Home Fluid Flow and Solute Transfer in a Tube with Variable Wall Permeability
Article Publicly Available

Fluid Flow and Solute Transfer in a Tube with Variable Wall Permeability

  • M. Varunkumar EMAIL logo and P. Muthu
Published/Copyright: August 1, 2019

Abstract

We considered a steady flow of viscous incompressible fluid and solute transfer in an axisymmetric tube of uniform cross section with variable wall permeability, which is relevant to the study of movement of solute across the glomerular capillaries. The solutions for the nonlinear governing equations of the fluid flow and solute transfer are obtained by analytical/numerical methods. The combined effect of variable wall permeability and flow parameters on the hydrostatic pressure, osmotic pressure, velocity profiles, concentration profiles, and the total solute clearance are investigated and are presented in this paper. It is found that an increase in the variable permeability parameter increases the solute concentration at the wall.

1 Introduction

The formation of urine involves filtration and reabsorption processes in the flow through nephrons which are permeable to water and solutes of small molecules. The glomerulus is the place of urine formation, which filter fluid and solute compounds from blood. Blood enters the glomerulus by way of the afferent arteriole and leaves through the efferent arteriole. Glomerular filtration is a bulk process where water and dissolved components can move freely between the vascular space and Bowman’s capsule. The glomerular filtrate is formed due to the Starling’s forces in the glomerular capillaries. Under these conditions, there is a net movement of fluid from the glomerular capillaries into Bowman’s capsule [1], [2], [3].

Many researchers have developed theoretical and mathematical models to study the fluid flow in renal tubules and ultrafiltration process in the glomerular capillaries. These models which are related to the fluid flow through the permeable boundary of the tubules have wide range of applications in understanding the mechanics of physiological fluid flows [4], [5], [6], [7].

The ultrafiltration process and movement of the solute particles across the glomerular capillaries have been analysed by Brenner et al. [8], [9]. Deen et al. [10], [11] presented the concept of concentration distribution in a capillary using the relation between the transmembrane fluid flux and the differences of hydrostatic pressure with osmotic pressure. They also employed the relation between the concentration of proteins and osmotic pressure by considering the pressure gradient is constant at the capillary wall. A mathematical model for the glomerular ultrafiltration process by taking pressure gradient as a function of axial distance of the capillary was developed by Marshal and Trowbridge [12] and Papenfuss and Gross [13].

The above studies have not taken the movement of the solute particles across the membrane of the capillary. But in reality, the movement is significant in the physiological situation as mentioned by Guyton [1]. Papenfuss and Gross [14] and Salathe [15] have analysed the capillary tissue fluid exchange and the transport of solute across the wall, assuming uniform concentration at each cross-section. By neglecting osmotic pressure, Ross [16] investigated the mass transport in fluid flow through a permeable tube. Tyagi and Abbass [17] studied the distribution of solute concentration in a blood passage in a hallow-fibre artificial kidney.

The transcapillary movement of water across the glomerular capillary is governed by the Starling’s forces that control the fluid movement across the capillary wall. Chaturani and Ranganatha [18] mentioned that ultrafiltration occurs because of both hydrostatic and osmotic pressures and the movement of solutes across the permeable membrane by diffusion and convection under the assumption of constant wall permeability. It may be noted that the assumption of constant permeability of glomerular capillary is of mathematical simplicity. There can be a significant changes in the filtration process of individual glomerulus in different regions of the kidney, where some nephrons are having larger volume compared with the other nephrons. The ultrafiltration process is dependent on the hydraulic conductivity and capillary surface area and this surface area may vary in the individual glomerular capillaries (Pollak et al. [19]). Thus, it can be concluded that permeability is not constant in a glomerular capillary.

Motivated by the article [19], an attempt has been made to describe the solute transfer and fluid flow through a tube by considering variable wall permeability. We explored the model, analysis, and solution procedure in detail, in the following Sections.

Figure 1: Geometrical representation of glomerular capillary.
Figure 1:

Geometrical representation of glomerular capillary.

2 Mathematical Model

We considered the axisymmetric transport of fluid in a permeable tube of radius R and the length L, as represented in Figure 1. The u(r, z) and v(r, z) are velocity components along axial (z) and radial (r) directions, respectively. Assuming the low Reynolds number flow (of order 103), the governing equations for incompressible Newtonian fluid flow of constant viscosity and the equation for solute transfer are,

(1)uz+1rr(rv)=0,
(2)Pz=μ[1rr(rur)],
(3)Pr=0,
(4)ucz+vcr=D[1rr(rcr)]

where P and μ are the pressure and dynamic viscosity coefficient of the fluid, respectively and c and D are the concentration and diffusivity of the solute, respectively.

