Skip to main content
Skip to article control options
Open AccessRegular Article

Viscous Mean Flow Approximations for Porous Tubes with Radially Regressing Walls

Published Online:https://doi.org/10.2514/1.J055949

Abstract

The mean gaseous motion in solid rocket motors has been traditionally described using an inviscid solution in a porous tube of fixed radius and uniform wall injection. This model, usually referred to as the Taylor–Culick profile, consists of a rotational solution that captures the bulk gaseous motion in a frictionless rocket chamber. In practice, however, the port radius increases as the propellant burns, thus leading to time-dependent effects on the mean flow. This work considers the related problem in the context of viscous motion in a porous tube and allows the radius to be time dependent. By implementing a similarity transformation in space and time, the incompressible Navier–Stokes equations are first reduced to a nonlinear fourth-order ordinary differential equation with four boundary conditions. This equation is then solved both numerically and asymptotically, using the injection Reynolds number Re and the dimensionless wall regression ratio α as primary and secondary perturbation parameters. In this manner, closed-form analytical solutions are obtained for both large and small Reynolds number with small-to-moderate α. The resulting approximations are then compared with the numerical solution obtained for an equivalent third-order ordinary differential equation in which both shooting and the irregular limit that affects the fourth-order formulation are circumvented. This code is found to be capable of producing the stable solutions for this problem over a wide range of Reynolds numbers and wall regression ratios.

Nomenclature
a

time-dependent radius

a˙

wall regression speed

b

scaling factor

F

characteristic mean flow function

G

transformed mean flow function

K

characteristic integration constant, K(Re)

nmax

total number of points in the integration domain

np

point where G(np)=0

nTaylor

number of points for which the Taylor series expansion is applied

Re

Reynolds number

r

radial coordinate

S

surface

t

time

Um

mean axial velocity

u

velocity vector

V

absolute injection velocity as seen by a stationary observer

V

volume

Vw

injection velocity with respect to the moving wall

z

axial coordinate

α

wall expansion ratio

β

scaled expansion ratio, α/b

γi

ith coefficient in the Taylor series expansion

ε

small real number used in numerical integration

ϵ

small perturbation parameter

η

radial transformation variable, (1/2)r2

λ

scaling factor, 2/Re

ν

kinematic viscosity

ξ

transformation variable, bη

ξmax

domain integration size

ξp

coordinate location corresponding to np

ρ

density

ψ

Stokes stream function

Ω

vorticity vector

I. Introduction

Viscous motion in cylindrical chambers with sidewall injection is of interest in a variety of applications, including mean flow modeling of solid [13] and hybrid rockets [4,5], sweat cooling [6,7], boundary-layer control [810], peristaltic pumping [11,12], gaseous diffusion, and isotope separation [1315]. It is the latter group of studies by Berman [13,14] that has actually provided the impetus to develop the first similarity transformation of the Navier–Stokes equations into a fourth-order nonlinear ordinary differential equation (ODE). The resulting ODE was later solved asymptotically over different ranges of the wall injection Reynolds number. Most notable examples include those by Yuan and Finkelstein [16], Terrill and Thomas [17], Terrill [18,19], Skalak and Wang [20], and others.

In the propulsion area, an inviscid rotational counterpart known as the Taylor [21], Culick [22], or Taylor–Culick solution occupies the central stage in modeling the bulk gaseous motion inside solid rocket motors (SRMs). This may be owed to its association with several studies involving hydrodynamic instability [2328], acoustic instability [2935], wave propagation [3639], particle-mean flow interactions [40], and rocket performance measurements [4143]. The Taylor–Culick solution was originally verified to be an adequate representation of the expected flowfield in SRMs both numerically by Sabnis et al. [44] and experimentally by Dunlap et al. [45,46], thereby confirming its character in a nonreactive chamber environment. It was extended by Majdalani and Akiki [4] to include effects of viscosity and headwall injection, by Saad et al. [42] and Sams et al. [43] to account for wall taper, by Kurdyumov [47] to capture effects of irregular cross sections, and by Majdalani and Saad [48] to allow for arbitrary headwall injection. Then, using variational calculus and the Lagrangian optimization principle, Saad and Majdalani [49] uncovered a continuous spectrum of Taylor-like solutions exhibiting increasing kinetic energy signatures, while ranging from the traditional Culick profile down to its predecessor, the irrotational mean flow known as the Hart–McClure profile [50,51]. This potential mean flow preceded the use of Culick’s model in several fundamental investigations of combustion instability [5254]. However, these studies have not been concerned with the time dependence that is inherent to the grain surface of a burning propellant.

In a practical analysis, the radius of a rocket motor is allowed to increase during wall regression, and the corresponding problem gives rise to a partial differential equation (PDE) in lieu of an ODE. Although the problem in a tube with expanding or contracting surfaces was treated by Uchida and Aoki [12] in the late 1970s, the effects of an injecting sidewall were incorporated more than a decade later by Goto and Uchida [55], in the context of a pulsating porous tube and by Majdalani et al. [1,2] in the context of a rocket chamber. In their work, the effect of wall regression was prescribed by a dimensionless wall expansion ratio, α, written as a viscous Reynolds number based on the radial regression speed of the sidewall. In the interim, Dauenhauer and Majdalani [5658] formulated the corresponding equation and numerical solution for the planar flow analog. The analytical solutions promoted through these efforts would later receive attention in follow-up studies aimed at devising asymptotic approximations over various ranges of the control parameters or at reconstructing the solution using alternative techniques such as the Lie-group theory [59,60] or the Homotopy-Analysis Method [6163].

In this article, we revisit the problem of viscous motion in a porous tube and allow the radius to transiently expand or contract [1,2]. As before, we employ a dual similarity transformation in space and time to reduce the Navier–Stokes equations into a nonlinear fourth-order ODE that can be solved both asymptotically and numerically. After reconstructing the analytical solution of Majdalani et al. [1,2] for large injection, we employ two perturbation parameters, the injection Reynolds number Re and the wall expansion ratio α to formulate an asymptotic series for the case of small Reynolds number and small-to-moderate α. We then introduce an efficient numerical approach that overcomes the singularity encountered in this model, specifically in the form of an intrinsically satisfied limit at the origin. To make headway, an equivalent third-order ODE is presented and solved using a technique that may be attributed to Terrill and Thomas [17]. Subsequent numerical predictions are obtained directly, with no shooting or iteration, and then compared with our analytical approximations. In this manner, the robustness of the numerical algorithm is used to demonstrate that the present approach can capture the stable solutions for this problem over a wide range of Reynolds number and α. Although the case of large Reynolds number is connected with modeling solid rocket motors, the small injection case is relevant to other problems involving sweat cooling, filtration, peristaltic pumping, and boundary-layer control.

The paper is organized as follows. First, we consider the mathematical model developed by Majdalani et al. [1,2] and reaffirm its validity. Second, we discuss the numerical technique that appears to be most practical. Third, we produce regular perturbation approximations for the large and small injection cases, the latter of which being are presented for the first time. Finally, we compare asymptotics and numerics over a common range of the control parameters.

II. Mathematical Model

The cylindrical propellant grain of a solid rocket motor is modeled as a long tube with one end closed at the headwall and the other end open. The circumference of the wall is assumed to be permeable so as to simulate propellant burning and normal gas injection. Furthermore, the wet area of the tube is allowed to radially expand at a speed equal to a˙. For this to occur, the mathematical model requires that the headwall area stretches accordingly to accommodate the expansion of the tube. This behavior is shown in Fig. 1, where an axisymmetric coordinate system is defined. By assuming an incompressible mean flow, the vorticity transport equation is written as

Ω*t*×(u*×Ω*)=ν*×*2u*(1)

Fig. 1
Fig. 1

Schematic of a cylindrical chamber used to illustrate sidewall injection and wall regression, as well as the control volume used to evaluate the average axial velocity.

