Reactor Modeling Part 4 (Petroleum Refining)

When a high-purity hydrogen stream without gas recycle is used, such as in the case of some laboratory and bench-scale HDT reactors, or when the gas recycle has been subject to the purification process in commercial units, values of partial pressure (pG) and liquid molar concentrations (CL) of H-S, NH3, and LHC at the entrance of the catalytic bed (z = 0 and z = LB for co – current and countercurrent operation, respectively) are equal or very close to zero. For commercial HDT reactors without the high purification of a gas recycle stream, values of partial pressures (pG) and liquid molar concentrations (CL) of H2S, NH3, and LHC at the entrance of the catalytic bed (z = 0) differ from 0 (Mederos et al., 2006).

The axial and radial dispersion terms of mass and heat result in a second-order differential equation for all phases; consequently, two boundary conditions are necessary. According to Danckwerts (Wehner and Wilhelm, 1956), the Danckwerts’ boundary condition at z = 0 is

tmp5C-460_thumb[2][2]

which can be simplified to


tmp5C-461_thumb[2][2]

These boundary conditions are true because the axial dispersion of mass and heat is relatively small and the concentration and temperature gradients at the reactor inlet are quite flat (Chen et al., 2001). The Danckwerts’ boundary condition at z = LB is is the real condition at the outlet of the reactor, but as it is an infinite type of boundary condition, it is difficult to apply under most numerical methods.

tmp5C-462_thumb[2][2]

The boundary condition

tmp5C-463_thumb[2][2]

TABLE 2.14. Boundary Conditions (t > 0) of Generalized Mass and Heat Balance Equations

Gas Phase

Liquid Phase

tmp364-464 tmp364-465 tmp364-466

No

No

Condition

Operation Mode

Dispersion

Dispersion

Dispersion

Dispersion

tmp364-467

Co – current

tmp364-468 tmp364-469 tmp364-470 tmp364-471
tmp364-472 tmp364-473 tmp364-474 tmp364-475

Countercurrent

tmp364-476 tmp364-477 tmp364-478 tmp364-479
tmp364-480

Co – current

tmp364-481 tmp364-482 tmp364-483

-

Countercurrent

tmp364-484 tmp364-485 tmp364-486
tmp364-487

Co – current Countercurrent

tmp364-488 tmp364-489 tmp364-490 tmp364-491
tmp364-492

Co – current/ countercurrent

tmp364-493 tmp364-494 tmp364-495 tmp364-496
tmp364-497

Co – current/ countercurrent

tmp364-498 tmp364-499 tmp364-500 tmp364-501

Because of this, many researchers prefer to use Danckwerts’ boundary conditions, where the temperature gradients at the exit of the reactor are zero. It must be pointed out that unless the reactor is infinitely long, this exit condition will not be true because the fluid will not reach equilibrium with its surroundings. This lack of equilibrium at the exit would be even more pronounced at the high temperatures frequently experienced in TBRs. The boundary conditions proposed by Young and Finlayson (1973) avoid this abnormality. They derived inlet and outlet conditions similar to those of Danckwerts for the case in which both axial and radial dispersion is present. However, the outlet condition was reported with a nonzero gradient, which is reduced to Danckwerts form when radial dispersion is neglected. The boundary condition dTS / dz = 0 for z > Lb isa Danckwerts type also, but as the solid phase is commonly nonexistent for z > LB, it is generally regarded as a valid exit condition. For the same reason, as the reaction does not take place without the solid catalyst, the concentration gradients can be assumed to be zero at the exit. Therefore, the Danckwerts boundary conditions are used when there is no mass (or heat) dispersion outside the reactor (Chao and Chang, 1987- Feyo De Azevedo et al., 1990).