Now, we considered the boundary conditions for the above system of equations. At the inlet cross-section, solute enters with uniformly distributed concentration and the fluid flow has a constant hydrostatic pressure and a constant flux. Mathematically, they are represented as:

(5)c=c0atz=0
(6)ΔP=ΔPaatz=0
(7)0R2πru(r,0)dr=Q0atz=0.

Fluid flow and solute concentration distribution are axisymmetric. Hence, we have

(8)ur=cr=v=0atr=0.

Fluid flow through the permeable tube wall is governed by Starling’s hypothesis [18], [19],

(9)v=k(z)(ΔPσΔπ)=VR(z),say,atr=R.

At the permeable wall, we shall assume the no-slip condition for axial velocity, given as,

(10)u=0atr=R.

Solute mass flux at the permeable wall is given as,

(11)Dcr=h(ccT)+(TR1)VRΦatr=R

where Φ={c;VR>0cT;VR<0, ΔP=PPT, ΔPa=PaPT, Δπ=ππT, (P,PT) – the hydrostatic pressure inside and outside the tube wall and (π,πT) – the osmotic pressure inside and outside the tube wall (PT and πT are taken as constant). Pa is hydrostatic pressure at z = 0 (entrance) and Q0 is volume flow rate at z = 0. σ is the reflection coefficient [20], c0 is the solute concentration at z = 0, cT is the concentration outside the tube, h is the permeability of solute at the wall and TR is transmittance coefficient [21]. The k(z) is the arbitrary function of axial distance z, denoting variable wall permeability.

The osmotic pressure of solute is usually a polynomial function of its concentration, as in [13],

(12)π(c)=2.1c(R,z)+0.16c(R,z)2+0.009c(R,z)3

The solute mass flux JS across the permeable tube wall [20] is given as,

(13)JS(z)=h(c(R,z)cT)+TRVRΦ

The total solute clearance JC is given as,

(14)JC=z=0z=L2πRJS(z)dz.

To analyze the fluid flow, the following dimensionless quantities (denoted by ^)

z=z^R,    r=r^R,    u=u^U0,    v=v^U0,VR=V^RU0,    c=    c^c0,    cT=c^Tc0,    ΔP=ΔP^ΔPa,Δπ=Δπ^ΔPa,    k=k^ka,    JS=J^Sc0D/a,    JC=Q0c0J^C,Q=Q0Q^,    Q0=πR2U0,

are introduced, in (1)–(4). The dimensionless equations are (after dropping ^)

(15)uz+1rr(rv)=0
(16)(ΔP)z=1Rp[1rr(rur)]
(17)(ΔP)r=0
(18)ucz+vcr=1Pe[1rr(rcr)]

Here, Rp=RΔPa/(μU0) is a nondimensional parameter and Pe=U0R/D is the Peclet number, which is a measure of the mass transfer by convection compared due to diffusion.

The dimensionless boundary conditions are

(19)c=1atz=0
(20)ΔP=1atz=0
(21)012ru(r,0)dr=1atz=0
(22)ur=cr=v=0atr=0
(23)v=εRpk(z)(ΔPσΔπ)=VR(z)atr=1
(24)u=0atr=1
(25)cr=Sh(cTc)+Pe(1TR)VRΦatr=1where Φ={c;VR>0cT;VR<0

The non-dimensional form of JS of (13) and JC of (14) are given as,

(26)JS=Sh(c(1,z)cT)+PeTRVRΦatr=1,
(27)JC=1Pe0L/RJSdz,

where, Sh = hR/D is Sherwood number and ε=kaμ/R is the ultrafiltration parameter.

3 Analysis

We solved (15) and (16), for u and v, using the conditions (22) and (24) and we got the axial and radial velocities as,

(28)u(r,z)=Rp4d(ΔP)dz(r21),
(29)v(r,z)=Rp16d2(ΔP)dz2(2rr3).

The second order ordinary differential equation governing the hydrostatic pressure (ΔP) is determined by the (29) and (23), is given by,

(30)d2(ΔP)dz216ϵk(z)(ΔPσΔπ)=0.

The boundary conditions (20) and (21) will take the form,

(31)ΔP=1atz=0
(32)dΔPdz=8Rpatz=0

where

(33)Δπ=ππT=2.1[c(1,z)cT]+0.16[c(1,z)2cT2]+0.009[c(1,z)3cT3].
Figure 2: Grid detail for Crank–Nicolson scheme around node (i, j).
Figure 2:

Grid detail for Crank–Nicolson scheme around node (i, j).