A. Boundary Conditions

These can be organized as follows:

r*=a(t):{ur*=Vsidewallinjectionuz*=0noslip(2)
r*=0:{ur*=0noflowacrossthecenterlineuz*r*=0axialvelocitysymmetry(3)
z*=0:uz*=0inertheadwall(4)
Here, V is the absolute fluid injection velocity at the wall. It is defined as the fluid velocity seen by an observer in a reference frame located outside of the tube. If Vw is used to denote the fluid velocity with respect to the wall, then the following relation holds:
V=Vwa˙(5)

Evidently, when the walls are stationary, the absolute and relative velocities become equal. But when the wall undergoes inward contraction, then a˙<0 and the fluid will appear as though it is being injected at a larger speed than Vw. Conversely, if the wall regresses outwardly, then a˙>0 and the fluid injection velocity will appear to be smaller than Vw.

Note that, at z*=0, we do not enforce any condition on the radial velocity profile. This slip condition on the radial velocity is what actually allows us to solve this problem by assuming a linear relation between the stream function and the axial coordinate z* [see Eq. (14)]. In reality, a small, rather negligible viscous layer develops at the headwall, whose analysis is described by Chedevergne et al. [24].

B. Similarity in Space

By inspection, one expects the axial velocity to vary linearly in z*. To show this, we follow Majdalani et al. [13] by considering the control volume delineated by the dashed lines in Fig. 1. Through a mass balance on the inlet and outlet surfaces, we require

Vt+Sexitu*ndS2πVwaz*=0(6)

The term containing the integral in Eq. (6) may be connected to the average axial velocity Um inside the volume using

Um=1πa2Sexitu*ndS(7)

Its substitution into Eq. (6) yields

Vt+πa2Um2πVwaz*=0(8)
and so, by using V=πa2z*, Eq. (8) reduces to
2πaa˙z*+πa2Um2πVwaz*=0(9)

Thus, based on mass conservation alone, we must have

Um=2(Vwa˙)z*a=2Vz*a(10)
which must be equated to the area integral for the average velocity
Um(z*,t)=1πa20auz*(r*,z*,t)r*dr*=2Vz*a(11)
which proves that the only admissible form of the axial velocity is a linear relation of the type
uz*(r*,z*,t)=z*f(r*,t)(12)

C. Vorticity Transport Equation

We introduce the Stokes stream function using

ur*=1r*ψ*z*;uz*=1r*ψ*r*(13)

Then, based on Eq. (12), the stream function may be written as

ψ*=νz*F(r,t);r=r*a(14)
where ν denotes the kinematic viscosity for a dimensionless F(r,t). The radial and axial velocities can now be expressed in terms of F. This operation yields
ur*=νarF,uz*=νz*a2rFr(15)

Because ur* remains independent of the axial coordinate, the vorticity displays a single nonzero component in the tangential direction, viz.,

Ω*=Ωθ*=ur*z*uz*r*=uz*r*(16)

Upon substitution of Eq. (16) into Eq. (1), one recovers

r*(uz*t)+r*(ur*uz*r*)+r*(uz*uz*z*)νr*[1r*r*(r*uz*r*)]=0(17)

By paying careful attention to partial differentiation, Eq. (17) is reduced to

{a2νFrtr+aa˙νr(Frr)r+2aa˙νFrr+Fr(Frr)r(Frr)2+1r(Frr)r(Frr)rr}r=0(18)
where the subscripts r and t denote partial differentiation with respect to r and t, respectively. At this juncture, we define the wall expansion ratio
α(t)aa˙ν(19)
and convert Eq. (18) into
{(Frr)rr+(1r+Fr+αr)(Frr)r(Frr2α)Frra2νFrtr}r=0(20)

The outcome can be readily integrated with respect to r and rearranged into

(Frr)rr+(1r+Fr+αr)(Frr)r(Frr2α)Frra2νFrtr=K(21)

The formidable PDE that we arrive at embodies the physics of viscous motion in a porous pipe with retracting walls. It is related to work by Uchida and Aoki [12], who studied a similar problem in the context of pipe flow with expanding or contracting impermeable walls. Goto and Uchida [55], Dauenhauer and Majdalani [5658], and Majdalani et al. [1,2] later extended Uchida and Aoki [12] to permit fluid injection or suction along the moving sidewall. Their analysis was carried out for both planar [5658] and cylindrical configurations [1,2]. Majdalani et al. [13] also derived asymptotic solutions for this equation over a practical range of Reynolds numbers and expansion ratios.

D. Similarity in Time

At first glance, Eq. (21) may seem intractable using any of the standard analytical techniques. This excludes the Homotopy Analysis Method, which has been shown to handle highly nonlinear equations quite admirably [61,6466]. By using a practical hypothesis that may be traced back to Uchida and Aoki [12], Eq. (21) may be reduced to a third-order nonlinear ODE. As shown by Majdalani et al. [1], this is accomplished by setting

α(t)=constant;F(r,t)F[r,α(t)](22)

The constancy of α ensures that

Frt=Frαdαdt=0(23)

One can then deduce the required wall regression speed for which α becomes a constant. According to this model

α=aa˙ν=a0a˙0ν(24)
where a0 and a˙0 correspond to the initial radius and regression speed, respectively. Upon integration, we get
a(t)=a01+2ναa02t(25)

Note that this model does not suppress time dependence, but rather embeds it within the solution implicitly. This is realized by specifying the regression speed in such a way that α(t) remains constant.

Before substituting the time similarity conditions into Eq. (21), we find it useful to introduce the following normalizations:

FFRe;η12r2(26)
where Re is the sidewall injection Reynolds number based on the absolute velocity of the fluid:
ReVaν=Vwaνα(27)

In this setting, the Reynolds number remains a function of time. To make further headway, we assume that Re=constant or

Vwaν=Vw0a0ν(28)

This requirement enables us to extract the proper form of Vw in our model, namely,

Vw=Vw0a0a=Vw0(1+2ναa02t)(1/2)(29)

Finally, backward substitution of Eq. (26) into Eq. (21) yields

ηF+F+12Re(FFF2)+α(ηF+F)=K(Re)(30)
where primes denote differentiation with respect to η. To eliminate the constant K(Re), one differentiates Eq. (30) with respect to η and writes
ηF+α(ηF+2F)+12Re(FFFF)+2F=0(31)

Finally, the boundary conditions may be assimilated into the following set

dF(1/2)dη=0(32a)
F(12)=1(32b)
F(0)=0(32c)
limη02ηd2Fdη2=0(32d)

Note that the constant K(Re) may be specified by evaluating Eq. (30) at any location inside the domain, such as the centerline. At r=0, one gets

K(Re)=F(0)+12ReF2(0)+αF(0)(33)

Either of the differential equations given by Eqs. (30) and (31) may be used to solve for the mean flow function F(η). The choice depends on the type of solution sought. For example, we elect to use Eq. (30) for the numerical solution because it requires less memory storage and fewer Runge–Kutta integrations. On the other hand, Eq. (31) will be used to initiate the regular perturbation expansion because of the implicit nonlinearity that the constant K(Re) embodies.

III. Numerical Technique

Being nonlinear ODEs, Eqs. (30) and (31) may be solved directly through a Runge–Kutta integration routine. However, a close inspection of the boundary conditions as well as the governing equations reveals several difficulties. First, the explicit presence of η in these equations requires a careful treatment by virtue of the singularity at the origin. This matter may be settled through a Taylor series expansion near the centerline. Second, the boundary conditions given by Eq. (32) result in a boundary value problem that requires a double infinity of integrations [17]. This problem may be overcome via a shooting method based on a Newton–Raphson root finding algorithm, such as the one proposed by Dauenhauer and Majdalani [58]; however, the fourth boundary condition in Eq. (32d) is difficult to implement. The reason is that, as long as F(0) remains finite, Eq. (32d) will be automatically satisfied. Indeed, one expects F(0) to be finite because it corresponds to the vorticity at the centerline. In short, we find Eq. (32d) to be impractical for numerical treatment.

