude oil. Energy Fuels 2007;21:653–61.
[24] Martínez J, Ancheyta J. Modeling the kinetics of parallel thermal and catalytic
been modeled in small-scale isothermal CSTR and SPR based on in- hydrotreating of heavy oil. Fuel 2014;138:27–36.
trinsic reaction rate coefficients involving catalyst deactivation and
catalyst effectiveness factors for each of the reactions. The dynamic

860

Вам также может понравиться