Skip to main content
Log in

A Single-Pole Approximation to Interfacial Mass Transfer in Porous Media Augmented with Bulk Reactions

  • Published:
Transport in Porous Media Aims and scope Submit manuscript

Abstract

We study interfacial mass transfer between two immiscible fluids in porous media augmented with bulk reactions. A closed-form analytic approximation derived through Laplace transformation is represented for Darcy–Brinkman flow in the presence of first-order bulk reactions represented by the Damköhler number (Da). By comparing the residues of all the singularities, it is shown that the main contributor to the solution is the simple pole of the equations. We show that the solution holds for a range of \(Da.Pe<2.5\times 10^{5}\), where Pe is the Péclet number. By determining the bulk flow mass transfer coefficient, we verify our approximation compared to the previous studies and show acceptable conformity of the solutions. We find that the results are applicable for potential bulk reactions, such as biodegradation, in a typical groundwater system.

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

Notes

  1. We assume that the thickness ratio is almost constant over the length at which the equilibrium concentration profile is developed (i.e. the entrance length scale). The validity of this assumption better holds for higher Da.Pe (Sookhak Lari et al. 2015).

Abbreviations

\(\acute{C_{1-2}}\) :

\(= C_{1-2}{/}\left\langle u_2\right\rangle \)

\(\acute{P_{2}}\) :

\(= P_{2}{/}\left\langle u_2\right\rangle \)

\(\alpha \) :

\(=-Pe (Da+s\acute{P_{2}})\)

\(\bar{c}\) :

\(=c{/}c_{\mathrm{eq}}\)

\(\bar{x}\) :

\(=x{/}h_2\)

\(\bar{y}\) :

\(=y{/}h_2\)

\(\beta \) :

\(=-Pe\acute{C_{1-2}}\)

\(\dot{Sh}\) :

Sh in Dykaar and Kitanidis (1996)

\(\ell \) :

Laplace transformation

\(\epsilon \) :

Porosity

\(\eta \) :

\(=\sqrt{\epsilon {/}k}\)

\(\varGamma \) :

Gamma function

\(\gamma _{1,2}\) :

Parameters defined in (31)–(32)

\(\hat{Pe}\) :

Pore-scale Pe in Dykaar and Kitanidis (1996)

\(\mu _\alpha \) :

Viscosity of fluid \(\alpha \) (\(=1, 2\))

\(\nu \) :

\(=2\sqrt{-\alpha }{/}\tau \)

\(\overline{v}\) :

Pore-scale velocity

\(\tau \) :

\(=\eta h_2\)

\(\theta \) :

Inclination angle

\(\tilde{c}\) :

Laplace-transformed concentration

\(\underline{\nu }_\alpha \) :

Kinematic viscosity of fluid \(\alpha \) (=1, 2)

a :

Specific area

b :

Variable

c :

Dissolved concentration

\(C_{1-2}\) :

Constant in the velocity profile equation

\(c_{\mathrm{eq}}\) :

Dissolved equilibrium concentration

D :

Effective diffusion coefficient

d :

Pore length scale

Da :

Damköhler number

f :

Function defined in (33)

g :

Function defined in (34)

h :

Thickness of fluid

J :

Bessel function (first kind)

k :

Permeability

\(k_b\) :

Bulk reaction coefficient

\(P_2\) :

Constant in the velocity profile equation

Pe :

Péclet number

Re(s):

Real part of s

Res :

Residue

s :

Laplace transform variable

Sh :

Sherwood number

u :

Bessel function (generalized first kind)

\(u_\alpha \) :

Velocity of fluid \(\alpha \) (\(=1,2\))

w :

Bessel function (generalized)

x :

Streamwise coordinate

y :

Crosswise coordinate

Z :

\(=\left( 2\sqrt{\beta }e^{0.5\tau \bar{y}}\right) {/}\tau \)

\({Sh_N}\) :

Normalized Sherwood number

References

  • Baricz, A.: Generalized Bessel Functions of the First Kind. Springer, Berlin (1994)