Figure 3: Flow sheet representing the iterative method used to solve the flow parameters.
Figure 3:

Flow sheet representing the iterative method used to solve the flow parameters.

The dimensionless solute concentration (18),

(34)ucz+vcr=1Pe[1rr(rcr)]

with corresponding initial/boundary conditions,

(35)c=1atz=0
(36)cr=0atr=0
(37)cr=Sh(cTc)+Pe(1TR)VRΦatr=1whereΦ={c;VR>0cT;VR<0
Figure 4: The hydrostatic pressure (ΔP) distributions for different values of U0.
Figure 4:

The hydrostatic pressure (ΔP) distributions for different values of U0.

Figure 5: The hydrostatic pressure (ΔP$\Delta P$) distribution for different values of TR and h.
Figure 5:

The hydrostatic pressure (ΔP) distribution for different values of TR and h.

The solutions of the (30) and (34), in closed form, are difficult to obtain due to the coupled nature of hydrostatic pressure and solute concentration. So, we need to solve (30) and (34) using boundary conditions (31), (32) and (35)–(37), respectively.

Figure 6: Effect of variable permeability on axial distribution of hydrostatic pressure (ΔP) for different values of D and ΔPa$\Delta{P_{a}}$.
Figure 6:

Effect of variable permeability on axial distribution of hydrostatic pressure (ΔP) for different values of D and ΔPa.

4 Numerical Solution of Transport Equation

Now, we solved (34) for solute concentration (c), which is coupled and dependent on u, v, ΔP and Δπ as given in (28), (29), (30) and (33), respectively. A solution for c in terms of an analytical expression not possible in this case involving u and v, explicitly. Hence, we obtained numerical solution of c, satisfying the initial/boundary conditions (19)–(25) using an iterative procedure, outlined below.

Equation (34) is discretized by finite difference method of Crank–Nicolson scheme. Let the tube inlet and outlet be denoted by index i = 0 and i = N, respectively. Similarly, the axis of the tube and the boundary are represented by the index j = 0 and j = M. Δz and Δr are the increments in z and r directions, respectively (Refer Fig. 2). The finite difference scheme corresponding to (34) at a general grid point with node index (i, j), is written as,

(38)Ajci+1,j1+Bjci+1,j+Ejci+1,j+1=Rj,for0<iN,0jM,

where Aj=λ(2rΔr+rPeΔrvi+1,j),Bj=2r(2λ+Pe(ui+1,j+ui,j)),Ej=λ(2r+ΔrrPeΔrvi+1,j) and Rj=λ(2rΔr+rPeΔrvi,j)ci,j1+2r(2λPe(ui+1,j+ui,j))ci,jλ(2r+ΔrrPeΔrvi,j)ci,j+1 with λ=Δz/(Δr)2.

Figure 7: The osmotic pressure (Δπ$\Delta\pi$) distribution for different values of U0.
Figure 7:

The osmotic pressure (Δπ) distribution for different values of U0.

Figure 8: The osmotic pressure (Δπ$\Delta\pi$) distribution for different values of TR and h.
Figure 8:

The osmotic pressure (Δπ) distribution for different values of TR and h.

Figure 9: Effect of variable permeability on osmotic pressure (Δπ$\Delta\pi$) for different values of D and ΔPa$\Delta{P_{a}}$.
Figure 9:

Effect of variable permeability on osmotic pressure (Δπ) for different values of D and ΔPa.

These coefficients Aj, Bj, Ej, and Rj are valid for interior nodes. For the nodes at the boundary and axis of the tube, we simplify (38) using the discretised form of conditions (36) and (37) and the details are given below:

Figure 10: Effect of variable permeability on axial velocity u with r, at z = 15, z = 30 and z = 50.
Figure 10:

Effect of variable permeability on axial velocity u with r, at z = 15, z = 30 and z = 50.

Figure 11: Effect of variable permeability on radial velocity (v)$(v)$ with r, at z = 15, z = 30 and z = 50.
Figure 11:

Effect of variable permeability on radial velocity (v) with r, at z = 15, z = 30 and z = 50.

Figure 12: Concentration profiles (c)$(c)$ with r at different cross-sections z = 15, z = 25 and z = 50 the tube for different values of TR and h.
Figure 12:

Concentration profiles (c) with r at different cross-sections z = 15, z = 25 and z = 50 the tube for different values of TR and h.