Solid Phase

tmp364-502

Dispersion

No Dispersion

Surface (i=All Compounds)

Interior (i = All Compounds)

tmp364-503 tmp364-504 tmp364-505
tmp364-506 tmp364-507 tmp364-508
tmp364-509 tmp364-510 tmp364-511
tmp364-512 tmp364-513 tmp364-514
tmp364-515 tmp364-516 tmp364-517
tmp364-518 tmp364-519 tmp364-520
tmp364-521 tmp364-522 tmp364-523
tmp364-524 tmp364-525 tmp364-526
tmp364-527 tmp364-528 tmp364-529
tmp364-530 tmp364-531 tmp364-532 tmp364-533
tmp364-534 tmp364-535 tmp364-536 tmp364-537

The boundary condition attmp5C-538_thumb[2][2]is sometimes considered to betmp5C-539_thumb[2][2]

(Shokri and Zarrinpashne, 2006; Shokri et al., 2007). According to Chen et al. (2001), the following boundary conditions will be employed for heat transfer in the thermowell:

tmp5C-540_thumb[2][2]

The general energy heat balance for the liquid phase is

tmp5C-541_thumb[2][2]

Assuming no temperature gradient within the catalyst particles (isothermal catalyst), the general heat balance for the solid phase is

tmp5C-542_thumb[2][2]

tmp5C-543_thumb[2][2]

The resulting set of PDEs coupled with the respective initial and boundary conditions are then solved simultaneously using an appropriate numerical method (Melis et al., 2004 ).

Example of Simplification of the Generalized Model Sometimes, to simplify heat transfer modeling in HDT reactors, the three processes involved (heat transfer in the solid, liquid, and gas phases) can be lumped into only one equation with a pseudohomogeneous heat balance. In the following, an example of how to obtain a pseudohomogeneous heat balance from the generalized heat balance presented in Table 2.12 is shown.

The general heat balance for the gas phase is When modeling commercial HDT reactors it is generally accepted that axial dispersion can be omitted; then HA3 = HB3 = HC3 = 0. If adiabatic operation is also assumed, then HA4 = HB4 = HC4 = 0 (isothermal reactor in the radial direction) and HA8 = HB8 = 0 (no heat transfer between the fluid phase and the reactor wall).

The pseudohomogeneous model is based on the fact that the temperature difference among the gas, liquid, and catalyst at any particular axial position of the reactor is negligible. Hence, TG = TI = TL = TSS = T, and in consequence the following terms can be neglected: HA5 = HA7 = HB5 = HB7 = HC9 = HC10 = 0. The temperature gradients TG – TI and TI – TL in the HA6 and HB6 terms are also neglected. The final heat balance equation for the gas, liquid, and solid phases is then

tmp5C-544_thumb[2][2]

Equation (2.137) in terms of model parameters for co-current operation gives

tmp5C-545_thumb[2][2]

The final simplified pseudohomogeneous heat balance is

tmp5C-546_thumb[2][2]

If the catalyst wetting efficiency is also assumed to be complete, fw = 1 (e.g., when properly designed liquid distributors and high liquid velocities in commercial reactors are used). Therefore, reactions occur only in the liquid phase; then Eq. (2.139) is rewritten as

tmp5C-547_thumb[2][2]

According to Feyo De Azevedo et al. (1990), some researchers have included the radiation effects in nonisothermal reactors in the radial direction (HA4 = HB4 = HC4 ^ 0) by means of the effective radial thermal conductivity

tmp5C-548_thumb[2][2]

wheretmp5C-549_thumb[2][2]is the conductive plus convective contributions to the effective radial conductivity,tmp5C-550_thumb[2][2]the radiant contribution,tmp5C-551_thumb[2][2]the Stephan-Boltzmann constant, andtmp5C-552_thumb[2][2]a radiant transfer factor defined as

tmp5C-553_thumb[2][2]

where e is the particle emissivity. Therefore, rearranging Eq. (2.151) as

tmp5C-555_thumb[2][2], the pseudohomogeneous radial heat dispersion term (HA4 + HB4 + HC4 withtmp5C-556_thumb[2][2]is expressed as

tmp5C-554_thumb[2][2]

where the last term on the right side of Eq. (2.144) represents the radioactive heat transfer, and its exclusion may produce errors in estimation of some other heat transfer parameters.

Estimation of Model Parameters

To solve the set of ordinary differential equations (ODEs) (for the steady-state regime) or the set of PDEs (for the dynamic regime), it is necessary to evaluate several parameters and chemical properties of the system. Those parameters can be estimated with existing correlations, whose accuracy is of great importance for the entire state of robustness of the reactor model.