    Google Scholar 

  • Batchelor, G.K.: Mass transfer from a particle suspended in fluid with a steady linear ambient velocity distribution. J. Fluid Mech. 95, 369–400 (1979)

    Article  Google Scholar 

  • Bird, R., Stewart, W., Lightfoot, E.: Transport Phenomena, 2nd edn. Wiley, New York (2002)

    Google Scholar 

  • Brinkman, H.C.: A calculation of the viscous force exerted by a flowing fluid in a dense swarm of particles. Appl. Sci. Res. 1, 27–34 (1949)

    Article  Google Scholar 

  • Chrysikopoulos, C.V., Hsuan, P., Fyrillas, M.M., Lee, K.Y.: Mass transfer coefficient and concentration boundary layer thickness for a dissolving NAPL pool in porous media. J. Hazard. Mater. B97, 245–255 (2003)

    Article  Google Scholar 

  • Chrysikopoulos, C.V., Kim, T.: Local mass transfer correlations for nonaqueous phase liquid pool dissolution in saturated porous media. Transp. Porous Media 38, 167–187 (2000)

    Article  Google Scholar 

  • Das, D., Hassanizadeh, S.M. (eds.): Upscaling Multiphase Flow in Porous Media From Pore to Core and Beyond. Springer, Berlin (2005)

    Google Scholar 

  • Dykaar, B.B., Kitanidis, P.K.: Macrotransport of a biologically reacting solute through porous media. Water Resour. Res. 32, 307–320 (1996)

    Article  Google Scholar 

  • Fyrillas, M.M.: Advection-dispersion mass transport associated with a non-aqueous-phase liquid pool. J. Fluid Mech. 413, 49–63 (2000)

    Article  Google Scholar 

  • Hildebrand, F.B.: Advanced Calculus for Applications. Prentice-Hall, Inc, Upper Saddle River (1962)

    Google Scholar 

  • Huntley, D., Beckett, G.D.: Persistence of LNAPL sources: relationship between risk reduction and LNAPL recovery. J. Contam. Hydrol. 59, 3–26 (2002)

    Article  Google Scholar 

  • Johnston, C.D., Trefry, M.G.: Characteristics of light nonaqueous phase liquid recovery in the presence of fine-scale soil layering. Water Resour. Res. 45(5), W05412 (2009)

    Google Scholar 

  • Kim, T., Chrysikopoulos, C.V.: Mass transfer correlations for nonaqueous phase liquid pool dissolution in saturated porous media. Water Resour. Res. 35, 449–459 (1999)

    Article  Google Scholar 

  • Krantz, S.G.: Complex Variables: A Physical Approach with Applications and MATLAB. Chapman & Hall/CRC, London (2008)

    Google Scholar 

  • Lawrence, S.J.: Description, properties, and degradation of selected volatile organic compounds detected in ground water: a review of selected literature. Technical report, USGS (2006)

  • Lekmine, G., Bastow, T., Johnston, C., Davis, G.: Dissolution of multi-component LNAPL gasolines: the effects of weathering and composition. J. Contam. Hydrol. 160, 1–11 (2014)

    Article  Google Scholar 

  • Lu, G., Clement, T.P., Zheng, C., Wiedemeier, T.H.: Natural attenuation of BTEX compounds: model development and field-scale application. Ground Water 37, 707–717 (1999)

    Article  Google Scholar 

  • Miller, C.T., Dawson, C.N., Farthing, M.W., Hou, T.Y., Huang, J., Kees, C.E., Kelley, C., Langtangen, H.P.: Numerical simulation of water resources problems: models, methods, and trends. Adv. Water Resour. 51, 405–437 (2013)

    Article  Google Scholar 

  • Neale, G., Nader, W.: Practical significance of Brinkman’s extension of Darcy’s law: coupled parallel flows within a channel and a bounding porous medium. Can. J. Chem. Eng. 52, 475–478 (1974)

    Article  Google Scholar 

  • Poularikas, A.D. (ed.): Transforms and Applications Handbook, 3rd edn. CRC Press, Boca Raton (2010)

