Home Kinetic theory based multiphase flow with experimental verification
Article Publicly Available

Kinetic theory based multiphase flow with experimental verification

  • Dimitri Gidaspow

    Dimitri Gidaspow was introduced to multiphase flow in 1972 by his first former PhD student, Dr. Charles W. Solbrig, who was directing the development of computer codes for licensing nuclear reactors by the Atomic Energy Commission in Idaho. Earlier at IIT, Gidaspow had taught a numerical methods course with Solbrig which was the beginning of the 642 page book, Computational Techniques, by D. Gidaspow and V. Jiradilok, Nova Science, 2009. In 1984 he received the AIChE Kern Award for his earlier work on energy conversion. His lecture on fluidization and heat transfer was published in Applied Mechanics Review. In 1994 he completed his highly cited book, Multiphase flow and fluidization. He received three other awards in fluidization from AIChE.

    EMAIL logo
    and Marcelo S. Bacelos

    Marcelo S. Bacelos received his PhD in chemical engineering in 2006 from Federal University of Sao Carlos, Brazil. He has been teaching transport phenomena in the Engineering and Technology Department, Federal University of Espirito Santo, Brazil. His research has been focused on computational multiphase fluid dynamics to describe fluidized and spouted bed reactors. In the last 5 years, he has used gas-solid fluidized reactors for recovering waste, such as post-consumer carton packages. He has used fluidized and spouted beds to verify his CFD designs. Starting in 2016, he has been working with Prof. Gidaspow at IIT using kinetic theory to describe the gas-solid flow in fluidized and spouted bed reactors.

Published/Copyright: April 21, 2017
Become an author with De Gruyter Brill

Abstract

This review is an extension of our 2014 circulating fluidized bed (CFB) plenary lecture. A derivation of multiphase mass, momentum and energy balances is presented, with a review of elementary kinetic theory, to explain the concepts of granular temperature and pressure and the core-annular flow regime commonly observed in CFB. The kinetic theory shows that the particle concentration is given by the reciprocal of a fourth order parabola of dimensional tube radius, in agreement with experiments. Computed flow regimes and heat and mass transfer coefficients in fluidization are also discussed.

1 Introduction

Multiphase flow models were reviewed in 1982 by Lyczkowski et al. (1982). At that time Soo’s (1967) book by the then unique title of multiphase flow treated the multiphase pressure as the product of fluid pressure times the phase volume fraction. Such a model has the huge unrealistic force of pressure times the phase volume fraction gradient in the momentum equation. Professor Wallis’s (1969) book on one-dimensional two-phase flow has the fluid pressure times the phase volume fraction in both the gas and the liquid momentum balances. As a part of the study of safety of nuclear reactors and their licensing, we had shown that these equations were ill-posed as an initial value problem. For the 5th International Heat Transfer held in Tokyo, Japan, in 1974, Gidaspow solicited the opinions of the world’s experts on the subject. He showed that for incompressible fluids, the characteristics, the paths along which information is propagated of the four equations, and the mass and momentum balances for each phase have two imaginary roots for unequal phase velocities. For fluidized beds where the particle and fluid velocities differ a great deal, this term can become large and lead to instabilities.

In his 1994 book Gidaspow shows that the four characteristics for these equations are the phase velocities and plus and minus sonic type velocities for the momentum balances derived using granular flow type kinetic theory. The kinetic theory shows that each phase has its own pressure. The fluid pressure must appear only in the fluid phase in the fluid momentum balance. Unfortunately, most commercial computational fluid dynamics (CFD) codes have the fluid pressure in both phases, as in the book by Wallis (1969).

The pressure in the particulate phase is due to oscillations and particle collisions. It is related to a new concept pioneered by Professor Savage (1983), the granular temperature of the particles. Kinetic theory shows that this granular pressure is related to its granular pressure in the particulate phase momentum balance. However, this new theory clearly needs experimental confirmation, which was done in the last three decades, as partially summarized in the Computational Techniques book by Gidaspow and Jiradilok (2009).

Today, CFD is an emerging tool for the design of circulating fluidized beds (CFB). In the oil industry CFBs replaced bubbling beds when active fluidized catalytic cracking (FCC) catalysts were developed four to five decades ago. The oil industry and the refinery designers, such as universal oil products (UOP), did not know that their risers operate in the core-annular flow regime (Gidaspow and Jiradilok 2009) until the fluidized beds were modeled using kinetic theory of granular flow (Sinclair and Jackson 1989). The kinetic theory approach was reviewed 10 years ago in the 2002 Flour Daniel lecture (Gidaspow et al. 2004), and the CFD approach was described in Arastoopour’s 1999 lecture (Arastoopour 2001). Since then the kinetic theory approach was extended to flow of mixtures of particles of various sizes by several groups, including rotation (Songprawat and Gidaspow 2010, Shuai et al. 2012), to anisotropic flow using the method of moments (Strumendo et al. 2005, Juhui et al. 2012, Huang et al. 2013) and to continuous particle size distributions using population balances (Strumendo and Arastoopour, 2010). Recently, it was shown that the core-annular flow can be eliminated (Khongprom and Gidaspow 2010, Kashyap et al. 2011).

In 1995, Berruti et al. (1995) published an excellent review of fluidized bed risers. Werther (2005) reviewed CFB combustors and Chen and Williams (2005) the FCC technologies. Computational approaches to fluidization are discussed in a recent book by Pannala et al. (2011).

Our review is based on the 2014 Circulating Fluidized Bed plenary lecture (Gidaspow and Arastoopour 2014).

2 Multiphase conservation laws

For modelling of multiphase flow we need two concepts: conservation laws and constitutive equations. The conservation laws can be obtained from the Reynolds Transport Theorem, schematically shown in Figure 1.

Figure 1: Reynolds transport theorem.
Figure 1:

Reynolds transport theorem.

2.1 Mass balances

The mass of phase i in a volume element V moving with phase i velocity is as follows:

(1)mi=εiρidV

where εi is the volume fraction of phase i, and ρi is defined as the density of phase i. This volume fraction is the only new variable in multiphase flow. Traditionally, this volume fraction, εi, was derived by volume averaging. For multicomponent systems, εi is 1, because it is assumed that molecules of component i occupy the same space at the same time. Such an approximation cannot be made in multiphase flow.

The balance on mass mi moving with the velocity vi is

(2)dmidti=ddtiεiρidV=midV

where mi′ is the rate of production of phase i; d/dti is the substantial derivative moving with phase i velocity.

The Reynolds Transport Theorem gives the continuity equation for phase i:

(3)(εiρi)t+(εiρivi)=mi

Conservation of mass requires that

(4)i=1nmi=0

The volume fraction of phase i was defined as follows:

(5)Vi=εidV

where i=1nεi=1

For incompressible fluid with no phase change, the continuity equation becomes

(6)εit+(εivi)=0

where i=1, 2, …n phases

Such an equation (6) does not appear in conventional transport phenomena theory.

2.2 Momentum balances

The rate of change of momentum of phase i moving with the velocity vi equals the forces acting on the system and can be mathematically expressed by the following equation:

(7)ddtρiviεidV=fi

where force fi is given by

(8)fi=Tida+ρiFiεidV+pidV+mividV

The first term represents the surface forces acting on the differential area da; the second term is the external forces; the third term is the interaction forces between phases, and the last term is the force due to the phase change.

The stress tensor Ti for phase i is composed of nine components and can be written as follows:

(9)Ti=(TixxTixyTixzTiyxTiyyTiyzTizxTizyTizz)

Figure 2 represents the stress tensor acting on surfaces of volume of phase i.

Figure 2: Stress tensor acting on surfaces of volume of phase i.
Figure 2:

Stress tensor acting on surfaces of volume of phase i.

The Reynolds Transport Theorem is applied to the left side of equation (7) and divergence theorem (Tida=TidV) to the first term of equation (8). The result is the momentum balance for phase i:

(10)(ρiεivi)t+(ρiεivivi)=Ti+ρiεiFi+pi+mivi

Syamlal et al. (2016) used a different treatment of momentum transfer in their MFIX code.

Using the continuity equation (3), the momentum balance for phase i moving with phase i velocity becomes

(11)ρiεidvidti=Ti+ρiεiFi+pi

where ρiεidvidti is acceleration of phase i; ∇·Ti is the momentum in flow due to the surface forces; ρiεiFi is the body force, and pi is the interaction force between phases.

where i=1n phasepi=0

The interaction forces between phases (pi) account for the drag between the phases. Note that moving with the velocity vi, there is no force due to the phase change, because the constant mass in the balance (7) moved with the velocity vi.