Effective Diffusivity The description of steady-state diffusion and reaction of a multicomponent liquid mixture in a porous catalyst particle requires appropriate definition of the species fluxes to the particle. Fick’s law for equi-molar counterdiffusion through an ideal cylindrical pore is given by

tmp5C-557_thumb[2][2]

where Df is the effective diffusivity, by means of which the structure (porosity and tortuosity) of the pore network inside the particle is taken into consideration in the modeling. Table 2.15 shows Bosanquet’s formula to estimate the effective diffusivity inside the catalyst particle (Bosanquet, 1944), which consists of two diffusion contributions: Knudsen diffusivity (DfK) and molecular diffusivity (DfMi). Both Bosanquet’s formula and Fick’s law can also be applied with sufficient accuracy to cases involving a narrow unimodal pore-size distribution and very dilute mixtures. The restrictive factor F(Xg) accounts for additional friction between the solute and the pore walls. The exponent Z is 4 for Xg < 0.2 (Iliuta et al., 2006). The tortuosity factor of the pore network, t, is used in the calculation of Dfei because the pores are not oriented along the normal direction from the surface to the center of the catalyst particle, and its value generally varies between 3 and 7, but for HDT process it is commonly assumed to be equal to 4. It is also possible to estimate the tortuosity factor assuming valid the upper bound correlation given by Weissberg (1963) for a packing of random spheres as shown in Table 2.15.

Effectiveness Factor The effectiveness factor of independent reactions can be defined as the ratio of the volumetric average of the reaction rate into the particle to the reaction rate at the surface of the particle as proposed by Thiele (1939) and Zeldovich (1939) :

tmp5C-558_thumb[2][2]

Analytical solutions for Eqs. (2.146) and (2.147) are possible only for single reactions and for zero- and first-order rate expressions. The various correlations used in the literature to estimate the catalyst effectiveness factor for isothermal and irreversible reactions are shown in Table 2.16.

For kinetic models other than the power-law approach, such as Langmuir-Hinshelwood-Hougen-Watson (LHHW)-type kinetic expressions, there is no analytical solution of Eq. (2.146) – Therefore, an alternative method to avoid the numerical integration of Eq. (2.146) is the Bischoff generalized modulus approach (Bischoff, 1965), which enables an analytical solution to any type of rate equation and single reaction.

TABLE 2.15. Estimation of Effective Diffusivity

Parameter

Gas Phase Liquid Phase

Molecular diffusivity coefficient

tmp364-559

Knudsen diffusivity coefficient

tmp364-560

Effective diffusivity

tmp364-561

Restrictive factor

tmp364-562

Ratio of radius of gyration over pore radius

tmp364-563

Mean pore radius

tmp364-564

Tortuosity factor

tmp364-565

Binary diffusion

coefficient (P in atm)

tmp364-566

Dynamic liquid viscosity

tmp364-567

Molar volume of solute (i) in liquid phase and liquid solvent (L)

tmp364-568

Solvent critical specific volume

tmp364-569

As mentioned previously, the effectiveness factor in commercial HDS catalysts has been reported to be in the range 0.4 to 0.8.Expressions for isothermal first-order reactions with irregularly shaped catalysts lead under steady-state conditions to acceptable results with errors not exceeding 20% (Aris, 1975; Dudukovic, 1977). Bischoff (1965) proposed a general modulus to predict the effectiveness factor for any reaction type within a relatively narrow region. If reactions of order less than one-half are excluded, the spread between all the various curves is about 15%. The mean deviation of values of n calculated from the empirical correlation proposed by Papayannakos and Georgiou (1988) is less than 2.4% from those predicted with the normalized modulus for simple order reactions proposed by Froment and Bischoff (1990) in the range 0.05 < n < 0.99.

TABLE 2.16. Estimation of Catalyst Effectiveness Factor

Kinetic Model

Reaction Order

Shape

Thiele Modulus

Effectiveness Factor

Power law

tmp364-570

Spheres and crushed

tmp364-571 tmp364-572
tmp364-573 tmp364-574 tmp364-575 tmp364-576
tmp364-577

Any geometry

tmp364-578 tmp364-579
tmp364-580

Any geometry

tmp364-581 tmp364-582
tmp364-583

Any geometry

tmp364-584 tmp364-585

LHHW

tmp364-586

Spheres and crushed

tmp364-587 tmp364-588
tmp364-589

Pellet, cylinder, 2-, 3 – lobe, etc.

tmp364-590 tmp364-591

Global Gas-Liquid Mass Transfer The gas-liquid interphase mass transfer flux is described in terms of the simple two-film theory:

tmp5C-592_thumb[2][2]

The overall external resistance to mass transfer (KLi) is composed by the resistance to mass transfer in the gas ( kG) and liquid ( kt) films. Estimates of the compressibility factor Z sometimes give values close to 1; therefore, ideal gas law could be used (Mejdell et al., 2001).

