Mechanics Research Communications

9
Mechanics Research Communications 76 (2016) 32–40 Contents lists available at ScienceDirect Mechanics Research Communications journa l h om epa ge: www.elsevier.com/locate/mechrescom Semi-analytical solution for laminar particle-laden flow in curved lumina with permeable walls Semion Shaul , Tarek I. Zohdi Department of Mechanical Engineering, 6102 Etcheverry Hall, University of California, Berkeley, CA, 94720-1740, USA a r t i c l e i n f o Article history: Received 28 March 2016 Received in revised form 21 June 2016 Accepted 22 June 2016 Available online 29 June 2016 Keywords: Lumina tortuosity Effective permeability Creeping flow Nephrons Lymphatic capillaries a b s t r a c t In this work, a semi-analytical approach presenting a laminar particle-laden flow in curved lumen with permeable walls is developed. In order to illustrate its practical use, the fluid dynamics modeling of a human Juxtamedullary nephron is presented. Close agreement between the calculated values and the medical literature is found. The presented modeling approach provides an essential guide for researchers and designers to rapidly model and analyze the behavior of laminar particle-laden flows in lumen with permeable walls. © 2016 Elsevier Ltd. All rights reserved. 1. Introduction For the past few decades, bioengineering studies have made extensive use of Computational Fluid Dynamics (CFD) analysis as a proven tool for study, diagnosis, treatment and design of numerous bio fluid dynamics applications [1–5]. One of the essential ingre- dients of those CFD models is the lumen geometry, often derived from X-ray computed tomography. Other ingredients are the pres- sure and flow rates, often assumed (although occasionally based on experimental measurements), or available from ultrasonic Doppler imaging technics [6]. In this bio modeling, a complex geometry is discretized into a large number of smaller but regular (typically tetrahedral or hexahedral) elements. By assuming the shape of the velocity field within these elements, it is possible to solve the gov- erning Navier–Stokes equations at the connecting nodes of these elements. From this solution, it is straightforward to extract impor- tant fluid dynamics quantities. This has made CFD a particularly attractive tool for modeling fluid flow in lumina [7–11]. However, there are numerous bio-applications where using CFD is not prac- tical to achieve the designer and researcher needs. For example, for the single simulation of particle-laden con- tinua, the spatio-temporal discretization grids must be extremely fine, with several thousand numerical unknowns needed per parti- Corresponding author. E-mail addresses: [email protected] (S. Shaul), [email protected] (T.I. Zohdi). cle length-scale, and using extremely small time-steps, in order to obtain numerically accurate results. Thus, only for several hundreds of thousands of particles in a single lumen, a proper discretization would require several billion numerical unknowns (see i.e. [12]). Although, theoretically, simulations of single particle-laden con- tinua are possible in high-performance computing centers, their usefulness for rapid design and analysis is questionable for the amount of effort involved. Moreover, a simulation of the flow dynamics within the human kidney, containing about one million lumen of nephrons, or analysis of a human lymphatic system, which has at least 600 lymphatic nodes and much higher number of assigned lymphatic capillaries [13,14] becomes a simply unreachable objective for traditional CFD. However, a method that is able to rapidly simulate particle-laden fluid flow in a single lumen could certainly be applied to much larger networks and would be extremely useful for such purposes. It has been shown that various analytical models are applica- ble for rapid large-scale solutions (see i.e. [15–17]). These models are found to be applicable to describe laminar hemodynamics of various capillary systems. However, those models do not take into account the contributions of particles and lumen tortuosity to lam- inar flow dynamics, and therefore are not able to be used in certain cases. In response to this need, the present study suggests a novel theoretical approach as an essential guide for future research stud- ies to rapidly model and analyze particle-laden flows in either curved or straight lumina with permeable walls. http://dx.doi.org/10.1016/j.mechrescom.2016.06.004 0093-6413/© 2016 Elsevier Ltd. All rights reserved.

Transcript of Mechanics Research Communications

Page 1: Mechanics Research Communications

Sl

SD

a

ARRAA

KLECNL

1

epbdfseidtveetatt

tfi

(

h0

Mechanics Research Communications 76 (2016) 32–40

Contents lists available at ScienceDirect

Mechanics Research Communications

journa l h om epa ge: www.elsev ier .com/ locate /mechrescom

emi-analytical solution for laminar particle-laden flow in curvedumina with permeable walls

emion Shaul ∗, Tarek I. Zohdiepartment of Mechanical Engineering, 6102 Etcheverry Hall, University of California, Berkeley, CA, 94720-1740, USA

r t i c l e i n f o

rticle history:eceived 28 March 2016eceived in revised form 21 June 2016ccepted 22 June 2016vailable online 29 June 2016

a b s t r a c t

In this work, a semi-analytical approach presenting a laminar particle-laden flow in curved lumen withpermeable walls is developed. In order to illustrate its practical use, the fluid dynamics modeling of ahuman Juxtamedullary nephron is presented. Close agreement between the calculated values and themedical literature is found. The presented modeling approach provides an essential guide for researchersand designers to rapidly model and analyze the behavior of laminar particle-laden flows in lumen with

eywords:umina tortuosityffective permeabilityreeping flowephrons

permeable walls.© 2016 Elsevier Ltd. All rights reserved.

ymphatic capillaries

. Introduction

For the past few decades, bioengineering studies have madextensive use of Computational Fluid Dynamics (CFD) analysis as aroven tool for study, diagnosis, treatment and design of numerousio fluid dynamics applications [1–5]. One of the essential ingre-ients of those CFD models is the lumen geometry, often derivedrom X-ray computed tomography. Other ingredients are the pres-ure and flow rates, often assumed (although occasionally based onxperimental measurements), or available from ultrasonic Dopplermaging technics [6]. In this bio modeling, a complex geometry isiscretized into a large number of smaller but regular (typicallyetrahedral or hexahedral) elements. By assuming the shape of theelocity field within these elements, it is possible to solve the gov-rning Navier–Stokes equations at the connecting nodes of theselements. From this solution, it is straightforward to extract impor-ant fluid dynamics quantities. This has made CFD a particularlyttractive tool for modeling fluid flow in lumina [7–11]. However,here are numerous bio-applications where using CFD is not prac-ical to achieve the designer and researcher needs.