The aforementioned issues may be overcome if one uses a transformation that is nearly identical to that employed by Terrill and Thomas [17]. These researchers encountered difficulties in an analogous problem involving flow through a porous tube. In our model, however, one needs to incorporate the effect of α. By allowing both the Reynolds number and α to be determined a posteriori (i.e., at the end of integration), a shooting method will no longer be required and the numerical solution may be arrived at directly, using a single Runge–Kutta integration. To illustrate this procedure, we begin by introducing the dual transformation

F=λG(ξ);ξ=bη(34)
where both λ and b are scaling factors that may be determined once the numerical integration is complete. Upon substitution into Eq. (30), we recover a third-order ODE in G(ξ), namely,
ξG+G+12λRe(GGG2)+αb(ξG+G)=Kλb2=K1(35)
where the primes denote differentiation with respect to ξ. If we further set
λ=2Reandβ=αb(36)
then Eq. (35) reduces to the convenient form
ξG+G+GGG2+β(ξG+G)=K1(37)

Finally, by applying this procedure to the conditions encapsulated in Eq. (32), they transform into

dG(b/2)dξ=0(38a)
G(12b)=1λ=Re2(38b)
G(0)=0(38c)
limξ0ξd2Gdξ2=0(38d)

By initializing the solution with arbitrary values for G(0), G(0), and β, the integration may be carried out until G(b/2)=0. In practice, to produce a desired Reynolds number and wall expansion ratio, these are chosen based on the asymptotic solutions (to be derived in the next section). The point at which G vanishes enables us to explicitly deduce b with no need for iteration. Subsequently, the Reynolds number and α may be calculated from Eq. (38b) as Re=2G(b/2) and α=βb. The constant K1 can also be extracted from the initial guesses by putting

K1=G(0)G(0)2+βG(0)(39)

Finally, near centerline (ξ0), a Taylor series expansion of the governing equation is required. Using

G(ξ)=i=0γiξi(40)
substitution into Eq. (37) yields
γ0=G(0)=0;γ1=G(0);γ2=12G(0)(41)

Note that these coefficients are known because the values of the first and second derivatives at the centerline are used to seed the numerical solution. Then, by substituting Eq. (40) into Eq. (30), coefficients of the same order may be equated to the extent of establishing a recurrence formula for the coefficients of the Taylor series expansion; we get

(n+2)(n+1)2γn+2=j=0n(j+1)(nj+1)γj+1γnj+1j=0n(j+1)(j+2)γj+2γnjβ(n+1)2γn+1;n1(42)

Algorithm 1 lists the necessary steps for our numerical implementation. Results of the numerical simulation will be presented in subsequent sections, where they will be compared with the asymptotic approximations.

IV. Asymptotic Treatment

A. Large Injection

In this case, which is appropriate for modeling gaseous injection in solid rocket motors, the Reynolds number is sufficiently large to warrant a regular perturbation in the inverse of the Reynolds number. We follow Majdalani et al. [1] and write

F=F0+ϵF1+O(ϵ2);ϵ1Re(43)

Upon substitution into Eq. (31), we recover the ODEs for the leading and first-order solutions. These are

O(0):F0F0F0F0=0(44a)
O(1):F0F1+F1F0F0F1F1F0+2ηF0+(2αη+4)F0+4αF0=0(44b)

Note that we limit our analysis to the leading and first-order approximations partly due to the excellent agreement that we later show between the present solution and numerical calculations.

1. Leading-Order Solution

At leading order, we have

F0F0F0F0=0(45)

This can be solved by inspection to obtain

F0=sin(πη)=sinθ;θ=πη(46)

Equation (46) corresponds to the Taylor–Culick mean flow with no wall regression [22].

2. First-Order Solution

By substituting Eq. (46) into Eq. (44b), the ODE for the first-order term may be determined:

sinθF1cosθF1cosθF1+sinθF1=2ϑsinθ+(2π1αθ+4)cosθ+4π1αsinθ(47)

The solution of Eq. (47) requires the identification of a homogeneous part F1h. This can be obtained by solving

sinθF1hcosθF1hcosθF1h+sinθF1h=0(48)
for which a general homogeneous solution can be defined as
F1h=C(θ)cosθ(49)

Upon substitution of Eq. (49) into Eq. (48), one obtains the differential equation governing C(θ), namely,

Csinθcosθ2Csin2θC=0(50)

The solution of Eq. (50) may be managed through division by Csinθcosθ before integration. This operation yields

C(θ)=12K0tanθ+K1θ+K2(51)

Finally, the general homogeneous solution of Eq. (48) emerges as

F1h=C(θ)cosθ=K0sinθ+K1θcosθ+K2cosθ(52)

To obtain the total solution for Eq. (47), we use variation of parameters and rewrite

F1=K0(θ)sinθ+K1(θ)θcosθ+K2(θ)cosθ(53)
then, upon differentiation, we get
F1=K0cosθ+K1(cosθθsinθ)K2sinθ(54a)
F1=K0sinθ+K1(2sinθθcosθ)K2cosθ(54b)
F1=2K1sinθK0cosθ+K1(2cosθcosθ+θsinθ)+K2sinθ(54c)
along with the auxiliary conditions
K0sinθ+K1θcosθ+K2cosθ=0(55a)
K0cosθ+K1(cosθθsinθ)K2sinθ=0(55b)

Inserting Eq. (54) into Eq. (47) leads to

2K1sinθ=(2π1αθ+4)cotθ+4π1α2θ(56)

At this juncture, one can solve Eq. (56) in conjunction with Eq. (55) simultaneously for K0, K1, and K2. The resulting linear system of equations produces the following solution

K0=απ(cosθ+3lntan12θθsinθθcscθ)θcosθsinθ2cscθS(θ)+C0(57)
K1=απ(θcscθ3lntan12θ)+2cscθ+S(θ)+C1(58)
K2=απ[3S(θ)sinθθ2cscθθcosθ]cosθ+θsinθ2θcscθS1(θ)+C2(59)
where
S(θ)=0θϕcscϕdϕ;S1(θ)=0θϕ2cscϕdϕ(60)
and C0, C1, and C2 are constants that can be determined by invoking the boundary conditions in Eq. (32). The first-order solution given by Eq. (53) is now at hand, specifically,
F1=απ[3(sinθθcosθ)ln(tanθ2)2θ]+3παcosθS(θ)3S1(θ)cosθ+(θcosθsinθ)S(θ)+C0sinθ+C1θcosθ+C2cosθ(61)

What is left is the application of the boundary conditions, and these require that

F1(12π)=0orC0=3+α+S(12π)(62)
in conjunction with
limθ0F1(θ)=0;orC2=3(63)
and finally
dF1(1/2)dη=πdF1(π/2)dθ=0(64)
from which we extract
C1=16π+2π2α+2πS1(12π)S(12π)(6π2α+1)(65)

3. Verification

The total solution may now be assembled into

F=F0+ϵF1+O(ϵ2)=sinθ+ϵ{απ[3(sinθθcosθ)ln(tanθ2)2θ]+3παcosθS(θ)3S1(θ)cosθ+(θcosθsinθ)S(θ)+C0sinθ+C1θcosθ+C2cosθ}+O(ϵ2)(66)

This result is identical to the one obtained by Majdalani et al. [1,2], where corresponding flow attributes such as velocity, pressure, and vorticity are described. In this study, we limit ourselves to the detailed mathematical derivation and verification using a robust numerical solution. To this end, Eq. (66) is compared with results from the Runge–Kutta integration for Reynolds numbers of 100, 500, and 1000 taken at different values of the wall expansion ratio, namely, at α=10, 20, 10, and 20.

