Abstract

New exact solutions for the motion of a fractionalized (this word is suitable when fractional derivative is used in constitutive or governing equations) second grade fluid due to longitudinal and torsional oscillations of an infinite circular cylinder are determined by means of Laplace and finite Hankel transforms. These solutions are presented in series form in term of generalized 𝐺𝑎,𝑏,𝑐(,𝑡) functions and satisfy all imposed initial and boundary conditions. In special cases, solutions for ordinary second grade and Newtonian fluids are obtained. Furthermore, other equivalent forms of solutions for ordinary second grade and Newtonian fluids are presented and written as sum of steady-state and transient solutions. The solutions for Newtonian fluid coincide with the well-known classical solutions. Finally, by means of graphical illustrations, the influence of pertinent parameters on fluid motion as well as comparison among different models is discussed.

1. Introduction

In recent years, the non-Newtonian fluids have received considerable attention by scientist and engineers. Such interest is inspired by practical applications of non-Newtonian fluids in industry and engineering applications. The shear stress and shear rate in non-Newtonian fluids are connected by a relation in a nonlinear manner. Because of diverse fluids characteristics in nature, all the non-Newtonian fluids cannot be described by a single constitutive relation [16]. Thus, among the several existing non-Newtonian fluid models, there is one which is most famous model called second grade fluid [7]. Although the constitutive equation of second grade fluid is simpler than that for the rate type fluids (those fluids which encounter viscoelastic and memory effects), it has been shown by Walters [8] that, for many types of problems in which the flow is slow enough in the viscoelastic sense, the results given using Oldroyd fluid will be substantially similar to those obtained for second grade fluid. Therefore, if we discuss the result in this manner, it is reasonable to use the second grade fluid to carry out the calculations as compared to other non-Newtonian fluids. This fact seems to be true, not only for exact analytic solutions but even for numerical solutions. The second grade fluid is the simplest subclass of non-Newtonian fluids for which one can reasonably hope to obtain exact analytic solutions. Some recent attempts regarding exact analytic solutions for the flow of a second grade fluid are present in [916].

Linear viscoelasticity is certainly the field of most extensive applications of fractional calculus, in view of its ability to model hereditary phenomena with long memory. During the twentieth century, a number of authors have (implicitly or explicitly) used the fractional calculus as an empirical method of describing the properties of viscoelastic materials [17]. A motivation for using fractional order operators in viscoelasticity is that a whole spectrum of viscoelastic mechanisms can be included in a single internal variable [18]. The stress relaxation spectrum for the fractional order model is continuous with the relaxation constant as the most probable relaxation time, while the order of the operator plays the role of a distribution parameter. Note that the spectrum is discrete for the classical model that is based on integer order derivatives. By a suitable choice of material parameters for the classical viscoelastic model, it is observed both numerically and analytically that the classical model with a large number of internal variables (each representing a specific viscoelastic mechanism) converges to the fractional model with a single internal variable [18, 19]. In other cases, it has been shown that the governing equations employing fractional derivatives are also linked to molecular theories [20]. The use of fractional derivatives within the context of viscoelasticity was firstly proposed by Germant [21]. Later, Bagley and Torvik [22] demonstrated that the theory of viscoelasticity of coiling polymers predicts constitutive relations with fractional derivatives, and Makris et al. [23] achieved a very good fit of the experimental data when the fractional derivative Maxwell model has been used instead of the Maxwell model for the silicon gel fluid. Some important recent attempts of fractional derivative approach to non-Newtonian fluids to obtain exact analytic solutions are listed here [2430].

The oscillating flow of the viscoelastic fluid in cylindrical pipes has been applied in many fields, such as industries of petroleum, chemistry, and bioengineering. In the field of bioengineering, this type of investigation is of particular interest since blood in veins is forced by a periodic pressure gradient. In the petroleum and chemical industries, there are also many problems which involve the dynamic response of the fluid to the frequency of the periodic pressure gradient. An excellent collection of papers on oscillating flow can be found in the paper by Yin and Zhu [31]. We also include some important studies of non-Newtonian fluids, where oscillating boundary value problems are used in cylindrical region [3240]. Consequently, for completeness and motivated by the above remarks, we solve our problem for fractionalized second grade fluid. The aim of this paper is to find some new and closed-form exact solutions for the oscillating flows of fractionalized second grade fluid. More precisely, our objective is to find the velocity field and the shear stresses corresponding to the motion of a fractionalized second grade fluid through a cylinder due to longitudinal and torsional oscillations of an infinite circular cylinder. The general solutions are obtained using the discrete Laplace and finite Hankel transforms. They are presented in series form in term of the 𝐺𝑎,𝑏,𝑐(,𝑡) functions in simpler forms as comparison to known results from literature. The solutions for similar motion of ordinary second grade and Newtonian fluids are obtained as spacial cases from general solutions. Equivalent forms of the solutions for ordinary second grade and Newtonian fluids are also constructed and presented as a sum between steady-state and transient solutions. The equivalent forms of general solutions for Newtonian fluid coincide with the well known classical solutions from the literature. Finally, the influence of material and fractional parameters on the motion of fractionalized second grade fluid is underlined by graphical illustrations. The difference among fractionalized, ordinary second grade and Newtonian fluid models is also spotlighted.