Hence, our form of phase change momentum is correct. This representation was verified experimentally and computationally for detonation of solid particles, such as TNT, in which the solid particles are rapidly converted into gases (Aldis and Gidaspow 1990, Pape and Gidaspow 1998).

The simplest expression for the stress is

(12)Ti=PiI

Each phase i has its own pressure Pi similar to Euler equation for single-phase flow.

2.2.1 Incompressible viscous flow

To meet the requirement of objectivity for each phase k,

(13)Tk=Tk(svk)

where ∇svk is the symmetrical gradient of vk.

Linearization of the Tk gives

(14)Tk=AkI+Bk(svk)

For incompressible fluids, Ak is chosen to be the negative of the pressure of fluid k and the derivative of the traction with respect to the symmetric gradient is the viscosity of fluid k, as shown below.

Ak=Pk

Bk=2μk=Tksvk

The traction for phase k

(15)Tk=PkI+2μksvk

where

svk=12[vk+(vk)T]

2.2.2 Incompressible Navier-Stokes equation

Substitution of stresses tensor for each phase k into the momentum balance (10) gives the incompressible equation of Navier-Stokes for each phase k.

(16)(εkρkvk)t+εkρkvkvk=εkρkFk+jβj(vjvk)+mkvkPk+μk2vk

where interaction forces between phases was expressed in terms of friction coefficients (βj). The friction coefficients can be expressed in terms of standard drag coefficients (Gidaspow 1994). In this model, the viscosity of each phase k (μk) is an input into the model. It can be either directly measured or obtained from kinetic theory (Gidaspow and Jiradilok 2009).

2.2.3 Compressible viscous flow

For compressible viscous flow, there is another parameter, bulk viscosity of phase k (λk).

(17)Ak=PkI+λkvk

Then the traction for phase k can be expressed as follows:

(18)Tk=PkI+2μksvk+λkIvk

where 2μk+3λk=0, similar to that for single-phase flow (Aris 1962).

Hence, the compressible Navier-Stokes equation for each phase k in substantial derivative form becomes as follows:

(19)εkρkdvkdtk=εkρkFk+jβj(vjvk)Pk+(2μksvk+23μkIvk)

2.3 Energy balance

The energy balance for an open system with phase change can be written as follows (Gidaspow 1994):

(20)dUidti=dQidtPidVidti+Diss+(Uin+Pin/ρin)dmidti

where Ui is the internal energy per unit mass of phase i; Qi is the heat per unit mass of phase i flowing into the system and the only work done by the system is the mechanical work due to the volume changes (Pi dVi). In the open system with phase change there is inflow due to phase change, consisting of the energy inflow (Uin) and pressure Pin. The energy dissipation by the system is Diss.

The rate of heat transfer is related to the flux, qi by the following relations:

(21)dQidt=qiεida=(εiqi)dV

In Equation (21) the heat flux qi need not be multiplied by the volume fraction of phase i.

The Reynolds Transport Theorem applied to the energy balance (19) produces energy equation for phase i.

(22)(εiρiUi)t+(εiρiviUi)=εiqiPiεitPiεivi+hinmi+ui

where Diss=uidV,uit internal energy per unit of volume, hin is the enthalpy of phase i flowing into the system; qi is the rate of heat transfer by conduction.

In the above equation there appears a strange term, the work due to the volume fraction changes of phase i, not found in single-phase flow.

2.3.1 Enthalpy representation

The energy balance expressed in terms of enthalpy of phase i per unit of mass (hi) in substantial derivative form becomes as follows:

(23)εiρidhidti=εiqi+εidPidti+Ui+mi(hinhi)

Equation (22) reduces itself to that found in standard transport phenomena text books for εi equal to 1 and (hin=hi). In enthalpy form there is no strange work term found in equation (22).

2.3.2 Entropy representation

The entropy balance for phase i moving with the velocity vi is as follows:

(24)ddtεiρiSidV+qiεidaTimiSidV=σidV

where ddtεiρiSidV is accumulation of entropy in the system moving with the velocity vi; qiεidaTi is the rate of entropy outflow due to flow of heat across area; miSidV is the rate of entropy at equilibrium due to the phase change; σidV is the rate of entropy production in phase i.

The combination of entropy balance (24) with the energy balance (20) generates the following equation for entropy production in the system, which is 0 for reversible processes and greater than 0 for irreversible processes.

(25)σi=1Tiεiqi+qiεiTi+UiTi=qiεiTiTi2+UiTi0

3 Elementary multiphase kinetic theory

3.1 Frequency distributions

The frequency distribution of velocities of particles, f, is a function of position, r and the instantaneous velocity, c, as well as time, t.

(26)f=f(t,r,c)

The six coordinates, the position, r, and the velocity, c, are sufficient to determine the location of a particle, since Newton’s second law has six integration constants.

The number of particles per unit volume, n, is the integral over the velocity space, c.

(27)n=fdc

The mean values of a quantity, ϕ, such as mass, momentum, energy and stress is defined to be in the usual way as

(28)nϕ=ϕfdc

Hence, the hydrodynamic velocity, v, is the integral over all the velocity space, as shown below:

(29)v=1ncf(c)dc

3.2 Peculiar velocity and transport

The transport of a quantity, ϕ, such as heat, must be invariant under a change of frame. Hence, it cannot be a function of the velocity, c. Otherwise, it will have different values in different frames of reference. But cv is independent of the frame of reference. Hence, we define the difference between the instantaneous (c) and the hydrodynamic velocities (v) as

(30)C=cv

In kinetic theory (Chapman and Cowling 1961), this difference is called the peculiar velocity. Its mean is 0, as shown below, since the mean of c is v:

(31)<C>=<cv>=vv=0

This property is the same as that of the turbulent velocity, defined as the instantaneous minus the time average velocity.

The flux vector of ϕ is defined as n <Cϕ (C)>. For example, if ϕ=E, the internal energy, then the conduction flux, q, becomes q=n <EC>.

Since momentum is the mass, m times the velocity, C, the kinetic stress tensor, Pk, is as follows:

(32)Pk=n<CmC>=ρ<CC>

since the bulk density, ρ=nm. Table 1 shows the components of the stress tensor.

Table 1:

Kinetic stress tensor.

Pk=ρCC=(ρCx2ρCxCyρCxCzρCyCxρCy2ρCyCzρCzCxρCzCyρCz2)
Since nCC=CCfdC,
Pk=(ρCx2fdCρCxCyfdCρCxCzfdCρCyCxfdCρCy2fdCρCyCzfdCρCzCxfdCρCzCyfdCρCz2fdC)
where ρ=Bulk density=εsρs and where f=f(C)
Note that if f(C)=fx (Cxfy (Cyfz (Cz ) as in Maxwellian distribution, then
Pxy=ρCxCyf(C)dC=ρfzdCzCyfydCyCxfxdCx
But CxfxdCx=Cx=cxvx=vxvx=0
Hence, for a Maxwellian distribution, Pxy =Pzx =Pij =0(ij)

For a Maxwellian distribution, the kinetic viscosity is 0. Particle viscosity is nonzero due to collisions and due to a non-Maxwellian distribution.

The hydrostatic pressure, p is the mean of the sum of the normal components of the stress tensor p:

(33)p=1/3(pxx+pyy+pzz)

3.3 Granular temperature and the equation of state

In kinetic theory of gases the thermal temperature, T is defined as the average of the random kinetic energy, with the conversion factor of the Boltzmann constant from joules to degrees kelvin, as shown below.

(34)kBT=1/3m<C2x+C2y+C2z>

where the subscript of the peculiar velocity C indicates the component of C in the x, y and z directions, respectively. The Boltzmann constant has the value

(35)kB=1.3805×1023J/K

The ideal gas law constant equals the very small Boltzmann constant, due to the small mass of the molecule, m times the large value of the Avogadro’s number, 6.023×1023, the number of molecules per mole. Converting from joules to calories gives the gas law constant of 1.987 cal/g mole ° K.

Elimination of the squares of the peculiar velocities in

(36)P=1/3nm<C2>

and in the definition of temperature, Equation (34), gives the ideal gas law equation of state,

(37)P=nkBT=(N/V)RT

where N is the number of moles, V is the volume, and R is the gas constant.

