Skip to main content
Log in

Effect of radial velocity profiles on axial dispersion in packed beds: asymptotic behaviour

  • Original Paper
  • Published:
Brazilian Journal of Chemical Engineering Aims and scope Submit manuscript

Abstract

Axial dispersion of a solute in a flow through packed beds arises from molecular diffusion, velocity variations at the interstitial-pore scale and systematic radial velocity profiles. Currently, there is not a widely accepted procedure for estimating the last contribution. One possible reason is the fact that radial velocity profiles in packed beds are intimately related on the radial porosity profile, which is not a deterministic property, and velocity variations are relatively mild and confined to a region close to the vessel walls. A widespread notion is that only in the zone about a particle radius from the wall is the fluid velocity significantly larger (wall channelling) than in the remaining of the bed core. Based on literature results and effective-model evaluations, the impacts of porosity profiles and related velocity profiles are explored in this manuscript. It is found that the dispersion of a solute can be increased several times in magnitude solely due to the wall channelling when the detailed porosity profile is considered. This can be explained in terms of small velocity differences between a second region (up to around five particle diameters) and the innermost bed region. Guidelines are also discussed for predicting likely levels of dispersion. A further aspect concerns the simplification of the treatment in the wall-zone.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Abbreviations

LBS:

Lattice Boltzmann Simulation

RWM:

Random Walk Method

TAMM:

Taylor-Aris method of moments

TZM:

Two-zone model

a = d t/d p :

Tube to particle diameter ratio

Bi p = k d p/D R :

Biot number

C :

Molar concentration of a solute, kmol m−3

d p, d t :

Particle and tube diameters, m

\(\overline{D}_{ch}\) :

Cross-section averaged superficial channel dispersion, m2 s1

D ax :

Superficial axial dispersion coefficient, m2 s1

D ch(ρ):

Local superficial channel dispersion coefficient, m2 s1

D m :

Molecular diffusivity, m2 s1

D R :

Characteristic superficial radial dispersion coefficient, m2 s1

D rad(ρ):

Local superficial radial dispersion coefficient, m2 s1

D T , int :

Interstitial transversal dispersion coefficient, m2 s1

D v :

Superficial trans-column dispersion coefficient, m2 s1

D v , :

Asymptotic value of Dv at long distances from the tracer input, m2 s1

k :

Mass exchange coefficient between wall and core zones, m s1

L = \(\tfrac{1}{4}\) a 2 Pe R :

Parameter defined in Eq. (31)

Pe j = u d p/D j :

Peclet number for superficial dispersion coefficient Dj

Pe m = u d p/(εD m):

Molecular Peclet number

Pe R :

Characteristic radial Peclet number

r :

Radial coordinate, m

r p, r t :

Particle and tube radii, m

Re p = u d p/υ :

Reynolds number

Re pc = u c d p/υ :

Reynolds number in the core zone

Sc = υ/D m :

Schmidt number

u = ε \(\overline{v}\) :

Cross-section averaged superficial velocity, m s1

u c, u w :

Average superficial velocities in the core and wall zone, m s1

v(ρ):

Local interstitial axial velocity, m s1

\(\overline{v}\) :

Cross-section averaged interstitial axial velocity, m s1

v c, v w :

Average interstitial velocities in the core and wall zone, m s1

v i, v m :

Average interstitial velocities in the inner and middle regions, m s1

V(ρ):

Variable defined in Eq. (8)

y = (r tr)/d p :

Dimensionless distance from the wall

z :

Axial coordinate, m

β :

Coefficient relating v and ψ, according to (Eq. 20)

χ = \((v - \overline{v})/\overline{v}\) :

Relative excess velocity

χ mi = (v m v i)/v i :

Relative excess velocity accounting for the middle region effect

χ wc = (v w v c)/v c :

Relative excess velocity accounting for the wall zone effect

ε :

Cross-section averaged porosity

ε 0 :

Porosity value eventually reached far away from the wall (Eq. 19)

ε c, ε w :

Average porosities in the core and wall zones

ε i, ε m :

Average porosities in the inner and middle regions

ρ = r/r t :

Dimensionless radial coordinate

ρ c = 1–1/a :

Dimensionless radial coordinate bounding core and wall zones

υ :

Kinematic viscosity, m2 s1

ψ(ρ):

Local porosity

:

Asymptotic value at long time or distance

I :

Initial

References

Download references

Acknowledgements

The authors are grateful for the financial support from the following argentine institutions: ANPCyT-MINCyT (PICT'15-3546), CONICET (PIP 0018) and UNLP (PID I226). C.D. Luzi, O.M. Martinez and G.F. Barreto are research members of CONICET.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Carlos Daniel Luzi.