2. Governing Equations for Fractionalized Second Grade Fluid

The Cauchy stress 𝑇 in an incompressible homogeneous fluid of second grade is related to the fluid motion in the following manner:𝑇=𝑝𝐼+𝑆,𝑆=𝜇𝐴1+𝛼1𝐴2+𝛼2𝐴21,(2.1) where 𝑝𝐼 is the indeterminate part of the stress due to the constraint of incompressibility, S is the extra-stress tensor, 𝜇 is the dynamic viscosity, 𝛼1 and 𝛼2 are the normal stress moduli, and 𝐴1 and 𝐴2 are the kinematic tensors defined through𝐴1=(𝑉)+(𝑉)𝑇,𝐴2=𝑑𝐴1𝑑𝑡+𝐴1(𝑉)+(𝑉)𝑇𝐴1.(2.2) In the above equations, 𝑉 is the velocity field, is the gradient operator, and 𝑑/𝑑𝑡 denotes the material time derivative. Since the fluid is incompressible, it can undergo only isochoric motion, and the equations of motion are𝑉=0,𝑇=𝜌𝑑𝑉𝑑𝑡+𝜌𝑏,(2.3) where 𝜌 is the constant density of the fluid and 𝑏 is the body force. If the model (2.1) is required to be compatible with thermodynamics in the sense that all motions satisfy the Clausius-Duhem inequality and the assumption that the specific Helmholtz free energy is a minimum in equilibrium, then the material moduli must meet the following restrictions [41]:𝜇0,𝛼10,𝛼1+𝛼2=0.(2.4) The sign of the material moduli 𝛼1 and 𝛼2 has been the subject of much controversy. A comprehensive discussion on the restrictions given in (2.4) as well as a critical review on the fluids of differential type can be found in the extensive work by Dunn and Rajagopal [42].

For the problem under consideration, we shall assume a velocity field and an extra-stress of the form𝑉=𝑉(𝑟,𝑡)=𝑤(𝑟,𝑡)𝑒𝜃+𝑣(𝑟,𝑡)𝑒𝑧,(2.5) where 𝑒𝜃 and 𝑒𝑧 are unit vectors in the 𝜃 and 𝑧-directions of the cylindrical coordinate system 𝑟,𝜃 and 𝑧. For such flows the constraint, of incompressibility is automatically satisfied. If the fluid is at rest up to the moment 𝑡=0, then𝑉(𝑟,0)=0,(2.6) and (2.1) implies 𝑆𝑟𝑟=0 and the meaningful equations𝜏1(𝑟,𝑡)=𝜇+𝛼1𝜕𝜕𝜕𝑡1𝜕𝑟𝑟𝑤(𝑟,𝑡),𝜏2(𝑟,𝑡)=𝜇+𝛼1𝜕𝜕𝑡𝜕𝑣(𝑟,𝑡)𝜕𝑟,(2.7) where 𝜏1=𝑆𝑟𝜃 and 𝜏2=𝑆𝑟𝑧 are the shear stresses that are different of zero.

The equation of motion (2.3) 2, in the absence of a pressure gradient in the axial direction and neglecting body forces, leads to the relevant equations (𝜕𝜃𝑝=0 due to the rotational symmetry)𝜌𝜕𝑤(𝑟,𝑡)=𝜕𝜕𝑡+2𝜕𝑟𝑟𝜏1(𝑟,𝑡),𝜌𝜕𝑣(𝑟,𝑡)=𝜕𝜕𝑡+1𝜕𝑟𝑟𝜏2(𝑟,𝑡).(2.8) Eliminating 𝜏1 and 𝜏2 between (2.7) and (2.8), we attain the governing equations𝜕𝑤(𝑟,𝑡)=𝜕𝜕𝑡𝜈+𝛼𝜕𝜕𝑡2𝜕𝑟2+1𝑟𝜕1𝜕𝑟𝑟2𝑤(𝑟,𝑡),𝑟(0,𝑅),𝑡>0,𝜕𝑣(𝑟,𝑡)=𝜕𝜕𝑡𝜈+𝛼𝜕𝜕𝑡2𝜕𝑟2+1𝑟𝜕𝜕𝑟𝑣(𝑟,𝑡),𝑟(0,𝑅),𝑡>0,(2.9) where 𝜈=𝜇/𝜌 is the kinematic viscosity and 𝛼=𝛼1/𝜌 is the material parameter of the fluid. The governing equations corresponding to an incompressible fractionalized second grade fluid, performing the same motion, are𝜕𝑤(𝑟,𝑡)=𝜕𝑡𝜈+𝛼𝐷𝛽𝑡𝜕2𝜕𝑟2+1𝑟𝜕1𝜕𝑟𝑟2𝑤(𝑟,𝑡),𝑟(0,𝑅),𝑡>0,(2.10)𝜕𝑣(𝑟,𝑡)=𝜕𝑡𝜈+𝛼𝐷𝛽𝑡𝜕2𝜕𝑟2+1𝑟𝜕𝜕𝑟𝑣(𝑟,𝑡),𝑟(0,𝑅),𝑡>0,(2.11)𝜏1(𝑟,𝑡)=𝜇+𝛼1𝐷𝛽𝑡𝜕1𝜕𝑟𝑟𝑤(𝑟,𝑡),(2.12)𝜏2(𝑟,𝑡)=𝜇+𝛼1𝐷𝛽𝑡𝜕𝑣(𝑟,𝑡)𝜕𝑟,(2.13) where 0<𝛽<1 is the fractional parameter. Of course, the new material constant 𝛼1, although for simplicity we keep the same notation, tends to the original 𝛼1 as 𝛽1. The fractional differential operator so-called Caputo fractional operator 𝐷𝛽𝑡 defined by [43, 44]𝐷𝛽𝑡1𝑓(𝑡)=Γ(1𝛽)𝑡0𝑓(𝜏)(𝑡𝜏)𝛽𝑑𝜏,0𝛽<1,𝑑𝑓(𝑡)𝑑𝑡,𝛽=1,(2.14) and Γ() is the Gamma function.

3. Oscillating Flows of Fractionalized Second Grade Fluids

Let us consider an incompressible fractionalized second grade fluid at rest, in an infinitely long cylinder of radius 𝑅  as shown in Figure 1. At time 𝑡=0+, the cylinder starts to oscillate according to𝑉𝑊(𝑅,𝑡)=1𝐻𝜔(𝑡)cos1𝑡+𝑊2𝜔sin1𝑡𝑒𝜃+𝑉1𝐻𝜔(𝑡)cos2𝑡+𝑉2𝜔sin2𝑡𝑒𝑧,(3.1) where 𝜔1 and 𝜔2 are the frequencies of the velocity of the cylinder and 𝑉1,𝑉2,𝑊1, and 𝑊2 are constant amplitudes. Owing to the shear, the fluid in cylinder is gradually moved, its velocity being of the form (2.5). The governing equations are given by (2.10)–(2.13) while the associated initial and boundary conditions are𝑤(𝑟,0)=𝑣(𝑟,0)=0,𝑟(0,𝑅),(3.2) respectively, and𝑤(𝑅,𝑡)=𝑊1𝐻𝜔(𝑡)cos1𝑡+𝑊2𝜔sin1𝑡,𝑣(𝑅,𝑡)=𝑉1𝐻𝜔(𝑡)cos2𝑡+𝑉2𝜔sin2𝑡,𝑡0,(3.3) where 𝐻(𝑡) is the Heaviside function [45]. In the following, the system of fractional partial differential equations (2.10)–(2.13), with appropriate initial and boundary conditions, will be solved by means of Laplace and finite Hankel transforms. In order to avoid lengthy calculations of residues and contour integrals, the discrete inverse Laplace transform method will be used [2430].