The granular temperature is defined as the random kinetic energy of the particles without the conversion of joules to calories. Equation (34) suggests that it can be defined in two ways: similarly to Equation (34) or as kinetic energy per unit mass. Let θ be the granular temperature, the random kinetic energy per unit mass. Then

(38)θ=1/3C2=1/(3n)(Cx2+Cy2+Cz2)dCxdCydCz

in three dimensions. In two dimensions, we would have only two random velocities, and we would divide <C2> by 2. In one dimension we have only one random velocity, and the granular temperature is then simply the variance of the measured instantaneous velocities. However, its behavior is not the same as the three-dimensional granular temperature in Equation (38) (Strumendo et al. 2005). The units of the granular temperature are (m/s)2.

These units are convenient. The alternate definition with mass multiplying θ is not that convenient for a single particle size mixture. It may, however, be useful for a mixture of particles (Gidaspow 1994). For a gaseous mixture of molecules, where there is no dissipation of energy, mass times velocity square of molecules is the same, since there is only one temperature.

The equation of state for particles is obtained by eliminating <C2> between Equations (36) and (38). This gives

(39)ps=nmθ

where the subscript s was added to emphasize that it is the solids pressure. But nm is the bulk density. In terms of the volume fraction of solids, εs and the solids density, ρs, the ideal equation of state for particles becomes as follows:

(40)ps=εsρsθ

The more complete equation of state for particles containing the collisional contribution has been verified experimentally by Gidaspow and Huilin (1998). Its comparison to theory is discussed in the experimental chapter.

3.4 FCC equation of state

The experimental granular temperature for flow of FCC particles is presented in Figure 3. As the particle concentration increases, the granular temperature (turbulent kinetic energy of particles) increases, similar to the rise of thermal temperature upon compression of a gas. The decrease of granular temperature in the collisional regime is due to the decrease of the mean free path, which becomes 0 in the packed state. The Sandia National Laboratory (Bhusarapu et al. 2006) measured the granular temperatures in a large riser with a splash plate using a radioactive particle tracer technique for flow of 150 μm glass beads at velocities of 4.5–7.7 m/s and fluxes of 102–145 kg/m2 s. Their granular temperatures were of the order of magnitude of those shown in Figure 3.

Figure 3: Granular temperature for 75 μm FCC in CFB. Granular temperature=2/3 σx2+1/3 σy2 =2/3 turbulent kinetic energy. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.
Figure 3:

Granular temperature for 75 μm FCC in CFB. Granular temperature=2/3 σx2+1/3 σy2 =2/3 turbulent kinetic energy. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.

The particulate pressure measured with a specially designed transducer (Gidaspow and Huilin 1998) is shown in Figure 4. To construct a complete equation of state for FCC particles, we have used the statistical mechanics of liquid theory and our CCD camera system to determine radial distribution functions as a function of solids volume fraction.

Figure 4: Time-average solid pressure for FCC particles. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.
Figure 4:

Time-average solid pressure for FCC particles. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.

Figure 5 presents the equation of state. It shows that the kinetic theory of granular flow is valid for flow of FCC particles below about 5%. For dilute flow, we have an analogy of the ideal gas law: The ratio of solids pressure to bulk density multiplied by the granular temperature is 1, as a limit. For volume fractions above about 5%, the standard granular flow theory had to be corrected for a cohesive pressure, obtained from measurements of the radial distribution functions and granular temperature, using a modified Boltzmann relation, as explained by Gidaspow and Huilin (1998).

Figure 5: Equation of state for 75 μm FCC particles determined in IIT CFB. Dimensionless solid pressure: Psρsεsθ=1+2(1+e)εsgo−(0.73εs+8.957εs2).${{{P_s}} \over {{\rho _s}{\varepsilon _s}\theta }} = 1 + 2(1 + e){\varepsilon _s}{g_o} - (0.73{\varepsilon _s} + 8.957{\varepsilon _s}^2).$ Reprinted from Gidaspow and Huilin (1998) with permission from ©Wiley.
Figure 5:

Equation of state for 75 μm FCC particles determined in IIT CFB. Dimensionless solid pressure: Psρsεsθ=1+2(1+e)εsgo(0.73εs+8.957εs2). Reprinted from Gidaspow and Huilin (1998) with permission from ©Wiley.

3.5 Particle and molecular velocities

To estimate the average molecular velocity, the concept of thermal temperature (Equation 34) is used. Thus, Equation (34) is multiplied by the Avogadro’s number, A, resulting in the following expression:

(41)AkBT=1/3mA<C2>

where mA (kg/molecule×molecules/mole) represents the molecular mass (M), and kBA is the gas law constant (R).

Based on Equation (41), the equation for molecular velocity (C2) can be obtained as a function of thermal temperature:

(42)<C2>=3RT/M

where R=8314 J/mol K; 3R=158

The molecular velocity can be expressed in terms of molecular mass and temperature as follows:

(43)(<C2>)1/2=158(T/M)1/2

For gases as CO2 and H2 at 273 K, the velocity obtained by Equation (43) gives 393 m/s and 1845 m/s. The critical or the sonic velocity is evaluated at a constant entropy and hence is about 20% higher due to the ratio of specific heats at constant temperature to constant volume in the square root relation in Equation (43).

For particles, the average velocity, the hydrodynamic velocity, cannot be 0, since an energy input is required to keep the particles in motion due their inelasticity. So Equation (38) shows that the granular temperature is of the order of the hydrodynamic velocity square. Thus, for fluidization of small particles in a bubbling bed the velocity is of the order of cm/s, while for fluidization in risers it is of the order of m/s. Due to dissipation of energy, the granular temperatures are not inversely proportional to the square root of masses, as shown for gases in Equation (43), although they are smaller for large particles in a binary mixture of small and large particles (Gidaspow et al. 2004).

3.6 Maxwellian distribution

The Maxwellian distribution using the granular temperature can be shown to be as follows (Gidaspow and Jiradilok 2009):

(44)f(r,c)=n(2πθ)3/2exp((cv)22θ)

This is the expression found in the kinetic theory of gases (Chapman and Cowling 1961) with kB/m=1. By comparing Equation (44) with that originally reported by Chapman and Cowling (1961), it can be noted that the thermal temperature was replaced by granular temperature, which is expressed in the units of velocity square.

3.7 Restitution coefficients

Figure 6 shows typical restitution coefficients based on data reported in the book by Johnson (1985). During impact of particles, the work of deformation can be expressed in terms of the elastic pressure Pe and the plastic pressure Pp and the deformation volume by means of the usual relation, as

Figure 6: Restitution coefficients for various materials.
Figure 6:

Restitution coefficients for various materials.

(45)Work of deformation=0V(Pe+Pp)dV

This work equals the relative velocity of impact square times half the mass. However, the relative kinetic energy equals only the integral of the elastic pressure. Hence, the restitution coefficient e can be expressed as

(46)e2=v2v2=0VPedV0V(Pe+Pp)dV

where v is the relative velocity before impact and v′ is the rebound velocity. Equation (46) suggests that for low velocities, where the plastic deformation is small, the restitution coefficient will be nearly 1. It also clearly shows that the restitution coefficient is a function of the material properties, as well as the dynamic properties associated with plastic flow. Indeed, data summarized by Johnson (1985) show that for hard materials the restitution coefficients are nearly 1 for impact velocities of 0.1 m/s and less (see Figure 6).

3.8 Frequency of binary collisions

Based on Chapter 16 in Chapman and Cowling (1961), the classical binary frequency of collisions corrected for the dense packing effect (g0) is obtained. A collisional pair distribution function, f(2), is introduced, which is analogous to the single frequency distribution given by Equation (26):

(47)f(2)=f(2)(c1,r1,c2,r2)

The probability of finding a pair of particles in the volume dr1dr2 centered on points r1, r2 and having velocities within the ranges of c1 and c1+dc1, and c2 and c2+dc2 are defined as follows:

(48)f(2)dc1dc2dr1dr2

Figure 7 shows the geometry of collisions and the spherical coordinates and solid angle. It is that given by Savage and Jeffrey (1981) with a generalization to two rigid spheres of unequal diameters. Equation (48) is integrated over the volume of collision cylinder as fully described in Gidaspow (1994) to obtain the collision frequency:

Figure 7: Geometry of a collision of two spheres of diameter, d1 and d2.
Figure 7:

Geometry of a collision of two spheres of diameter, d1 and d2.