    Google Scholar 

  • Seagren, E.A., Rittmann, B.E., Valocchi, A.J.: An experimental investigation of NAPL pool dissolution enhancement by flushing. J. Contam. Hydrol. 37, 111–137 (1999)

    Article  Google Scholar 

  • Sookhak Lari, K., Safavi, H.: A simulation-optimization model for ‘air sparging’ and ‘pump and treat’ groundwater remediation technologies. J. Environ. Inform. 12, 44–53 (2008). doi:10.3808/jei.200800123

  • Sookhak Lari, K., Johnston, C.D., Davis, G.B.: Interfacial mass transport in porous media augmented with bulk reactions: analytical and numerical solutions. Transp. Porous Media 106, 405–423 (2015)

    Article  Google Scholar 

  • Sookhak Lari, K., van Reeuwijk, M., Maksimović, Č.: Simplified numerical and analytical approach for solutes in turbulent flow reacting with smooth pipe walls. J. Hydraul. Eng. ASCE 136, 626–632 (2010)

    Article  Google Scholar 

  • Sookhak Lari, K., van Reeuwijk, M., Maksimović, Č.: The role of geometry in rough wall turbulent mass transfer. Heat Mass Transf. 49, 1191–1203 (2013)

    Article  Google Scholar 

  • Sookhak Lari, K., van Reeuwijk, M., Maksimović, Č., Sharifan, S.: Combined bulk and wall reactions in turbulent pipe flow: decay coefficients and concentration profiles. J. Hydroinform. 13(3), 324–333 (2011)

    Article  Google Scholar 

  • Trefry, M., McLaughlin, D., Metcalfe, G., Lester, D., Ord, A., Regenauer-Lieb, K.: On oscillating flows in randomly heterogeneous porous media. Proc. R. Soc. A 368, 197–216 (2010)

    Google Scholar 

  • Vafai, K., Tien, C.L.: Boundary and inertia effects on flow and heat transfer in porous media. Int. J. Heat Mass Transf. 24, 195–203 (1981)

    Article  Google Scholar 

  • van Reeuwijk, M., Sookhak Lari, K.: Asymptotic solutions for turbulent mass transfer at high Schmidt number. Proc. R. Soc. A 468, 1676–1695 (2012a)

    Article  Google Scholar 

  • van Reeuwijk, M., Sookhak Lari, K.: Asymptotic solutions for turbulent mass transfer augmented by a first order chemical reaction. Int. J. Heat Mass Transf. 55, 6485–6490 (2012b)

    Article  Google Scholar 

  • Vayenas, D.V., Michalopoulou, E., Constantinides, G.N., Pavlou, S., Payatakes, A.C.: Visualization experiments of biodegradation in porous media and calculation of the biodegradation rate. Adv. Water Resour. 25, 203–219 (2002)

    Article  Google Scholar 

  • Wang, C.Y.: Darcy Brinkman flow past a two dimensional screen. Eur. J. Mech. B Fluids 28, 321–327 (2009)

    Article  Google Scholar 

  • Warhaft, Z.: Passive scalars in turbulent flows. Ann. Rev. Fluid Mech. 32, 203–240 (2000)

    Article  Google Scholar 

  • Wilson, E.J., Geankoplis, C.J.: Liquid mass transfer at very low Reynolds numbers in packed beds. Ind. Eng. Chem. Fundam. 5, 9–14 (1966)

    Article  Google Scholar 

  • Wood, B.D., Golfier, F., Quintard, M.: Dispersive transport in porous media with biofilms: local mass equilibrium in simple unit cells. Int. J. Environ. Waste Manag. 7, 24–48 (2011)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kaveh Sookhak Lari.

Appendix

Appendix

“Lemma 1” section shows that \(\nu \) does not take nonzero integers in our study. “Lemma 2” section proves the problem has a simple pole at \(s=0\) and a number of identified singularities. “Lemma 3” section shows that except for the pole, the effect of residues of other singularities is negligible in inverse Laplace transformation.