Starting with the mean flow function F, corresponding graphs are shown in Figs. 2 and 3. The first noticeable feature is the accuracy of the solution for large Reynolds number regardless of the wall expansion ratio. In fact, even for the relatively small value of Re=100, the curves appear to be visually indistinguishable unless one uses magnification on certain areas of the graphs. The other noteworthy feature is the effect of decreasing the wall expansion ratio. Specifically, the curves become increasingly more difficult to separate as one scans the plots sequentially, starting with the largest value of α in Figs. 2b and 3b.

Fig. 2
Fig. 2

Comparison between analytical and numerical solutions for the mean flow function F using a) α10 and b) α20. Curves are shown for Re100 (black), 500 (red), and 1000 (blue). Everywhere, inset frames are used with symbols for added clarity.

Fig. 3
Fig. 3

Comparison between analytical and numerical solutions for the mean flow function F using a) α10 and b) α20. Curves are shown for Re100 (black), 500 (red), and 1000 (blue).

The second set of comparisons is provided in Figs. 4 and 5 for F. We recall that the derivative of the mean flow function is directly connected to the axial velocity and, as such, the importance of arriving at a reliable analytical representation for F(η) cannot be underrated. In this case, similar trends are observed as those corresponding to F. The most prominent feature in these graphs may be the effect of α on the centerline velocity. As the wall expansion ratio is decreased, so is the centerline velocity. Altogether, the curves are practically inseparable, a satisfactory observation given that our solution is of first-order accuracy.

Fig. 4
Fig. 4

Comparison between analytical and numerical solutions for F using a) α10 and b) α20. Curves are shown for Re100 (black), 500 (red), and 1000 (blue).

Fig. 5
Fig. 5

Comparison between analytical and numerical solutions for F using a) α10 and b) α20. Curves are shown for Re100 (black), 500 (red), and 1000 (blue).

Additional confirmatory comparisons between numerics and asymptotics are provided in Tables 1 and 2 where F, F, and F are listed side-by-side along with their numerical predictions at three injection Reynolds numbers of 100, 500, and 1000, and either α=10 (Table 1) or 10 (Table 2). Corresponding values are provided at five characteristic radial positions corresponding to η=0,0.125,0.25,0.375, and 0.5. Irrespective of whether the injecting surface is expanding or contracting, it may be seen in both tables that the accuracy in F,F, and F for the Re=1000 case extends to 4, 3, and 2 significant digits, respectively. A slight degradation in precision occurs as the Reynolds number is lowered first to Re=500 and then to 100. The overall agreement, however, continues to fall approximately within the universally accepted engineering tolerance of 5%.

Given the error entailed in the model itself, the first-order approximation presented here appears to be sufficiently adequate, especially in the modeling of SRM flowfields for which the Reynolds number can be quite large [24]. There may be situations, however, where higher accuracy is desired at intermediate ranges of Reynolds number, for example, between 1 and 100. For such cases, higher-order approximations will be required and these could be obtained by resorting to other techniques, such as Lie-group theory [59,60], Adomian decomposition [64], or Homotopy Analysis Method (HAM) [61].

B. Small Injection and Suction with Low Regression

In this case, both the Reynolds number and the expansion ratio are low. This necessitates rewriting Eq. (31) as

2ηF+α(2ηF+4F)+Re(FFFF)+4F=0(67)

To solve Eq. (67) using a perturbation expansion, one needs to expand in both small parameters, Reynolds number and α. This is done by setting

F=F0+ReF1+O(Re2)(68)
F0=F00+αF01+O(α2)(69)
F1=F10+αF11+O(α2)(70)

1. Leading-Order Solution

At leading order in Reynolds number (i.e., Re=0), we collect

2ηF0+α(2ηF0+4F0)+4F0=0(71)
then, given
F0=F00+αF01+O(α2)(72)
the leading and first-order equations in α may be found to be
O(0):  ηF00+2F00=0(73a)
O(α):  ηF01+2F01=ηF002F00(73b)

The solution of Eq. (73a) may be readily found to be

F00=4η2+4η(74)

Subsequently, Eq. (73b) becomes

ηF01+2F01=16(75)

Because the boundary conditions are fulfilled by F00, the conditions on all remaining orders must be null. This may be secured by taking

{F0i(12)=0;F0i(12)=0F0i(0)=0;limη02ηF0i(η)=0;i=1,2,3,(76)

The solution of Eq. (75) returns

F01=13(4η34η2+η)(77)

Finally, the leading-order solution may be combined into

F0=F00+αF01=4η2+4η+13α(4η34η2+η)(78)

2. First-Order Solution

At first order in Reynolds number, we recover

2ηF1+α(2ηF1+4F1)+4F1=F0F0F0F0(79)

As before, the leading and first-order equations in α may be collected at successive orders such that

O(0):  ηF10+2F10=32η16(80a)
O(α):  ηF11+2F11=ηF102F10+49(72η2+48η15)(80b)

The corresponding solutions may be expressed as

F10=19(4η412η3+9η22η)(81)
and
F11=19(2η5+173η48η3+194η22324η)(82)

The first-order solution may be extracted to be

F1=19(4η412η3+9η22η)19α(2η5173η4+8η3194η2+2324η)(83)

3. Verification

The total solution for the small injection case with low regression becomes

F=4η(η1)+13αη(4η24η+1)+19η(4η312η2+9η2)Re19αη(2η4173η3+8η2194η+2324)Re(84)

Comparisons with numerical solutions are presented in Figs. 6 and 7 for Re=±1 and representative values of α=1, 0.5, 0.5, and 1. In general, the curves corresponding to both analytical and numerical results may be seen to exhibit substantial agreement. As one would expect, the agreement between numerics and asymptotics depreciates with increasing values of the Reynolds number, as well as increasing values of the wall expansion ratio. Larger values of the Reynolds number will not be suitable for comparison because the perturbation expansion is based on a small Reynolds number.

Fig. 6
Fig. 6

Comparison between analytical and numerical solutions for F using a) Re1 and b) Re1. Curves are shown for α0.5 (black), 1 (red), 0.5 (green), and 1 (blue).

Fig. 7
Fig. 7

Comparison between analytical and numerical solutions for F using a) Re1 and b) Re1. Curves are shown for α0.5 (black), 1 (red), 0.5 (green), and 1 (blue).

As before, side-by-side comparisons between numerics and asymptotics are provided in Tables 3 and 4, where approximate representations of F, F, and F are showcased against their numerical estimates for two wall expansion ratios of 1 and +1, and either Re=1 (Table 3) or 1 (Table 4). For these small injection or suction cases, computed values are given at five radial positions corresponding to η=0,0.125,0.25,0.375, and 0.5. Recalling that the asymptotic expansion is carried out in orders of Reynolds number, it is not surprising that the accuracy of the analytical approximation is lower than its large injection and suction counterparts, specifically where the perturbation series rather depends on ϵ=1/ReO(103102); the present ϵ1 is considerably larger in comparison. In view of this conservative value of |ϵ=Re|, tabulated values reflect diminishing accuracy going from the parent function F to its derivatives F and F. Furthermore, it may be seen that a higher degree of precision may be achieved for radial positions corresponding to η1/4. Those closer to the centerline exhibit lower precision, down to a single digit accuracy, whereas those closer to the sidewall reflect either double or triple digit accuracy.