For example, for the single simulation of particle-laden con-inua, the spatio-temporal discretization grids must be extremelyne, with several thousand numerical unknowns needed per parti-

∗ Corresponding author.E-mail addresses: [email protected] (S. Shaul), [email protected]

T.I. Zohdi).

ttp://dx.doi.org/10.1016/j.mechrescom.2016.06.004093-6413/© 2016 Elsevier Ltd. All rights reserved.

cle length-scale, and using extremely small time-steps, in order toobtain numerically accurate results. Thus, only for several hundredsof thousands of particles in a single lumen, a proper discretizationwould require several billion numerical unknowns (see i.e. [12]).Although, theoretically, simulations of single particle-laden con-tinua are possible in high-performance computing centers, theirusefulness for rapid design and analysis is questionable for theamount of effort involved.

Moreover, a simulation of the flow dynamics within the humankidney, containing about one million lumen of nephrons, or analysisof a human lymphatic system, which has at least 600 lymphaticnodes and much higher number of assigned lymphatic capillaries[13,14] becomes a simply unreachable objective for traditional CFD.However, a method that is able to rapidly simulate particle-ladenfluid flow in a single lumen could certainly be applied to muchlarger networks and would be extremely useful for such purposes.

It has been shown that various analytical models are applica-ble for rapid large-scale solutions (see i.e. [15–17]). These modelsare found to be applicable to describe laminar hemodynamics ofvarious capillary systems. However, those models do not take intoaccount the contributions of particles and lumen tortuosity to lam-inar flow dynamics, and therefore are not able to be used in certaincases. In response to this need, the present study suggests a noveltheoretical approach as an essential guide for future research stud-

ies to rapidly model and analyze particle-laden flows in eithercurved or straight lumina with permeable walls.
Page 2: Mechanics Research Communications

S. Shaul, T.I. Zohdi / Mechanics Research

2

2

oantgl

o�avsflpoi

pe∑wpp

P

lE

wflg

U

wtrti

U

¯

Fig. 1. Forces acting on a control volume of a lumen.

. Theoretical modeling

.1. Model of particle-laden flow in a lumen

This analysis is focused on determining the forces, which actn a control volume of lumen. The essential assumption for thisnalysis is that the fluid properties, such as density and viscosity doot change within a given control volume. In addition, it is assumedhat the helical vortex secondary flow effect due to curvature of theeometry is negligible. However, this assumption will be revisitedater.

Fig. 1 presents a force depiction acting on a control volumef lumen given by geometry of the hydraulic diameter, DH (A =D2

H/4) and the containing permeable walls, of thickness t. Theverage pressure, P, the mass flow rate, m, and the average axialelocity, U, characterize the inlet or outlet contributions. The wallhear stress is marked by �w . Due to the pressure difference of theuid inside a lumen and the interstitial fluid, Pif , part of the fluidasses through the lumen walls with mass flow rate of mw into orut of the capillary (a case of fluid suction from lumen is presentedn Fig. 1).

By assuming steady state conditions, fully developed and incom-ressible laminar flow inside lumen, the integrated momentumquation in the z direction becomes:

F�z = moutUout − minUin, (1)

here the sum of the forces acting on control volume,∑

F�z , isroduced by fluid viscous shear stress with lumen walls whichroduces the average pressure difference. Eq. (1) becomes:

¯ inAin − PoutAout − �w�DH�z = moutUout − minUin. (2)

Assuming that change in lumen’s hydraulic diameter within theength of the control volume can be neglected (i.e. DH |�z = const),q. (2) can be rewritten as:

P�z = Pout − Pin = �eff (U2in − U2

out) − 4�w�z

DH, (3)

here �eff is the effective density of the biological particle-ladenuid which will be defined later in this study. Applying the inte-rated continuity equation on the control volume results in:

¯ out = Uin − 4Uw�z

DH, (4)

here Uw represents the average velocity of the fluid flowinghrough the lumen side wall. By applying Darcy’s law for perfo-ated flow and noting that the wall volumetric flux is equivalent tohe average wall velocity (Q

′′w = Qw/(�DH�z) = Uw ∝ �P/(t�eff )),

t is possible to define the average wall velocity as:

¯ w = keff

t�eff(Pin − Pif ), (5)

Communications 76 (2016) 32–40 33

where keff is the effective permeability coefficient of the lumenwalls. This coefficient must be modeled theoretically and inves-tigated experimentally for different wall properties: looking atporosity (or filtration ability), diffusivity and degree of interstitialfluid pressure.

As an example into a lumen’s permeability measurement, theliterature provides some information for different biological appli-cations. For example, the works of Verniory et al. [18], Miyamotoet al. [19] and Wang et al. [20] analyzed wall permeability coef-ficients for biological membranes, while the work of Kokko andTisher [21] examined the permeability in different segments ofnephron with respect to the filtration and diffusion properties.According to the results of Kokko and Tisher study, it is possible todefine the method for analysis of effective permeability coefficientfor nephrons as a functional of:

keff = f(

kh, g (kc, knc) , h(

T, Pif

)), (6)

where kh is hydrostatic component which must be derived fromquantitative analysis of lumen pores, andg (kc, knc) and h (T, Pif) arethe osmotic and the diffusion functions, respectively. The osmoticfunction depends on colloid osmotic, kc, and noncolloid osmotic,knc terms, while the diffusion function is characterized by tem-perature, T, and interstitial fluid pressure, Pif. It should be notedthat this coefficient must be examined appropriately with respectto the well-defined problem, and therefore, is a subject of futureresearch works. Since this coefficient depends on many parameters,the reader may find the work of Hagen et al. [22] useful, as it pro-vides an approach for accurate parameter estimation. However, forthe purposes of the present study, its value will be approximated byusing the reverse engineering method that is based on comparisonof the characteristic properties of different lumen’s segments.