(49)N12=4n1n2d122g0πθ

For dilute gases (m0/m1=2 and kB/m2=1), Equation (49) is similar to that obtained by classical kinetic theory of gases proposed by Chapman and Cowling’s (1961), which can be written as follows:

(50)N12=2n1n2d122(2πkBTm0m1m2)1/2

where

(51)m0/m1=2

and

(52)kB/m2=1

Analyzing Equations (49) and (50), the difference between the granular and the thermal temperature can be demonstrated. In Equation (50), the Boltzmann constant kB serves to assign the scale of the temperature, whereas in Equation (49) g0 becomes infinite at maximum packing. However, at this point, the granular temperature is expected to approach 0, leaving N12 undefined.

3.9 Mean free path

In the kinetic theory of gases described by Chapman and Cowling (1961), molecules are considered as rigid spheres and exert no intermolecular forces. In addition, the collisions between the molecules are assumed to be completely elastic. Based on this theory, the concepts of transport coefficients such as viscosity and diffusion coefficient can be obtained.

The mean time between successive collisions, called the collision time (τ), is obtained as shown below for dilute flow:

(53)τ=n1/N11

where g0=1

Hence,

(54)τ=1/(4n1d122πθ)

The mean free path of particles can be understood as the distance between successive collisions and can be expressed as the product of the average velocity and the collision time:

(55)l=cτ

where

(56)c=(8θπ)1/2

Substituting Equation (54) into Equation (55), the mean free path becomes independent of θ:

(57)l=1π2n1dp2=0.707πndp2g0

Using the relation

(58)εs=16πdp3n

the mean free equation can be defined in terms of diameter and bed porosity as

(59)l=162dpεs

3.10 Elementary treatment of transport coefficients

3.10.1 Diffusion coefficients

This derivation is similar to that used in Prandtl’s mixing length theory of turbulence. In this theory we assume some quantity Q is preserved between two points a distance l apart. Then the change in Q is represented by ΔQ. Next we expand Q in a Taylor series and drop terms higher than the first. For diffusion, Q equals the mass flux of species A which is equal to the density of species A times an average velocity of species A. Hence, for diffusion of species A the change in Q becomes as follows:

(60)ΔQ=lvAdρAdx

Fick’s law of diffusion in mass units is

(61)ρA(vAv)=DdρAdx

where D is diffusion coefficient and v the mass average velocity. In this case, v is equal to 0.

Therefore, the following definition of diffusion coefficient can be obtained by comparing Equations (60) and (61):

(62)D=lvA

Substituting Equations (56) and (59) into (62), the diffusion coefficient can be expressed in terms of granular temperature, particle diameter and porosity as follows:

(63)D=13πθdpεs

The average of oscillating velocity is related to the granular temperature

(64)vA=c=(8θπ)1/2

For molecules, the fluctuating velocity can be estimated from

(65)c212=3RTM

Hence,

(66)D=l3RTM

For example, at 273 K, the diameter of a CO2 molecule is 3×10−10 m, and the mean free path is 2×10−7 m with the fluctuating velocity of 394 m/s. Based on these data, the diffusion coefficient can be estimated using Equation (66), resulting in D=7.88×10−5 m2/s. Comparing the estimating of D with the experimental value (1.4×10−5 m2/s), it can be noted that they are in the same order of magnitude for D. The diffusion coefficient of gas into liquids is 3 orders of magnitude smaller. For particles, the diffusivity can be calculated by direct measurement of particle velocity and computation using the autocorrelation method (Jiradilok et al. 2007).

In a multiphase CFD approach, dispersion coefficients are not an input into the codes, as they are in the convection dispersion model. But we have computed them for fluidization in agreement with data (Kashyap and Gidaspow 2012a,b).

3.10.2 Viscosity

To obtain the viscosity of the particulate phase, let the momentum flux be

(67)Q=ρvv(Momentum/Volume×Average of oscillating velocity)

Then for a constant density ρ, the change in momentum flux is

(68)ΔQ=lρvdvdx

The viscosity for fully developed incompressible flow is defined by

(69)Shear=μdvdx

The momentum transport ΔQ equals the force per unit area

(70)ΔQ=Shear

Therefore, the viscosity assumes the form

(71)μ=lρv

From Equation (62) it can be seen that

(72)μ=ρD

as in the kinetic theory of gases. Therefore, using Equation (63), a simple formula for the collisional viscosity is

(73)μ=(13π)ρpdpθ

Kinematic viscosity=Mean free path×Fluctuating velocity

For smooth rigid spherical molecules of diameter dp (Chapman and Cowling 1961)

(74)μl=516dp2(kBmTπ)1/2

where m=ρpπ6dp3

To convert T to granular temperature θ, let kBm=1. Thus,

(75)DILUTE  μs=5π96ρpdpθ1/2

Multiplying and dividing Equation (75) by εs and using the definition of bulk density (ρp.εs), the following physical interpretation of viscosity can be obtained:

(76)μs=5π96(ρpεs)(dpεs)θ1/2

Viscosity=Constant×Bulk density×Mean free path×Oscillation velocity

For critical flow experiments (Gidaspow 1994), it can be found that θ≅1 m/s. Considering critical flow, for 100 μm particles with 1500 kg/m3 specific density, the viscosity estimated by Equation (76) results in 1.384×10−3 kg/m s≅15 cP (15 centipoises). Such a value is 15 times that of water at 20°C temperature. It is close to the measurements (Gidaspow and Huilin 1996). The collisional viscosity includes the effect of particle concentration.

Figure 8 shows the measured viscosity in the Illinois Institute of Technology (IIT) riser for transport of FCC particles. In the figure the dilute viscosity is that given by Equation (75). This dilute viscosity expression as well as the kinetic and the collisional contributions are those derived in Chapman and Cowling’s (1961) book for molecules. The dilute expression for viscosity agrees with that derived here using the elementary kinetic theory, except for the constant, as is evident by comparing Equations (73) and (75). The derivation of the collisional and the kinetic terms are given in Gidaspow’s (1994) book. In Figure 8 the circles are the measurements obtained using the CCD camera method, where the viscosity was obtained by substituting the measured granular temperature and the solids volume fractions into the equations for viscosity obtained from kinetic theory. Figure 8 shows that these viscosities agree with the other measurements obtained earlier by Miller and Gidaspow (1992) where the triangles are for data obtained using the shear pressure drop method, and the squares are for the data using Brookfield viscometer (Miller and Gidaspow 1992).

Figure 8: FCC powder viscosity measured in three different ways.
Figure 8:

FCC powder viscosity measured in three different ways.

3.10.3 Thermal conductivity

(77)Here,  ΔQ=ldQdx
(78)Heat flux:  Q=ρUv

(Density×Internal energy×Average of fluctuating velocity)

(79)ΔQ=lvρdUdx

But specific heat cv is defined as

(80)cv=(UT)v

By chain rule

(81)ΔQ=lvρcvdTdx

Fourier’s law of heat conduction

(82)q=kdTdx

Since qQ, Equations (81) and (82) show that

(83)k=lvρcv

Conductivity=Mean free path×Oscillation velocity×density×Specific heat

If we use enthalpy in place of U, we get then

(84)k=μcp

The thermal diffusivity is

(85)α=(kρcp)=ρv

There is a rough agreement between this simplified theory and measurements, as demonstrated by Gidaspow and Huilin (1996) in the IIT riser with a heated section.

4 Multiphase flow experimental verification

4.1 Experimental

To test the validity and the accuracy of the kinetic theory model, a two-story riser was built at IIT with a splash plate on top of the riser to obtain symmetry (Gidaspow and Jiradilok 2009). Figure 9 shows the IIT CFB with splash plate and measurement equipment. The γ-ray source was used to measure the particle concentration. The particle velocities were measured using kinetic theory based particle image velocity (PIV).

Figure 9: IIT circulating fluidized bed with splash plate.
Figure 9:

IIT circulating fluidized bed with splash plate.