3.1. Calculation of the Velocity Field

Applying the Laplace transform to (2.10) and (2.11) and having in mind the initial and boundary conditions (3.2) and (3.4), we find that𝑞𝑤(𝑟,𝑞)=𝜈+𝛼𝑞𝛽𝜕2𝜕𝑟2+1𝑟𝜕1𝜕𝑟𝑟2𝑤(𝑟,𝑞),𝑟(0,𝑅),(3.4)𝑞𝑣(𝑟,𝑞)=𝜈+𝛼𝑞𝛽𝜕2𝜕𝑟2+1𝑟𝜕𝜕𝑟𝑣(𝑟,𝑞),𝑟(0,𝑅),(3.5) where the image functions 𝑤(𝑟,𝑞) and 𝑣(𝑟,𝑞) of 𝑤(𝑟,𝑡) and 𝑣(𝑟,𝑡) have to satisfy the conditions𝑊𝑤(𝑅,𝑞)=1𝑞+𝑊2𝜔1𝑞2+𝜔21,𝑉𝑣(𝑅,𝑞)=1𝑞+𝑉2𝜔2𝑞2+𝜔22.(3.6)

Multiplying now both sides of (3.4) and (3.5) by 𝑟𝐽1(𝑟𝑟𝑚) and 𝑟𝐽0(𝑟𝑟𝑛), respectively, integrating them with respect to 𝑟 from 0 to 𝑅 and taking into account the conditions (3.6) and the known relations [46, 47]𝑅0𝑟𝜕2𝑤𝜕𝑟2+1𝑟𝜕𝑤𝜕𝑟𝑤𝑟2𝐽1𝑟𝑟𝑚𝑑𝑟=𝑅𝑟𝑚𝐽2𝑅𝑟𝑚𝑤(𝑅,𝑡)𝑟2𝑚𝑤𝐻𝑟𝑚,,𝑡𝑅0𝑟𝜕2𝑣𝜕𝑟2+1𝑟𝜕𝑣𝐽𝜕𝑟0𝑟𝑟𝑛𝑑𝑟=𝑅𝑟𝑛𝐽1𝑅𝑟𝑛𝑣(𝑅,𝑡)𝑟2𝑛𝑣𝐻𝑟𝑛,,𝑡(3.7) we find that𝑤𝐻𝑟𝑚,𝑞=𝑅𝑟𝑚𝐽2𝑅𝑟𝑚𝑊1𝑞+𝑊2𝜔1𝑞2+𝜔21𝜈+𝛼𝑞𝛽𝑞+𝛼𝑟2𝑚𝑞𝛽+𝜈𝑟2𝑚,𝑣𝐻𝑟𝑛,𝑞=𝑅𝑟𝑛𝐽1𝑅𝑟𝑚𝑉1𝑞+𝑉2𝜔2𝑞2+𝜔22𝜈+𝛼𝑞𝛽𝑞+𝛼𝑟2𝑛𝑞𝛽+𝜈𝑟2𝑛,(3.8) where [46, 47]𝑤𝐻𝑟𝑚=,𝑞𝑅0𝑟𝑤(𝑟,𝑞)𝐽1𝑟𝑟𝑚𝑑𝑟,𝑣𝐻𝑟𝑛=,𝑞𝑅0𝑟𝑣(𝑟,𝑞)𝐽0𝑟𝑟𝑛𝑑𝑟,𝑚,𝑛=1,2,3,(3.9) are the Hankel transforms of 𝑤(𝑟,𝑞) and 𝑣(𝑟,𝑞), while 𝑟𝑚 and 𝑟𝑛 are the positive roots of the transcendental equations 𝐽1(𝑅𝑟)=0 and 𝐽0(𝑅𝑟)=0, respectively. In order to determine 𝑤(𝑟,𝑞) and 𝑣(𝑟,𝑞), we must apply the inverse Hankel transforms. However, for a more suitable presentation of final results, we firstly rewrite in (3.8), in the equivalent forms:𝑤𝐻𝑟𝑚=,𝑞𝑅𝐽2𝑅𝑟𝑚𝑟𝑚𝑊1𝑞+𝑊2𝜔1𝑞2+𝜔21𝑅𝐽2𝑅𝑟𝑚𝑟𝑛𝑊1𝑞+𝑊2𝜔1𝑞2+𝜔21𝑞𝑞+𝛼𝑟2𝑚𝑞𝛽+𝜈𝑟2𝑚,𝑣𝐻𝑟𝑛=,𝑞𝑅𝐽1𝑅𝑟𝑛𝑟𝑛𝑉1𝑞+𝑉2𝜔1𝑞2+𝜔21𝑅𝐽1𝑅𝑟𝑛𝑟𝑛𝑉1𝑞+𝑉2𝜔2𝑞2+𝜔21𝑞𝑞+𝛼𝑟2𝑛𝑞𝛽+𝜈𝑟2𝑛,(3.10) and apply the inverse Hankel transform formulae [46, 47]2𝑤(𝑟,𝑞)=𝑅2𝑚=1𝑤𝐻𝑟𝑚𝐽,𝑞1𝑟𝑟𝑚𝐽22𝑅𝑟𝑚,2𝑣(𝑟,𝑞)=𝑅2𝑛=1𝑣𝐻𝑟𝑛𝐽,𝑞0𝑟𝑟𝑛𝐽21𝑅𝑟𝑛.(3.11) Taking into account the following results [47]:𝑅0𝑟2𝐽1𝑟𝑟𝑚𝑅𝑑𝑟=2𝑟𝑚𝐽2𝑅𝑟𝑚,𝑅0𝑟𝐽0𝑟𝑟𝑛𝑅𝑑𝑟=𝑟𝑛𝐽1𝑅𝑟𝑛,(3.12) we find that𝑅𝑤(𝑟,𝑞)=𝑟𝑊1𝑞+𝑊2𝜔1𝑞2+𝜔212𝑅𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚𝑊1𝑞+𝑊2𝜔1𝑞2+𝜔21𝑞𝑞+𝛼𝑟2𝑚𝑞𝛽+𝜈𝑟2𝑚,𝑉𝑣(𝑟,𝑞)=1𝑞+𝑉2𝜔2𝑞2+𝜔222𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽2𝑅𝑟𝑛𝑉1𝑞+𝑉2𝜔2𝑞2+𝜔22𝑞𝑞+𝛼𝑟2𝑛𝑞𝛽+𝜈𝑟2𝑛.