Figure 13: Effect of variable permeability on concentration profiles c with r.
Figure 13:

Effect of variable permeability on concentration profiles c with r.

At r = 0, we have,

(39)B0ci+1,0+E0ci+1,1=R0forj=0

where B0=2λPe(ui+1,0+ui,0),E0=2λ; and R0=(2λPe(ui+1,0+ui,0))ci,02λci,1.

At r = 1, we have,

ci,M=2ΔrShcT+4ci,M1ci,M23+2ΔrSh2ΔrPe(1TR)VR  if VR>0 and
(40)ci,M=2ΔrcT(Sh+Pe(1TR)VR)+4ci,M1ci,M23+2ΔrShifVR<0.

Equations (38), (39), and (40) will generate a system of linear, simultaneous equations for unknown ci,j for each node (i,j). The coefficients of the matrix can be represented in a tridiagonal form and is solved by well known Thomas algorithm to get c(r,z) at discretised grid points.

The c, ΔP and Δπ values are known, at i = 0 level. To get the c values at i = 1 level, we applied the following iterative procedure. For this, we assumed an approximate value of c, say c1,M0, at the boundary and it needs to be corrected based on discretised scheme of (38) along with condition of (37). The corrected c value (say c1,M1 at the boundary) is obtained by three point backward difference formula corresponding to the condition (37). The iterative process is continued to get the correct value of c at the boundary, at the level i = 1, with the tolerance error value of 106. With this c value, i = 2 level values of c at the boundary is calculated in a similar way and this procedure was repeated until i=N level. The various steps of the algorithm of the iterative process is shown in flow chart as in Figure 3.

5 Results and Discussion

In this analysis, we considered the tube wall with variable permeability (k(z)) for 0z1, which is more significant in physiological situations [1], [19], [22]. So, we took a linear function form [23],

(41)k(z)=1+θz

where θ is the permeability parameter.

The computational results are obtained for a given set of parameters relevant to the physiological situation [11], [13]. The flow quantities such as osmotic and hydrostatic pressures, velocity, solute concentration and solute clearance have been computed from analytical and numerical solutions based on these parameter values.

5.1 Hydrostatic Pressure (ΔP) and Osmotic Pressure (Δπ)

The axial distributions of hydrostatic pressure (ΔP) and osmotic pressure (Δπ) are shown in the Figures 49. The hydrostatic pressure decreases linearly along the axial direction and the osmotic pressure rises nonlinearly with axial length, in general. It is observed that the ΔP distribution decreases as U0 increases (Fig. 4). ΔP profiles are not much affected by solute wall permeability h and transmittance coefficient TR, as shown in Figure 5. Figure 6 depicts the influence of variable wall permeability θ on the hydrostatic pressure. The ΔP decreases along the axial direction as θ increases. If the diffusion coefficient D increases, the hydrostatic pressure increases (Fig. 6a and b). For ΔPa, the ΔP rises at the discharge end (Fig. 6b and c).

From Figure 7, the Δπ distribution decreases as U0 increases. The osmotic pressure profile is nonlinear, due to the quadratic relationship between the solute concentration and the osmotic pressure (Vander et al. [3] and Pollak et al. [9]). The rise in values of h means that the increased movement of solute across the permeable wall and due to this osmotic pressure decreases along the axial direction. Figure 8 depicts the effect of variable wall permeability (θ) on Δπ. As θ increases, the Δπ increases. If there is a decrease in diffusion coefficient D then the osmotic pressure decreases. The Δπ increases as ΔPa increases, refer Figure 9a–c.

5.2 Axial Velocity (u) and Radial Velocity (v)

Figures 10 and 11 indicate the axial and radial velocity profiles for various set of parameters. Figure 10 represents the axial velocity profiles with reference to variation in θ at different cross sections of the tube. We noted the significant changes in the behaviour of these profiles for varying permeability. Keeping the other parameters fixed and varying the values of θ, u decreases in a region close to the axis of the tube.

Figure 11 shows the effect of wall permeability (θ) on radial velocity v at different locations of the tube. The velocity v raises at the boundary as θ increases (Fig. 11). At z = 15, the velocity v is more s compared to v at z = 30 and z = 50. Both the axial and radial velocities, calculated from this analysis are matching well in the limiting case of θ → 0, which corresponds to the no-slip case study [18]. At the outset, we observed that the variable permeability (θ) has significant effect on axial velocity (u) and radial velocity(v).