Ethics declarations

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary file1 (DOC 99 kb)

Appendices

Appendix 1: Taylor-Aris formulation

The mass conservation Eq. (1) for a species in a packed bed is re-written here:

$$\psi (\rho )\frac{\partial C}{{\partial t}} = \frac{1}{{r_{t}^{2} }}\frac{\partial }{\rho \,\partial \rho }\left( {\rho D_{rad} (\rho )\,\frac{\partial C}{{\partial \rho }}} \right) + D_{ch} (\rho )\frac{{\partial^{2} C}}{{\partial \,z^{2} }} - \psi (\rho )\,v(\rho )\frac{\partial C}{{\partial z}};\quad \frac{\partial C}{{\partial \rho }} = 0\quad \rho \, = \,0,{1}$$
(40)

Equation (40) will be used for a bed infinitely extended in the axial direction with a tracer initially present (t = 0) according to an axi-symmetric distribution CI (ρ, z), and without being fed at t > 0. For convenience, a change of variables from z to ζ it is made, with ζ = z-\(\overline{v}\) t, where \(\overline{v}\) = \((2/\varepsilon )\int_{0}^{1} {\psi (\rho )\,v(\rho )\,\rho \,d\rho }\) is the average interstitial velocity and ζ is the mobile axial coordinate. Equation (40) then becomes:

$$\psi (\rho )\frac{\partial C}{{\partial t}} = \frac{1}{{r_{t}^{2} }}\frac{\partial }{\rho \,\partial \rho }\left( {\rho D_{rad} (\rho )\,\frac{\partial C}{{\partial \rho }}} \right) + D_{ch} (\rho )\frac{{\partial^{2} C}}{{\partial \,\zeta^{2} }} - \psi (\rho )\,\overline{v}\,\chi (\rho )\,\frac{\partial C}{{\partial \zeta }};\quad \frac{\partial C}{{\partial \rho }}\, = \,0,\rho \, = \,0,{1}$$
(41)

where the relative excess velocity \(\chi = (v - \overline{v})/\overline{v}\) verifies:

$$\int_{0}^{1} {\psi (\rho )\chi (\rho )\,\rho \,d\rho } = 0$$
(42)

Note that in Eq. (41) (\(\partial C/\partial t\)) is the partial derivate of C at constant ζ.

The normalized local spatial moment of order n is defined as:

$$c_{n} (t,\rho )\, = \,\tfrac{1}{{M_{I} }}\int_{ - \infty }^{\infty } {\zeta^{n} \,C(t,\zeta ,\rho )\,d\zeta } ;$$
(43)

where MI = N/(ε π \(r_{t}^{2}\)) and N is the initial moles of tracer, which will be kept inside the bed at all times. The cross-section averaged spatial moment of order n is defined as:

$$m_{n} (t)\, = \,\left( {2/\varepsilon } \right)\int_{0}^{1} {\psi (\rho )\;c_{n} (t,\rho )\,\rho \,d\rho }$$
(44)

It is assumed that CI (z,ρ) is such that cn(0,ρ) is finite and that the origin z = 0 is chosen in such a way that m1(0) = 0. Multiplying Eq. (41) by ζn and integrating on ζ =  ± ∞ gives:

$$\psi (\rho )\frac{{\partial c_{n} }}{\partial t} = \frac{1}{{r_{t}^{2} }}\frac{\partial }{\rho \,\partial \rho }\left( {\rho \,D_{rad} (\rho )\,\frac{{\partial c_{n} }}{\partial \rho }} \right) + n(n - 1)D_{ch} (\rho )\,c_{n - 2} + n\,\psi (\rho )\,\overline{v}\,\chi (\rho )\,c_{n - 1}$$
(45)

where \(\partial c_{n} /\partial \rho = 0\), ρ = 0,1. Differentiating Eq. (44) with respect to t and using Eq. (45) for \(\partial c_{n} /\partial t\):

$$\frac{{dm_{n} }}{dt} = n(n - 1)\frac{2}{\varepsilon }\int_{0}^{1} {D_{{^{ch} }} (\rho )\;c_{n - 2} (t,\rho \,)\rho \,d\rho } + n\;\overline{v}\frac{2}{\varepsilon }\int_{0}^{1} {\psi (\rho )\chi (\rho )\,c_{n - 1} (t,\rho \,)\;\rho \,d\rho }$$
(46)