(3.13) Finally, in order to obtain 𝑤(𝑟,𝑡)=1{𝑤(𝑟,𝑞)} and 𝑣(𝑟,𝑡)=1{𝑣(𝑟,𝑞)} and to avoid the lengthy calculations of residues and contour integrals, we will apply the discrete inverse Laplace transform method [2430]. For this, we firstly write (3.13) in series form𝑅𝑤(𝑟,𝑞)=𝑟𝑊1𝑞+𝑊2𝜔1𝑞2+𝜔212𝑅𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚𝑗=0𝜔21𝑗𝑘=0𝛼𝑟2𝑚𝑘×𝑊1𝑖=0(𝑘)𝑖(𝜈/𝛼)𝑖1𝑖!𝑞(𝑖𝑘)𝛽+𝑘+2𝑗+1+𝑊2𝜔1𝑖=0(𝑘)𝑖(𝜈/𝛼)𝑖1𝑖!𝑞(𝑖𝑘)𝛽+𝑘+2𝑗+2,𝑉𝑣(𝑟,𝑞)=1𝑞+𝑉2𝜔2𝑞2+𝜔222𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽1𝑅𝑟𝑛𝑗=0𝜔22𝑗𝑘=0𝛼𝑟2𝑛𝑘×𝑉1𝑖=0(𝑘)𝑖(𝜈/𝛼)𝑖1𝑖!𝑞(𝑖𝑘)𝛽+𝑘+2𝑗+1+𝑉2𝜔2𝑖=0(𝑘)𝑖(𝜈/𝛼)𝑖1𝑖!𝑞(𝑖𝑘)𝛽+𝑘+2𝑗+2,(3.14) where we used the fact that𝑘𝑖=(1)𝑖(𝑘)𝑖𝑖!,(3.15) and (𝑘)𝑖 is the Pochhammer symbol(𝑘)𝑖=1,𝑖=0,𝑘(𝑘+1)(𝑘+𝑖1),𝑖N.(3.16) In particular (0)0=1, (𝑘)0=1 and (0)𝑖=0, for 𝑖N. Applying the discrete inverse Laplace transform, we get𝑟𝑤(𝑟,𝑞)=𝑅𝑊1𝜔𝐻(𝑡)cos1𝑡+𝑊2𝜔sin1𝑡2𝑅𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚𝑗=0𝜔21𝑗𝑘=0𝛼𝑟2𝑚𝑘×𝑊1𝐻(𝑡)𝑖=0(𝑘)𝑖(𝜈/𝛼)𝑖𝑡(𝑖𝑘)𝛽+𝑘+2𝑗𝑖!Γ((𝑖𝑘)𝛽+𝑘+2𝑗+1)+𝑊2𝜔1𝑖=0(𝑘)𝑖(𝜈/𝛼)𝑖𝑡(𝑖𝑘)𝛽+𝑘+2𝑗+1,𝑖!Γ((𝑖𝑘)𝛽+𝑘+2𝑗+2)𝑣(𝑟,𝑞)=𝑉1𝜔𝐻(𝑡)cos2𝑡+𝑉2𝜔sin2𝑡2𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽1𝑅𝑟𝑛𝑗=0𝜔22𝑗𝑘=0𝛼𝑟2𝑛𝑘×𝑉1𝐻(𝑡)𝑖=0(𝑘)𝑖(𝜈/𝛼)𝑖𝑡𝑘((𝑖𝑘)𝛽+𝑘+2𝑗)𝑖!Γ((𝑖𝑘)𝛽+𝑘+2𝑗+1)+𝑉2𝜔2𝑖=0(𝑘)𝑖(𝜈/𝛼)𝑖𝑡(𝑖𝑘)𝛽+𝑘+2𝑗+1.𝑖!Γ((𝑖𝑘)𝛽+𝑘+2𝑗+2)(3.17) In terms of the generalized 𝐺𝑎,𝑏,𝑐(,𝑡) functions [48], we rewrite the above equations in simple forms:𝑟𝑤(𝑟,𝑞)=𝑅𝑊1𝜔𝐻(𝑡)cos1𝑡+𝑊2𝜔sin1𝑡2𝑅𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚𝑗=0𝜔21𝑗𝑘=0𝛼𝑟2𝑚𝑘×𝑊1𝐻(𝑡)𝐺𝛽,𝑘2𝑗1,𝑘𝜈𝛼,𝑡+𝑊2𝜔1𝐺𝛽,𝑘2𝑗2,𝑘𝜈𝛼,,𝑡(3.18)𝑣(𝑟,𝑞)=𝑉1𝜔𝐻(𝑡)cos2𝑡+𝑉2𝜔sin2𝑡2𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽1𝑅𝑟𝑛𝑗=0𝜔22𝑗𝑘=0𝛼𝑟2𝑛𝑘×𝑉1𝐻(𝑡)𝐺𝛽,𝑘2𝑗1,𝑘𝜈𝛼,𝑡+𝑉2𝜔2𝐺𝛽,𝑘2𝑗2,𝑘𝜈𝛼,,𝑡(3.19) where the generalized 𝐺𝑎,𝑏,𝑐(,𝑡) function is defined by [48]𝐺𝑎,𝑏,𝑐(𝑑,𝑡)=𝑗=0(𝑐)𝑗𝑑𝑗𝑡𝑗!(𝑐+𝑗)𝑎𝑏1Γ[]||||𝑑(𝑐+𝑗)𝑎𝑏,Re(𝑎𝑐𝑏)>0,Re(𝑞)>0,𝑞𝑎||||<1.(3.20)