1.1 Lemma 1

Lemma 1

For the range of Da . Pe in the study, the index of Bessel function defined by Eq. (28) does not take nonzero integers. That is, \(\nu (s) \notin \{\ldots ,-3,-2,-1,1,2,3,\ldots \}\).

Proof

Suppose \(\nu =\frac{2 \sqrt{-\alpha }}{\tau } \in \{ \ldots ,-3,-2,-1,1,2,3,\ldots \}\). We know that \(Da.Pe=\frac{\tau ^2 \nu ^2}{4}-Pe.\acute{P_2}.s\) (Eq. 28). Considering \(P_2 < 0\) (Sookhak Lari et al. 2015), the term \(Pe.\acute{P_2}.s\) is negative, which leads to \(Da.Pe > \frac{1}{4}\tau ^2 \nu ^2\). For this study, we suppose the upper bound of \(Da . Pe < \tau ^2 / 4\). This indicates that \(-1<\nu <1\), and therefore, \(\nu \) \(\notin \{ \ldots ,-3,-2,-1,1,2,3,\ldots \}\). Considering the most limiting range of parameters in Sookhak Lari et al. (2015) (i.e. \(\eta = 10^3\)), our study then holds for \(Da. Pe < 2.5 \times 10^5\).

1.2 Lemma 2

Lemma 2

The singularities of \(\tilde{c}\) are the pole \(s=0\), the removable singularity \(s=-\frac{Da}{\acute{P_2}}\) and the roots of function \( K(s)=e^{\frac{1}{2}\tau \nu }(1+\frac{1}{\nu +1}\frac{\beta }{\tau ^2})+e^{\frac{-1}{2}\tau \nu }(1+\frac{1}{-\nu +1}\frac{\beta }{\tau ^2})\). Except for these singularities, \(\tilde{c}\) is holomorphic.

Proof

Since s is a factor in the denominator of Eq. (35) and \(s.\tilde{c}\) is bounded in the neighbourhood of \(s=0\) (see Fig. 8), \(s=0\) is a first-order pole (Krantz 2008, p. 116).

We now consider the generalized Bessel function \(w_p\)

$$\begin{aligned} w_p(z)=\frac{z^p }{2^p \varGamma (p+\frac{b+1}{2})}u_p(z^2); \quad \forall z \in \mathbb {C}, \end{aligned}$$
(45)

where

$$\begin{aligned} u_p(z)=\sum _{n\ge 0} \frac{(-\frac{c}{4})^n z^n \varGamma (p+\frac{b+1}{2})}{n!\varGamma (p+\frac{b+1}{2}+n) };\quad p+\frac{b+1}{2}\ne 0,-1,-2,\ldots . \end{aligned}$$
(46)

is the generalized and normalized Bessel function of first kind and is analytic in \(\mathbb {C}\) (Baricz 1994, p. 11). \(\Gamma \) is also the Gamma function. By letting \(b=1\) and \(c=1\), the generalized Bessel function in Eq. (45) reduces to \(J_p (z)\) (Baricz 1994, p. 10). Therefore, the only possible non-analytic points for \(J_p (z)\) are the roots of the denominator of \(w_p(z)\), i.e. \((p+1) \in \{0,-1,,-2,\ldots \}\) (properties of the Gamma function) (Hildebrand 1962, p. 80).

As a result of this, by referring to the transformed solute concentration defined in Eq. (35), all possible non-analytic points (except for the roots of denominator) occur when \((\pm \nu (s) \vee \pm \nu (s) \pm 1) \in \{0,-1,-2,\ldots \}\). This is equal to \(\nu (s) \in \mathbb {Z}\). However, we know \(\nu (s)\notin \mathbb {Z}-\{0\}\) (see Lemma 1). For the case of \(\nu (s)=0\), Eq. (28) gives \(s=-\frac{Da}{\acute{P_2}}\). Figure 9 shows that \(\tilde{c}\) remains bounded at the neighbourhood of this point, and therefore, it is a removable singularity (Krantz 2008, p. 116).