5.3 Concentration Profiles (c(r, z))

Figure 12 shows the radial concentration profiles at two different cross sections of the tube, with effect of TR and h. At any given cross section, the value of the solute concentration at the wall is greater than the value at the centerline, as required by the balance between convective transport of protein towards the wall and diffusion away from it. The solute concentration increases with radial distance, as a result of removal of solute by ultrafiltration. If TR = 0 and h = 0 then there is no solute transport across the wall [10]. The effect of variable permeability θ on concentration profiles at different locations of the tube is shown in Figure 13. As θ increases the concentration of solute increases significantly, at various cross sections z = 15, z = 25, and z = 50. This indicates that the spread of the solute concentration is more as the fluid moves along the length of the tube.

5.4 Concentration at the Wall (cw(z))

Figure 14 illustrates the axial distribution concentration at wall, for different values of TR and h. It is clearly seen that the effects of transmittance coefficient and solute wall permeability are significant on the wall concentration. That is, the cw decreases as TR and h increases. It may be noted that the concentration at the wall at any axial position increases with the increase in variable permeability θ (See Fig. 15).

Figure 14: Distribution of wall concentration profiles cw for different values of TR and h.
Figure 14:

Distribution of wall concentration profiles cw for different values of TR and h.

Figure 15: Effect of variable permeability on the axial distribution of wall concentration profiles cw.
Figure 15:

Effect of variable permeability on the axial distribution of wall concentration profiles cw.

Figure 16: Axial distribution of concentration at the wall cw for (a) Δπ=0$\Delta\pi=0$ and (b) Δπ≠0$\Delta\pi\neq 0$ for different values of ε.
Figure 16:

Axial distribution of concentration at the wall cw for (a) Δπ=0 and (b) Δπ0 for different values of ε.

Figure 16 represents the concentration at the wall cw with axial distance, for the two cases Δπ=0 (Fig. 16a) and Δπ0 (Fig. 16b) for different values of ultrafiltration parameter ε. It may be noted that Δπ is a function of z as given in (12). Δπ=0 refers the case of the absence of osmotic pressure across the permeable boundary, which can be enforced explicitly in the (30).

It is observed from both the Figure 16a and b that the values of concentration at the wall are close to each other when ε=108. As ε increases, the concentration at the wall (cw) increases in both cases of Δπ. The numerical values of cw are comparatively large in the case of Δπ=0 than in the case of Δπ0, as noted in [16]. We mentioned that the osmotic pressure and ultrafiltraion have greater impact than the hydrostatic pressure, in this analysis.

Table 1:

Total solute clearance (JC) for TR=0.2, U0=0.07cm/sec, D=1×106cm2/sec and ΔPa=40×103dyn/cm2

ε=7.98×107ε=1.98×106
θSh = 0.005Sh = 0.035Sh = 0.005Sh = 0.035
0.00.02149910.03490160.04425320.0586395
0.050.04122480.05543410.07351180.0894347
0.10.05560310.07052420.08050.0973311
0.20.0595380.0750440.08806070.10602

5.5 Total Solute Clearance (JC)

The effects of θ, h and ε on solute clearance JC are shown in Table 1. It is clearly observed that the effect of permeability (θ) is significant on the clearance of solute. From Table 1, when the θ value is increased from 0.0 to 0.2, the solute clearance increases. The clearance of solute is improved by increasing the ultrafiltration parameter (ε) and solute wall permeability (h).

6 Conclusions

In this article, we have described the idea of fluid flow and solute transfer in a straight tube with wall permeability as a linear function of axial distance. Numerical solutions have been obtained for the governing coupled fluid flow and solute concentration equations. The effects of the variable permeability on hydrostatic pressure, osmotic pressure, velocity, concentration profiles, and solute clearance have been obtained.

  1. The effect of variable permeable wall on hydrostatic pressure, shows that as θ increases the ΔP curves decreases linearly. The pressure drop variations show that in the variable permeable case, different parametric values are greater than the corresponding constant permeability case. Also the pressure drop rapidly decreases in a variable permeable tube for increase in θ or the axial distance z, in comparison to the constant permeable tube.

  2. The osmotic pressure curves increases nonlinearly with axial distance.

  3. The axial and radial velocity values at the boundary are increasing along the axial direction, as θ increases.

  4. Concentration profiles have been observed and are affected by TR and h. Solute concentration rises as the wall variable permeability increases. The wall concentration decreases with the increase in TR and h and increases with θ.