3.2. Calculation of the Shear Stress

Applying the Laplace transform to (2.12) and (2.13), we find that𝜏1(𝑟,𝑞)=𝜇+𝛼1𝑞𝛽𝜕1𝜕𝑟𝑟𝑤(𝑟,𝑞),𝜏2(𝑟,𝑞)=𝜇+𝛼1𝑞𝛽𝜕𝑣(𝑟,𝑞),𝜕𝑟(3.21) where𝜕𝑤(𝑟,𝑞)1𝜕𝑟𝑟2𝑤(𝑟,𝑞)=𝑅𝑚=1𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚𝑊1𝑞+𝑊2𝜔1𝑞2+𝜔21𝑞𝑞+𝛼𝑟2𝑚𝑞𝛽+𝜈𝑟2𝑚,𝜕𝑣(𝑟,𝑞)=2𝜕𝑟𝑅𝑚=1𝐽1𝑟𝑟𝑚𝐽1𝑅𝑟𝑚𝑉1𝑞+𝑉2𝜔2𝑞2+𝜔22𝑞𝑞+𝛼𝑟2𝑛𝑞𝛽+𝜈𝑟2𝑛,(3.22) have been obtained form (3.13), using the identities𝑟𝑟𝑚𝐽1𝑟𝑟𝑚𝐽1𝑟𝑟𝑚=𝑟𝑟𝑚𝐽2𝑟𝑟𝑚,𝐽0𝑟𝑟𝑛=𝑟𝑛𝐽1𝑟𝑟𝑛.(3.23) Substituting (3.22) into (3.21), respectively, and applying again the discrete inverse Laplace transform method, we find that the shear stresses 𝜏1(𝑟,𝑡) and 𝜏2(𝑟,𝑡) have the following forms:𝜏1(𝑟,𝑡)=2𝛼𝜌𝑅𝑚=1𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚𝑗=0𝜔21𝑗𝑘=0𝛼𝑟2𝑚𝑘𝑊1𝐻(𝑡)𝐺𝛽,𝑘2𝑗1,𝑘1𝜈𝛼,𝑡+𝑊2𝜔1𝐺𝛽,𝑘2𝑗2,𝑘1𝜈𝛼,𝜏,𝑡2(𝑟,𝑡)=2𝛼𝜌𝑅𝑛=1𝐽1𝑟𝑟𝑛𝐽1𝑅𝑟𝑛𝑗=0𝜔22𝑗𝑘=0𝛼𝑟2𝑛𝑘𝑉1𝐻(𝑡)𝐺𝛽,𝑘2𝑗1,𝑘1𝜈𝛼,𝑡+𝑉2𝜔2𝐺𝛽,𝑘2𝑗2,𝑘1𝜈𝛼.,𝑡(3.24)