Fig. 8
figure 8

Bounded behaviour of \(s.\tilde{c}\) near \(s=0\)

Another set of singular points is the roots for denominator of \(\tilde{c}\) (the part without the factor s). In order to obtain a better interpretation of how the roots are distributed and where their locations are, we first denote \(J_{\nu (s)}(Z(s,0)) f(\nu (s) , Z(s,-1))+J_{-\nu (s)}(Z(s,0))g(\nu , Z(s,-1))\) (in the denominator of Eq. 35) by \(J_\nu f+J_{-\nu } g \) for convenience. We then assume \(Z(s,-1)\) is a small argument for Bessel function of first kind. Indeed, this assumption is valid since \(\tau \) is a large quantity and makes \(Z(s,-1)=\frac{2\sqrt{\beta }}{\tau }e^{\frac{-1}{2}\tau }\) small enough to perform the approximation \(J_\nu (Z(s,-1)) \sim \frac{ Z(s,-1)^\nu }{2^\nu \varGamma (\nu +1)}\) (Hildebrand 1962, p. 151). Therefore,

$$\begin{aligned} J_\nu f+J_{-\nu } g= & {} \left( \frac{1}{2^{-\nu -1} \varGamma (-\nu )}Z(s,-1)^{-\nu -1}\!-\!\frac{1}{2^{-\nu +1}\varGamma (-\nu \!+\!2)}Z(s,-1)^{-\nu +1}\right) J_{\nu }(Z(s,0)\nonumber \\&+\left( \frac{1}{2^{\nu +1} \varGamma (\nu +2)}Z(s,-1)^{\nu +1}-\frac{1}{2^{\nu -1}\varGamma (\nu )}Z(s,-1)^{\nu -1}\right) J_{-\nu }(Z(s,0)).\nonumber \\ \end{aligned}$$
(47)

With the aid of asymptotic notation for Bessel function of first kind (by letting \(b=c=1\) in Eqs. 45 and 46) and substituting \(Z(s,-1)\) for \(\frac{2\sqrt{\beta }}{\tau }e^{\frac{-1}{2}\tau }\), Eq. (47) gets the form

$$\begin{aligned} J_\nu f+J_{-\nu } g= & {} e^{\frac{1}{2}\tau \nu }u_{\nu }\left( \frac{4\beta }{\tau ^2}\right) \left( \frac{\tau . e^{\frac{1}{2}\tau }}{\sqrt{\beta }.\varGamma (-\nu )\varGamma (\nu +1)}-\frac{\sqrt{\beta } }{\tau . e^{\frac{1}{2}\tau } \varGamma (-\nu +2)\varGamma (\nu +1)}\right) \nonumber \\&+\,e^{\frac{-1}{2}\tau \nu }u_{-\nu }\left( \frac{4\beta }{\tau ^2}\right) \left( -\frac{\tau . e^{\frac{1}{2}\tau }}{\sqrt{\beta }.\varGamma (\nu )\varGamma (-\nu +1)}+\frac{\sqrt{\beta } }{\tau . e^{\frac{1}{2}\tau } \varGamma (\nu +2)\varGamma (-\nu +1)}\right) .\nonumber \\ \end{aligned}$$
(48)

The second terms in the two parentheses above can be neglected since \(\tau \) is large and Gamma functions are bounded (note that \(\nu (s)\notin \mathbb {Z}\)). Moreover, application of reflection relationship between Gamma functions defined as \(\varGamma (\nu +1) \varGamma (-\nu )=\frac{-\pi }{\sin {\pi \nu }}\) \((\nu \notin \{0,\pm 1,\pm 2,\ldots \})\) (Hildebrand 1962, p. 80) and the recurrence formula \(\varGamma (\nu +1 )= \nu \varGamma (\nu )\) (Hildebrand 1962, p. 78) changes Eq. (48) into

$$\begin{aligned} J_\nu f+J_{-\nu } g=e^{\frac{1}{2}\tau \nu }u_{\nu }\left( \frac{4\beta }{\tau ^2}\right) \times \frac{\tau }{\sqrt{\beta }} \frac{e^{\frac{1}{2}\tau }}{\frac{- \pi }{\sin {\pi \nu }}}+ e^{\frac{-1}{2}\tau \nu }u_{-\nu }\left( \frac{4\beta }{\tau ^2}\right) \times \frac{-\tau }{\sqrt{\beta }} \frac{e^{\frac{1}{2}\tau }}{\frac{ \pi }{\sin {\pi \nu }}}. \end{aligned}$$
(49)
Fig. 9
figure 9