It should be noted that this solution is not connected with solid rocket propulsion due to the large injection rates that characterize propellant burning. However, it is relevant to other problems involving sweat cooling, filtration, peristaltic pumping, boundary-layer control, and so on, where the injection or suction Reynolds number at the slowly moving surface remains relatively small.

V. Conclusions

This paper revisits the problem of a viscous incompressible fluid in a porous tube with an expanding or contracting sidewall. The method used is based on conventional, regular perturbation expansions in two distinct physical settings: large injection (Re), and small injection or suction with weak regression. One of the key aspects of this study stands in the development of a robust numerical algorithm that obviates the need to use predictor–corrector or shooting schemes. Numerical results are obtained directly, in fact, nearly instantaneously, given an assortment of three guesses: the first and second derivatives of G(ξ) at the origin and the control parameter β. Given that G(0)=0 remains fixed, it is found best for the cases at hand to set G(0)>0 and G(0)<0, and then carefully vary β to the extent of canvassing the solution domain for the desired parameters. Once convergence is achieved, the values of Re=2G(b/2) and α=bβ can be deduced.

In principle, this approach outperforms the Dauenhauer–Majdalani technique [58], which requires the direct specification of Reynolds number and α, although iteration on the initial guesses is still required until convergence is achieved. In actuality, all existing techniques involve some level of iteration, and the ability to prescribe Reynolds number and α beforehand remains the main advantage of the Dauenhauer–Majdalani approach, despite its reliance on guesswork and the Newton–Raphson scheme to accelerate convergence.

In the present algorithm, the effort lies in solving as many cases as possible near the desired values of the Reynolds number and wall expansion ratio. Evidently, to minimize labor and eliminate trial-and-error in pursuit of convergence to a desired pair of (Re,α), generating a lookup table in which the practical range of Reynolds number and α is finely canvassed would be most efficient. The development of such a lookup table will, however, require a systematic procedure for varying the two initial guesses and β. The procedure described here will be quite suitable in filling the lookup table, given a predefined matrix for initial guesses. However, although a lookup table offers numerous advantages, the solutions reported in this article are determined manually.

In summary, the numerical simulations presented in this article were carried out at representative values of the Reynolds number and the wall expansion ratio. It was determined that, so long as the control parameters were chosen from their prescribed asymptotic range, the corresponding analytical approximations and their numerical predictions agreed substantially. In comparison with other series approximations that have since appeared in the literature, the expressions associated with a perturbation expansion remain relatively compact, thus adding physical insight to the problem under investigation.

Other series expressions, such as those obtained through the Homotopy-Analysis Method, also offer distinct advantages, including improved accuracy, controlled convergence to the desired solution (in the case of multiple solutions), and extended ranges of applicability, especially in that their precision remains independent of the size of Reynolds number and α. Their main drawback, however, may be connected to their typical reliance on such a large number of terms to make them nearly equivalent to an elegant numerical routine. As for the asymptotic approximations, their compactness often leads to the identification of key similarity parameters and scales. From an engineering perspective, their accuracy is often sufficient with the retention of only one term beyond the leading-order approximation. In future work, it is hoped that both homotopy analysis method and perturbation tools will continue to receive attention given their clearly visible benefits and complementary characters.

M. M. ChoudhariAssociate Editor

Acknowledgments

This material is based on work supported partly by the National Science Foundation, and partly by Auburn University, Department of Aerospace Engineering, through the Hugh and Loeda Francis Chair of Excellence.

References

  • [1] Majdalani J., Vyas A. B. and Flandro G. A., “Higher Mean–Flow Approximation for a Solid Rocket Motor with Radially Regressing Walls,” AIAA Journal, Vol. 40, No. 9, 2002, pp. 1780–1788. doi:https://doi.org/10.2514/2.1854 AIAJAH 0001-1452 LinkGoogle Scholar

  • [2] Majdalani J., Vyas A. B. and Flandro G. A., “Higher Mean–Flow Approximation for a Solid Rocket Motor with Radially Regressing Walls–Erratum,” AIAA Journal, Vol. 47, No. 1, 2009, pp. 286–286. doi:https://doi.org/10.2514/1.40061 AIAJAH 0001-1452 LinkGoogle Scholar

  • [3] Majdalani J. and Zhou C., “Moderate-to-Large Injection and Suction Driven Channel Flows with Expanding or Contracting Walls,” Journal of Applied Mathematics and Mechanics, Vol. 83, No. 3, 2003, pp. 181–196. doi:https://doi.org/10.1002/zamm.200310018 JAMMAR 0021-8928 CrossrefGoogle Scholar

  • [4] Majdalani J. and Akiki M., “Rotational and Quasiviscous Cold Flow Models for Axisymmetric Hybrid Propellant Chambers,” Journal of Fluids Engineering, Vol. 132, No. 10, 2010, Paper 101202. doi:https://doi.org/10.1115/1.4002397 JFEGA4 0098-2202 CrossrefGoogle Scholar

  • [5] Majdalani J., “Analytical Models for Hybrid Rockets,” Fundamentals of Hybrid Rocket Combustion and Propulsion, edited by Kuo K. and Chiaverini M. J., Progress in Astronautics and Aeronautics, AIAA, Reston, VA, 2007, pp. 207–246, Chap. 5. Google Scholar

  • [6] Peng Y. and Yuan S. W., “Laminar Pipe Flow with Mass Transfer Cooling,” Journal of Heat Transfer, Vol. 87, No. 2, May 1965, pp. 252–258. doi:https://doi.org/10.1115/1.3689085 JHTRAO 0022-1481 CrossrefGoogle Scholar

  • [7] Yuan S. W. and Finkelstein A. B., “Heat Transfer in Laminar Pipe Flow with Uniform Coolant Injection,” Jet Propulsion, Vol. 28, No. 3, 1958, pp. 178–181. doi:https://doi.org/10.2514/8.7264 JETPAV 0095-8751 LinkGoogle Scholar

  • [8] Acrivos A., “The Asymptotic Form of the Laminar Boundary-Layer Mass-Transfer Rate for Large Interfacial Velocities,” Journal of Fluid Mechanics, Vol. 12, No. 3, 1962, pp. 337–357. doi:https://doi.org/10.1017/S0022112062000257 JFLSA7 0022-1120 CrossrefGoogle Scholar

  • [9] Libby P. A., “The Homogeneous Boundary Layer at an Axisymmetric Stagnation Point with Large Rates of Injection,” Journal of the Aerospace Sciences, Vol. 29, No. 1, 1962, pp. 48–60. doi:https://doi.org/10.2514/8.9302 LinkGoogle Scholar

  • [10] Libby P. A. and Pierucci M., “Laminar Boundary Layer with Hydrogen Injection Including Multicomponent Diffusion,” AIAA Journal, Vol. 2, No. 12, Dec. 1964, pp. 2118–2126. doi:https://doi.org/10.2514/3.2752 AIAJAH 0001-1452 LinkGoogle Scholar

  • [11] Fung Y. C. and Yih C. S., “Peristaltic Transport,” Journal of Applied Mechanics, Vol. 35, No. 4, Dec. 1968, pp. 669–675. doi:https://doi.org/10.1115/1.3601290 JAMCAV 0021-8936 CrossrefGoogle Scholar

  • [12] Uchida S. and Aoki H., “Unsteady Flows in a Semi–Infinite Contracting or Expanding Pipe,” Journal of Fluid Mechanics, Vol. 82, No. 2, 1977, pp. 371–387. doi:https://doi.org/10.1017/S0022112077000718 JFLSA7 0022-1120 CrossrefGoogle Scholar

  • [13] Berman A. S., “Laminar Flow in Channels with Porous Walls,” Journal of Applied Physics, Vol. 24, No. 9, 1953, pp. 1232–1235. doi:https://doi.org/10.1063/1.1721476 JAPIAU 0021-8979 CrossrefGoogle Scholar