4. Limiting Cases

4.1. Ordinary Second Grade Fluid (𝛽1)

Making 𝛽1 into (3.18), (3.19), (3.24), we obtain the solutions𝑤OSG(𝑟𝑟,𝑞)=𝑅𝑊1𝜔𝐻(𝑡)cos1𝑡+𝑊2𝜔sin1𝑡2𝑅𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚𝑗=0𝜔21𝑗𝑘=0𝛼𝑟2𝑚𝑘×𝑊1𝐻(𝑡)𝐺1,𝑘2𝑗1,𝑘𝜈𝛼,𝑡+𝑊2𝜔1𝐺1,𝑘2𝑗2,𝑘𝜈𝛼,𝑣,𝑡OSG(𝑟,𝑞)=𝑉1𝜔𝐻(𝑡)cos2𝑡+𝑉2𝜔sin2𝑡2𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽1𝑅𝑟𝑛𝑗=0𝜔22𝑗𝑘=0𝛼𝑟2𝑛𝑘×𝑉1𝐻(𝑡)𝐺1,𝑘2𝑗1,𝑘𝜈𝛼,𝑡+𝑉2𝜔2𝐺1,𝑘2𝑗2,𝑘𝜈𝛼,𝜏,𝑡1OSG(𝑟,𝑡)=2𝛼𝜌𝑅𝑚=1𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚𝑗=0𝜔21𝑗𝑘=0𝛼𝑟2𝑚𝑘×𝑊1𝐻(𝑡)𝐺1,𝑘2𝑗1,𝑘1𝜈𝛼,𝑡+𝑊2𝜔1𝐺1,𝑘2𝑗2,𝑘1𝜈𝛼,𝜏,𝑡2OSG(𝑟,𝑡)=2𝛼𝜌𝑅𝑛=1𝐽1𝑟𝑟𝑛𝐽1𝑅𝑟𝑛𝑗=0𝜔22𝑗𝑘=0𝛼𝑟2𝑛𝑘×𝑉1𝐻(𝑡)𝐺1,𝑘2𝑗1,𝑘1𝜈𝛼,𝑡+𝑉2𝜔2𝐺1,𝑘2𝑗2,𝑘1𝜈𝛼,,𝑡(4.1) corresponding to an ordinary second grade fluid, performing the same motion. Other equivalent forms of solutions for ordinary second grade fluids can be directly obtained from (3.10) by substituting 𝛽=1, and performing the inverse Laplace transform. The expressions for velocity field are given by𝑤OSG(𝑟,𝑡)=𝑤OSS(𝑟,𝑡)+𝑤OST(𝑟,𝑡),𝑣𝑆(𝑟,𝑡)=𝑣OSS(𝑟,𝑡)+𝑣OST(𝑟,𝑡),(4.2) where𝑤OSS=𝑟𝑅𝑊1𝜔𝐻(𝑡)cos1𝑡+𝑊2𝜔sin1𝑡2𝜔1𝑅𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚×𝑊1𝜔𝐻(𝑡)11+𝛼𝑟2𝑚𝜔cos1𝑡𝜈𝑟2𝑚𝜔sin1𝑡𝜈2𝑟4𝑚+𝜔211+𝛼𝑟2𝑚2+𝑊2𝜈𝑟2𝑚𝜔cos1𝑡+𝜔11+𝛼𝑟2𝑚𝜔sin1𝑡𝜈2𝑟4𝑚+𝜔211+𝛼𝑟2𝑚2,𝑤OST=2𝜈𝑅𝑚=1𝑟𝑚𝐽1𝑟𝑟𝑚𝐽2𝑅𝑟𝑚𝑊1𝐻(𝑡)𝜈𝑟2𝑚𝑊2𝜔11+𝛼𝑟2𝑚1+𝛼𝑟2𝑚𝜈2𝑟4𝑚+𝜔211+𝛼𝑟2𝑚2exp𝜈𝑟2𝑚𝑡1+𝛼𝑟2𝑚,𝑣OSS=𝑉1𝜔𝐻(𝑡)cos2𝑡+𝑉2𝜔sin2𝑡2𝜔2𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽1𝑅𝑟𝑛×𝑉1𝜔𝐻(𝑡)21+𝛼𝑟2𝑛𝜔cos2𝑡𝜈𝑟2𝑛𝜔sin2𝑡𝜈2𝑟4𝑛+𝜔221+𝛼𝑟2𝑛2+𝑉2𝜈𝑟2𝑛𝜔cos2𝑡+𝜔21+𝛼𝑟2𝑛𝜔sin2𝑡𝜈2𝑟4𝑛+𝜔221+𝛼𝑟2𝑛2,𝑤OST=2𝜈𝑅𝑛=1𝑟𝑛𝐽0𝑟𝑟𝑛𝐽1𝑅𝑟𝑛𝑉1𝐻(𝑡)𝜈𝑟2𝑛𝑉2𝜔21+𝛼𝑟2𝑛1+𝛼𝑟2𝑛𝜈2𝑟4𝑛+𝜔221+𝛼𝑟2𝑛2exp𝜈𝑟2𝑛𝑡1+𝛼𝑟2𝑛,(4.3) are the steady-state and transient solutions. Introducing (4.2) into (2.7), we find that𝜏1OSG(𝑟,𝑡)=𝜏1OSS(𝑟,𝑡)+𝜏1OST(𝑟,𝑡),𝜏2OSG(𝑟,𝑡)=𝜏2OSS(𝑟,𝑡)+𝜏2OST(𝑟,𝑡),(4.4) where the steady-state and transient components are given by𝜏1OSS=2𝜌𝜔1𝑅𝑚=1𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚1𝜈2𝑟4𝑚+𝜔211+𝛼𝑟2𝑚2×𝑊1𝐻(𝑡)𝜈𝜔1𝜔cos1𝑡𝜈2𝑟2𝑚+𝛼𝜔211+𝛼𝑟2𝑚𝜔sin1𝑡+𝑊2𝜈2𝑟2𝑚+𝛼𝜔211+𝛼𝑟2𝑚𝜔cos1𝑡+𝜈𝜔1𝜔sin1𝑡,(4.5)𝜏1OST=2𝜌𝜈2𝑅𝑚=1𝑟2𝑚𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚𝑊1𝐻(𝑡)𝜈𝑟2𝑚𝑊2𝜔11+𝛼𝑟2𝑚1+𝛼𝑟2𝑚2𝜈2𝑟4𝑚+𝜔211+𝛼𝑟2𝑚2exp𝜈𝑟2𝑚𝑡1+𝛼𝑟2𝑚,(4.6)𝜏2OSS=2𝜌𝜔2𝑅𝑛=1𝐽1𝑟𝑟𝑛𝐽1𝑅𝑟𝑚1𝜈2𝑟4𝑛+𝜔221+𝛼𝑟2𝑛2×𝑉1𝐻(𝑡)𝜈𝜔2𝜔cos2𝑡𝜈2𝑟2𝑛+𝛼𝜔221+𝛼𝑟2𝑛𝜔sin2𝑡+𝑉2𝜈2𝑟2𝑛+𝛼𝜔221+𝛼𝑟2𝑛𝜔cos2𝑡+𝜈𝜔2𝜔sin2𝑡,(4.7)𝜏2OST=2𝜌𝜈2𝑅𝑚=1𝑟2𝑛𝐽1𝑟𝑟𝑚𝐽1𝑅𝑟𝑛𝑉1𝐻(𝑡)𝜈𝑟2𝑛𝑉2𝜔21+𝛼𝑟2𝑛1+𝛼𝑟2𝑛2𝜈2𝑟4𝑛+𝜔221+𝛼𝑟2𝑛2exp𝜈𝑟2𝑛𝑡1+𝛼𝑟2𝑛.(4.8)

In practice, the steady-state solutions for unsteady motions of Newtonian or non-Newtonian fluids are important for those who need to eliminate transients from their rheological measurements. Consequently, an important problem regarding the technical relevance of these solutions is to find the approximate time after which the fluid is moving according to the steady-state. More exactly, in practice it is necessary to know the required time to reach the steady-state.