The shear stress on the viscous particle-laden fluid is given interms of the effective viscosity, �eff , by the following:

� = �eff∂u

∂r. (7)

For a laminar flow and control volume length that is much smallerthan hydraulic diameter of the lumen (�z � DH), under a parabolicvelocity profile (for details see Appendix A), the shear stress at thewalls of the lumen is calculated from:

�w = �eff8Uin

DH. (8)

By substituting Eqs. (4) and (8) into Eq. (3) the average pressureloss of particle-laden fluid is estimated by:

�P�z = (8�eff Uw

DH− 32�eff

D2H

− 16�eff U2w

UinD2H

�z)Uin�z. (9)

Considering that the control volume length is small enough toneglect the high order terms, this solution can be also written interms of flow rate of the fluid, which is given as:

�P�z = (Qw − �eff

�eff4��z)

32�eff Qin

�2D4H

. (10)

For the possible case of the flow in lumen with impermeable walls(where Qw = 0), the developed relationship presented in Eq. (10)simplifies to the well-known form of Hagen-Poiseuille solution:

128�eff Qin�z

�P�z = −

�D4H

. (11)

It should be pointed out that solution presented in Eqs. (9) and (10)represents particle-laden flow through lumen for the cases where

Page 3: Mechanics Research Communications

3 earch Communications 76 (2016) 32–40

ln

D

i[efl

2l

�w

w

C

Aim

C

wa

R

To

elfc

C

C

fastbDrctc

Fig. 2. A comparison of correlations of corrected wall friction coefficient for laminarflow in pipe after White [23] and Ito [25].

Table 1Characteristics of analyzed fluid and lumen properties.

Parameter Value

Case I (De < 1) Case II (De ≥ 1)

Hydraulic diameter, DH, �m 50 100Volume fraction of particles, vp 0 − 0.1 0.05Particle density, �p, kg/m3 2500 2500Wall thickness, t, �m 2 2Effective permeability coeff., keff , m2 1 · 10−19 − 1 · 10−17 1 · 10−16

Dean number, De − 1 − 2000Fluid viscosity, �f , Pa · s 8 · 10−3 4 · 10−3

Fluid density, �f , kg/m3 1000 1000Initial flow rate, Qin, m3/s 1 · 10−12 5 · 10−10

Initial Reynolds number, Re ≈ 0.006 1.5

4 S. Shaul, T.I. Zohdi / Mechanics Res

umen geometry can be assumed as non-curved, or where the Deanumber, which is defined as:

e = 4ReDH

(DH

Rc

)1/2, (12)

s relatively small. According to the works of White [23] and Ito24,25], this assumption is valid for values of De < 11.6. The mod-ling approaches for the other cases of laminar particulate-ladenow are presented in the next section.

.2. Correction of the laminar particulate-laden flow due toumen curviness

By substituting Eq. (4) into Eq. (3) and dividing the result by

eff U2in/2, one can show that the normalized average pressure loss

ithin the control volume of small length is:

P�z = Pout − Pin

�eff U2

in/2= 2 2 − Qw

Qin

Qw

Qin− 4�z

DHCf , (13)

here Cf is the wall friction coefficient defined by:

f = �w

�eff U2

in/2. (14)

ccordingly to work of Raithby [26], describing the laminar flownside a non-curved permeable channel, this coefficient is esti-

ated as:

f ReDH=(

0.0481 + 0.0494

(Rew + 4.70)0.800

)−1

, (15)

here, Rew is the Reynolds number based on average wall velocitynd is presented as:

ew = ± UwDH�eff

�eff(16)

his number may be positive or negative, depending on the naturef the flow (injection or suction respectively).

For cases where the curviness of lumen affects the flow prop-rties significantly (resulting in secondary recirculating flow), theiterature presents some empirical relationships to correct the wallriction coefficient. Within the laminar flow when De > 11.6 theorrection presented by White [23] is:

∗f = Cf

(1 −(

1 − 11.6De

0.45) 1

0.45

)−1

, (17)

and

∗f = Cf

(21.5De

(1.56 + log (De))5.73

), (18)

or the range 13.5 < De < 2000 according to Ito [25]. Fig. 2 presents comparison between these empirical correlations. It is clearlyeen from this figure that there is no significant difference betweenhe ratio of the wall friction coefficients defined by Dean num-er smaller than 2000. Thus, for the range of laminar flow wheree < 2000, both correlations may be used to determine the cor-

ected wall friction coefficient. Hence, by replacing the wall frictionoefficient in Eq. (12) by the corrected one from Eq. (17) or Eq. (18),he average pressure loss in curved lumen of the length �zc can be

alculated by:

P�zc = ((Qin − 12

Qw)Qw − �zc

DHC*

f Q2in)

32�eff

�2D4H

. (19)

in

Initial average pressure, Pin, Pa 2000 4000Interstitial fluid pressure, Pif , Pa 800 800

3. Parametric analysis

Various parameters affect the fluid dynamics properties of theparticle-laden fluid. In this analysis, the following parameters arevaried: volume fraction of particles, effective permeability coeffi-cient and different level of lumen curvature, characterized by Deannumber. All the properties for fluid and lumen are given in Table 1.The parameters in Table 1 are divided into two cases, which areselected according to the Dean number range. This is done witha purpose to simulate a variety of possible cases of the laminarflow in a lumen. It is known that De ∝ Re, and for the case Re � 1,the laminar flow region is characterized as creeping flow (for moredetailed overview of creeping flow, the reader is referred to the stud-ies [27–29]). In addition to those variations in this analysis, otherparameter sets can be easily simulated and studied as well.