  • [14] Berman A. S., “Effects of Porous Boundaries on the Flow of Fluids in Systems with Various Geometries,” Proceedings of the Second United Nations International Conference on the Peaceful Uses of Atomic Energy, Vol. 4, United Nations, Geneva, Switzerland, 1958, pp. 351–358. Google Scholar

  • [15] Berman A. S., “Laminar Flow in an Annulus with Porous Walls,” Journal of Applied Physics, Vol. 29, No. 1, 1958, pp. 71–75. doi:https://doi.org/10.1063/1.1722948 JAPIAU 0021-8979 CrossrefGoogle Scholar

  • [16] Yuan S. W. and Finkelstein A. B., “Laminar Pipe Flow with Injection and Suction Through a Porous Wall,” Journal of Applied Mechanics, Vol. 78, No. 3, May 1956, pp. 719–724. Google Scholar

  • [17] Terrill R. M. and Thomas P. W., “On Laminar Flow Through a Uniformly Porous Pipe,” Applied Scientific Research, Vol. 21, No. 1, 1969, pp. 37–67. doi:https://doi.org/10.1007/BF00411596 ASRHAU 0003-6994 CrossrefGoogle Scholar

  • [18] Terrill R. M., “On Some Exponentially Small Terms Arising in Flow Through a Porous Pipe,” Quarterly Journal of Mechanics and Applied Mathematics, Vol. 26, No. 3, 1973, pp. 347–354. doi:https://doi.org/10.1093/qjmam/26.3.347 QJMMAV 0033-5614 CrossrefGoogle Scholar

  • [19] Terrill R., “Laminar Flow in a Porous Tube,” Journal of Fluids Engineering, Vol. 105, No. 3, 1983, pp. 303–307. doi:https://doi.org/10.1115/1.3240992 CrossrefGoogle Scholar

  • [20] Skalak F. M. and Wang C.-Y., “On the Nonunique Solutions of Laminar Flow Through a Porous Tube or Channel,” SIAM Journal on Applied Mathematics, Vol. 34, No. 3, May 1978, pp. 535–544. doi:https://doi.org/10.1137/0134042 SMJMAP 0036-1399 CrossrefGoogle Scholar

  • [21] Taylor G. I., “Fluid Flow in Regions Bounded by Porous Surfaces,” Proceedings of the Royal Society of London, Series A: Mathematical and Physical Sciences, Vol. 234, No. 1199, March 1956, pp. 456–475. doi:https://doi.org/10.1098/rspa.1956.0050 CrossrefGoogle Scholar

  • [22] Culick F. E. C., “Rotational Axisymmetric Mean Flow and Damping of Acoustic Waves in a Solid Propellant Rocket,” AIAA Journal, Vol. 4, No. 8, 1966, pp. 1462–1464. doi:https://doi.org/10.2514/3.3709 AIAJAH 0001-1452 LinkGoogle Scholar

  • [23] Casalis G., Avalon G. and Pineau J.-P., “Spatial Instability of Planar Channel Flow with Fluid Injection Through Porous Walls,” Physics of Fluids, Vol. 10, No. 10, 1998, pp. 2558–2568. doi:https://doi.org/10.1063/1.869770 CrossrefGoogle Scholar

  • [24] Chedevergne F., Casalis G. and Féraille T., “Biglobal Linear Stability Analysis of the Flow Induced by Wall Injection,” Physics of Fluids, Vol. 18, No. 1, 2006, Paper 014103. doi:https://doi.org/10.1063/1.2160524 CrossrefGoogle Scholar

  • [25] Griffond J. and Casalis G., “On the Dependence on the Formulation of Some Nonparallel Stability Approaches Applied to the Taylor Flow,” Physics of Fluids, Vol. 12, No. 2, Feb. 2000, pp. 466–468. doi:https://doi.org/10.1063/1.870323 CrossrefGoogle Scholar

  • [26] Griffond J. and Casalis G., “On the Nonparallel Stability of the Injection Induced Two-Dimensional Taylor Flow,” Physics of Fluids, Vol. 13, No. 6, June 2001, pp. 1635–1644. doi:https://doi.org/10.1063/1.1367869 CrossrefGoogle Scholar

  • [27] Griffond J., Casalis G. and Pineau J.-P., “Spatial Instability of Flow in a Semiinfinite Cylinder with Fluid Injection Through its Porous Walls,” European Journal of Mechanics-B/Fluids, Vol. 19, No. 1, 2000, pp. 69–87. doi:https://doi.org/10.1016/S0997-7546(00)00105-9 CrossrefGoogle Scholar

  • [28] Abu-Irshaid E. M., Majdalani J. and Casalis G., “Hydrodynamic Stability of Rockets with Headwall Injection,” Physics of Fluids, Vol. 19, No. 2, 2007, Paper 024101. doi:https://doi.org/10.1063/1.2434797 CrossrefGoogle Scholar

  • [29] Majdalani J., Flandro G. A. and Roh T. S., “Convergence of Two Flowfield Models Predicting a Destabilizing Agent in Rocket Combustion,” Journal of Propulsion and Power, Vol. 16, No. 3, 2000, pp. 492–497. doi:https://doi.org/10.2514/2.5595 JPPOEL 0748-4658 LinkGoogle Scholar

  • [30] Fabignon Y., Dupays J., Avalon G., Vuillot F., Lupoglazoff N., Casalis G. and Prévost M., “Instabilities and Pressure Oscillations in Solid Rocket Motors,” Journal of Aerospace Science and Technology, Vol. 7, No. 3, April 2003, pp. 191–200. doi:https://doi.org/10.1016/S1270-9638(02)01194-X CrossrefGoogle Scholar

  • [31] Flandro G. A. and Majdalani J., “Aeroacoustic Instability in Rockets,” AIAA Journal, Vol. 41, No. 3, Feb. 2003, pp. 485–497. doi:https://doi.org/10.2514/2.1971 AIAJAH 0001-1452 LinkGoogle Scholar

  • [32] Majdalani J., Flandro G. A. and Fischbach S. R., “Some Rotational Corrections to the Acoustic Energy Equation in Injection-Driven Enclosures,” Physics of Fluids, Vol. 17, No. 7, 2005, Paper 074102. doi:https://doi.org/10.1063/1.1920647 CrossrefGoogle Scholar

  • [33] Majdalani J., Fischbach S. R. and Flandro G. A., “Improved Energy Normalization Function in Rocket Motor Stability Calculations,” Journal of Aerospace Science and Technology, Vol. 10, No. 6, Sept. 2006, pp. 495–500. doi:https://doi.org/10.1016/j.ast.2006.06.002 CrossrefGoogle Scholar

  • [34] Flandro G. A., Fischbach S. R. and Majdalani J., “Nonlinear Rocket Motor Stability Prediction: Limit Amplitude, Triggering, and Mean Pressure Shift,” Physics of Fluids, Vol. 19, No. 9, Sept. 2007, Paper 094101. doi:https://doi.org/10.1063/1.2746042 CrossrefGoogle Scholar

  • [35] Fischbach S. R., Majdalani J. and Flandro G. A., “Acoustic Instability of the Slab Rocket Motor,” Journal of Propulsion and Power, Vol. 23, No. 1, Jan.–Feb. 2007, pp. 146–157. doi:https://doi.org/10.2514/1.14794 JPPOEL 0748-4658 LinkGoogle Scholar

  • [36] Majdalani J. and Roh T., “The Oscillatory Channel Flow with Large Wall Injection,” Proceedings of the Royal Society of London, Series A: Mathematical and Physical Sciences, Vol. 456, No. 1999, July 2000, pp. 1625–1657. doi:https://doi.org/10.1098/rspa.2000.0579 CrossrefGoogle Scholar