4.2. Newtonian Fluids (𝛼10)

Making the limit 𝛼1 and then 𝛼0 into (3.13), (3.22), and proceeding as in the last section, the solutions for a Newtonian fluid𝑤𝑁(𝑟𝑟,𝑞)=𝑅𝑊1𝜔𝐻(𝑡)cos1𝑡+𝑊2𝜔sin1𝑡2𝑅𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚𝑗=0𝜔21𝑗×𝑊1𝐻(𝑡)𝐺1,2𝑗,1𝜈𝑟2𝑚,𝑡+𝑊2𝜔2𝐺1,2𝑗1,1𝜈𝑟2𝑚,𝑣,𝑡𝑁(𝑟,𝑞)=𝑉1𝜔𝐻(𝑡)cos2𝑡+𝑉2𝜔sin2𝑡2𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽1𝑅𝑟𝑛𝑗=0𝜔21𝑗×𝑉1𝐻(𝑡)𝐺1,2𝑗,1𝜈𝑟2𝑛,𝑡+𝑉2𝜔2𝐺1,2𝑗1,1𝜈𝑟2𝑛,𝜏,𝑡1𝑁(𝑟,𝑡)=2𝜇𝑅𝑚=1𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚×𝑗=0𝜔21𝑗𝑊1𝐻(𝑡)𝐺1,2𝑗,1𝜈𝑟2𝑚,𝑡+𝑊2𝜔1𝐺1,2𝑗1,1𝜈𝑟2𝑚,𝜏,𝑡2𝑁(𝑟,𝑡)=2𝜇𝑅𝑛=1𝐽1𝑟𝑟𝑛𝐽1𝑅𝑟𝑛×𝑗=0𝜔22𝑗𝑉1𝐻(𝑡)𝐺1,2𝑗,1𝜈𝑟2𝑛,𝑡+𝑉2𝜔2𝐺1,2𝑗1,1𝜈𝑟2𝑛,,𝑡(4.9) are obtained. Similarly by making 𝛼1 and then 𝛼0 into (4.2)–(4.8), the corresponding solutions𝑤𝑁(𝑟,𝑡)=𝑤𝑁𝑆(𝑟,𝑡)+𝑤𝑁𝑇(𝑟,𝑡),𝑣𝑁(𝑟,𝑡)=𝑣𝑁𝑆(𝑟,𝑡)+𝑣𝑁𝑇(𝑟,𝑡),(4.10)𝜏1𝑁(𝑟,𝑡)=𝜏1𝑁𝑆(𝑟,𝑡)+𝜏1𝑁𝑇(𝑟,𝑡),𝜏2𝑁(𝑟,𝑡)=𝜏2𝑁𝑆(𝑟,𝑡)+𝜏2𝑁𝑇(𝑟,𝑡),(4.11) where𝑤𝑁𝑆=𝑟𝑅𝑊1𝜔𝐻(𝑡)cos1𝑡+𝑊2𝜔sin1𝑡2𝜔1𝑅𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚×𝑊1𝜔𝐻(𝑡)1𝜔cos1𝑡𝜈𝑟2𝑚𝜔sin1𝑡+𝑊2𝜈𝑟2𝑚𝜔cos1𝑡+𝜔1𝜔sin1𝑡𝜈2𝑟4𝑚+𝜔21,𝑤(4.12)𝑁𝑇=2𝜈𝑅𝑚=1𝑟𝑚𝐽1𝑟𝑟𝑚𝐽2𝑅𝑟𝑚𝑊1𝐻(𝑡)𝜈𝑟2𝑚𝑊2𝜔1𝜈2𝑟4𝑚+𝜔21𝑒𝜈𝑟2𝑚𝑡,𝑣(4.13)𝑁𝑆=𝑉1𝜔𝐻(𝑡)cos2𝑡+𝑉2𝜔sin2𝑡2𝜔2𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽1𝑅𝑟𝑛×𝑉1𝜔𝐻(𝑡)2𝜔cos2𝑡𝜈𝑟2𝑛𝜔sin2𝑡+𝑉2𝜈𝑟2𝑛𝜔cos2𝑡+𝜔2𝜔sin2𝑡𝜈2𝑟4𝑛+𝜔22,𝑣(4.14)𝑁𝑇=2𝜈𝑅𝑛=1𝑟𝑛𝐽0𝑟𝑟𝑛𝐽1𝑅𝑟𝑛𝑉1𝐻(𝑡)𝜈𝑟2𝑛𝑉2𝜔2𝜈2𝑟4𝑛+𝜔22𝑒𝜈𝑟2𝑚𝑡,𝜏(4.15)1𝑁𝑆=2𝜌𝜔1𝑅𝑚=1𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚1𝜈2𝑟4𝑚+𝜔21×𝑊1𝐻(𝑡)𝜈𝜔1𝜔cos1𝑡𝜈2𝑟2𝑚𝜔sin1𝑡+𝑊2𝜈2𝑟2𝑚𝜔cos1𝑡+𝜈𝜔1𝜔sin1𝑡,𝜏(4.16)1𝑁𝑇=2𝜌𝜈2𝑅𝑚=1𝑟2𝑚𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚𝑊1𝐻(𝑡)𝜈𝑟2𝑚𝑊2𝜔1𝜈2𝑟4𝑚+𝜔21𝑒𝜈𝑟2𝑚𝑡,𝜏(4.17)2𝑁𝑆=2𝜌𝜔2𝑅𝑛=1𝐽1𝑟𝑟𝑛𝐽1𝑅𝑟𝑚1𝜈2𝑟4𝑛+𝜔22×𝑉1𝐻(𝑡)𝜈𝜔2𝜔cos2𝑡𝜈2𝑟2𝑛𝜔sin2𝑡+𝑉2𝜈2𝑟2𝑛𝜔cos2𝑡+𝜈𝜔2𝜔sin2𝑡,𝜏(4.18)2𝑁𝑇=2𝜌𝜈2𝑅𝑚=1𝑟2𝑛𝐽1𝑟𝑟𝑚𝐽1𝑅𝑟𝑛𝑉1𝐻(𝑡)𝜈𝑟2𝑛𝑉2𝜔2𝜈2𝑟4𝑛+𝜔22𝑒𝜈𝑟2𝑛𝑡(4.19) for Newtonian fluids are obtained. Substituting 𝑊1=𝑅Ω,𝑊2=0,𝑉1=𝑈,𝑉2=0, 𝜔1=0, and 𝜔2=0 in (4.9) and using the definition of generalized 𝐺𝑎,𝑏,𝑐(,𝑡) functions, the solutions𝑤𝑁(𝑟,𝑡)=𝑟Ω2Ω𝑚=1𝐽1𝑟𝑟𝑚𝑟𝑚𝐽2𝑅𝑟𝑚𝑒𝜈𝑟2𝑚𝑡,𝑣𝑁(𝑟,𝑡)=𝑈2𝑈𝑅𝑛=1𝐽0𝑟𝑟𝑛𝑟𝑛𝐽1𝑅𝑟𝑛𝑒𝜈𝑟2𝑛𝑡,𝜏1𝑁(𝑟,𝑡)=2𝜇Ω𝑚=1𝐽2𝑟𝑟𝑚𝐽2𝑅𝑟𝑚𝑒𝜈𝑟2𝑚𝑡,𝜏2𝑁(𝑟,𝑡)=2𝜇𝑈𝑅𝑛=1𝐽1𝑟𝑟𝑛𝐽1𝑅𝑟𝑛𝑒𝜈𝑟2𝑛𝑡,(4.20) obtained in [49, equations (36)–(39)] by a different technique are recovered. Of course the above expressions can also be obtained form (4.10)–(4.19).