3.1. Volume fraction of particles, vp

Fig. 3 presents the results of the effect of the particulate volumefraction on the pressure loss and the flow rate of particle-ladenfluid moving in a small curvature lumen (see Table 1 -case I). Itcan be seen from Fig. 3 that whenever the volume fraction of par-ticles increases, the pressure loss increases while the flow rate

decreases. This happens because both the effective fluid densityand viscosity increase. Higher fluid viscosity, in turn, increasesthe wall shear stress and leads to the higher kinetic energy dis-sipation of the particulate-laden fluid. The curves of the modified
Page 4: Mechanics Research Communications

S. Shaul, T.I. Zohdi / Mechanics Research Communications 76 (2016) 32–40 35

Fig. 3. A comparison of: (a) – pressure loss, and (b) – flow rate, as a function of thevolume fraction of particles, vp. Showing the case of small curvature (De < 11.6)l

cFcteaHtcdsag

bE(wcriec

Fig. 4. A comparison of: (a) – pressure loss, and (b) – flow rate, as a function of

−18

umen lengths and keff = 5 · 10−18m2.

lassic expression of Hagen-Poiseuille (see Eq. (11)) were plotted inig. 3(a). This was done for the purpose of comparing pressure lossurves, calculated by using the modified classic expression, withhose developed in this study (see Eq. (10)). Thus, this comparisonxamines the effects of the permeable lumen walls. For example,ccording to Fig. 3(a), the pressure loss calculated by using modifiedagen-Poiseuille expression has a linear tendency to increase with

he lumen length. Contrary to this linear tendency, the pressure lossurves calculated from the model of this study have small parabolicependence with respect to the lumen length. This tendency is noturprising since part of the fluid passes through lumen walls, ands a result, both the flow rate and the shear stress at lumen wallsradually decreases.

Moreover, by demonstrating the variation of pressure lossounds in Fig. 3(a), the flow rate for the lumen length of 10, 000�mq. (10) is used as the initial flow rate (which is constant) in Eq.11). It can be seen from the presented bound comparison thathenever particulate volume ratio is relatively high or low, the

alculated pressure loss from Eq. (10) is closer to the upper boundesulting from Eq. (11) along the entire lumen length. This resultndicates that the contribution of the shear stress on pressure loss

stimation, within simulated parameters, is much higher than theontrary effect arising from fluid passing through lumen walls.

effective permeability coefficient, keff , for small curvature (De < 11.6) lumen lengthsand vp = 0.05.

3.2. Effective permeability coefficient, keff

As mentioned in Section 2.1, the effective permeability coeffi-cient is a characteristic parameter of the lumen and defines theability of the fluid to pass through its walls. It is a key parame-ter in the presented modeling approach, and represents differentmechanical properties (as well as possible bio-chemical reac-tive contributions) that affect the lumen walls and their closestsurroundings. Hence, definition of this parameter must be well-defined and objective; therefore, leave it as a subject of futureresearch works. The goal of the parametric analysis in the studyis to present an example of its contribution to the properties of thelaminar particulate-laden flow in lumen. The next analysis will usecharacteristics of case I in Table 1 to simulate flow.

Fig. 4 compares the different values of effective permeabil-ity coefficients and the affected properties of particle-laden fluidmovement in small curvature lumen. It is clearly seen in the fig-ure that for cases where lumen length is greater than 2000�m andthe range of the effective permeability coefficients is 1 · 10−19 −1 · 10−17, the variation of the pressure loss curves cannot beneglected. Moreover, for the entire range of the lumen lengths, a

relatively high effective permeability coefficient (keff ≥ 5 · 10 )has a large effect on the flow rate prediction. However, small val-ues of the coefficient (keff = 1 · 10−19) results in very low effective
Page 5: Mechanics Research Communications

36 S. Shaul, T.I. Zohdi / Mechanics Research

Fl

prflmobmpl

3

inoTfll

pDccD

ig. 5. A comparison of: (a) – pressure loss, and (b) – flow rate as a function of lumenength for different Dean numbers, De.

ermeability of the lumen walls, which does not change the flowate (see Fig. 4(b)). It is shown that for an analyzed particulate-ladenuid, the different values of the effective permeability coefficientsay dramatically change the flow rate predictions and the shape

f the pressure loss (characterized as linear or parabolic). On thisasis, it can be concluded that the accuracy of the effective per-eability coefficient, or its variation, plays an important role in

rediction of the pressure and flow rate values, especially for largeumen lengths.

.3. Curviness of the lumina

The possible curviness of the lumen geometry may stronglynfluence the particle-laden flow, even if the flow regime is lami-ar (Re < 2000). In order to present an example of the contributionf the curvature, at varying Dean numbers, physical parameters inable 1 -case II are used. Fig. 5 presents the results of pressure andow rate losses as functions of the Dean number and the lumen

ength.As seen in Fig. 5(a), there is a parabolic relationship between the

ressure loss and the lumen length within a full range of analyzed

ean numbers. In addition, the parabolic shape of the pressure lossurve changes with respect to the Dean number. This relationshipannot be neglected for the whole range of lumen length wheree > 100. However, for De < 100 flow rate is almost independent

Communications 76 (2016) 32–40

of the lumen curvature (see Fig. 5(b)). Also, for this range of Deannumbers, the flow rate can be assumed to have a union dependencewith the respect to lumen length. Otherwise for a relatively largerange of curvature values (where De > 1000) the flow rate of theparticle-laden fluid has a parabolic dependency with lumen lengthand has a moderate decrease with increasing length. This happensdue to the increase in the wall friction coefficient, which is due tothe kinematic energy loss of the fluid mainstream from the effectof the secondary flow interruption.

4. Case study

4.1. Problem definition

The modeling approach developed in this study is applied topresent the fluid dynamics analysis of the flow inside a typicalhuman Juxtamedullary nephron. According to the literature [30],each human kidney contains an average of 1,000,000 nephrons.Based on nephron depth of the penetration into medulla, thenephron is characterized as two possible types: Cortical or Jux-tamedullary. The Cortical nephrons are short, lie in the cortex ofthe kidney, and form about 80–85% of the kidney nephrons. TheJuxtamedullary nephrons have their Bowman’s capsule very closeto the function of the cortex and the medulla, and have long loopsof Henle extending deep into the medulla.

Unfortunately, there is no available information in the litera-ture regarding comprehensive quantitative analysis of the urineproperties specifically for Juxtamedullary type of nephrons. Conse-quently, it will be assumed that the quantitative properties of urineare characteristic for both types of nephrons. Since there are no twoidentical sized nephrons in the human body, the flow properties ofJuxtamedullary nephron will use available mean values providedin the literature.

