Taylor–Couette system in the narrow-gap limit, revisited, with a corrigendum to the paper by Nagata (2023, Philosophical Transaction A.)
Abstract
We study transition from circular Couette flow to Taylor–Couette flow in the narrow-gap limit. We find that subcritical bifurcations of a nonlinear axisymmetric flow occur at the linear critical point when the angular velocity ratio for the cylinders, , is less than approximately −0.5. We also demonstrate that two nonlinear axisymmetric flow branches with the same wavenumber as that for the critical mode exist for . Introducing the modified Taylor number , analyses for any are made possible, so that this study supplements the recent analysis of Nagata M. 2023 Taylor-couette flow in the narrow-gap limit. Philos. Trans. R. Soc. A 381, 20220134. (doi:10.1098/rsta.2022.0134), in which the case was excluded, because the critical value of the conventional Taylor number used therein approaches zero as . The present analysis takes into account all the contributions from the inertia term, including a component which was neglected in [1]. The effect of the inclusion of this previously neglected component is surveyed at the end.
1. Introduction
Taylor–Couette flow is one of the most heavily explored flow configurations in fluid mechanics that serves as a theoretical and experimental testing ground to analyse the problem of flow instabilities and the transition to turbulence. In the seminal experimental work [1] by Andereck et al. [2], various types of flow state were identified in the parameter space based on the Reynolds numbers, and , which are defined with respect to the outer and the inner cylinder angular velocities and their radii. The radius ratio used in this experiment was 0.883, and one necessary question to answer is how the experimental observations of the transitions from one flow to another change when the gap is reduced to zero, that is, when the radii are increased to infinity. In this so-called narrow-gap limit, the circular geometry of the cylinder walls approaches a plane geometry.
The similarity between Taylor–Couette flow and Rayleigh–Bénard convection is well-known ([3,4] and [5]) and, in fact, axisymmetric Taylor–Couette flow in the narrow-gap limit becomes mathematically equivalent to two-dimensional Rayleigh–Bénard convection of a fluid with a unit Prandtl number (see, for instance, [6]).
Recently, while large-Rayleigh-number thermal convection has been investigated numerically [7] and experimentally [8], high-Reynolds-number Taylor–Couette turbulence has also gathered attention ([9] and [10], also see references in [11]).
The analogy between Taylor–Couette flow and Rayleigh–Bénard convection is utilized for the comparison of physical quantities, such as momentum transport and energy dissipation, occurring in the turbulent states of the two flows. In particular, the similarity has been investigated using plane Couette flow and rotating plane Couette flow, as these flows are believed to correspond to Taylor–Couette flow in the limit of a narrow gap between the cylinders ([12] and [13]).
Noting that there exists a distinct difference between Taylor–Couette flow and plane Couette flow, namely, that Taylor–Couette flow becomes unstable to axisymmetric disturbances, whereas plane Couette flow is always stable to infinitesimal disturbances, Nagata [1] reviewed the way of representing the Taylor–Couette system in the narrow gap in a Cartesian coordinate system.
The experimental observations by [2,14] and [15] for and , respectively, where is the radius ratio of the inner cylinder of radius to the outer cylinder of radius , show that the onset of instability of circular Couette flow occurs at a Reynolds number (based on the inner cylinder motion) of order 100 when the outer cylinder is at rest, for example. Given such moderate Reynolds numbers, the study of [1] neglected the nonlinear perturbation term, , resulting from the inertia term of the Navier–Stokes equations expressed in the original cylindrical coordinate system (see the last paragraph of this section). This is in contrast to [16,17], who considered the case when this inertia term remains large in this coordinate system. Furthermore, the assumption of high-Reynolds-number or rapid-rotation flow considered by [18] were also not applied in [1]. With this modelling assumption, the order of magnitude of the nonlinear perturbation term resulting from the inertia in the original cylindrical coordinate system, which influences the development of the flow structures, is determined only after the linear stability analysis as described below.