In particular for n = 2:

$$\tfrac{1}{2}\varepsilon \frac{{dm_{2} }}{dt} = 2\int_{0}^{1} {D_{{^{ch} }} (\rho )\;c_{0} (t,\rho \,)\rho \,d\rho } + \;2\,\overline{v}\,\int_{0}^{1} {\psi (\rho )\chi (\rho )\,c_{1} (t,\rho \,)\;\rho \,d\rho }$$
(47)

The previous formulation is now applied for the so-called axial dispersion plug flow model, which assumes transversally uniform properties, porosity ε, velocity \(\overline{v}\), axial dispersion coefficient Dch = Dax, along with an initial distribution CI(z) also uniform on ρ. The terms in Eqs. (40, 41, 45) involving radial derivatives vanish. Under these conditions, it becomes, in particular, cn = mn, \(\chi (\rho )\) = 0 and c0 = 1 (at any t). Equation (47) is then reduced to:

$$D_{ax} \, = \,\tfrac{1}{2}\varepsilon \left( {dm_{2} /dt} \right)$$
(48)

Based on this result, for the two-dimensional model Eq. (40), an effective axial dispersion coefficient is analogously defined at a given t, with dm2/dt evaluated from Eq. (47). The moments c0 and c1 needed for this end are to be evaluated from Eq. (45):

$$\psi (\rho )\frac{{\partial c_{0} }}{\partial t} = \frac{1}{{r_{t}^{2} }}\frac{\partial }{\rho \,\partial \rho }\left( {\rho D_{rad} (\rho )\,\frac{{\partial c_{0} }}{\partial \rho }} \right);\quad \partial c_{0 } /\partial \rho \, = \,0\,{\text{ en }}\,\rho \, = \,0,{1}$$
(49)
$$\psi (\rho )\frac{{\partial c_{1} }}{\partial t} = \frac{1}{{r_{t}^{2} }}\frac{\partial }{\rho \,\partial \rho }\left( {\rho D_{rad} (\rho )\,\frac{{\partial c_{1} }}{\partial \rho }} \right) + \psi (\rho )\,\overline{v}\chi (\rho )\;c_{0};\quad \partial c_{{1}} /\partial \rho \, = \,0\,{\text{ en }}\,\rho \, = \,0,{1}$$
(50)

The two integrals in Eq. (47) are defined as:

$$\tfrac{1}{2}\varepsilon \,\left( {dm_{2,ch} /dt} \right) = \overline{D}_{ch,I} (t) = 2\int_{0}^{1} {D_{{^{ch} }} (\rho )\,c_{0} (t,\rho \,)\rho \,d\rho }$$
(51)
$$\tfrac{1}{2}\varepsilon \,\left( {dm_{2,v} /dt} \right)\, = \,D_{v} \left( t \right)\, = \,u\frac{2}{\varepsilon }\int_{0}^{1} {\psi (\rho )\chi (\rho )\,c_{1} (t,\rho \,)\;\rho \,d\rho }$$
(52)

where \(\overline{v}\) = \(u/\varepsilon\) was used. The second moment m2 has been split as m2 = m2,ch + m2,v and the overall dispersion coefficient Dax (Eq. 48) as:

$$D_{ax} \, = \,\overline{D}_{ch,I} \, + \,D_{v}$$
(53)

\(\overline{D}_{ch,I}\) is the average superficial coefficient of channel dispersion and Dv is the superficial trans-column dispersion coefficient caused by the non-uniform profile v(ρ). Both \(\overline{D}_{ch,I}\) and Dv depend on the initial distribution CI(ρ, z) and time t. However, for any CI(ρ, z), Eq. (49) indicates that, when \(t \to \infty\), c0 becomes uniform on ρ, and from the definition in Eq. (43) it follows that the asymptotic value is c0 = 1. Using this result in Eq. (50), it can be concluded that c1 reaches a stationary profile (i.e. \(\partial c_{1} /\partial t = 0\)) as t → ∞. Denoting as \(c_{1}^{\infty } (\rho )\) the stationary profile of c1, it is obtained from Eq. (50):

$$c_{1}^{\infty } (\rho ) = c_{1}^{\infty } (0) - \overline{v}\,r_{t}^{2} \int_{0}^{\rho } {V(\rho^{\prime})} \frac{{\quad d\rho^{\prime}}}{{D_{rad} (\rho^{\prime})\rho^{\prime}}}$$
(54)

where \(c_{1}^{\infty } (0)\) is the value at ρ = 0, and

$$V(\rho ) = \int_{0}^{\rho } {\psi (\rho ^{\prime})\,\chi (\rho ^{\prime})\,\rho ^{\prime}\,d\rho ^{\prime}}$$
(55)