It is well known that each day about 180 L of glomerular filtratefluid is formed in adult human kidneys from the blood [31]. Underthe net filtration pressure condition, which is the difference betweenglomerular capillary blood pressure (55mmHg ≈ 7, 333Pa) andplasma-colloid osmotic pressure (30mmHg ≈ 4, 000Pa) in additionto Bowman’s capsule hydrostatic pressure (15mmHg ≈ 2, 000Pa),the glomerular filtrate is passing into the Proximal ConvolutedTubules (PCTs) of nephrons. The normal total glomerular filtrationrate is 125 mL/min. By assuming that the flow rate is divided equallyto all nephrons in both kidneys, the average initial volumetric flowrate of filtrate in each nephron PCT is QPCT ≈ 1.0417 · 10−12, m3/s.If the entire amount of filtrate were excreted as urine, death wouldoccur quickly from dehydration. Actually, only about 1–3 L (whichis about 0.5-1.7% from glomerular filtrate) is excreted each day asurine [31]. The rest is reabsorbed back into the circulatory system.Reabsorption occurs from the filtrate across the tubular lumen ofthe nephron and into the blood of the peritubular capillaries. ThePCTs have a twisted shape and reabsorb about 60–65% of the totalglomerular filtrate [31–33]. From the PCT, the filtrate passes to rel-atively straight Descending Limb of Henle (DLH), which connectsfurther to the ascending limb of Henle with a hairpin shaped twist.There are two segments of the Ascending Limb of the loop of Henle:the thin (tALH) and the thick (TALH). From the thick ascendingsegment of the loop, through the twisted Distal Convoluted Tubule(DCT), almost final urine filtrate passes to the collecting duct tubule.The DCT section of nephron is responsible for the fine adjustmentof urine filtrate by resorption of water and sodium and secretionof hydrogen potassium [34]. For more detailed descriptions of the

human urine system see for example [31,33,35].

Gurm and Kooiman [32] reported that the viscosity of the fil-trate increases significantly when the filtrate fluid passes from onesegment of nephron to another, however, the quantitative values

Page 6: Mechanics Research Communications

S. Shaul, T.I. Zohdi / Mechanics Research Communications 76 (2016) 32–40 37

Table 2Characteristics of fluid and lumen analyzed properties.

Proximal convolutedtubule (PCT)

Descending loop ofHenle (DLH)

Thin ascending loop ofHenle (tALH)

Thick ascending loop ofHenle (TALH)

Distal convolutedtubule (DCT)

Hydraulic diameter,DH, �m

50 20 20 50 50

Length, z, �m 15,000 11,000 4,000 12,000 5000Wall thickness, t, �m 2∗∗ 2∗∗ 2∗∗ 2∗∗ 2∗∗

Effective viscosity,�eff , Pa · s

0.0011∗∗ 0.0028∗∗ 0.0034∗∗ 0.0063∗∗ 0.008

Effective density,�eff , kg/m3

1020 1020 1020 1020 1020

Interstitial fluidpressure, Pif , Pa

800 800 800 800 800

Flow rate, Q · 10−14 ±5 · 10−17, m3/s

42.31∗ − 104.17 2.22∗ − 42.31∗ 1.41∗ − 2.22∗ 1.41∗ − 1.45∗ 1.45∗ − 1.49∗

Average pressure,P ± 0.5, Pa

1921∗ − 2, 000 855∗ − 1921∗ 797∗ − 855∗ 790∗ − 797∗ 786∗ − 790∗

Effective permeability 0.5∗∗ 9.2∗∗ 9.2∗∗ 0.5∗∗ 0.5∗∗

*

fitvv8

otdpotwe

fiahthimtf

4

esr

tl

P

coeff., keff · 10−18, m2

Calculated value; ** assumed value.

or viscosity in different segments are not provided. On this basis,t will be assumed in this example that effective viscosity of the ini-ial urine filtrate is 1.1 · 10−3Pa · s (which is almost equal to plasmaiscosity of the blood [36,37]). In addition, it is well known that theiscosity of the final urine in the human body temperature averages

· 10−3Pa · s [38,39].There is no comprehensive study of the quantitative analysis

f interstitial fluid among different nephron segments. Thus, forhe purposes of expletory calculations in the modeling approacheveloped in this study, it is assumed that the interstitial fluidressure does not vary between the different segments of wallsf Juxtamedullary nephron. This fluid pressure was assumed to behe average interstitial fluid pressure in animals (6mmHg ≈ 800Pa),hich has been experimentally evaluated in the work of Guyton

t al. [40].The last parameter defined is the effective permeability coef-

cient. As it was mentioned in the modeling and parametricalnalysis sections of the study, this coefficient must be defined withigh accuracy with respect to the analyzed problem. Unfortunately,his parameter has not been studied in the different segments ofuman’s nephron lumina. However, some research works available

n the literature [18,20,21,41–44] indicate properties of wall per-eability in the different nephron’s segments, which are useful in

he calculations of the presented example. By comparing the resultsrom literature, the permeability properties of nephron walls are:

Wall permeability within an analyzed biological membrane ornephron section is often found to be a constant property.TALH and DCT nephron’s sections have almost impermeablewalls.The permeability of DLH and tALH sections can be assumed to besimilar and much higher than other nephron sections.

.2. Results and discussion

Fig. 6 presents the results from the analysis of the flow prop-rties in a typical human Juxtamedullary nephron while Table 2ummarizes the set of initialized parameters and calculationesults.