The equations presented in [1] are reproduced in §§2 and 3 below. The difference from the conventional representation of rotating plane Couette flow, where the control parameters are the Reynolds number and the system rotation , is the existence of an additional parameter , which results from the consideration of the inertia term in the equations of motion expressed in a cylindrical coordinate system. It is shown that both and are -order smaller than . At first glance this may appear to suggest all Taylor–Couette flows in the narrow-gap limit, , are reduced to plane Couette flow. This is, however, true only when is of order 1. As pointed out in [1] , when is of order so that is of order , the product of and can be finite, and the Taylor number defined by can take a finite critical value at the onset of axisymmetric instabilities. The numerical results for the critical Taylor number calculated by [1] were found to be in excellent agreement with previous studies.
It was also shown in [1] that may be expressed by , where is the ratio of the angular velocities of the outer to the inner cylinders, with angular velocities and , respectively. Therefore, is not determined when , that is, when . If is instead written in terms of , we have that for , and so , making it impossible to determine its critical state , and so this case was excluded from the analysis of [1]. We stress here that the case does not correspond to an inherent physical singularity of this flow. The inability to obtain for is due to an artefact where is expressed by means of . We add here that Davey et al. [16] (page 22) note that the replacement of by is more appropriate when the cylinders rotate in the opposite direction, so that may take a finite value for .
In this study, we consider the cases for , which supplements the analysis of Nagata [1], by including the case . We find that the critical state for the onset of axisymmetric instability for is determined by the product of and , i.e. the instability sets in at . This means that although our system approaches that of plane Couette flow as , the critical state exists even in the limit as . We call the modified Taylor number; its relation with is given by where , when .
The aim of the current paper is to investigate transition from circular Couette flow to Taylor-vortex flow in the narrow-gap limit for the for case, including the case which was excluded in the study by [1]. This investigation contributes towards understanding further transitions to non-axisymmetric flows.
We review briefly the analysis of the narrow gap limit by Nagata (2023) in §§2 and 3, followed by an analysis in §4, where we show that among two contributions, and , derived from the inertia term, , the latter, which was neglected in Nagata (2023) in the limit as , becomes important in the region of , and so is here taken into account in the nonlinear analysis after proper scalings for variables are introduced. (The former contribution leads to the term with in the equation of motion.) After describing the numerical procedures in §§5 and 6, we present the results for the counter-rotating case and the case with the outer cylinder at rest in §7; discussion is provided in §8. Finally, an appendix devoted to Nagata (2023) is included.
2. Formulation of the problem
We consider incompressible motion of a fluid with kinematic viscosity and density , which is confined between two coaxial cylinders of infinite extent in the axial direction, where the inner and outer cylinders, with radii and , rotate with angular velocities, and , respectively. We assume that .
The equations of motion in a cylindrical coordinate frame that rotates with an angular velocity are given by
where is the pressure and is the total velocity with components in the corresponding rotating coordinates .
In the above equations, denotes the material time derivative,
while denotes the Laplacian operator,
The basic angular velocity distribution in the rotating frame of reference, , satisfying no-slip boundary conditions at and , takes the form
where
To analyse the stability of the basic state and the ensuing nonlinear states, we superimpose disturbances , where includes the factor , to the basic state , where . The equations of motion for these disturbances are obtained by substituting for into equations (2.1)–(2.4) and by recalling that the basic state solves these equations:
where
The angular velocity and the azimuthal velocity of the basic state in the rotating frame take the form
and
respectively.
The third and fourth terms on the right-hand side of equation (2.10) result from with , where has been absorbed in the pressure term , whereas the third and fourth terms on the right-hand side of equation (2.11) result from in and from , respectively.
3. Transformation to the Cartesian system
We now focus on the narrow gap limit. It can be found from equations (2.8) and (2.14) that
where with . This shows that for all when , and that at or , whereas at or , when .
From equations (2.15), (2.8) and (2.14) we obtain
where is a half-gap between the cylinders.
The transformation of the equations of motion for the narrow gap limit in the cylindrical coordinates to those in non-dimensional Cartesian coordinates is accomplished by the identification and and , as . Here, we have chosen as the length scale, as the time scale and as the velocity scale. Then, the non-dimensional form of the equations of motion for non-dimensional disturbances is transformed from equation (2.9) to equation (2.12) as by neglecting and in equation (2.10), and and in equation (2.11), followed by multiplying both sides of equations (2.10)–(2.12) by , as follows.
where
Here, and are the non-dimensional forms of and , that is, and , respectively. We have transformed in equation (2.10) by using equation (3.1) in equation (3.4), and in equation (2.11) by using equation (3.2) in equation (3.5). For clarity, we drop the hat (∧) on , hereafter. In addition, we identify in equation (3.4) as .