Sketch of the Laplace-transformed function near the removable singularity \(s=\frac{-Da}{\acute{p_2}}\). For \(Da.Pe=0.01\), \(\frac{-Da}{\acute{p_2}}=1.02\times 10^{-12}\) and for \(Da.Pe=100\), \(\frac{-Da}{\acute{p_2}}=1.02\times 10^{-8}\)

We are now willing to find the roots of function (49). Since \(\frac{\tau }{\sqrt{\beta }} \frac{e^{\frac{1}{2}\tau }}{\frac{ \pi }{\sin {\pi \nu }}}\) is not zero, then we solve equation \(e^{\frac{1}{2}\tau \nu }u_{\nu }(\frac{4\beta }{\tau ^2})+ e^{\frac{-1}{2}\tau \nu }u_{-\nu }(\frac{4\beta }{\tau ^2})=0\) instead. Lets denote \(K(s)=e^{\frac{1}{2}\tau \nu }u_{\nu }(\frac{4\beta }{\tau ^2})+ e^{\frac{-1}{2}\tau \nu }u_{-\nu }(\frac{4\beta }{\tau ^2})\) for convenience. The roots for this function cannot be found explicitly, so we continue to approximate K(s) in order to clarify the distribution and location of the roots.

Considering the asymptotic representation \(u_p(z)=\sum _{n\ge 0} \frac{(-\frac{1}{4})^n z^n \varGamma (p+1)}{n!\varGamma (p+1+n) }\) (Eq. 46 by letting \(b=c=1\)) and replacing in K(s) gives

$$\begin{aligned} K(s)=\sum _{n\ge 0}{\frac{\left( \frac{1}{4}\right) ^n \left( \frac{4\beta }{\tau ^2}\right) ^n}{n!}}\left( \frac{e^{\frac{1}{2}\tau \nu }\varGamma (\nu +1)}{\varGamma (n+\nu +1)}+\frac{e^{-\frac{1}{2}\tau \nu }\varGamma (-\nu +1)}{\varGamma (n-\nu +1)}\right) . \end{aligned}$$
(50)

Again, by applying repetitive reflection and recurrence formula to the product of Gamma functions defined as \(\varGamma (\nu +1)\varGamma (n-\nu +1)=\nu \frac{\pi }{\sin {\pi \nu }}(-\nu +1)(-\nu +2)\ldots (-\nu +n)\), \(\varGamma (-\nu +1)\varGamma (n+\nu +1)=-\nu \frac{-\pi }{\sin {\pi \nu }}(\nu +1)(\nu +2)\ldots (\nu +n)\) and \(\varGamma (n+\nu +1)\varGamma (n-\nu +1)=\nu \frac{\pi }{\sin {\pi \nu }}\{(\nu +1)(\nu +2)\ldots (\nu +n)\}\{(-\nu +1)(-\nu +2)\ldots (-\nu +n)\}\), K(s) can be written as

$$\begin{aligned} K(s)=\sum _{n\ge 0}{\frac{\left( \frac{1}{4}\right) ^n \left( \frac{4\beta }{\tau ^2}\right) ^n}{n!}}\left( \frac{e^{\frac{1}{2}\tau \nu }}{\prod \limits _{i=1}^{n}(\nu +i)}+\frac{e^{-\frac{1}{2}\tau \nu }}{\prod \limits _{i=1}^{n}(-\nu +i)}\right) . \end{aligned}$$
(51)