For slightly soluble gases such as H2 , the value of Henry’ s constant (Hi ) exceeds unity, and then mass transfer resistance in the gas film can be neglected (Zhukova et al., 1990). Therefore, the total mass transfer is approximately equal to the liquid-side mass transfer coefficient:

tmp5C-593_thumb[2][2]

The liquid film mass transfer coefficient ( k;L) is calculated using the correlations reported in Table 2.17 .

Gas-Liquid Equilibrium The gas-liquid equilibrium along the catalyst bed is represented in the mass balance equations by the Henry’s law constant. The constants related to this law for different chemical species available in the system may be defined in two ways, described below:

1. Solubility coefficients. Employing solubility coefficients, the following expression is used to estimate Henry’s constant:

tmp5C-594_thumb[2][2]

where stands for the component – solubility and vN is the molar volume under normal conditions. Using this expression implies knowledge of the gaseous component solubility in the liquid phase considering the process temperature effect. Korsten and Hoffmann (1996) have reported the next correlations to evaluate this parameter only for H2 and H2S: for hydrogen sulfide. However, this last correlation may not be no adequate to evaluate the solubility of H2S in oil fractions at the complete temperature range used in commercial units. In that case, it is possible to use an EoS to estimate this Henry’s constant.

TABLE 2.17. Correlations to Estimate Model Parameters

Parameter

Symbol

References

Parameter

Symbol

References

Holdup gas

tmp364-595

Calculated from £b = £l + £g

Density of the liquid

tmp364-596

Ahmed (1989)

tmp364-597

phase

tmp364-598

Holdup liquid

tmp364-599

Charpentier and Favier (1975), Satterfield (1969),

Catalyst bulk density

tmp364-600

ASTM (2003)

tmp364-601

Specchia and Baldi (1977), Ellman et al. (1990)

tmp364-602

Bed void fraction

tmp364-603

Haughey and Beveridge (1969), Carberry and

Dynamic viscosity of

tmp364-604

Glaso (1980) , Ahmed (1989) ,

tmp364-605

Varma (1987), Froment and Bischoff (1990)

liquid

tmp364-606

Brule and Starling (1984)

Catalyst wetting efficiency (or

tmp364-607

Ring and Missen (1991), Al-Dahhan and

Dynamic viscosity of

tmp364-608

Ahmed (1989), Brule and Starling

contacting effectiveness)

tmp364-609

Dudukovic (1995)

gas

tmp364-610

(1984)

Two – phase pressure drop

tmp364-611

Larkins et al. (1961) , Ellman et al. (1988)

Diffusion coefficients

tmp364-612

Wilke and Chang (1955)

Gas-liquid interfacial area

tmp364-613

Iliuta et al. (1999)

for gases Diffusion coefficients

tmp364-614

Tyn and Calus (1975)

tmp364-615

for liquids

tmp364-616

Gas-solid interfacial area

tmp364-617

Puranik and Vogelpohl (1974), Onda et al. (1967)

Specific heat of gas

tmp364-618

Lee and Kesler (1975), Perry et al.

tmp364-619

phase

tmp364-620

(2004)

Mass transfer gas-solid

tmp364-621

Petrovic and Thodos (1968) , Dwivedi and

Specific heat of

tmp364-622

Lee and Kesler (1975, 1976),

coefficient

tmp364-623

Upadhyay (1977)

liquid phase

tmp364-624

Perry et al. (2004)

Mass transfer liquid-solid

tmp364-625

Dudukovic et al. (2002), Dwivedi and Upadhyay

Specific heat of solid

tmp364-626

Perry et al. (2004)

coefficient

tmp364-627

(1977), Bird et al. (2002), Evans and Gerald

phase

tmp364-628
tmp364-629

(1953) , Wilson and Geankoplis (1966) , Goto

tmp364-630
tmp364-631

and Smith (1975), Satterfield et al. (1978),

tmp364-632
tmp364-633

Specchia et al. (1974)

tmp364-634

Mass transfer coefficient of

tmp364-635

Goto and Smith (1975)

Gas-liquid heat

tmp364-636

Marroqum de la Rosa et al.

liquid side at G-L interface

tmp364-637

transfer coefficient

tmp364-638

(2002), Chilton and Colburn

tmp364-639 tmp364-640

(1939)

Mass transfer coefficient of

tmp364-641

Goto and Smith (1975), Reiss (1967),Yaici et al.