Then in a vector form, the above system of equations are written as
where , and are the unit vectors in the -, - and -directions, and
It can be found that is the non-dimensional form of by equation (2.14). The non-dimensional form of is found to be as , since by equation (3.1).
The reason why the term has been neglected in equation (2.10) is that the non-dimensional form of using the length scale is of order , and it is supposed that is not as large as . However, we show in the next section that this model, which was also used in [1], cannot express the general case. Note however, that since this term is nonlinear the above system of equations, equations (3.3)–(3.6) or equations (3.8)–(3.9), remain valid for linear analysis.
4. Analysis
From equation (3.10), the values of and are constrained by as
where is used.
The constraint (equation (4.1)) suggests that the system, equations (3.8) and (3.9), is governed by a single parameter for a given .
The choice of
is used by Krueger [19] for .
One can make an alternative choice for this single parameter.
The choice of
corresponds to the conventional Taylor number, used by, for instance, Lin [20] and Chandrasekhar [21] for . This choice of parameter was also made in [1]. Since , this choice would be unsuitable when . It should be stressed, however, that the case with is not a physical singularity as discussed below. The Reynolds numbers,
based on the inner and the outer cylinders in the real space, can be expressed by , as follows:
Therefore, we find that a physical singularity occurs only when , not when .
When , the ratio, , can only be of as by equation (4.1), so that and can take finite values, but should be neglected. So, this case reduces to rotating plane Couette flow in Cartesian geometry. Here, does not appear in the analysis. This case corresponds to the study by [6] of flow between almost co-rotating cylinders. Axisymmetric flow in this case is governed by the Taylor number, , (equation (4.3)) with .
When is not close to 1, is -order smaller than . Since is also -order smaller than , one might suppose that the system is reduced to that of plane Couette flow. However, this is true only when is of . In this case, is of , because by the -component of equation (3.9). On the other hand, when is as large as or , the balance is achieved in the -component of equation (3.9) with , provided is of or zero, i.e. the axisymmetric case (). Then, we find by equation (4.1). Since , as described above, we find ) by comparing the two order evaluations for above. In this case, and ; however these parameters are retained since even though they are -order smaller than the products of and are of . In fact, the linear stability analysis conducted by [1] showed that the product of , and so the Taylor number (equation (4.3)), takes a finite value at the linear criticality in good agreement with [20] and [21]. Thus, the scaling discussion below follows as a consequence of the linear stability results rather than being imposed as a premise, such as has been used for high-rotation Taylor–Couette flow [18].
In any case, we stress here that Taylor–Couette flow in the narrow gap limit must be distinguished from plane Couette flow () by the fundamental difference between them: the former experiences axisymmetric instability at , and as , whereas the latter is linearly stable for all .
The choice of parameter in the present study will be given in §6.
In the region above the axisymmetric linear criticality at and , the nonlinear disturbance velocity is . Then, the curvature disturbance term in the wall-normal direction, which has been ignored so far, is , becoming comparable with the other quadratic nonlinear terms, . Hence, we need to include on the left-hand side of equation (3.9).
It can be found that the following scaling is adequate for this higher number region,
where . From the above scalings, we see that
Then, the equations to be solved become
Note that some unnecessary terms with -derivatives have been included in the above for the sake of formality, such as in which is before scaling and should not be included. For clarity, asterisks on the spatial coordinates and their derivatives are omitted hereafter.
The nonlinear analysis of [1] ignored the term, in equation (2.10). The short-coming owing to this neglection is rectified in the appendix .
5. Numerical analysis
We outline the numerical procedure to describe the axisymmetric case in this section. We separate the velocity perturbation into the modification, , of the mean flow and a periodic (in space) part, . Then, we express the latter solenoidal vector field in terms of poloidal and toroidal parts,
where . By doing so, the continuity equation is satisfied automatically.
Operations and on equation (4.9) lead to
where the dash () denotes the partial differentiation with respect to . The two terms of the second line in equation (5.2) corresponds to the correction to the previous formulation of [1]. Since is an odd function in , the first term breaks the -symmetry.