It is noted that V(1) = 0, by virtue of Eq. (42). Equation (53) when t → ∞ is written as:

$$D_{ax,\infty } \, = \,\overline{D}_{ch} \, + \,D_{v,\infty }$$
(56)

With c0 = 1 in Eq. (51):

$$\overline{D}_{ch} = 2\int_{0}^{1} {D_{{^{ch} }} (\rho )\rho \,d\rho }$$
(57)

Substituting \(c_{1}^{\infty }\) ≡ c1 from Eq. (54) in Eq. (52), integrating by parts, having into account definition (55) and replacing \(\overline{v}\) = \(u/\varepsilon\), it is obtained for Dv,∞:

$$D_{v,\infty } \, = \,\left( {\frac{{u\,r_{t} }}{\varepsilon }} \right)^{2} 2\int_{0}^{1} {\;\frac{{V^{2} (\rho )\quad d\rho }}{{D_{rad} (\rho )\rho }}}$$
(58)

where the value \(c_{1}^{\infty } (0)\) (Eq. 54) has no effect on account of Eq. (42).

It is concluded that \(\overline{D}_{ch}\) (Eq. 57) and Dv,∞ (Eq. 58) become independent of CI(ρ, z).

Appendix 2: Dispersion in a bed with uniform oscillatory transversal porosity

Consider a packed bed between parallel plates with transversal porosity profile, far from the plates, given by ψ(ρ) = ε [1 + 0.5 cos(π ρ)], where ρ is the transverse coordinate in units of particle radius, dp/2. The amplitude of the waves is similar to the maximum of the waves in Fig. 2a and its wavelength is dp, also close to those of the actual profiles. The periodic velocity profile v(ρ) induced by ψ(ρ) will generate an asymptotic axial dispersion coefficient, Dwave, that can be evaluated from the TAMM. To this end, it is only necessary to consider an extension of half a wavelength, i.e., the range 0 < ρ < 1. Assuming the relation between v(ρ) and ψ(ρ) given by Eqs. (2023) in the main text, applied to the Cartesian coordinate rather than the cylindrical radius, it is obtained \(\psi_{c}^{2} = \int_{0}^{1} {\,\psi^{2} (\rho )\;\,d\rho }\) = \(\tfrac{9}{8}\) ε2, and v(ρ)/\(\overline{v}\) = χ(ρ) + 1 = \(\varepsilon \,\psi (\rho )/\psi_{c}^{2}\) = \(\tfrac{8}{9}\)[1 + 0.5 cos(π ρ)].

The equivalent expressions for V(ρ) and Dv,∞ = Dwave in the slab geometry (cfr. Eqs 8 and 4) with Drad (ρ) = DR (uniform) are:

\(V(\rho )\) = \(\int_{0}^{\rho } {\,\chi (\rho ^{\prime})\,\psi (\rho ^{\prime})\,d\rho } ^{\prime}\,\);

Dwave = \(\left( {\tfrac{1}{2}u\,d_{p} /\varepsilon } \right)^{2} (1/D_{R} )\int_{0}^{1} {\;V^{2} (\rho )d\rho }\).

It is obtained Dwave \(\cong \tfrac{1}{512}\,\,(u\,d_{p} )^{2} /D_{R}\), or 1/Pewave \(\cong\)\(\tfrac{1}{512}\) PeR. Considering very large values of Pem, PeR = 9 and 1/Pech = \(\overline{D}_{ch}\)/(u dp) = 0.5 can be taken as a reference. The value 1/Pewave \(\cong\) 0.0176 thus obtained is significantly lower than 1/Pech = 0.5, and it can be concluded that the waves of a stationary profile v(ρ) do not contribute significantly.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Luzi, C.D., Martinez, O.M. & Barreto, G.F. Effect of radial velocity profiles on axial dispersion in packed beds: asymptotic behaviour. Braz. J. Chem. Eng. 38, 865–885 (2021). https://doi.org/10.1007/s43153-021-00141-2

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s43153-021-00141-2

Keywords

Navigation