5. Numerical Results and Conclusions

The velocity fields and the adequate shear stresses corresponding to the unsteady motions of an incompressible fractionalized second grade fluid due to longitudinal and torsional oscillations of an infinite circular cylinder have been determined by means of the Laplace and finite Hankel transforms. The general solutions are written in series form in term of generalized 𝐺𝑎,𝑏,𝑐(,𝑡) functions and satisfy all imposed initial and boundary conditions. The solutions for ordinary second grade and Newtonian fluids performing the same motion are obtain as special cases of general solutions. Furthermore, another equivalent solutions for ordinary second grade and Newtonian fluids are presented, in terms of steady-state and transient solutions. They describe the motion of the fluid sometime after its initiation. After that time, when the transients disappear, they tend to the steady-state solutions, which are periodic in time and independent of the initial conditions. It is also shown that for 𝑊1=𝑅Ω, 𝑊2=0, 𝑉1=𝑈, 𝑉2=0, 𝜔1=0 and 𝜔2=0, (4.9) reduce to the well-known classical solutions [49, equations (36)–(39)]. The similar solutions corresponding to the sine and cosine oscillations of the boundary are immediately obtained by making 𝑉2=𝑊2=0, respectively, 𝑉1=𝑊1=0 into general solutions.

Now, in order to reveal some relevant physical aspects of the obtained results, the diagrams of the velocity components 𝑤(𝑟,𝑡) and 𝑣(𝑟,𝑡) are depicted against 𝑟 for different values of 𝑡 and the pertinent parameters. Figure 2 contains the diagrams of the velocity components for four different times 𝑡; it is obvious to see the impact of rigid boundary of the cylinder on the motion of the fluid. Further, the amplitude of oscillations for the two components of the velocity decreases with increasing values of 𝑡. However, this conclusion cannot be generalized. The influence of the frequency of oscillations 𝜔1 and 𝜔2, on fluid motion is shown in Figure 3. The amplitudes of both components of velocity are decreasing functions of frequency of oscillations 𝜔1 and 𝜔2 respectively. The effect of material parameter 𝛼 on fluid motion is discussed in Figure 4. The influence of material parameter 𝛼 is quite similar to that of Figures 2 and 3. Nowadays fractional derivative approach in viscoelastic fluid plays an important role to describe the behavior of complex fluid. Therefore, it is important to see the effect of fractional parameter on oscillating fluid. Figure 5 depict the influence of fractional parameter 𝛽 on fluid motion. It is again clear that the amplitude of fluid oscillations decreases with respect to fractional parameter 𝛽. The viscosity is an important property of the fluid. It is observed that the amplitude of oscillations is an increasing function of kinematic viscosity 𝜈 in this geometry as shown in Figure 6. The influence of radius 𝑟 against time 𝑡 is shown in Figure 7. The oscillating behavior of fluid motion clearly results from these figures. As expected, the amplitude of oscillations increases with the increasing values of 𝑟. The influence of 𝑊1 and 𝑊2 on rotational component 𝑤(𝑟,𝑡) and the effect of 𝑉1 and 𝑉2 on longitudinal component 𝑣(𝑟,𝑡) are presented in Figures 8 and 9. The influence of 𝑊1 and 𝑊2 on 𝑤(𝑟,𝑡) is quite opposite. For instance, 𝑤(𝑟,𝑡) is a decreasing function with respect to 𝑊1 and an increasing one of 𝑊2. The longitudinal component 𝑣(𝑟,𝑡) is an increasing function of 𝑉1 near the center of cylinder and of 𝑉2 on the whole flow domain.

Finally, for comparison, the diagrams of 𝑤(𝑟,𝑡) and 𝑣(𝑟,𝑡) corresponding to the three models, fractionalized second grade (𝛽=0.3 and 𝛽=0.6), ordinary second grade (𝛽=1), and Newtonian fluids (𝛼=0 and 𝛽=1) are presented in Figures 1013. It is clearly seen from Figure 10 that the fractionalized second grade fluid is the swiftest and the Newtonian one is the slowest. However, the behavior of these models is quite opposite at time 𝑡=4s as shown by Figure 11. The large time effect on oscillating fluid is shown in Figures 12 and 13. It is observed that for large time the non-Newtonian effects can be neglected for rotational component of velocity 𝑤(𝑟,𝑡). This seems to be not true for the longitudinal component 𝑣(𝑟,𝑡) of the velocity. The units of the material constants in Figures 213 are SI units and the roots 𝑟𝑚 and 𝑟𝑛 have been approximated by (4𝑚+1)𝜋/4𝑅 and (4𝑛1)𝜋/4𝑅, respectively.

Acknowledgments

The author M. Jamil is extremely grateful and thankful to the Abdus Salam School of Mathematical Sciences, GC University, Lahore, Pakistan; Department of Mathematics & Basic Sciences, NED University of Engineering & Technology, Karachi, Pakistan; also Higher Education Commission of Pakistan for supporting and facilitating this research work. The author N. A. Khan is highly thankful and grateful to the Dean of Faculty of Sciences, University of Karachi, Karachi, Pakistan for supporting and facilitating this research work. The author A. Rauf is highly thankful and grateful to the Abdus Salam School of Mathematical Sciences, GC University, Lahore, Pakistan and also Higher Education Commission of Pakistan for generous support and facilitating this research work.