4.2 Kinetic theory based PIV

Figure 10 illustrates our improved PIV method of obtaining instantaneous velocities for a binary mixture of glass beads. This technique was recently described fully for flow of 530 μm glass beads by Tartan and Gidaspow (2004). In the CFB regime, a curtain of solids at the wall of the pipe restricts the use of laser-Doppler velocity meters for obtaining velocities in the core of the riser, vertical pipe. Hence, a probe, shown in Figure 10, was used. Figure 11 shows the typical steak images captured by CCD camera. The velocity is the length of the streak divided by exposure time. The order of the colors on the rotating transparency establishes the direction. The study of Tartan and Gidaspow (2004) was generalized to a mixture of two sizes of particles. Large particles form thicker streaks than the small particles. To obtain radial profiles, a probe was inserted into the riser. The size of the probe was varied to establish an optimum balance between its hydrodynamic interference and sufficient number of streaks, 10–30, in a picture to obtain meaningful statistics of velocity averages and their variances. Figure 12 shows typical instantaneous axial and radial velocities measured by particle image velocity meter in the riser.

Figure 10: Particle image velocity measurements system with probe. Reprinted from Tartan and Gidaspow (2004) with permission from ©Wiley.
Figure 10:

Particle image velocity measurements system with probe. Reprinted from Tartan and Gidaspow (2004) with permission from ©Wiley.

Figure 11: Schematically typical streak images captured by CCD camera.
Figure 11:

Schematically typical streak images captured by CCD camera.

Figure 12: Typical axial (A) and radial (B) particle velocities. Reprinted from Tartan and Gidaspow (2004) with permission from ©Wiley.
Figure 12:

Typical axial (A) and radial (B) particle velocities. Reprinted from Tartan and Gidaspow (2004) with permission from ©Wiley.

The hydrodynamic velocity v was calculated from measurement of the instantaneous velocity c as follows:

(86)vi(r,t)=1nk=1ncik(r,t)

The kinetic stresses are calculated as follows:

(87)<CiCi>(r,t)=1nk=1n(cik(r,t)vi(r,t))(cik(r,t)vi(r,t))

where n is the total number of streaks in each frame and Ci=civi. We see that the particle stresses, CzCz¯, in the direction of flow are much larger than the tangential and the radial stresses, similar to turbulent flow of gases in a pipe (Schlichting 1960, Kim et al. 1987), but are an order of magnitude larger. The orders of magnitude of larger particle weight for fluidization are at expense of an order of magnitude larger pressure drop.

The particle Reynolds stresses are calculated from hydrodynamic velocity v as follows:

(88)vivi¯(r)=1mk=1m(vik(r,t)v¯i(r))(vik(r,t)v¯i(r))

where vi was calculated from Equation (86)

The laminar granular temperature is the average of normal stresses shown below.

(89)θ(r,t)=13<CzCz>+13<CrCr>+13<CθCθ>

4.3 Core-annular flow regime explanation

Recently, Benyahia et al. (2007) studied the ability of multiphase continuum models to predict the core-annulus flow. Figure 13 shows that the time-averaged particle velocity was parabolic and that the particle concentration was uniform in the center of the 7.6 cm tube and high at the wall. Such a concentration distribution is known as the core-annular flow.

Figure 13: Developed time average particulate phase axial velocity for 530 μm glass beads in the IIT two-story riser (Ug=4.9 m/s; Ws=14.2 kg/m2·s). Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.
Figure 13:

Developed time average particulate phase axial velocity for 530 μm glass beads in the IIT two-story riser (Ug=4.9 m/s; Ws=14.2 kg/m2·s). Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.

The granular temperate energy balance is similar to the energy balance (22) discussed earlier and is shown below:

(90)32[(ρsεsθ)t+(ρsεsθvs)]=(PsI+τs):vs+(ksθ)γ

For steady state and fully developed flow, the production of granular temperature equals the conduction of granular energy and inelastic dissipation:

(91)0=(PsI+τs):vs+(κsθ)γ

The production of granular temperature due to oscillation of particles reduces itself to μ(vr)2. In transport phenomena texts, this term corresponds to production of heat due to viscous dissipation. In a suspension the particle collisions can be assumed to be elastic. During collisions, the fluid between the particles has to be pushed out, requiring a large force. Therefore, the dissipation of granular energy (γ) is 0.

In cylindrical coordinates, the balance of granular energy then becomes

(92)κddr(rdθdr)=μ(vr)2

The measured velocity distribution in our riser as shown in Figure 13 is

(93)v=2v¯[1(rR)2]

Similar parabolic distributions have been reported by Berruti et al. (1995).

Substituting Equation (93) into Equation (92), the balance of granular temperature can be expressed by

(94)κddr(rdθdr)=16μv¯2r2R4

Integrating Equation (94) with zero wall granular temperature, the solution of the above equation is as follows:

(95)θ=v¯2(μκ)[1(rR)4]

As reported by Gidaspow (1994), Figure 14 shows that measured granular temperatures can be approximated by the fourth order parabolic equation obtained by solving the granular temperature derived above.

Figure 14: A comparison of measured granular temperature in the IIT two-story riser to the analytical solution. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.
Figure 14:

A comparison of measured granular temperature in the IIT two-story riser to the analytical solution. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.

For dilute flow, the granular pressure shown in Figure 5 can be approximated by the kinetic term only:

(96)ps=εsρsθ

This assumption corresponds to the use of ideal gas law for molecules, where the particles are far apart from each other.

In developed flow, the radial variation of this pressure is approximately 0. This approximation allows us to obtain the very simple expression for the particle volume fraction distribution shown below:

(97)εs=Psκρsμsv¯21[1(rR)4]

where μs is the particle viscosity, κ the granular conductivity and v̅2 is the average particle velocity.

The above Equation (97) explains the core-annular particle distribution in Figure 13. This expression is not valid at the tube wall. The kinetic theory shows that the core annular regime is independent of the pipe radius, not believed to be so by the oil industry, until they finally made the measurements (Mohammad-Reza 2002).

With a few simplifications (Gidaspow and Chandra 2014), this equation can give the ratio of the number of particles per unit volume (n) to its inlet into the system.

(98)nninlet=1[1(rR)4]

where n=6εsπdp3;

These simplifications are as follows: (i) from kinetic theory, for dilute flow μsκs415 and (ii) the inlet granular temperature can be approximated by θinlet=μsκsv¯2 since μsκs1.

Equation (98) describes the platelets distribution for blood flow in a fully developed regime (Gidaspow and Chandra 2014).

4.4 Turbulent granular temperature

There are two granular temperatures. A laminar granular temperature that is computed using the granular temperature equation in the CFD codes, such as FLUENT (Syamlal et al. 1993), MFIX (Syamlal et al. 1993) or the IIT code (Gidaspow and Jiradilok 2009) shown in Table 2, and a turbulent granular temperature computed from the normal Reynolds stresses per unit bulk density. For dilute riser flow, the granular kinetic theory agrees well with CFB experiments, as shown by Tartan and Gidaspow (2004).

Table 2:

A comparison of laminar and turbulent granular temperatures (m2/s2) in dilute and dense regimes of risers for flow 54 μm FCC (vg=3.5 m/s) and 1094 μm ceramic alumina particles (vg=19 m/s).

SectionHeight (m)Solids volume fraction (−)Laminar granular temperature (m2/s2)Turbulent granular temperature (m2/s2)
Geldart D particles
 Bottom20.320.11.2
 Top60.021.11.25
Geldart A particles
 Bottom20.2020.0010.558
 Top60.0480.1421.675

Unfortunately, for the commercially useful dense flow, the turbulent granular temperatures exceed the laminar granular temperatures, as shown in Table 2. In Table 2 the values of solids volume fractions and granular temperatures for flow of Geldart D particles are from the paper of Kashyap et al. (2011) done for UOP for high solids flux in the IIT riser. The flow was in a solids slugging regime. The computations were done using the standard drag model, as described in Gidaspow’s 1994 book. The values for Geldart A particles are for the riser of Wei et al. (1998), with computations done using a correction for the drag, derived using the energy minimization principle invented by Li and Kwauk (1994).

Figure 15 shows the dimensionless turbulent granular temperature in the dense bubbling bed, called “bubble-like” in Jung et al. (2005). It is represented by solid circles and is almost an order of magnitude larger than the laminar or particle granular temperature. For the dilute risers, the turbulent granular temperatures for both 156 and 530 μm particles were smaller than the laminar granular temperatures which agree with the theoretical analytical solution for the granular temperature equation for elastic particles shown in Figure 14.

Figure 15: Measured dimensionless laminar and turbulent granular temperatures in the IIT two-story riser in dilute flow and in rectangular bubbling beds. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.
Figure 15:

Measured dimensionless laminar and turbulent granular temperatures in the IIT two-story riser in dilute flow and in rectangular bubbling beds. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.