From the results of filtrate flow in PCT lumen, it is observed thathe pressure loss of 3.95% is relatively small compared to the fluid

oss of 59.4%, while the loss ratio of the property defined by:

roperty "X"change, % = Xinitial − Xfinal

Xinitial· 100. (20)

The value of fluid loss is close to 60% and can be compared to theapproximated range of 60–65% (see Section 4.1) as reported in theliterature. From the calculated values corresponding to DLH sectionit is clear that the fluid pressure significantly decreases along thelumen length (more than half), while the fluid loss in this section isabout 94.75%. The pressure of the fluid is slightly decreases in theother sections of nephron: tALH, TALH and DCT, while the flow ratechanges in a different manner from the previous sections. Approx-imately in the middle part of tALH, when the pressure of the fluidmoving inside lumen reduces to the value of interstitial pressure,the fluid flows into the lumen. This process remains active untilthe fluid reaches the collecting duct tubule (via TALH and DCT sec-tions) and is also referred to as filtrate secretion [19,34,42]. It shouldbe noted that due to the relatively small difference between theinterstitial pressure and the pressure in the lumen, the modelingapproach is sufficient in describing the fluid dynamics of the fil-trate in those sections, while neglecting the possible bio-chemicalreactions between the urine filtrate and interstitial fluid. Thus, inthose sections of nephron, the fluid may pass the lumen wallsin both directions, and consequently some fluid exchange mayoccur. Nonetheless, the modeling example results in 98.6% fluidloss from initial fluid filtrate, which is close to an average value of99% reported in the literature [13,30–33,35].

From the presented analysis of Juxtamedullary nephron, it canbe concluded that:

• Since the laminar flow regime of urine filtrate has characteristicsof a creeping flow region, it has been shown that unless the cur-vature of nephron lumen become high (for example, as hairpinpin shape between DLH and tALH), the fluid dynamics propertiesof urine filtrate are not affected.

• The filtrate pressure variation along the lumen length is char-acteristic of each nephron section. It was shown that the highestpressure loss occurs in PCT, while the lowest pressure loss occursin the DCT section. Moreover, the pressure loss in descending limbof Henle is much higher compared with the whole ascending loop,55.5% and 6.78%, respectively.

• The most significant reabsorption of filtrate liquid occurs fromPCT up to the end of DLH sections of nephron (about 97.8%).

• It is possible that filtrate flow rate will neither increase nor

decrease by passing through nephron segments of tALH up tothe end of DCT. These sections have specific lumen characteris-tics, which are able to provide reabsorption and secretion of thefiltrate.
Page 7: Mechanics Research Communications

38 S. Shaul, T.I. Zohdi / Mechanics Research Communications 76 (2016) 32–40

F side ld

5

apeppf

lotlact

attetlieoanm

sJttal

ig. 6. The results from the computation of: (a) the fluid pressure and (b) flow rate iniameter of the lumina increased by half.

. Summary

The study presents a modeling approach for fast analysis, designnd prediction of laminar particulate-laden flow in lumen withermeable walls. The presented semi-analytic modeling approachnables researchers and designers to rapidly model, analyze andredict the behavior of laminar particle-laden flows in lumen withermeable walls for various biological applications as dictated byuture research needs.

First, the study developed a theoretical modeling approach ofaminar flow in a lumen with permeable walls. The possible effectf lumen curvature on the flow characteristics are presented usinghe Dean number. In addition, a method of examining the effect ofumen curvature on flow properties of a particle-laden fluid wasnalyzed. Also a case of De < 1 was reviewed and it was found thatharacteristic to the creeping flow region of the laminar flow regime,he curvature of the lumen becomes negligible.

Second, the study contains a parametric analysis for the key vari-bles used in the modeling approach. From the results, it is observedhat as the volume fraction of the particles of the fluid increases,he pressure loss also increases while the flow rate decreases. Theffective permeability coefficient plays an important role in predic-ion of the pressure and flow rate losses, especially for relativelyarge lumen lengths. Therefore, this coefficient must be exam-ned and defined appropriately for future research problems tonsure high accuracy. The possible cases of how different degreesf lumen curvature affect the flow properties of particle-laden fluidre reviewed. For this case, it is shown that the range of Deanumbers is De > 1000 and De < 100 where the curvature becomesore or less contributive.Finally, an example for using the modeling approach of this

tudy to examine the fluid dynamics properties in a typical humanuxtamedullary nephron is presented. The major achievement ofhis analysis is to present particle-laden fluid dynamics charac-eristic analysis for different sections of nephron lumina. A good

greement of calculated results with those presented in medicaliterature is found.

Fig. A1. Fully developed velocity profile in cha

umen of human Juxtamedullary nephron. *For illustration purpose only, the hydraulic

Appendix A.

A.1 Velocity profile and shear stress of the fluid

It is possible, analytically, to examine the shape of a fully devel-oped velocity profile for the above problem by an approach similarto that used to define a thermally fully developed condition. For athermally fully developed region, it can be shown that:

d

dz[Tw(z) − T(r, z)

Tw(z) − T(z)] = 0, (A1)

where r is the radial coordinate, z is the axial coordinate, T is thetemperature, Tw is the wall surface temperature, and T is the meantemperature for a particular cross-section. Analogously, by replac-ing the temperature in Eq. (A1) with axial velocity, u(r, z) and meanaxial velocity for a particular cross-section, U, results in:

d

dz[uw(z) − u(r, z)

uw(z) − U(z)] = 0. (A2)

Applying the no-slip boundary condition, uw = 0, gives:

∂u(r, z)∂z

= 0. (A3)

where u(r, z) is the normalized axial velocity defined by:

u(r, z) = u(r, z)

U(z). (A4)

Therefore, when the normalized axial velocity u(r, z) is constant(see Eq. (A3)), the flow in the porous channel is fully developed.This effectively means that the shape of the axial velocity profileis the same for every cross-section. For the idealized constantlypermeable channel with a circular cross-section of area,A = �R2, alaminar velocity profile can be given by a classical channel-flow ofthe form [27,43,45]:

u(r, z) = umax,i(1 − (r

R)2), (A5)

where umax,i is the centerline velocity for given cross-section i (seeFig. A1).

nnel with a constant radial permeability.