  • [37] Majdalani J., “The Oscillatory Channel Flow with Arbitrary Wall Injection,” Journal of Applied Mathematics and Physics (ZAMP), Vol. 52, No. 1, 2001, pp. 33–61. doi:https://doi.org/10.1007/PL00001539 CrossrefGoogle Scholar

  • [38] Majdalani J. and Flandro G. A., “The Oscillatory Pipe Flow with Arbitrary Wall Injection,” Proceedings of the Royal Society of London, Series A, Vol. 458, No. 2023, July 2002, pp. 1621–1651. doi:https://doi.org/10.1098/rspa.2001.0930 CrossrefGoogle Scholar

  • [39] Majdalani J., “Multiple Asymptotic Solutions for Axially Travelling Waves in Porous Channels,” Journal of Fluid Mechanics, Vol. 636, No. 1, Oct. 2009, pp. 59–89. doi:https://doi.org/10.1017/S0022112009007939 JFLSA7 0022-1120 CrossrefGoogle Scholar

  • [40] Féraille T. and Casalis G., “Channel Flow Induced by Wall Injection of Fluid and Particles,” Physics of Fluids, Vol. 15, No. 2, Feb. 2003, pp. 348–360. doi:https://doi.org/10.1063/1.1530158 CrossrefGoogle Scholar

  • [41] Clayton C. D., “Flow Fields in Solid Rocket Motors with Tapered Bores,” 32nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference, AIAA Paper  1996-2643, July 1996. LinkGoogle Scholar

  • [42] Saad T., Sams O. C. and Majdalani J., “Rotational Flow in Tapered Slab Rocket Motors,” Physics of Fluids, Vol. 18, No. 10, 2006, Paper 103601. doi:https://doi.org/10.1063/1.2354193 CrossrefGoogle Scholar

  • [43] Sams O. C., Majdalani J. and Saad T., “Mean Flow Approximations for Solid Rocket Motors with Tapered Walls,” Journal of Propulsion and Power, Vol. 23, No. 2, March–April 2007, pp. 445–456. doi:https://doi.org/10.2514/1.15831 JPPOEL 0748-4658 LinkGoogle Scholar

  • [44] Sabnis J. S., Gibeling H. J. and McDonald H., “Navier-Stokes Analysis of Solid Propellant Rocket Motor Internal Flows,” Journal of Propulsion and Power, Vol. 5, No. 6, 1989, pp. 657–664. doi:https://doi.org/10.2514/3.23203 JPPOEL 0748-4658 LinkGoogle Scholar

  • [45] Dunlap R., Willoughby P. G. and Hermsen R. W., “Flowfield in the Combustion Chamber of a Solid Propellant Rocket Motor,” AIAA Journal, Vol. 12, No. 10, 1974, pp. 1440–1442. doi:https://doi.org/10.2514/3.49513 AIAJAH 0001-1452 LinkGoogle Scholar

  • [46] Dunlap R., Blackner A. M., Waugh R. C., Brown R. S. and Willoughby P. G., “Internal Flow Field Studies in a Simulated Cylindrical Port Rocket Chamber,” Journal of Propulsion and Power, Vol. 6, No. 6, 1990, pp. 690–704. doi:https://doi.org/10.2514/3.23274 JPPOEL 0748-4658 LinkGoogle Scholar

  • [47] Kurdyumov V. N., “Steady Flows in the Slender, Noncircular, Combustion Chambers of Solid Propellants Rockets,” AIAA Journal, Vol. 44, No. 12, Dec. 2006, pp. 2979–2986. doi:https://doi.org/10.2514/1.21125 AIAJAH 0001-1452 LinkGoogle Scholar

  • [48] Majdalani J. and Saad T., “The Taylor-Culick Profile with Arbitrary Headwall Injection,” Physics of Fluids, Vol. 19, No. 9, Sept. 2007, Paper 093601. doi:https://doi.org/10.1063/1.2746003 CrossrefGoogle Scholar

  • [49] Saad T. and Majdalani J., “On the Lagrangian Optimization of Wall-Injected Flows: From the Hart-McClure Potential to the Taylor-Culick Rotational Motion,” Proceedings of the Royal Society of London, Series A, Vol. 466, No. 2114, Feb. 2010, pp. 331–362. doi:https://doi.org/10.1098/rspa.2009.0326 CrossrefGoogle Scholar

  • [50] Hart R. W. and McClure F. T., “Combustion Instability: Acoustic Interaction with a Burning Propellant Surface,” Journal of Chemical Physics, Vol. 30, No. 6, June 1959, pp. 1501–1514. doi:https://doi.org/10.1063/1.1730226 CrossrefGoogle Scholar

  • [51] McClure F., Hart R. and Cantrell R., “Interaction Between Sound and Flow: Stability of T-Burners,” AIAA Journal, Vol. 1, No. 3, 1963, pp. 586–590. doi:https://doi.org/10.2514/3.54846 AIAJAH 0001-1452 LinkGoogle Scholar

  • [52] McClure F. T., Hart R. W. and Bird J. F., “Acoustic Resonance in Solid Propellant Rockets,” Journal of Applied Physics, Vol. 31, No. 5, May 1960, pp. 884–896. doi:https://doi.org/10.1063/1.1735713 JAPIAU 0021-8979 CrossrefGoogle Scholar

  • [53] Cantrell R. H., Hart R. W. and McClure F. T., “Acoustic Energy Losses in Rocket-Engine Cavities,” Journal of the Acoustical Society of America, Vol. 35, No. 5, 1963, pp. 773–773. doi:https://doi.org/10.1121/1.2142356 JASMAN 0001-4966 CrossrefGoogle Scholar

  • [54] Hart R. W., Bird J. F., Cantrell R. H. and McClure F. T., “Nonlinear Effects in Instability of Solid Propellant Rocket Motors,” AIAA Journal, Vol. 2, No. 7, 1964, pp. 1270–1273. doi:https://doi.org/10.2514/3.55069 AIAJAH 0001-1452 LinkGoogle Scholar

  • [55] Goto M. and Uchida S., “Unsteady Flows in a Semi–Infinite Expanding Pipe with Injection Through Wall,” Journal of the Japan Society for Aeronautical and Space Sciences, Vol. 38, No. 434, 1990, pp. 131–138. doi:https://doi.org/10.2322/jjsass1969.38.131 NKGAB8 0021-4663 CrossrefGoogle Scholar

  • [56] Dauenhauer E. C. and Majdalani J., “Unsteady Flows in Semi-Infinite Expanding Channels with Wall Injection,” 30th AIAA Fluid Dynamics Conference, AIAA Paper  1999-3523, June–July 1999. LinkGoogle Scholar

  • [57] Dauenhauer E. C. and Majdalani J., “Exact Self-Similarity Solution of the Navier-Stokes Equations for a Deformable Channel with Wall Suction or Injection,” 37th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, AIAA Paper  2001-3588, July 2001. LinkGoogle Scholar

  • [58] Dauenhauer E. C. and Majdalani J., “Exact Self-Similarity Solution of the Navier–Stokes Equations for a Porous Channel with Orthogonally Moving Walls,” Physics of Fluids, Vol. 15, No. 6, 2003, pp. 1485–1495. doi:https://doi.org/10.1063/1.1567719 CrossrefGoogle Scholar

  • [59] Boutros Y. Z., Abd-el Malek M. B., Badran N. A. and Hassan H. S., “Lie-Group Method for Unsteady Flows in a Semi-Infinite Expanding or Contracting Pipe with Injection or Suction through a Porous Wall,” Journal of Computational and Applied Mathematics, Vol. 197, No. 2, 2006, pp. 465–494. doi:https://doi.org/10.1016/j.cam.2005.11.031 JCAMDI 0377-0427 CrossrefGoogle Scholar