We then approximate the summation into its two main terms corresponding to \(n=0\) and \(n=1\). The result is

$$\begin{aligned} K(s)=e^{\frac{1}{2}\tau \nu }\left( 1+\frac{1}{\nu +1}\frac{\beta }{\tau ^2}\right) +e^{\frac{-1}{2}\tau \nu }\left( 1+\frac{1}{-\nu +1}\frac{\beta }{\tau ^2}\right) . \end{aligned}$$
(52)

Figure 10 shows an acceptable conformity between the roots of K(s) and the poles of the transformed function \(\tilde{c}\). Equation 52 suggests that there is an infinite number of roots that occur periodically. They do not make a continuous set, but there is the smallest root (greater than zero). This enables us to consider paths of integration in the complex plane which surrounds separately each singular point, and as a result, the reduction in complex integral (36) to the summation (37) is possible. Except for \(s=0\), \(s=\frac{-Da}{\acute{p_2}}\), and the roots of K(s), \(\tilde{c}\) is holomorphic (analytic) elsewhere. \(\square \)

Fig. 10
figure 10

Coincidence between the roots of K(s) and the poles of \(\tilde{c}\) for \(Da.Pe=0.01\) and 100

1.3 Lemma 3

Lemma 3

The residue truncation is acceptably small.

Proof

Let \(Res_0\) be the residue of \(e^{s\bar{x}}\tilde{c}(s)\) at s=0 and \(Res_i\) be the residue at singularity \(s=s_i\). Then,

$$\begin{aligned} \frac{Res_0}{Res_i}= & {} \frac{\lim _{s \rightarrow 0}(s-0)\tilde{c}(s)e^{s\bar{x}}}{\lim _{s \rightarrow s_i}(s-s_i)\tilde{c}(s)e^{s\bar{x}}}=\frac{\lim _{s \rightarrow 0}(s)\tilde{c}(s)e^{s\bar{x}}}{\lim _{t \rightarrow 0}(t)\tilde{c}(s_i+t)e^{(s_i+t)\bar{x}}}\nonumber \\= & {} \frac{\lim _{s \rightarrow 0}(s)\tilde{c}(s)e^{s\bar{x}}}{\lim _{s \rightarrow 0}(s)\tilde{c}(s_i+s)e^{(s_i+s)\bar{x}}}, \end{aligned}$$
(53)

where \(t=s-s_i\). We now take the limit sign out and rewrite Eq. (53) as

$$\begin{aligned} \frac{Res_0}{Res_i}=e^{-s_i\bar{x}}\lim _{s \rightarrow 0}\frac{\tilde{c}(s)}{\tilde{c}(s+s_i)}. \end{aligned}$$
(54)
Fig. 11
figure 11

Behaviour of the function \(\frac{\tilde{c}(s,\bar{y})}{\tilde{c}(s+s_i,\bar{y})}\) near \(s=0\) and \(s_i \in (10^{-12},10^{-7})\), for \(Da.Pe=0.01\)

Fig. 12
figure 12

Behaviour of the function \(\frac{\tilde{c}(s,\bar{y})}{\tilde{c}(s+s_i,\bar{y})}\) near \(s=0\) and \(s_i\in (10^{-10},10^{-7})\), for \(Da.Pe=100\)

We plotted the ratio \(\frac{\tilde{c}(s)}{\tilde{c}(s+s_i)}\) for \(Da.Pe=0.01\) and 100 in Figs. 11 and 12, respectively. It is observed that near \(s=0\), this ratio tends to infinity. Therefore, \(\frac{Res_0}{Res_i}\) is large enough to let us consider the residue of the simple pole \(s=0\) as the main contributor. It must be mentioned that for \(Da.Pe=0.01\), \(s_i \in (10^{-12},10^{-7})\) (to cover all other singularities) while this range is \(s_i \in (10^{-10},10^{-7})\) for \(Da.Pe=100\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lari, K.S., Moeini, M. A Single-Pole Approximation to Interfacial Mass Transfer in Porous Media Augmented with Bulk Reactions. Transp Porous Med 109, 781–797 (2015). https://doi.org/10.1007/s11242-015-0552-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11242-015-0552-3

Keywords

Navigation