Page 8: Mechanics Research Communications

earch

w

U

oftz

A

ea[wteoIo

v

i

� p)

0)

wc

wasfoZ

yat

R

[

[

[

[[

[

[

[

[

[

[

[

[

[

[

[[

[

[

[

[[

[

[

[

[

[

S. Shaul, T.I. Zohdi / Mechanics Res

By substituting Eq. (A5) in Eq. (7), the shear stress of the fluidith perforated channel walls in i cross-section is:

w,i = −�eff2umax,i

R, (A6)

while the mean axial velocity is:

¯ i = 1A

∫ ∫u(r, z)dA = umax,i

2, (A7)

Substituting Eq. (A7) in (A6) results in:

w,i = −�eff4Ui

R. (A8)

It should be pointed that using Eq. (A8) to estimate shear stressf the fluid with constantly permeable channel walls is reasonableor the negligible difference of Ui . In this study it was assumedhat it is valid for the cases of laminar flow where �z = zi+1 − zi =i − zi−1 � 2R.

.2 Effective properties of particle-laden fluids

A flow regime that is thoroughly mixed can usually be mod-led as a single-phase flow by defining its effective properties suchs: density, viscosity, etc. [46–48]. For example, the work of Wallis46] presents an analysis of homogenous flow friction correlationshich is based on effective flow parameters for both laminar and

urbulent flow regimes. The recent study of Zohdi [48] provides anxample of using effective properties of particle-laden fluid flows inrder to estimate the pumping pressures for different applications.n this study, the effective properties were presented as a functionf volume fraction of fluid, vf , and particles, vp, such that:

f + vp = 1. (A9)

The effective density of the mixture, defined as volume averagesn the z direction is:

eff ≡< �(z) >V ≡ 1V

∫ ∫ ∫�(z)dV = 1

V(

∫ ∫ ∫�f dVf +

∫ ∫ ∫�pdV

= vf �f + vp�p. (A1

To estimate the effective viscosity of the particle-laden fluidithin extremely low volume fraction of particles (below one per-

ent), the expression developed by Einstein [49] is used:

eff = �f (1 + 2.5vp), (A11)

here �f is the viscosity of the surrounding incompressible fluidnd the particles are assumed to be rigid. More empirical expres-ions for effective viscosity are typically used at higher volumeractions [50]. As an example, for the cases where volume fractionf the particles in the fluid can reach a value of 0.25, the work byohdi [48] is used:

eff = �f (1 + 2.5vp

1 − vp). (A12)

The expression presented in Eq. (A12) is derived from the anal-sis of the isotropic material responses in the well-known Hashinnd Shtrikman [51,52] works, where the lower bound of the effec-ive response was taken into consideration.

eferences

[1] A.A. Lazakidou, 2010. Biocomputation and Biomedical Informatics: CaseStudies and Applications Medical Information Science Reference.

[2] G. Janiga, C. Rössl, M. Skalej, D. Thévenin, Realistic virtual intracranial stentingand computational fluid dynamics for treatment analysis, J. Biomech. 46(2013) 7–12.

[

[

Communications 76 (2016) 32–40 39

[3] a.K. Niemann, J. Udesen, S. Thrysoe, J.V. Nygaard, E.T. Fründ, S.E. Petersen,et al., Can sites prone to flow induced vascular complications in a-v fistulas beassessed using computational fluid dynamics? J. Biomech. 43 (2010)2002–2009.

[4] D.J. Quam, T.J. Gundert, L. Ellwein, C.E. Larkee, P. Hayden, R.Q. Migrino, et al.,Immersive visualization for enhanced computational fluid dynamics analysis,J. Biomech. Eng. 137 (2015) 31004.

[5] F. Levi, E. Mosekilde, D.a. Rand, Advancing systems medicine and therapeuticsthrough biosimulation, Interface Focus 1 (2011) 3–6.

[6] D. Evans, J. Jensen, M. Nielsen, Ultrasonic colour doppler imaging, InterfaceFocus 1 (2011) 490–502, http://dx.doi.org/10.1098/rsfs.2011.0017.

[7] V.V. Gokhale, R.I. Tanner, B.K. Bischoff, Finite element solution of theNavier-Stokes equations for two dimensional steady flow through a section ofa canie aorta model, J. Biomech. 11 (1978) 241–249.

[8] P.M. McGah, D.F. Leotta, K.W. Beach, J.J. Riley, A. Aliseda, A longitudinal studyof remodeling in a revised peripheral artery bypass graft using 3D ultrasoundimaging and computational hemodynamics, J. Biomech. Eng. 133 (2011)41008.

[9] D. Groen, J. Borgdorff, C. Bona-Casas, J. Hetherington, R.W. Nash, S.J. Zasada,et al., Flexible composition and execution of high performance, high fidelitymultiscale biomedical simulations, Interface Focus 3 (2013) 20120087.

10] Y. He, C.M. Terry, C. Nguyen, S.a. Berceli, Y.T.E. Shiu, A.K. Cheung, Serialanalysis of lumen geometry and hemodynamics in human arteriovenousfistula for hemodialysis using magnetic resonance imaging andcomputational fluid dynamics, J. Biomech. 46 (2013) 165–169.

11] C.a. Taylor, T.a. Fonte, J.K. Min, Computational fluid dynamics applied tocardiac computed tomography for noninvasive quantification of fractionalflow reserve: scientific basis, J. Am. Coll. Cardiol. 61 (2013) 2233–2241, http://dx.doi.org/10.1016/j.jacc.2012.11.083.

12] T.I. Zohdi, F. a Kuypers, Modelling and rapid simulation of multiple red bloodcell light scattering, J. R. Soc. Interface 3 (2006) 823–831.

13] L.R. Johnson, Essential Medical Physiology, 3rd ed., Elsevier, 2003.14] E.N. Marieb, P.B. Wilhelm, J.B. Mallatt, Human Anatomy, 7th ed., Pearson,

2013.15] S. Oka, T. Murata, A theoretical study of the flow of blood in a capillary with

permeable wall, Jpn. J. Appl. Phys. 9 (1970) 345–352.16] N.K. Marimma, S.N. Majhi, Flow of a Newtonian fluid in a blood vessel with

permeable wall – a theoretical model, Comput. Math. 40 (2000) 1419–1432.17] C. Pozrikidis, Stokes flow through a permeable tube, Arch. Appl. Mech. 80

(2010) 323–333.18] A. Verniory, R. Du Bois, P. Decoodt, J.P. Gassee, P.P. Lambert, Measurement of

the permeability of biological membranes. Application to the glomerular wall,J. Gen. Physiol. 62 (1973) 489–507.