The equation for is obtained by averaging the -component of equation (4.9) in -space:
This shows that is created by the Reynolds stress. Note that and .
The no-slip boundary conditions become
Note again that unnecessary terms with -derivatives have been included in equations (5.1)–(5.4).
To examine an axisymmetric perturbation we let in equations (5.2), (5.3) and (5.4), and thus obtain
It is clear from equations (5.6), (5.7) and (5.8) that the axisymmetric nonlinear solutions satisfy the following symmetry:
To study the stability of the axisymmetric nonlinear solutions we superimpose infinitesimal perturbations on them. Here, we consider only axisymmetric perturbations, which satisfy
Taking into account the symmetry (equation (5.9)) of the nonlinear solutions and , we can find from equations (5.10) and (5.11) that the axisymmetric perturbations satisfy the same symmetry as
6. Numerical method
The perturbations , and are expanded by Fourier series in the (spanwise) directions for and , and by Chebyshev polynomials , in the wall-normal direction , which are multiplied by for and for and so as to satisfy no-slip boundary conditions.
Then, the perturbations, truncated at , are expressed as
where is the wavenumber in the spanwise directions.
Similarly, the infinitesimal axisymmetric perturbations, and , imposed on the nonlinear axisymmetric solutions (equations (6.1) and (6.2)) are expanded as follows:
where is the Floquet exponent (note that the present system is periodic in ). Also note that the case for is included in equations (6.4) and (6.5), which act as a perturbation imposed on .
Upon discretizing the equations (5.6), (5.7) and (5.8) in the wall-normal direction using the collocation method, the resulting algebraic equations for the amplitudes, , are solved as an eigenvalue problem for the complex growth rate of perturbation in the stability analysis, and by the Newton method for nonlinear analysis.
For calculations, the parameters, , and , are varied with the constraints
which corresponds to the relations in equation (4.1) for the unscaled , and .
Since equation (4.3) can be written as
and expressing in terms of from equation (6.6), we obtain
where
Then, equation (6.7) is simply rewritten as
The fact that the axisymmetric instabilities depend on only for a given () implies here that the flows ought to depend only on
for that given . Now, it is possible to determine the critical state owing to axisymmetric instability for , the case excluded in [1], by . The involvement of in results from in equation (5.6). We call the modified Taylor number. Since , the linear stability analysis for the original system, equations (3.8) and (3.9), gives the same critical value of the modified Taylor number as the one for the scaled system, equations (4.8) and (4.9). Note that one drawback is that the case must be excluded for this choice of parameter, .
7. Results
(a) The counter rotating case
Let us start describing linear and nonlinear results when the cylinders are rotating in the opposite directions with , which has been excluded in [1].
(i) Stability analysis
Table 1 lists the convergence of the real part of the eigenvalue for axisymmetric perturbations with , with respect to the truncation level of the Chebyshev polynomials. The parameter values, and , are chosen. The imaginary part, , is found to be zero in this case, and also for other cases with different values of and . The critical value, , is plotted against in figure 1a. Detailed calculations show the minimum occurs at . This value can be compared with that of Krueger et al. [19]. The linear critical value they obtained1 by solving an eigenvalue problem for a relevant six-order system by the Runge–Kutta method was 18 669 () for the critical wavenumber 3.999 (/2 = 1.9995) (their figures, based on our scaling are shown in the parentheses).
7 | 0.157362382950 |
9 | 0.157356630439 |
11 | 0.157355610920 |
13 | 0.157355638457 |
15 | 0.157355638527 |
17 | 0.157355638539 |
19 | 0.157355638532 |
The critical values for some given are compared with for at in table 2. Note that is not determined when , whereas is not determined when . It is clear from the consideration described for equations (4.3) and (6.11) that is identical to for . The important thing is that is continuous for , and so the linear critical state for exists, and the existence of a nonlinear solution branch bifurcating from the critical state is assured by the implicit function theorem.