Liquid-solid heat

tmp364-642

Chilton and Colburn (1939)

gas side at G-L interface

tmp364-643

(1988)

transfer coefficient

tmp364-644

Axial dispersion of gas

tmp364-645

Hochman and Effron (1969), Sater and

Chilton-Colburn

tmp364-646

Froment and Bischoff (1990) , Hill

tmp364-647

Levenspiel (1966) , Demaria and White (1960)

‘-factor for energy

tmp364-648

(1977), Bird et al. (2002), Gupta

tmp364-649

transfer

tmp364-650

et al. (1974)

Axial dispersion of liquid

tmp364-651

Gierman (1988) , Hochman and Effron (1969) ,

Effective thermal

tmp364-652

Hashimoto et al. (1976)

tmp364-653

Sater and Levenspiel (1966), Tsamatsoulis and

conductivity radial

tmp364-654
tmp364-655

Papayannakos (1998)

tmp364-656

Radial dispersion of gas

tmp364-657

Fahien and Smith (1955)

Effective thermal

tmp364-658

Tarhan (1983) , Dixon (1985)

tmp364-659

conductivity axial

tmp364-660

Radial dispersion of liquid

tmp364-661

Fahien and Smith (1955), De Ligny (1970),

Thermal conductivity

tmp364-662

API (1997) , Chung et al. (1988)

tmp364-663

Herskowitz and Smith (1978a,b)

of f phase

tmp364-664

Binary interaction parameter

tmp364-665

Moysan et al. (1983), Ronze et al. (2002), Riazi

Heat of reaction

tmp364-666

Tarhan (1983)

for H2-oil using PR EoS

tmp364-667

(2005), Lal et al. (1999), Magoulas and Tassios

tmp364-668
tmp364-669

(1990)

tmp364-670

Binary interaction parameter

tmp364-671

Feng and Mather (1993a,b), Carroll and Mather

Heat of vaporization/

tmp364-672

Soave (1972) , Peng and Robinson

for H2S – oil using PR EoS

tmp364-673

(1995)

condensation

tmp364-674

(1976)

Binary interaction parameter

tmp364-675

API (1997)

Thiele modulus

tmp364-676

Bischoff (1965)

for NH3- oil using SRK EoS

tmp364-677 tmp364-678

Density of the gas phase

tmp364-679

Soave (1972) , Peng and Robinson (1976)

Effectiveness factor

tmp364-680

Froment and Bischoff (1990) , Aris (1975)

 

tmp5C-681_thumb[2][2]

for hydrogen and

tmp5C-682_thumb[2][2]

2. Equation ofstate. It is also possible to obtain Henry-s constant assuming local equilibrium at the liquid-gas interface:

tmp5C-683_thumb[2][2]

where the equilibrium constant (Ke q,i ) is calculated using an adequate EoS (e.g., Peng-Robinson (PR), Soave-Redlich-Kwong (SRK), t-van der Waals, Grayson-Streed). For this expression, it was necessary to calculate a two-phase thermodynamic equilibrium previously at each local point along the catalytic bed to estimate the interface temperature and the molar compositions in liquid and gas phases. The main advantage of this expression is that it takes into account the volatility of the feedstock; however, it also increases the computing time too greatly.