4.5 Flow regime computation

An excellent review of flow regimes before wide use of CFD was given by Berruti et al. (1995). One-dimensional two-phase models of three decades ago required a specification of measured flow regimes (Lyczkowski et al. 1982). The two- and three-dimensional models used today can successfully compute these flow regimes. Tables 35 summarize our kinetic theory model.

Table 3:

Kinetic theory CFD model.

Continuity equations with no phase change
 Gas phase
t(εgρg)+(εgρgvg)=0
 Solid phases (k=1,2)
t(εkρk)+(εkρkvk)=0
Momentum equations without bulk viscosity
 Gas momentum
t(εgρgvg)+(εgρgvgvg)=P+k=1Nβgk(vkvg)+2εgμgsvg+εgρgg
 Particulate phases, k=(1, …, N)
t(εkρkvk)+(εkρkvkvk)=βgk(vgvk)+l=1Nβkl(vlvk)Ps+2εkμksvk+εkρkg
Energy equations
 Gas phase
t(εgρgHg)+(εgρgHgvg)=(Pt+vgP)+k=1Nhvk(TkTg)+(KgεgTg)
 Solid phases
t(εkρkHk)+(εkρkHkvk)=hvk(TfTk)+(KkεkTk)
Granular temperature
32[(ρsεsθ)t+(ρsεsθvs)]=(PsI+τs):vs+(ksθ)γ
Table 4:

Constitutive equations.

Particle pressure
Ps=εsρsθ+2ρs(1+e)εs2g0θ
Shear particle viscosities
μs=45εsρsdpg0(1+e)(θπ)1/2+10ρsdpθπ96εs(1+e)g0[1+45g0εs(1+e)]2
Bulk particleviscosity
ζs=43εsρsdpg0(1+e)(θπ)1/2
Radial distribution function
g0=[1(εsεs,max)1/3]1
Granular conductivity
ks=150ρsdpθπ384(1+e)g0[1+65εsg0(1+e)]2+2ρsεs2dp(1+e)g0θπ
Particle stress
τs=εsμs(vs+vsT)+εs(ξs23μs)vsI
Dissipation
γ=12(1e2)g0dpπρsεs2θ3/2
Table 5:

Drag and heat transfer coefficients.

Gas-solid drag coefficients k=(1, …, N)
β=150εs2μgεg2dp2+1.75ρgεsεgdp|vgvs| when ε<0.74
β=34Cdρgεs|vgvs|dpω(ε) when ε≥0.74
 with Cd=24Re(1+0.15Re0.687) Re<1000
Cd =0.44 Re≥1000
 where Re is the Reynolds number, Re=εgρgdp|vgvs|μg
For fast fluidization
ω(ε)={0.5760+0.02144(ε0.7463)2+0.0044   (0.74<ε0.82)0.0101+0.00384(ε0.7789)2+0.0040   (0.82<ε0.97)31.8295+32.8295ε   (ε>0.97)
Particle-particle drag coefficients
βklk,lf=32α(1+e)ρkρlεkεl(dk+dl)2ρkdk3+ρldl3|vkvl|
Gas-particle heat transfer, k (=s)
 for εf<0.8
Nuk=(2+1.1Re0.6Pr1/3)Sk=0.123(4Redk)0.183Sk0.17=0.61Re0.67Sk
 for εf>0.8
Nuk=(2+0.16Re0.67)Sk=8.2Re0.6Sk=1.06Re0.457Sk
 where
Re=ρf|vfvk|dkμf
Sk=εk6dk;Nu=hvkdkkf

Figure 16 shows batch fluidization flow regimes for nanoparticles and Geldart A (aerated), B (bubbling) and C (cohesive) particles. The batch fluidization flow of 10 nm Tullanox nanoparticles occurs without bubbles due to the formation of clusters (Jung and Gidaspow 2002). Geldart C particles are small cohesive particles, which fluidize with the formation of small bubbles. Group A particles form aerated beds with small bubbles, whereas Group B particles fluidize with the formation of large bubbles. Details can be found in Gelderbloom et al. (2003). As the gas velociy is increased, the particles are blown out of the bed.

Figure 16: A summary of basic batch fluidization flow regimes computed at Illinois Institute of Technology. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.
Figure 16:

A summary of basic batch fluidization flow regimes computed at Illinois Institute of Technology. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.

To obtain continuos particle flow, the particles are fed into the bed through one or more jets. Figure 17 summarizes our computed fluidization flow regimes. Turbulent fluidization is the name given to the flow regime in which there exists a dense phase at the bottom of the bed and a dilute particle phase on the top of the bed. The volume fractions of the particles in the bottom and top sections of the bed were estimated by Matsen (2000) using the drift flux model, knowing the gas and particles flow rates. At IIT we had seen a sharp interphase between the dilute and dense portions of the bed. Our code describes the axial experimental measurements of solid volume fractions of Wei et al. (1998). In the fast fluidization regime, large clusters are formed, which descend near the wall. Dense suspension flow is formed at high gas velocities and high solid fluxes. In this regime there is a core-annular flow. For large particles (dp=1093 μm) and very high velocities, slugging fluidization was observed and computed. In a reactor, high velocities are needed to obtain high production rates. But if the velocity is too high, we will be in the pneumatic transport flow regime with a low catalyst concentration.

Figure 17: A summary of continuous basic fluidization flow regimes computed at Illinois Institute of Technology. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.
Figure 17:

A summary of continuous basic fluidization flow regimes computed at Illinois Institute of Technology. Reprinted from Gidaspow et al. (2004) with permission from ©Elsevier.

4.6 Heat transfer

To design better fluidized bed reactors for new processes, such as for CO2 capture from flue gases or for the production of pure silicon for solar collectors, wall to bed heat transfer must be improved. The values of the wall to bed heat transfer coefficients limit the size of the fluidized bed reactors. Since the temperatures inside such reactors are nearly constant, the wall to bed heat transfer coefficients are simply the thermal conductivity divided by the small boundary layer thickness. To obtain a high thermal conductivity, the turbulent granular temperature must be as high as possible. Since the turbulent heat flux=ρkcpkεkTv¯, we must also have the solids volume fraction as high as possible. Our kinetic theory based CFD code (Gidaspow and Jiradilok 2009, Chaiwang et al. 2014) computes the thermal temperature and its standard deviations. The standard deviation is the highest near the walls, in the thermal boundary layer, in agreement with measurements done in the IIT CFB. The average standard deviations of the temperature and velocities are 0.75° and 0.15 m/s, respectively. Then the heat transfer coefficient obtained by this method and from an overall energy balance is 0.39kwm2K. If we increase the standard deviation of velocity from 0.15 m/s to 1 m/s, the heat transfer coefficient becomes 2.6, still well below boiling heat transfer coefficients but much larger than the usual CFB heat transfer coefficients. This example demonstrates the potential of CFD to improve fluidized bed reactors.

4.7 Mass transfer

For half a century it was known that the mass transfer coefficients in fluidized beds when expressed in conventional dimensional way as the Sherwood number, where Sherwood number=mass transfer coefficient×particle diameter/gas diffusivity, are orders of magnitude below the theoretical value of 2 for diffusion to a sphere in stagnant fluid. Typical 1970 data obtained in C. Y. Wen’s laboratory in West Virginia (Kato et al. 1970) and our CFD calculation is shown below in Figure 18.

Figure 18: Sherwood number as a function of riser height.
Figure 18:

Sherwood number as a function of riser height.

In Figure 18 the behavior of the Sherwood number is similar to that of the diffusion in a channel in fully developed laminar flow, called the Graetz problem, except for the much lower Sherwood number. There the Sherwood number is high in the diffusion boundary layer and reaches a constant value of about 2 for large heights. For the fluidized bed riser in Figure 18, our explanation (Chalermsinsuwan et al. 2008a,b, Kashyap and Gidaspow 2012a,b) is the formation of clusters, as shown for the fast fluidization regime in Figure 17.

If we define the Sherwood number based on the cluster size rather than particle diameter, the low Sherwood number in Figure 18 becomes close to the theoretical value of two, since the ratio of cluster size to particle diameter can be as much as 1000. This theory suggests that we reduce the cluster size to eliminate potential mass transfer limitation which may require much taller reactors than those computed from reaction rate data.