1.0 | 106.7346 | 1.5585 | — | ||
0.5 | 106.6455 | 1.56 | 3/2 | 159.9683 | 159.9683 |
0.0 | 105.9345 | 1.56 | 1/2 | 211.8708 | 211.8708 |
−0.5 | 100.2996 | 1.56 | 1/6 | 601.7976 | 601.7976 |
−0.9 | 48.3777 | 1.56 | 1/38 | 1838.3526 | 1838.3526 |
−0.99 | 6.1335 | 1.56 | 1/398 | 2441.133 | 2441.133 |
−1.0 | 0 | 1.56 | 0 | — | 2518.388 |
(ii) Nonlinear analysis
The convergence of the momentum transport
at is checked with respect to the truncation levels for , and the optimal wavenumber in table 3. The critical value is , and so this is approximately higher than the critical value. For consistency with the linear analysis, the truncation level is chosen in the following nonlinear calculations.
(9, 6) | −0.1426510720228113D+01 |
(11, 7) | −0.1427378424590911D+01 |
(13, 8) | −0.1426697986553903D+01 |
(15, 9) | −0.1426804715949070D+01 |
(17, 10) | −0.1426807574622539D+01 |
(19, 11) | −0.1426799709544118D+01 |
It is confirmed that the bifurcation of the nonlinear state takes place subcritically at as shown in figure 1b.
As described above, the two terms of the second line in equation (5.2) are due to the coordinate transformation. Since is an odd function in , the first term breaks the -symmetry. As a result, an even component of the mean flow distortion is produced. Both anti-symmetric and symmetric components, and , of the mean flow increase gradually as is increased as displayed in figure 2.
The total mean flow is plotted in figure 3a. The degree of asymmetry is strengthened as increases. The asymmetry also appears in the flow field: the Taylor vortices move closer to the inner cylinder as increases, as seen from the stream function on the -cross-section in figure 3b.
(b) The case when the outer cylinder is at rest
There is experimental evidence [2] for the case of that stable axisymmetric Taylor-vortex flow exists near the axis in the physical -space, where and are Reynolds numbers based on the outer and the inner cylinder motions, respectively. When these Reynolds numbers are scaled in the same way as , they can be expressed by , as follows:
(see equation (4.5)). To compare the results in this section with the experimental observations, it would be convenient to switch the controlling parameter from to .
The basic state becomes unstable with respect to an axisymmetric perturbation with at as seen from the neutral curve in figure 4a. Figure 4b shows the supercritical nonlinear solution branch with which bifurcates from the basic state at this . Also shown in figure 4b is another supercritical nonlinear solution branch for , bifurcating at . We confirm that both branches are stable against perturbations with the same wavenumbers, that is, with the Floquet parameter in equations (6.4) and (6.5), as those of the nonlinear solutions. This observation is consistent with the so called exchange of stability.
We call the axisymmetric flows on the nonlinear solution branches with and the primary mode and the harmonic mode, respectively. The reason why the branch with is focused on here is that the wavenumber of the prime mode could be regarded as subharmonic, i.e. , from the view point of the harmonic mode, and that we anticipate that the harmonic mode might become unstable to subharmonic perturbations giving rise to solution branches with the same wavenumber as that of the primary mode. In fact, this anticipation is realized as shown later in this section.
The stability of the axisymmetric solutions against axisymmetric perturbations with other Floquet parameter values is examined. Instability caused by axisymmetric perturbations, in our case, for flow near the neutral curve is the well-known as Eckhaus instability [22]. The domain of the stable axisymmetric flow is bounded towards larger or smaller wavenumbers away from the neutral curve. Figure 5 shows the largest real part, , of the eigenvalues as a function of for various at . Recall that the value of the marginal for is given by , so that the neutral stability point for takes place at a wavenumber slightly larger than . As is clear, the axisymmetric flow is unstable at this value of when , with for all . Along the line of , decreases as is decreased. Figure 5 shows that the axisymmetric flow gains stability as decreases through . Thus, the stable domain within the neutral curve for the axisymmetric flow is reduced wavenumber-wise. (A similar reduction of the stable domain near the neutral curve for smaller wavenumbers should occur.)
It should be recognized that the eigenvalues are symmetric about . We see from the symmetries (equations (5.9) and (5.12)) that the system of the perturbation (equations (5.10) and (5.11)) is symmetric in . With this symmetry, and the fact that the system is periodic in , we can conclude that the system is symmetric about . Therefore, the eigenvalues as a function of are also symmetric about , which is verified numerically as seen in figure 5. Also, it should be recognized that the reaches a local minimum or maximum at . When it takes a positive local maximum, perturbations with the wavenumber , that is, growing subharmonic perturbations, would be created.