  • [60] Boutros Y. Z., Abd-el Malek M. B., Badran N. A. and Hassan H. S., “Lie-Group Method Solution for Two-Dimensional Viscous Flow Between Slowly Expanding or Contracting Walls with Weak Permeability,” Applied Mathematical Modelling, Vol. 31, No. 6, 2007, pp. 1092–1108. doi:https://doi.org/10.1016/j.apm.2006.03.026 AMMODL 0307-904X CrossrefGoogle Scholar

  • [61] Xu H., Lin Z.-L., Liao S.-J., Wu J.-Z. and Majdalani J., “Homotopy Based Solutions of the Navier-Stokes Equations for a Porous Channel with Orthogonally Moving Walls,” Physics of Fluids, Vol. 22, No. 5, May 2010, Paper 053601. doi:https://doi.org/10.1063/1.3392770 CrossrefGoogle Scholar

  • [62] Dinarvand S., Rashidi M. and Doosthoseini A., “Analytical Approximate Solutions for Two-Dimensional Viscous Flow Through Expanding or Contracting Gaps with Permeable Walls,” Central European Journal of Physics, Vol. 7, No. 4, 2009, pp. 791–799. doi:https://doi.org/10.2478/s11534-009-0024-x 1895-1082 Google Scholar

  • [63] Dinarvand S. and Rashidi M. M., “A Reliable Treatment of a Homotopy Analysis Method for Two-Dimensional Viscous Flow in a Rectangular Domain Bounded by Two Moving Porous Walls,” Nonlinear Analysis: Real World Applications, Vol. 11, No. 3, 2010, pp. 1502–1512. doi:https://doi.org/10.1016/j.nonrwa.2009.03.006 NARWCQ 1468-1218 CrossrefGoogle Scholar

  • [64] Liao S., Beyond Perturbation: Introduction to the Homotopy Analysis Method, 1st ed., CRC Press, Boca Raton, FL, 2003. CrossrefGoogle Scholar

  • [65] Liao S., “On the Homotopy Analysis Method for Nonlinear Problems,” Applied Mathematics and Computation, Vol. 147, No. 2, Jan. 2004, pp. 499–513. doi:https://doi.org/10.1016/S0096-3003(02)00790-7 AMHCBQ 0096-3003 CrossrefGoogle Scholar

  • [66] Liao S. and Tan Y., “A General Approach to Obtain Series Solutions of Nonlinear Differential Equations,” Studies in Applied Mathematics, Vol. 119, No. 4, Oct. 2007, pp. 297–354. doi:https://doi.org/10.1111/sapm.2007.119.issue-4 SAPMB6 0022-2526 CrossrefGoogle Scholar

Tables

Algorithm 1: RK4INTEGRATE()a

allocate memory for arrays
process user input: G(0), G(0), G(0), β, nmax, nTaylor, ξmax
calculate K1
calculate Taylor series coefficients γi
fori0tonTaylor
do calculate G(ξ) near the centerline
forinTaylortonmax
do integrate governing equation
fori0tonmax
do find npG(np)=0
calculate ξp that corresponds to np
comment: now that ξp has been found, redo the integration from 0 to ξp for better accuracy
ξmaxξp+ε
forinTaylortonmax
do integrate governing equation
fori0 to nmax
do find npG(np)=0
calculate ξp that corresponds to np
b2ξp
αbβ
Re2G(np)

a Algorithm describing the numerical method used to solve Eq. (37). Here, ξmax denotes the domain size starting at ξ=0, nmax stands for the number of points that divide the domain, and nTaylor represents the number of points for which the Taylor series expansion must be applied. Note that np corresponds to the point where G(np)=0 and ε remains a small parameter. All subscripts with p refer to this particular point.

Table 1 Comparison between numerical and analytical solutions for the large injection case with α=10 and Re=100, 500, and 1000a

 FFF
ReηAnalyticalNumericalAnalyticalNumericalAnalyticalNumerical
10000.00000.00003.30333.29820.39480.6082
1000.1250.40860.39563.00322.94545.62254.7169
1000.250.72040.71922.17192.17407.49057.4199
1000.3750.93360.92831.07061.14108.71968.8978
1000.51.00001.00000.00000.0000049.07319.1578
50000.00000.00003.17393.17290.07890.0915
5000.1250.38540.38542.91252.91143.97743.9728
5000.250.70980.70972.21152.21107.08137.0745
5000.3750.92480.92481.18871.18869.07289.0684
5000.51.00001.0000.00000.00009.71009.7087
100000.00000.00003.15773.15690.03940.0430
10000.1250.38400.38402.90742.90663.87723.8746
10000.250.70840.70842.21642.21597.03017.0256
10000.3750.92430.92431.19541.19529.09559.0909
10000.51.00001.00000.00000.00009.78989.7857

a Samples are taken at representative radial positions of η={0,1/8,1/4,3/8,1/2}.

Table 2 Comparison between numerical and analytical solutions for the large injection case with α=10 and Re=100, 500, and 1000a

 FFF
ReηAnalyticalNumericalAnalyticalNumericalAnalyticalNumerical
10000.00000.00003.04743.05210.39470.2091
1000.1250.37230.37292.85152.85153.09263.1597
1000.250.69580.69602.25322.24936.45486.4480
1000.3750.91910.91911.26241.26179.23339.2008
1000.51.00001.00000.00000.000010.695810.7304
50000.00000.00003.12263.05210.07890.2091
5000.1250.38060.38062.89222.89173.63923.6408
5000.250.70480.69602.22782.24936.87356.4480
5000.3750.92290.92291.21431.21419.14169.1373
5000.51.00001.0000.00000.000010.035610.7304
100000.00000.00003.13213.13160.03940.0353
10000.1250.38160.38162.89732.89683.70813.7076
10000.250.70590.70592.22462.22426.92626.9238
10000.3750.92340.92341.20821.20809.12999.1265
10000.51.00001.00000.00000.0000079.95259.9496

a Samples are taken at representative radial positions of η={0,1/8,1/4,3/8,1/2}.

Table 3 Comparison between numerical and analytical solutions for the small injection case with Re1 and α=±1a

 FFF
αηAnalyticalNumericalAnalyticalNumericalAnalyticalNumerical
100.00000.00003.52723.56364.21684.5987
10.1250.40430.40642.91272.91455.62685.8030
10.250.72070.72212.11912.10947.07107.0892
10.3750.92660.92691.14591.13978.49058.4333
10.51.00001.00000.00000.0000029.82659.8035
100.00000.00003.98244.08667.38938.4808
10.1250.43820.44443.01743.02547.98078.4651
10.250.75230.75692.00341.97778.18448.2709
10.3750.93890.94030.98500.96378.05887.9303
10.51.00001.00000.00000.0000037.66227.4717

a Samples are taken at representative radial positions of η={0,1/8,1/4,3/8,1/2}.

Table 4 Comparison between numerical and analytical solutions for the small injection case with Re1 and α=±1a

 FFF
αηanalyticalnumericalanalyticalnumericalanalyticalnumerical
100.00000.00003.75853.85656.07367.1347
10.1250.42040.42612.95322.95886.80217.2564
10.250.73460.73862.05932.03487.49907.5599
10.3750.93160.93271.07731.05938.22318.0914
10.51.00001.00000.00000.0000039.03328.9109
100.00000.00004.68314.933313.910616.6442
10.1250.48360.49823.10843.125211.350812.4811
10.250.78940.80011.83011.77029.15829.3458
10.3750.95170.95500.80580.75657.27766.9838
10.51.00001.00000.00000.0000035.65415.2039

a Samples are taken at representative radial positions of η={0,1/8,1/4,3/8,1/2}.