5 Opportunities for multiphase cfd

Multiphase reactors are very common in the chemical and oil industries (Berruti et al. 1995). Multiphase CFD is a new tool that can be used to improve and design such reactors. For converting synthesis gas into liquids using slurry bubble column Fischer Tropsch reactors, Gidaspow et al. (2015) have shown how to eliminate the expensive heat exchangers tubes. In this concept the heat of reaction is removed by overflow of the liquid product.

In the nuclear industry, multiphase CFD has for many decades been used to license nuclear reactors (Lyczkowski et al. 1982). Gidaspow et al. (2013) have developed a preliminary CFD simulator for multiphase in reservoirs and pipes which was used to understand the British Petroleum oil spill in the Gulf of Mexico. To prevent future accidents, we believe that this problem requires better understanding of multiphase flow.

We believe that the multiphase CFD equations presented in this review (Tables 35) compute turbulence, without the need for an input of turbulence models, similarly to that of direct numerical simulation (DNS) of fully developed single-phase flow in channels done by NASA (Kim et al. 1987). At that time NASA was hoping that DNS will eliminate the need for tests in wind tunnels. Unfortunately, some multiphase CFD models in the literature contain untested turbulence models. The kinetic theory CFD model computes reasonably correct viscosities with reasonable input of restitution coefficients. Again, unfortunately some particle viscosities computed in the literature produced particle viscosities that differ by orders of magnitude from those measured. Hence, we believe that multiphase turbulence must be understood much better. Such an understanding will lead to improved wall to bed heat transfer coefficients, as illustrated in this review. It may also lead to our ability to accurately compute cluster sizes in fluidization. As reviewed in this study, large cluster sizes lead poor mass transfer in fluidization. This may result in excessively tall reactors. Perhaps these clusters may be made much smaller with production of high turbulence in the reactors.

About the authors

Dimitri Gidaspow

Dimitri Gidaspow was introduced to multiphase flow in 1972 by his first former PhD student, Dr. Charles W. Solbrig, who was directing the development of computer codes for licensing nuclear reactors by the Atomic Energy Commission in Idaho. Earlier at IIT, Gidaspow had taught a numerical methods course with Solbrig which was the beginning of the 642 page book, Computational Techniques, by D. Gidaspow and V. Jiradilok, Nova Science, 2009. In 1984 he received the AIChE Kern Award for his earlier work on energy conversion. His lecture on fluidization and heat transfer was published in Applied Mechanics Review. In 1994 he completed his highly cited book, Multiphase flow and fluidization. He received three other awards in fluidization from AIChE.

Marcelo S. Bacelos

Marcelo S. Bacelos received his PhD in chemical engineering in 2006 from Federal University of Sao Carlos, Brazil. He has been teaching transport phenomena in the Engineering and Technology Department, Federal University of Espirito Santo, Brazil. His research has been focused on computational multiphase fluid dynamics to describe fluidized and spouted bed reactors. In the last 5 years, he has used gas-solid fluidized reactors for recovering waste, such as post-consumer carton packages. He has used fluidized and spouted beds to verify his CFD designs. Starting in 2016, he has been working with Prof. Gidaspow at IIT using kinetic theory to describe the gas-solid flow in fluidized and spouted bed reactors.

Nomenclature

A

Avogadro’s number

C

peculiar velocity

c

instantaneous velocity

Cd

drag coefficient

D

diffusivity

dp

particle diameter

e

restitution coefficient

E

internal energy

f

frequency distribution of velocities of particles

g

radial distribution function

H

enthalpy

K

thermal conductivity

kB

the Boltzmann constant (1. 3805×10−23 J/K)

M

molecular weight

m

mass

N

number of moles

n

number of particles per unit volume

P

pressure

pk

kinetic stress tensor

p

stress tensor

q

conduction flux

Q

the mass flux of species A

R

gas constant

S

entropy

Sh

Sherwood number=mass transfer coefficient×dp/D

r

position

Ti

stress tensor

T

average of the random kinetic energy

t

time

V

volume

v

hydrodynamic velocity

v

turbulent velocity

Greek letters
β

friction coefficient

γ

dissipation of granular temperature

θ

granular temperature (the random kinetic energy per unit mass), (m/s)2

ϕ

mean values of a quantity

ρ

bulk density

εs

volume fraction of solids

ρs

solids density

μ

viscosity

Acknowledgments

We thank the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) for the financial support of this study.

References

Aldis DF, Gidaspow D. Two-dimensional analysis of a dust explosion. AIChE J 1990; 36: 1087–1096.10.1002/aic.690360715Search in Google Scholar

Arastoopour H. Numerical simulation and experimental analysis of gas/solid flow systems: 1999 Fluor-Daniel Plenary lecture. Powder Technol 2001; 119: 59–67.10.1016/S0032-5910(00)00417-4Search in Google Scholar

Aris DF. Vectors, tensors, and the basic equations of fluid mechanics. Englewood Cliffs, New Jersey, USA: Prentice-Hall, 1962.Search in Google Scholar

Bhusarapu S, Al-Dahhan MH, Duduković MP. Solids flow mapping in a gas-solid riser: mean holdup and velocity fields. Powder Technol. 2006; 163: 98–123.10.1016/j.powtec.2006.01.013Search in Google Scholar

Benyahia S, Syamlal M, O’Brein TJ. Study of ability of multiphase continuum models to predict core-annulus flow. AIChE J 2007; 53: 2549–2568.10.1002/aic.11276Search in Google Scholar

Berruti F, Chaouki J, Godfroy L, Pugsley TS, Patience GS. Hydrodynamics of circulating fluidized bed risers: a review. Canadian J Chem Eng 1995; 73: 579–602.10.1002/cjce.5450730502Search in Google Scholar

Chaiwang P, Gidaspow D, Chalermsinsuwan B, Piumsomboon P. CFD design of a sorber for CO2 capture with 75 and 375 micron particles. Chem Eng Sci 2014; 105: 32–45.10.1016/j.ces.2013.09.024Search in Google Scholar

Chalermsinsuwan B, Piumsomboon P, Gidaspow D. Kinetic theory based computation of PSRI riser – Part 1. Estimate of mass transfer coefficient. Chem Eng Sci 2008a; 64: 1195–1211.10.1016/j.ces.2008.11.010Search in Google Scholar

Chalermsinsuwan B, Piumsomboon P, Gidaspow D. Kinetic theory based computation of PSRI riser –Part II. Computation of mass transfer coefficient with chemical reaction. Chem Eng Sci 2008b; 64: 1212–1222.10.1016/j.ces.2008.11.006Search in Google Scholar

Chapman S, Cowling TG. The mathematical theory of non-uniform gases. Cambridge, UK: Cambridge University Press, 1961.Search in Google Scholar

Chen YM, Williams C. FCC technology – recent advances and new. In: Cen K, editor. Circulating fluidized bed technology VIII. China: International Academic Pub., 2005: 26–40.Search in Google Scholar

Gelderbloom S, Gidaspow D, Lyczkowski RW. CFD simulation of bubbling and collapsing fluidized bed experiments for three Geldart groups. AIChE J 2003; 49: 844–858.10.1002/aic.690490405Search in Google Scholar

Gidaspow D. Multiphase flow and fluidization, continuum and kinetic theory description. San Diego, California, USA: Academic Press, 1994.Search in Google Scholar

Gidaspow D, Arastoopour H. CFD modeling of CFB from kinetic theory to turbulence & heat transfer. In: Li J, Wei F, editors. International Conference on Fluidized Bed Technology. Beijing, China: Chemical Industry Press, 2014.Search in Google Scholar

Gidaspow D, Chandra V. Unequal granular temperature model for motion of platelets to the wall and red blood cells to the center. Chem Eng Sci 2014; 117: 107–113.10.1016/j.ces.2014.06.025Search in Google Scholar

Gidaspow D, Huilin L. Collisional viscosity of FCC particles in a CFB. AIChE J 1996; 42: 2503–2510.10.1002/aic.690420910Search in Google Scholar

Gidaspow D, Huilin L. Equation of state and radial distribution functions of FCC particles in a CFB. AIChE J 1998; 44: 279–291.10.1002/aic.690440207Search in Google Scholar

Gidaspow D, Jiradilok V. Computational techniques. The multiphase CFD approach to fluidization & green energy technologies. New York, USA: Nova Science Publishers, 2009.Search in Google Scholar

Gidaspow D, Jung J, Singh RJ. Hydrodynamics of fluidization using kinetic theory: an emerging paradigm 2002 Flour-Daniel Lecture. Powder Technol 2004; 148: 123–141.10.1016/j.powtec.2004.09.025Search in Google Scholar

Gidaspow D, Li F, Huang J. A CFD simulator for multiphase flow in reservoirs and pipes. Powder Technol 2013; 242: 2–12.10.1016/j.powtec.2013.01.047Search in Google Scholar

Gidaspow D, He Y, Chandra V. A new slurry bubble column reactor for diesel fuel. Chem Eng Sci 2015; 134: 784–799.10.1016/j.ces.2015.05.054Search in Google Scholar

Huang L, Juhui C, Guodong L, Huilin L, Dan L, Feixiang Z. Simulated second-order moments of clusters and dispersed particles in riser. Chem Eng Sci 2013; 101: 800–812.10.1016/j.ces.2013.06.036Search in Google Scholar

Jiradilok V, Gidaspow D, Breault RW. Computation of gas and solids dispersion coefficients in turbulent risers and bubbling beds. Chem Eng Sci 2007; 62: 3397–3409.10.1016/j.ces.2007.01.084Search in Google Scholar

Johnson KL. Contact mechanics. Cambridge, UK: Cambridge University Press, 1985.10.1017/CBO9781139171731Search in Google Scholar

Juhui C, Shuyan W, Dan S, Huilin L, Gidaspow D, Hongbin Y. A second-order moment method applied to gas-solid risers. AIChE J 2012; 58: 3653–3675.10.1002/aic.13754Search in Google Scholar

Jung J, Gidaspow D. Fluidization of nano-size particles. J Nanopart Res 2002; 4: 483–497.10.1023/A:1022807918247Search in Google Scholar

Jung J, Gidaspow D, Gamwo IK. Measurement of two kinds of granular temperatures, stresses and dispersion in bubbling beds. Ind Eng Chem Res 2005; 44: 1329–1341.10.1021/ie0496838Search in Google Scholar

Kashyap M, Gidaspow D. Dispersion and mass transfer coefficients in fluidized beds. Saarbruchen, Germany: Lap Lambert Academic Publishing, 2012a.Search in Google Scholar

Kashyap M, Gidaspow D. Measurements and computation of low mass transfer coefficients for FCC particles with ozone decomposition reaction. AIChE J 2012b; 58: 707–729.10.1002/aic.12615Search in Google Scholar

Kashyap M, Gidaspow D, Koves WJ. Circulation of Geldart D type particles: Part I – High flux solids measurements and computation under solids slugging conditions. Chem Eng Sci 2011; 66: 183–206.10.1016/j.ces.2010.10.012Search in Google Scholar

Kato K, Kubota H, Wen CY. Mass transfer in fixed and fluidized beds. Chem Eng Prog Symp Series 1970; 105: 87–99.Search in Google Scholar

Khongprom P, Gidaspow D. Compact fluidized bed sorber for CO2 capture. Particuology 2010; 8: 531–535.10.1016/j.partic.2010.09.001Search in Google Scholar

Kim J, Moin P, Moser R. Turbulence statistics in fully developed channel flow at low Reynolds number. J Fluid Mech 1987; 177: 133–166.10.1017/S0022112087000892Search in Google Scholar

Li J, Kwauk M. Particle-fluid two-phase flow. Beijing, China: Metallurgical Industry Press, 1994.Search in Google Scholar

Lyczkowski RW, Gidaspow D, Solbrig W. Multiphase flow-models for nuclear, fossil and biomass energy conversion. In: Majumdar AS, Mashelkar RA, editors. Advances in transport processes. New York: Wiley-Eastern Publisher, 1982: 198–351.Search in Google Scholar

Matsen JM. Drift flux representation of gas-particle flow. Powder Technol 2000; 111: 25–33.10.1016/S0032-5910(00)00236-9Search in Google Scholar

Miller A, Gidaspow D. Dense, vertical gas-solid flow in a pipe. AIChE J 1992; 38: 1801–1815.10.1002/aic.690381111Search in Google Scholar

Mohammad-Reza Mostofi A. CFD simulations of particulate two-phase and three phase flows. PhD thesis, Illinois Institute of Technology, 2002.Search in Google Scholar

Pannala S, Syamal M, O’Brien TJ, editors. Computational gas-solid flows and reacting systems. Theory, methods and practice. Hershey, PA, USA: IGI Global, 2011.10.4018/978-1-61520-651-3Search in Google Scholar

Pape R, Gidaspow D. Numerical simulation of intense reaction propagation in multiphase systems. AIChE J 1998; 44: 294–309.10.1002/aic.690440208Search in Google Scholar

Savage SB. Granular flows at high shear rates. In: Meyer RE, editor. Theory of dispersed multiphase flow. New York, USA: Academic Press Inc, 1983: 339–358.10.1016/B978-0-12-493120-6.50019-2Search in Google Scholar

Savage SB, Jeffrey DJ. The stress tensor in a granular flow at high shear rates. J Fluid Mech 1981; 110: 255–272.10.1017/S0022112081000736Search in Google Scholar

Schlichting HT. Boundary-layer theory, 4th ed., New York, USA: Mc Graw Hill, 1960.Search in Google Scholar

Shuai W, Zhenhua H, Huilin L, Guodong L, Jiaxing W, Pengfei X. A bubbling fluidization model using kinetic theory of rough spheres. AIChE J 2012; 58: 440–455.10.1002/aic.12590Search in Google Scholar

Sinclair JL, Jackson R. Gas-particle flow in a vertical pipe with particle-particle interactions. AIChE J 1989; 35: 1473–1486.10.1002/aic.690350908Search in Google Scholar

Songprawat S, Gidaspow D. Multiphase flow with unequal granular temperatures. Chem Eng Sci 2010; 65: 1134–1143.10.1016/j.ces.2009.09.068Search in Google Scholar

Soo SL. Fluid dynamics of multiphase systems, 1st ed., Waltham, Massachusetts, USA: Blaisdell Publishing Corporation, 1967.Search in Google Scholar

Strumendo M, Arastoopour H. Solution of population balance equations by the finite size domain complete set of trial functions method of moments (FCMOM) for inhomogeneous systems. Ind EngChem Res 2010; 49: 5222–5230.10.1021/ie901407xSearch in Google Scholar

Strumendo M, Gidaspow D, Canu P. Method of moments for gas-solid flows: application to the riser. In: Cen K, editor. Circulating fluidized bed technology VIII. New York: Intern. Acad. Pub., 2005: 936–942.Search in Google Scholar

Sun B, Gidaspow D. Computation of circulating fluidized-bed riser flow for the fluidization VIII benchmark test. Ind Eng Chem Res 1999; 38: 787–792.10.1021/ie9803669Search in Google Scholar

Syamlal M, Rogers W, O’Brien TJ. MFIX documentation; theory guide no. DOE/METC-94/1004 (DE94000087): Morgantown Energy Technology Center, 1993.10.2172/10145548Search in Google Scholar

Syamlal M, Musser J, Dietiker J-F. The two-fluid model in MFIX. In: Michaelides EE, Crowe CT, Schwarzkopf JD, editors. Multiphase flow handbook, 2nd ed., Boca Raton, Florida, USA: CRC Press, 2016: 242–274.Search in Google Scholar

Tartan M, Gidaspow D. Measurement of granular temperature and stresses in risers. AIChE J 2004; 50: 1760–1775.10.1002/aic.10192Search in Google Scholar

Tsuo YP, Gidaspow D. Computation of flow patterns in circulating fluidized beds. AIChE J 1990; 36: 885–896.10.1002/aic.690360610Search in Google Scholar

Wallis GB. One-dimensional two-phase flow, 1st ed., New York, USA: McGraw-Hill, 1969.Search in Google Scholar

Wei F, Lin H, Chang Y, Wang Z, Jin Y. Profiles of particle velocity and solids faction in a high density riser. Powder Technol 1998; 100: 183–189.10.1016/S0032-5910(98)00139-9Search in Google Scholar

Werther J. Fluid dynamics, temperature and concentration fields in large-scale CFB combustors. In: Cen K, editor. Circulating fluidized bed technology VIII, China: International Academic Pub., 2005: 1–25.Search in Google Scholar

Received: 2016-9-26
Accepted: 2017-2-26
Published Online: 2017-4-21
Published in Print: 2018-4-25

©2018 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 10.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/revce-2016-0044/html
Scroll to top button