  5. The total solute clearance increases with the values of θ, h, and ε. It may be noted that the present analysis studies the solute transfer and fluid flow in a straight tube. But in physiological situations, the capillaries are not uniform. So, this model can be made realistic by extending the above study to non-uniform tube.

References

[1] A. C. Guyton and E. H. John, Text Book of Medical Physiology, 11th Edition, Chapters 26 and 27, Pennsylvania: Elsevier Saunders Publishing 2011.Search in Google Scholar

[2] K. James and S. James, Mathematical Physiology, Chapter 17, 2nd Edition, New York: Springer 2009.Search in Google Scholar

[3] A. J. Vander, J. H. Sherman, and D. S. Luciano, Human Physiology – The Mechanisms of Body Function, Boston: McGraw-Hill, New York 2001.Search in Google Scholar

[4] A. Apelblat, A. K. Katchasky, and A. Silberberg., Biorheol. 11, 1 (1974).10.3233/BIR-1974-11101Search in Google Scholar

[5] A. S. Berman, J. Applied Physics. 24, 1232 (1953).10.1063/1.1721476Search in Google Scholar

[6] A. S. Berman, J. Applied Physics. 29, 71 (1958).10.1063/1.1722948Search in Google Scholar

[7] S. W. Yuan and A.B. Finkelstein, Trans. ASME. 78, 719 (1956).10.1115/1.4013794Search in Google Scholar

[8] B. M. Brenner, J. L. Troy, T. M. Daugharty, and W. M. Deen, Am. J. Physiol. 223, 1184 (1972).10.1152/ajplegacy.1972.223.5.1184Search in Google Scholar

[9] B. M. Brenner, C. Baylis, and W. M. Deen, Physiol. Rev. 56, 502 (1978).10.1152/physrev.1976.56.3.502Search in Google Scholar

[10] W. M. Deen, C. R. Robertson, and B. M. Brenner, Am. J. Physiol. 223, 1178 (1972).10.1152/ajplegacy.1972.223.5.1178Search in Google Scholar

[11] W. M. Deen, C. R. Robertson, and B. M. Brenner, Biophys. J. 14, 412 (1974).10.1016/S0006-3495(74)85924-2Search in Google Scholar

[12] E. A. Marshall and E. A. Trowbridge, J. Theor. Biol. 48, 389 (1974).10.1016/S0022-5193(74)80008-1Search in Google Scholar

[13] H. D. Papenfuss and J. F. Gross, Microvasc. Res. 16, 59 (1978).10.1016/0026-2862(78)90045-6Search in Google Scholar

[14] H. D. Papenfuss and J. F. Gross, Biorheology, 24, 319 (1987).10.3233/BIR-1987-24305Search in Google Scholar

[15] E. P. Salathe, Bull. of Mathematical Biology 50-3, 289 (1988).10.1016/S0092-8240(88)80069-7Search in Google Scholar

[16] M. S. Ross, J. Fluid Mechanics 63, 157 (1974).10.1017/S0022112074001066Search in Google Scholar

[17] V. P. Tyagi and M. Abbas, Bull. Math. Biol. 49-3, 697 (1987).10.1016/S0092-8240(87)90015-2Search in Google Scholar

[18] P. Chaturani and T. R. Ranganatha, Acta Mechanica. 96, 139 (1993).10.1007/BF01340706Search in Google Scholar

[19] M. R. Pollak, E. Q. Susan, P. H. Melanie, and D. D. Lance, Clin. J. Am. Soc. Nephrol. 9, 1461 (2014).10.2215/CJN.09400913Search in Google Scholar PubMed PubMed Central

[20] S. A. Regirer, J. Fluid dynamics 10, 442 (1975).10.1007/BF01015269Search in Google Scholar

[21] U. R. Shettinger, H. J. Prabhu, and D. N. Ghista, Med. Biol. Eng. Comput. 15, 32 (1977).10.1007/BF02441572Search in Google Scholar PubMed

[22] J. S. V. R. Krishnaprasad and P. Chandra, in: Bio Mechanics, (Eds. K. B. Sahay, B. K. Saxena), Willy Eastern Ltd, New Delhi, 1988, pp. 39–45.Search in Google Scholar

[23] P. Chaturani and T. R. Ranganatha, Acta Mechanica. 88, 11 (1991).10.1007/BF01170591Search in Google Scholar

Received: 2019-03-05
Accepted: 2019-07-11
Published Online: 2019-08-01
Published in Print: 2019-12-18

©2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 2.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/zna-2019-0071/html
Scroll to top button