The largest and the second (and the third, if necessary) largest real parts of the eigenvalues are plotted as a function of for two supercritical states for the primary mode in figure 6a and for two supercritical states for the harmonic mode in figure 6b. It is seen that the largest real parts start from zero at and end at zero at for the two critical states for both modes. In fact, the largest real part is zero at for any supercritical states. It is a well-known fact that the supercritical nonlinear solutions are always stable against infinitesimal translation, , in the periodic direction, i.e. in the -direction for our nonlinear solution, that is, .
The second largest real part at starts from zero when is critical and gradually decreases in the negative direction as increases. This second largest eigenvalue determines the stability of the nonlinear solutions against axisymmetric perturbations with the same wavenumber as that of the nonlinear solutions. Figure 6a,b indicates that both primary mode and harmonic mode are stable against an axisymmetric perturbation of the same wavenumber as that of each mode.
The real parts of the two least stable eigenvalues are negative when for the primary mode near the critical state (see figure 6a), while they are positive for the harmonic mode (see figure 6b). They decrease as is increased from the critical value. This means that the stable domain of the axisymmetric primary mode is shifted towards the neutral curve for larger .
For the harmonic mode, the sign of the second largest real part of the eigenvalue at changes from positive to negative as is increased from 18 to 19 as seen in figure 7a, while the change of the sign of the largest real part of the eigenvalue from positive to negative takes place as is increased from 22 to 23 as shown in figure 7b.
The reason why both the largest and the second largest parts of the eigenvalue at start from positive values, actually the same positive values, at the onset of the harmonic mode is due to the fact that the basic flow is already unstable against those perturbations, and that the wavenumber of the perturbations can be regarded as subharmonic, , from the view point of the harmonic mode, . Note that the amplitude of the harmonic mode vanishes at its onset, so that the state under consideration is on the neutral curve, while below the neutral curve flow is the basic state.
Figure 8a shows that the values of the positive real eigenvalue for at the onset of the harmonic mode are exactly the same as those from the stability result of the basic state against perturbation with at this point.
The eigenvalues for and for the subharmonic perturbation to the harmonic mode with are connected to those for perturbations with , (), and for perturbations with , () in figure 8a.
Both the largest and the second largest eigenvalues for subharmonic perturbations to the harmonic mode decrease as is increased and they cross the line of at and at , respectively, as seen in figure 8a. At these values of the Jacobian of the eigenvalue equations is singular and therefore the uniqueness of the solution branches is not assured by the implicit function theorem. As a result a solution branch with the same wavenumber as that of the primary mode bifurcates, as can be seen in figure 8b. It is found that these two branches continue to exist without merging for larger . We stopped the numerical continuation at . We call the solution on the branch bifurcating from the harmonic mode at the subharmonic mode-I and the one bifurcating at mode-II.
The cross-sectional flow patterns, in terms of the stream function , for the subharmonic mode-I and mode-II together with that for the primary mode are compared at the same value of in figure 9. From the values of , the vortex of the primary mode is stronger than those of both subharmonic mode-I and mode-II, although its asymmetric (in ) nature is weaker. The vortex strength does not differ significantly between the two subharmonic modes I and II. A secondary weak vortex pair can be seen near the inner cylinder wall, , for the subharmonic mode-I, while a similar secondary weak vortex pair is seen to exist near the outer cylinder wall, , for the subharmonic mode-II. The weaker and the stronger vortex pairs inside one periodic box for the subharmonic mode-I and mode-II originate from the two vortex pairs of equal strength inside the same periodic box size for the harmonic mode.
8. Summary
The transition from circular Couette flow to axisymmetric Taylor-vortex flow in the limit of a narrow gap is examined by applying the Cartesian representation of the Taylor–Couette flow system which has recently been studied by Nagata (2023), to the counter-rotating case . An extension is also made to examine the stability of the nonlinear axisymmetric flow to axisymmetric perturbations.
The narrow-gap Taylor–Couette flow experiments conducted by Andereck et al. [2] for , Dutcher et al. [14] for and Snyder [15] for report the transition from circular Couette flow to axisymmetric Taylor-vortex flow for the case of the outer cylinder at rest () when , 70 and 108, respectively, where is the Reynolds number based on the inner cylinder motion and a half gap between the cylinders (see figures 2 and 18 in [2], figure 8 in [14] and figure 18 in [15], respectively). Although may increase drastically as , these cited values do not seem sufficiently large as to make a premise for introducing scaled analyses based on high rotation rate or high-Reynolds-number flows, although some such asymptotic analyses ([17,18]) surprisingly succeeded in obtaining simplified equations of motion similar to equation (4.9). In the present study, the following steps were taken to reach equation (4.9) without requiring any parameter modelling assumptions such as a high rotation rate.
Reduction of the equations of motion from the cylindrical coordinate system to the Cartesian coordinate system. Identification of controlling parameters , and . The definitions of the parameter shows is of larger than or .
Stability of the basic flow shows that the linear critical state is determined by the product of and irrespective of the individual values of or .
The linear critical state is applied for and , so that and .
Introduction of scaled analysis for high-Reynolds-number Taylor–Couette flow by and .
In the present system, the curvature effect is expressed by , where is the circular Couette solution and is the azimuthal velocity disturbance. The analysis of [1] took into account only the first term, by assuming that the second term could be neglected as . This is true in the linear theory and would be valid only when is not very large. When , however, it is found that the second term becomes important in the region where the linear instability sets in. The current study also includes this second term in the nonlinear analysis.
Unfortunately, the analysis of Nagata (2023) had to exclude the case , owing to the expression as a function of , , which was used in that study. It should be stressed again that the inability to determine the critical state by means of the ordinary Taylor number is not due to a physical singularity, but the problem is simply a result of the substitution of in place of . To study the case for , or we instead express in terms of , that is, , and introduce the modified Taylor number in the current study, where . Our linear calculations show that the critical values of are continuous across , and they reproduce the criticalities in terms of for other .
Our nonlinear analysis for based on the scaled parameters, , and , shows a supercritical bifurcation of axisymmetric flows at . The mean flows are asymmetrically distorted by the additionally created symmetric components of the mean flow distortion. The mean flows would be anti-symmetric about the channel centreplane if were absent, so is the symmetry-breaking parameter. Moreover, the Taylor vortices are distributed asymmetrically, with the centres of the vortices positioned, not on the mid-plane () between the gap as for the ordinary two-dimensional roll-cells with in the rotating plane Couette flow system, but closer to the inner cylinder wall ().
In the case of , which describes the case when the outer cylinder is at rest, a supercritical bifurcation of nonlinear axisymmetric Taylor–Couette flow (the primary mode) from the basic flow is found at with the critical wavenumber . This value corresponds to the scaled Reynolds number based on the inner cylinder motion, by equation (4.5), and should be compared with those of the aforementioned experimental results by [2,14] and [15] for and , respectively. Those values are and , respectively, where .
Furthermore, we examined the stability of the nonlinear axisymmetric Taylor-vortex flow to axisymmetric perturbations when . We found that the domain of the stable Taylor-vortex flow above the neutral curve is bounded towards smaller wavenumbers away from the neutral curve by the Eckhaus instability. During the course of this investigation, two new solution branches were also found. One of them (mode-I) is characterized by the appearance of a weak secondary vortex pair near the inner cylinder wall, while for the other (mode-II) a similar weak secondary vortex pair appears on the outer cylinder wall. The wavenumber of both modes is the same as the primary mode. Even if the two modes are found to be unstable to three-dimensional perturbations, their existence should affect the appearance of successive flow states.
We now map the flow from -space, or -space, to the physical space . The solution for a fixed is represented by a point on the straight line in the -space for . Once the flow is determined by either and at , or and at , with the constraint of equation (6.6), the point on the line with the scaled coordinates, in the -space represents the physical flow state.
The bifurcation nature of Taylor-vortex flow for a given is shown in the space in figure 10.
The figure shows subcritical bifurcations for and with the saddle-node points indicated by red open circles. The subcritical region seems to expand as is decreased further from −1.5. The existence of the subcritical branches is consistent to the numerical results by [23] for . Their figure 3 shows a subcritical branch of the axisymmetric Taylor-vortex flow in a small interval of near at , although their primary interest is in non-axisymmetric flows.
It is known that the critical modes change at from the axisymmetric steady mode to a non-axisymmetric oscillatory mode in the narrow gap limit of Taylor–Couette flow ([19]). It may then be the case the solution obtained in the current study cannot be realized. Also, turbulent states prevail at large Reynolds number. However, the author believes that the recognition of the existence of axisymmetric flows as an underlying nonlinear state is still important.
For the current axisymmetric case, we set in equations (5.6), (5.7) and (5.8). We can confirm that the scalings introduced in §4 are adequate, i.e. balance for equation (5.6) and balance for equations (5.7) and (5.8) are attained with and , when and . For non-axisymmetric cases, if we consider a long azimuthal length scale of order , then, , and the same order balances as the axisymmetric case are still attained with . In fact, the choice of this longer azimuthal length scale was chosen by [19], and used later in the study of a Ginzburg–Landau type of equation to investigate transition between axisymmetric and non-axisymmetric modes by [24–26] and [27]. Furthermore, this choice of a longer length scale was recommended by [28] for analysing flow dynamics in the narrow gap Taylor–Couette flow.
With this choice of a longer azimuthal length scale, three-dimensional investigation, such as the stability of axisymmetric Taylor-vortex flow against three-dimensional perturbations, and bifurcating three-dimensional flows, is accomplished easily in our Cartesian representation, equations (5.2)–(5.5).
Finally, it would be beneficial to provide narrow-gap limit studies based on the Cartesian coordinate system, similar to equation (3.9), by [18] and [17], especially for those with interest in asymptotic analysis. The aim of the former study was to obtain a rational analytic description for a linear neutral curve based on asymptotic theory, while an eigenvalue problem is solved numerically and compared with wide gap cases with and , not with those of Lin [20] or Chandrasekhar [21]. Nonlinear steady solutions were obtained in the latter study by restricting the computation to the case at a few selected values of very large Taylor numbers of order and a large axial wavenumber, , to support the high-wavenumber asymptotic analysis. Note, however, that neither study is sufficient to describe the transition process from circular Couette flow to Taylor-vortex flow in the narrow-gap limit, which is the main objective of [1] and the current paper. We have shown that the transition proceeds through a supercritical or subcritical bifurcation depending on the angular velocities of the inner and the outer cylinder. As described above, the numerical study for by [23] supports our finding of the subcritical branches for .
Data accessibility
This article has no additional data.
Declaration of AI use
We have not used AI-assisted technologies in creating this article.
Authors’ contributions
M.N.: writing—original draft.
Conflict of interest declaration
We declare we have no competing interests.
Funding
No funding has been received for this article.
Appendix Corrigendum
First, the reason why the case was excluded was not due to a physical singularity, as has been described in the main text above, but simply an artefact related to how was expressed in terms of . The inability to obtain a critical Taylor number for cannot be resolved by the introduction of scaled analysis, contrary to the wishful expectation described in the summary of [1].
Second, as described in §4, one of the inertia terms, , which was omitted in [1], becomes important in the region of and , where the linear criticality occurs, and so should have been retained. With its inclusion the equations to be solved become
in place of eqn (5.1) of [1], where asterisks indicate the scaled quantities as expressed in §4.
For the axisymmetric case, equation (A 1) leads to
The difference from eqns (5.10), (5.11) and (5.12) of [1] is the second line of (A 2), resulting from the inclusion of in (A 1), which is nonlinear, and the absence of in eqn (5.1) of [1] in (A 3). The deletion of is necessary because is smaller than in this high-Reynolds-number region. It can be shown (and numerically confirmed) that a linear analysis based on (A 2) and (A 3) does not change the results of [1].
As for the nonlinear analysis, the results differ, as can be seen by comparison of figures 11 and 12 with figures 1 and 2 of [1]. We followed the bifurcation branch for each with the constraint
which is the scaled version of equation (4.1). Figure 11 shows the change in bifurcation nature from supercritical to subcritical at approximately as is decreased from 1, instead of the linear increases of the amplitude of the supercritical branches in [1]. Recall that we have seen the subcritical bifurcation of axisymmetric flow for in figure 1b.
Footnotes
1 Their definition of the Taylor number is given by where (see (6) in [19]). From equations (6.6) and (6.11) we get . Substitution of (4.4) leads to .