Another way to calculate the Henry’s constant of gaseous solute in a solvent is to use the next thermodynamic assumption:

tmp5C-684_thumb[2][2]

where yt is the fugacity coefficient of a gaseous compound i (solute) in the liquid phase (solvent), and its calculation using EoS is addressed below. The SRK and PR equations of state are the most widely used in HDT process modeling, being defined by the generalized expression

tmp5C-685_thumb[2][2]

For pure compounds, values of parameters a and b are given by

tmp5C-686_thumb[2][2]tmp5C-687_thumb[2][2]

The generalized temperature function a (Tf) was proposed by Soave (1972) to be an equation of the form with

tmp5C-688_thumb[2][2]

with

tmp5C-689_thumb[2][2]

When applied to mixtures, the classical mixing rules may be considered to evaluate parameters a and b:

tmp5C-690_thumb[2][2]

where ki k are the binary interaction parameters, which may be obtained from the references reported in Table 2.17. It is important to point out that the quality of Henry’s constant calculation depends enormously on the accuracy of these interaction parameters. The liquid-phase fugacity coefficient can be derived from Eq. (2.155) to give

tmp5C-691_thumb[2][2]

with

tmp5C-692_thumb[2][2]

TABLE 2.18. Parameters of Soave-Redlich-Kwong and Peng-Robinson Equations of State

Parameter

SRK

EoS

PR

tmp364-693

1

tmp364-694
tmp364-695

0

tmp364-696
tmp364-697

0.42748

tmp364-698
tmp364-699

0.08664

tmp364-700
tmp364-701

0.48

tmp364-702
tmp364-703

1.574

tmp364-704
tmp364-705

-0.176

tmp364-706
tmp364-707

1

tmp364-708
tmp364-709

-1

tmp364-710
tmp364-711 tmp364-712 tmp364-713
tmp364-714

-AB

tmp364-715

tmp5C-716_thumb[2][2]

where ZL is the compressibility factor of the liquid phase at saturation obtained from the solution of Eq. (2.155) expressed in its cubic compressibility factor form:

To evaluate the liquid-solid mass transfer coefficients, for example, Eq. (2.168) must be expressed as

tmp5C-717_thumb[2][2]

tmp5C-718_thumb[2][2]

The values of the universal parameterstmp5C-719_thumb[2][2] and DZ) are given in Table 2.18 .

Heat Transfer Coefficients Correlations employed for mass transfer can be used to calculate the parameters for energy transfer between phases by employing the Chilton and Colburn (1939) analogy. The Chilton-Colburn ‘-factor for mass transfer (jD) is given by

tmp5C-720_thumb[2][2]

where g(Re/) is a correlation that is a function of the Reynolds number of the respective phase f to be evaluated; some of these correlations are shown in Table 2.17 – If one needs to estimate the liquid-solid heat transfer coefficient, for example, Eq. (2.169) is rewritten in the expression

tmp5C-721_thumb[2][2]

There are various correlations to estimate the mass transfer coefficients, but the Chilton-Colburn analogy is not usually employed to evaluate them, whereas due to the lack of correlations to estimate the heat transfer coefficients in the gas or liquid film side at the gas-liquid interface and in the liquid film at the liquid-solid interface, it is common to use the Chilton-Colburn analogy, by equating Eqs. (2.168) and (2.170) (jD = jH), to estimate these coefficients. The physical and geometrical properties involved in the dimensionless numbers must be evaluated at conditions of reaction and for each phase in a heterogeneous reactor.

To use the Chilton-Colburn analogy, it is necessary to consider the following conditions (Bird et al., 2002):

• Constant physical properties

• Small net mass transfer rates

• No chemical reaction

• No viscous dissipation heating

• No absorption or emission of radian energy

• No pressure diffusion, thermal diffusion, or forced diffusion

Theoretical Calculations of Some Parameters Relative to the Catalyst

Bed Dilution of the catalyst bed with inert material is a common practice in experimental HDT reactors (Sie, 1996). The following simple formula is employed to calculate the dilution factor:

tmp5C-722_thumb[2][2]

where Vc is the catalyst volume and V is the volume of inert particles, both obtained experimentally.

An equivalent particle diameter (dpe), defined as the diameter of a sphere that has the same external surface (or volume) as the actual catalyst particle,

On the other hand, the Chilton-Colburn /-factor for energy transfer (jH) is given by

tmp5C-723_thumb[2][2]

is an important particle characteristic that depends on particle size and shape. For fixed beds with catalyst extrudates of commercial size, the equivalent particle diameter can be calculated according to Cooper et al. (1986):

tmp5C-724_thumb[2][2]

Bed void fraction (or bed porosity) for undiluted catalyst bed can be calculated with the following correlation reported by Froment and Bischoff (1990), Haughey and Beveridge (1969), and Carberry and Varma (1987):

tmp5C-725_thumb[2][2]