19] Y. Miyamoto, H. Yuasa, T. Iga, M. Hanano, Determination of the membranepermeability coefficient and the reflection coefficient by the two-dimensionallaminar flow model for intestinal perfusion experiments, Biochim. Biophys.Acta. 854 (1986) 191–197.

20] Y. Wang, P.E. Tomlins, A.G. a Coombes, M. Rides, On the determination ofDarcy permeability coefficients for a microporous tissue scaffold, Tissue Eng.Part C Methods 16 (2010) 281–289.

21] J.P. Kokko, C.C. Tisher, Water movement across nephron segments involvedwith the countercurrent multiplication system, Kidney Int. 10 (1976) 64–81.

22] D.R. Hagen, J.K. White, B. Tidor, Convergence in parameters and predictionsusing computational experimental design, Interface Focus 3 (2013) 20130008.

23] C.M. White, Streamline flow through curved pipe, Proc. R. Soc. A123 (1929)645–663.

24] H. Ito, Friction factors for turbulent flow in curved pipes, Trans. ASME J. BasicEng. 81D (1959) 123–134.

25] H. Ito, Laminar flow in curved pipes, Z. Angew Math Mech. 49 (1969) 653–663.26] G. Raithby, Laminar heat transfer in the thermal entrance region of circular

tubes and two-dimensional rectangular ducts with wall suction and injection,Int. J. Heat Mass Transf. 14 (1971) 223–243.

27] B. Atkinson, M.P. Brocklebank, C.C.H. Card, J.M. Smith, Low Reynolds numberdeveloping flows, AIChE J. 15 (1969) 548–553.

28] J. Happel, H. Brenner, Low Reynolds Number Hydrodynamics, MartinusNijhoff Publishers, 1983.

29] K. Sangtae, J.K. Seppo, Microhydrodynamics: Principles Selected Applications,Butterworth-Heinemann, 1991.

30] W.W. McCrory, Developmental Nephrology, Harvard University Press, 1972.31] D.D. Ignatavicius, M.L. Workman, Medical-Surgical Nursing: Patient-Centered

Collaborative Care, Elsevier/Saunders, Philadelphia, 2013.32] H.S. Gurm, J. Kooiman, Renal complications in the catheterization laboratory,

in: Interv Cardiol. Clin., Elsevier B.V, 2014, pp. 394–403.33] C.J. Lote, Principles of Renal Physiology, 4th ed., Kluwer Academic Publishers,

2000.34] A.R. Subramanya, D.H. Ellison, Distal convoluted tubule, Clin. J. Am. Soc.

Nephrol. (2014) 1–17.35] N.A. Brunzel, Fundamentals of Urine and Body Fluid Analysis, 3rd ed., Elsevier,

2013.36] A. Singh, K.U. Eckardt, A. Zimmermann, K.H. Götz, M. Hamann, P.J. Ratcliffe,

et al., Increased plasma viscosity as a reason for inappropriate erythropoietin

formation, J. Clin. Invest. 91 (1993) 251–256.

37] G. Késmárky, P. Kenyeres, M. Rábai, K. Tóth, Plasma viscosity: a forgottenvariable, Clin. Hemorheol. Microcirc. 39 (2008) 243–246.

38] D.F. Putnam, 1971. Composition and concentrative properties of human urine,Washington.

Page 9: Mechanics Research Communications

4 earch

[

[

[

[

[

[[

[[

[

[

[Springer, 2013.

0 S. Shaul, T.I. Zohdi / Mechanics Res

39] B.a. Inman, W. Etienne, R. Rubin, R.a. Owusu, T.R. Oliveira, D.B. Rodriques,et al., The impact of temperature and urinary constituents on urine viscosityand its relevance to bladder hyperthermia treatment, Int. J. Hyperth. 29(2013) 206–210.

40] A.C. Guyton, J. Prother, K. Scheel, J. Mcgehee, Interstitial fluid pressure, Circ.Res. 19 (1966) 1022–1030.

41] C. Crone, The permeability of capillaries in various organs as determined byuse of the indicator diffusion method, Acta Physiol. Scand. 58 (1963) 292–305.

42] M.B. Burg, Thick ascending limb of Henle’s loop, Kidney Int. 22 (1982)454–464.

43] V. Nassehi, Modelling of combined Navier-Stokes and Darcy flows incrossflow membrane filtration, Chem. Eng. Sci. 53 (1998) 1253–1265.

44] S. Coppock, H. Brady, MCAT Biology Review, Random House, 2010.45] T. Clemo, Flow in perforated pipes: a comparison of models and experiments,

SPE Prod. Oper. (2006) 302–311.

[

[

Communications 76 (2016) 32–40

46] G.B. Wallis, One-dimensional Two-phase Flow, McGraw-Hill, New York, 1969.47] T.I. Zohdi, Mechanically driven accumulation of microscale material at

coupled solid – fluid interfaces in biological channels, J. R. Soc. Interface 11(2014) 2–7.

48] T.I. Zohdi, On necessary pumping pressures for industrial process-drivenparticle-laden fluid flows, ASME J. Manuf. Sci. Eng. (2015) 1–14.

49] A. Einstein, A new determination of molecular dimensions, Ann. Phys. 19(1906) 289–306.

50] M. Kachanov, I. Sevostianov, Effective Properties of Heterogeneous Materials,

51] Z. Hashin, S. Shtrikman, On some variational principles in anisotropic andnonhomogeneous elasticity, J. Mech. Phys. Solids 10 (1962) 335–342.

52] Z. Hashin, S. Shtrikman, A variational approach to the theory of the elasticbehaviour of multiphase materials, J. Mech. Phys. Solids 11 (1963) 127–140.