This correlation was developed for undiluted packed beds of spheres; however, if the equivalent particle diameter concept is used, it can also be employed for nonsphere particles. Once the bed void fraction is determined, the particle density can be calculated as follows (Tarhan, 1983):

tmp5C-726_thumb[2][2]

Since the continuous models are based on the volume-average form of the transport equations for multiphase systems, equations expressing conservation of volume are (Whitaker, 1973)

tmp5C-727_thumb[2][2]

Relationships between phase holdups inside the catalyst solid are given in the following expressions:

tmp5C-728_thumb[2][2]

The external surface area of catalyst particles per unit of reactor volume for PBRs can be calculated as

tmp5C-729_thumb[2][2]

Catalyst porosity (eS) may be calculated with the following equation from the experimental data for total pore volume (Vg):

tmp6C5-1_thumb

In an extreme case where the experimental Sg parameter is not available in order to estimate the average pore radius, one can use the correlation proposed by Mace and Wei (1991):

tmp6C5-2_thumb

where

tmp6C5-3_thumb

Some parameters that account for bed characterization are experimentally measurable, others are experimental or can be obtained through simulations, and others are empirical. Of course, although it is better to obtain the local porosity experimentally, measurements required the use of advanced techniques. To do that, computational calculations are preferred.

Most of empirical correlations for predicting liquid saturation, pressure drop, and flow regimes are based on experiments performed at atmospheric conditions. Since industrial trickle-bed reactors are operated at high pressures and temperatures, the applicability of these correlations for such operating conditions needs to be investigated (Nguyen et al., 2006). The majority of correlations that have been developed on a laboratory scale may not work for large-scale reactors (operated at high pressure and temperature) due to significant changes in hydrodynamic characteristics with a reactor scale (Gunjal and Ranade, 2007). Although extensive studies in hydrodynamic correlations are available in the literature (Dudukovic et al., 2002) it seems that many works in modeling reactors for petroleum fractions still employ the classical correlations (i.e., those derived from reasonable assumptions with simple expressions). Some researchers have employed the same correlations of previous papers without checking the accuracy of these expressions or the range in which they are applicable. On the other hand, although the benefit of using correlations based on neural networks has been reported, many data are necessary to employ this approach. To overcome this feature, an alternative is to create a database with different types of crude oils and fractions in order to develop an online program able to determine the various parameters involved in modeling a HDT reactor. At present it is not available. The only valuable effort seems to be that of Larachi et al. (1999), who have developed a simulator based on neural networks for the prediction of some hydrodynamic parameters for trickle-bed reactors. Although neural networks are updated continuously, which favors its use for parameter predictions, it is necessary to develop a fundamental relationship that takes advantage of novel techniques for characterization of hydrodynamic parameters.

The limitations of the various models reported in the technical literature are closely related to the number of parameters involved and to the reliability of the data available. Therefore, appropriate correlations for mass and energy transfer should be employed in order to calculate each of the terms used in a model reactor (i.e., correlations developed under similar conditions, such as the same flow regime, pressure, and temperature, assuming similar liquid system and porous particles). Generalized correlations for the prediction of properties having a broad range of variation should be avoided when modeling a TBR in detail, because they could produce some miscalculations. Constant values assumed a priori, such as tortuosity factor, binary interaction parameters, heat of reaction, and specific heat, should be used as a reference when experimental data are available. To simplify a TBR model, some researchers have ignored the low heat of some reactions because its contribution is not significant in the energy balance, which seems to be a reasonable assumption.

Using different approaches to fluid dynamics, kinetics or thermodynamics can lead to very different conclusions in predictions of reactor performance. For example, Gunjal and Ranade (2007) have reported 15% more conversions considering all parameters to be constant and assuming only uniform porosity (i.e., no nonuniform porosity). Akgerman et al. (1985) has reported 24 to 38% higher conversions when considering volatiles with respect to nonvolatiles, and Inoue et al. (2000) has predicted the use of larger reactors employing an nth kinetic model when more accurate kinetic expression has been utilized. Another important finding is the consideration of a thermowell in the energy balance for a pilot plant, as pointed out by Chen et al. (2001). These are the reasons to account for detailed models which allow both for making accurate descriptions of the chemical phenomena and for reliable preliminary calculations when designing a TBR reactor.

Next post:

Previous post: