1. Introduction
The problem of the stability and transition to turbulence of pipe flows has witnessed immense interest in the scientific and engineering communities since the famous experimental work of Reynolds (Reference Reynolds1883). Flows through pipes are central to countless engineering applications, primarily because one must predict the occurrence of transition to turbulence as the majority of gas, oil, water and air pipe flows are turbulent even at moderate Reynolds numbers. While fluids engineers are mostly concerned with an accurate computation of the pressure gradient required for propelling the fluid through pipes, physicists and mathematicians have also been attracted by the fascinating physical phenomena observed in these flows and by the exceptional difficulties related to the theoretical analysis of these confined flows.
It is still not clear how a laminar pipe flow becomes unstable and evolves to turbulence. The first theoretical challenge is that numerical and theoretical studies have consistently reported that the Poiseuille laminar flow is always stable to small perturbations at every Reynolds number when the problem is investigated using classical stability theory (Drazin & Reid Reference Drazin and Reid2004). This result has therefore led researchers to analyse the dynamics of finite-amplitude disturbances and alternative routes to turbulence.
Research studies on pipe flows have mainly focussed on the dynamics of the fully developed laminar flow, while a limited number of investigations have reported results on the entrance problem, that is, where the laminar base flow evolves from the pipe mouth. Along the entrance region, the flow depends on the streamwise coordinate and the pressure gradient is unknown a priori, thus posing additional challenges to its analysis. Our main objectives are, therefore, to investigate how vortical disturbances superimposed on the oncoming base flow at the pipe entrance are influenced by the pipe confinement, and how they are convected and evolve downstream through the pipe along the entrance region, prior to the flow breakdown to turbulence. We have devoted particular attention to the flow specification at the pipe mouth because a realistic source of perturbation is fundamental as a first step towards a full understanding of the pipe-flow transition to turbulence. We have considered the pipe wall to be smooth, rigid and stationary, and therefore we have neglected two other realistic sources of disturbances, i.e. wall roughness and wall vibrations.
In the following, we first discuss the most relevant studies of stability and transition in pipe flows and then we present our research objectives and the structure of the paper.
1.1. Stability of fully developed pipe flows: small perturbations
Motivated by the impact of the pioneering experiments by Reynolds (Reference Reynolds1883), Rayleigh (Reference Rayleigh1892) conducted the first stability analysis of a cylindrical flow in the inviscid regime, finding only stable perturbations. The first viscous stability study was carried out years later by Sexl (Reference Sexl1927), who showed flow stability in the limits of small and large Reynolds numbers. Several studies in the past century, such as Pekeris (Reference Pekeris1948), Corcos & Sellars (Reference Corcos and Sellars1959), Gill (Reference Gill1965), Lessen, Sadler & Liu (Reference Lessen, Sadler and Liu1968), Crowder & Dalton (Reference Crowder and Dalton1971) and Salwen, Cotton & Grosch (Reference Salwen, Cotton and Grosch1980), have established that the fully developed Poiseuille pipe flow is stable to infinitesimal disturbances according to classical stability theory even at very large Reynolds number. It was therefore a natural step to analyse other mechanisms of instability, such as transiently growing disturbances. Bergström (Reference Bergström1993), Schmid & Henningson (Reference Schmid and Henningson1994), O'Sullivan & Breuer (Reference O'Sullivan and Breuer1994), Mayer & Reshotko (Reference Mayer and Reshotko1997), Trefethen, Trefethen & Schmid (Reference Trefethen, Trefethen and Schmid1999) and Meseguer (Reference Meseguer2003) studied the transient problem in time, while Reshotko & Tumin (Reference Reshotko and Tumin2001) focussed on the spatial transient evolution of the disturbances. The results of Mayer & Reshotko (Reference Mayer and Reshotko1997) are of note as they compare favourably with the experimental results by Kaskel (Reference Kaskel1961). These studies concur that small perturbations superimposed on the fully developed Poiseuille flow can grow algebraically in time or space and then decay because of viscous effects. The reader is referred to § 3.2 in Kerswell (Reference Kerswell2005) for an exhaustive discussion of stability studies of small perturbations superimposed on the Poiseuille flow.
Alongside the theoretical advancements, pipe-flow experimental campaigns followed the work of Reynolds (Reference Reynolds1883), as summarized in § 3.1 of Kerswell (Reference Kerswell2005). Fox, Lessen & Bhat (Reference Fox, Lessen and Bhat1968) experimentally tested the stability of small perturbations introduced in the fully developed region and reported that the flow became unstable for Reynolds numbers larger than approximately 2000 (based on the pipe diameter), i.e. comparable with the results found by Reynolds (Reference Reynolds1883), although no comparison was carried out because the perturbations of Reynolds (Reference Reynolds1883)'s experiments were at the pipe inlet. Other experimental campaigns, e.g. Draad, Kuiken & Nieuwstadt (Reference Draad, Kuiken and Nieuwstadt1998), demonstrated that the transition Reynolds number could be increased significantly by reducing the amplitude of the perturbations.
1.2. Stability and transition of fully developed pipe flows: finite-amplitude perturbations
As linearized dynamics only provided a limited understanding of the fully developed pipe-flow transition problem, considerable effort has been devoted to the nonlinear behaviour of perturbations. The dynamics of finite amplitude disturbances in the limit of high Reynolds number was first studied by Davey & Nguyen (Reference Davey and Nguyen1971) via weakly nonlinear theory. The equilibrium perturbation amplitude above which disturbances grow was predicted and the most unstable disturbance was found to be near the pipe axis. These results were contradicted by Itoh (Reference Itoh1977), who discovered no equilibrium states by a similar approach. Davey (Reference Davey1978) further analysed these studies and concluded that neither centre-mode results were reliable, although maintained that equilibrium states did exist. Smith & Bodonyi (Reference Smith and Bodonyi1982) indeed found neutral disturbances of finite amplitude by utilizing nonlinear critical layer theory. Patera & Orszag (Reference Patera and Orszag1981), using direct numerical simulations, did not find any axisymmetric equilibrium states and concluded that the weakly nonlinear approach can lead to invalid results.
Nonlinear travelling waves were first discovered numerically by Faisst & Eckhardt (Reference Faisst and Eckhardt2003) and Wedin & Kerswell (Reference Wedin and Kerswell2004), thus paving the way for the use of dynamical systems in the study of pipe-flow transition (Duguet, Willis & Kerswell Reference Duguet, Willis and Kerswell2008; Hof et al. Reference Hof, De Lozar, Kuik and Westerweel2008; Avila, Willis & Hof Reference Avila, Willis and Hof2010; Avila et al. Reference Avila, Mellibovsky, Roland and Hof2013). Section 4 in Kerswell (Reference Kerswell2005) discusses the main results on these travelling waves. Their existence was confirmed in the water pipe-flow experiments of Hof et al. (Reference Hof, Van Doorne, Westerweel, Nieuwstadt, Faisst, Eckhardt, Wedin, Kerswell and Waleffe2004, Reference Hof, van Doorne, Westerweel and Nieuwstadt2005).
Other useful papers that thoroughly discuss experimental, numerical and theoretical advances related to the pipe-flow problem in the nonlinear regime are those by Avila et al. (Reference Avila, Moxey, de Lozar, Avila, Barkley and Hof2011), Eckhardt et al. (Reference Eckhardt, Schneider, Hof and Westerweel2007), Eckhardt (Reference Eckhardt2009) and Mullin (Reference Mullin2011). The review paper by Eckhardt (Reference Eckhardt2007) offers an interesting list of open problems related to the pipe-flow nonlinear breakdown to turbulence.
1.3. Stability and transition of pipe-entrance flows
As Reynolds (Reference Reynolds1883) recognized that the perturbance dynamics from the pipe inlet played a key role in the flow breakdown to turbulence, researchers have turned their attention to the stability of the pipe entrance flow. The main conclusion is that the pipe entrance flow is linearly unstable according to classical stability theory (Tatsumi Reference Tatsumi1952; Huang & Chen Reference Huang and Chen1974a,Reference Huang and Chenb; Smith & Bodonyi Reference Smith and Bodonyi1980; Garg Reference Garg1981; Gupta & Garg Reference Gupta and Garg1981; da Silva & Moss Reference da Silva and Moss1994), although there is still controversy about the exact form of the neutral curve, inside which Tollmien–Schlichting (TS) waves appear. The results by Gupta & Garg (Reference Gupta and Garg1981) and da Silva & Moss (Reference da Silva and Moss1994) show good agreement for most of the downstream locations and Reynolds numbers, from which it has been concluded that the entrance flow is linearly stable for Reynolds numbers smaller than approximately 10 000, based on the pipe radius and bulk velocity. Much lower critical Reynolds numbers have been reported by the experimental study of Sarpkaya (Reference Sarpkaya1975), probably because of the finite size of the induced perturbation, and by the numerical investigation of Sahu & Govindarajan (Reference Sahu and Govindarajan2007). Duck (Reference Duck2005, Reference Duck2006) used asymptotic and numerical methods to study the algebraic growth of disturbances near the pipe mouth, in the form of the so-called Luchini mode (Luchini Reference Luchini1996), located within the boundary layer attached to the pipe wall.
Wygnanski & Champagne (Reference Wygnanski and Champagne1973)'s experimental study focussed on the initial growth of perturbations generated at the pipe mouth, with the objective of understanding how these disturbances seeded turbulent puffs and slugs farther downstream. More laboratory data have been reported by Zanoun, Kito & Egbers (Reference Zanoun, Kito and Egbers2009) in an attempt to quantitatively correlate the location and features of the transitional flow with the intensity of the inlet perturbations. The direct numerical simulations by Wu et al. (Reference Wu, Moin, Adrian and Baltzer2015) and Wu, Moin & Adrian (Reference Wu, Moin and Adrian2020) showed how localized perturbations imposed at the pipe entrance evolved downstream and led to the breakdown to turbulence. At the pipe inlet, the Poiseuille parabolic velocity profile was chosen as the base profile in most of the cases, and the uniform plug flow was selected in one case.
To the best of the authors’ knowledge, no theoretical studies exist on the entrainment of velocity and pressure perturbations at the pipe inlet, in an attempt to analyse how these disturbances evolve and grow downstream inside the confined space.
1.4. Objectives and structure of the paper
We aim to analyse the response of the entrance incompressible pipe flow to vortical disturbances convected by the inlet flow. The main motivation arose from the absence of theoretical results on the dynamics of pipe entrance flows perturbed by disturbances that can be created and measured in a laboratory, despite the central role played by these perturbations as harbingers of instability and transition, as recognized since the pioneering work of Reynolds (Reference Reynolds1883). We consider flows at Reynolds numbers for which the entrance pipe flow is stable according to classical stability theory, i.e. TS waves do not occur, although experimental results have shown these flows to transition to turbulence.
We adopt the reasonable assumptions of large Reynolds number and of velocity disturbances of small amplitude and low frequency. The last assumption is based on the observation that streamwise-stretched low-frequency perturbations are the most likely to penetrate viscous layers and amplify downstream to cause transition (Matsubara & Alfredsson Reference Matsubara and Alfredsson2001). We are particularly interested in explaining how the disturbances are altered as they enter the pipe confinement and experience the increasing base-flow pressure gradient, how they penetrate into the boundary layer attached to the pipe wall, and how they amplify inside the pipe as the base flow becomes fully developed. We report how the perturbation dynamics depends on the flow parameters, such as the Reynolds number, the frequency, and the wavelengths of the prescribed disturbance. We carry out visual and data-based comparisons with experimental data of Wygnanski & Champagne (Reference Wygnanski and Champagne1973), and interpret results from the direct numerical simulation study of Wu et al. (Reference Wu, Moin, Adrian and Baltzer2015) on a pipe flow perturbed by inlet disturbances.
Our mathematical approach and numerical results are a starting point for the understanding of the relationship between the transitional Reynolds number and the role of inlet vortical perturbations in laminar-to-turbulent transition in pipe flows. We have restricted the analysis to the linearized dynamics under the assumption of small perturbations and we are currently extending our study to the nonlinear case.
In § 2, the scaling and assumptions are presented, together with the governing equations and the numerical procedures. The initial flow development is studied in § 3 and the results of the downstream flow dynamics are discussed in § 4. The conclusions of our study are presented in § 5.
2. Scaling and equations of motion
We first present the asymptotic scaling in § 2.1 and the equations of motion in § 2.2. In § 2.3 we discuss how linear stability results are useful to extract the locations in the entrance pipe-flow region where exponentially growing waves may occur. The numerical procedures are discussed in the supplementary material S1 available at https://doi.org/10.1017/jfm.2021.1005.
2.1. Scaling and asymptotic regions
We consider the pressure-driven incompressible flow at the entrance of a pipe with circular cross-section. Dimensional quantities are hereafter indicated by the superscript $*$. The oncoming flow approaching the pipe entrance is assumed to be uniform, of velocity $U^*_\infty$ and aligned with the pipe axis. The pipe wall is assumed to be infinitely thin so that, similar to the case of uniform flow approaching a thin flat wall, the oncoming flow is undisturbed at leading order. In a laboratory, the upstream geometry and the finite thickness of the pipe wall may alter the uniformity of the oncoming flow. Our inlet flow, therefore, is a simplified version of that utilized by O. Reynolds in his seminal study (Reynolds Reference Reynolds1883), where the pipe entrance was bell shaped. As the oncoming flow enters the pipe, a boundary layer develops on the pipe wall. The thickness of this viscous layer increases downstream until the pipe flow becomes fully viscous and attains the classical Poiseuille parabolic profile.
The flow is conveniently represented in a cylindrical coordinate system, where the coordinates $x^*,r^*$ and $\theta$ are the streamwise direction, the radial direction and the azimuthal angle, respectively. The centreline of the pipe is at $r^*=0$ and the pipe wall is located at $r^*=R^*$, where $R^*$ is the pipe radius. Lengths are scaled by a reference length $\lambda ^*$, specified below. Velocities are normalized by $U^*_\infty$, the pressure is scaled by $\rho ^* U_\infty ^{*2}$, where $\rho ^*$ is the density of the fluid, and the time is scaled by $\lambda ^*/U^*_\infty$ (quantities without any symbols are non-dimensional).
Superimposed on the oncoming streamwise flow $\{U^*_\infty,0,0\}$ are gust-type vortical fluctuations advected by the base flow. In a technological or industrial system, these oncoming disturbances could be due to upstream vibrations or structural imperfections, and in a laboratory setting they could be generated by vibrating ribbons or rigid grids located near the pipe entrance, where the wavelengths and frequencies could be controlled accurately, as could the location of generation.
We first consider a region around the pipe axis where velocity fluctuations are not influenced by the pipe wall, either by the no-penetration condition on the radial velocity disturbance or by the no-slip condition on the streamwise and azimuthal velocity components. At these locations sufficiently near $x=0$ and sufficiently far away from the pipe wall, the velocity fluctuations can be expressed mathematically as a Fourier–Bessel series, where the Fourier expansions pertain to time, $x$ and $\theta$, while the Bessel expansion is used along the radial coordinate. As it is assumed that the amplitude of the perturbation is asymptotically smaller than $U^*_\infty$, the perturbation dynamics is linear. A single Fourier–Bessel coefficient of the full-spectrum series expansion is thus considered, as follows:
where $J_m$ is the Bessel function of the first kind of order $m$, $E=\textrm {e}^{\textrm {i}[k_x (x-t) + m \theta ]}$, $\varepsilon \ll 1$ is the amplitude of the gust, $\bar {r}=\xi _{m n}r/2R$, $\xi _{m n}$ are the zeros of the Bessel function, i.e. the real numbers that satisfy $J_m(\xi _{m n})=0$, c.c. stands for the complex conjugate, and the prime indicates the first derivative with respect to $\bar {r}$. In the most general case, the quantities $\{\hat {u}_{x,jm n}^\infty,\hat {u}_{r,jm n}^\infty,\hat {u}_{\theta,jm n}^\infty \}=O(1)$ are complex. The inviscid velocity field (2.1) satisfies the Euler equations and is generated at a circle of radius $2R$, centred at the pipe centreline and located in the plane $(r,\theta )$ at a small distance upstream of the pipe mouth at $x=0$. The scaling of the radial direction by $2R$ is used instead of $R$ to avoid the unrealistic scenario of vanishing streamwise and radial velocity components at $r=R$ where the disturbance is generated.
The reference length is $\lambda ^*=\theta _g R^*$ ($\lambda ^*/R^* = O(1)$), that is, the circumferential wavelength of the gust at $r^*=R^*$, where $\theta _g$ is the azimuthal angle corresponding to the wavelength $\lambda ^*$. As in the flat-plate case of Leib, Wundrow & Goldstein (Reference Leib, Wundrow and Goldstein1999) where the spanwise wavelength is used, the azimuthal wavelength $\lambda ^*$ at $r^*=R^*$ is chosen because downstream it becomes comparable with the boundary layer thickness and the pipe radius. If $\theta _g={\rm \pi} /2$, the wavelength $\lambda ^*$ spans a quarter of the circumference of the pipe. An obvious constraint is that the maximum $\lambda ^*$ coincides with the pipe circumference. The coefficient $m=2{\rm \pi} R \in \mathbb {N}$ denotes the number of wavelengths $\lambda ^*$ around the circumference (we restrict ourselves to cases for which $m\neq 0$ because we are interested in three-dimensional disturbances). The radial velocity is never singular at the centreline despite $J_m$ being divided by $\bar {r}$ because, given $J_m(\bar {r}) \sim (\bar {r}/2)^m/\varGamma (m+1)$ for $\bar {r}\ll 1$ (where $\varGamma$ is the gamma function), the radial average $\hat {u}_{r,0n}^\infty$ is zero.
The focus is on low-frequency (i.e. long wavelength) disturbances with a streamwise wavenumber $k_x=2{\rm \pi} \lambda ^*/\lambda _x^* \ll 1$, where $\lambda _x^*$ is the streamwise wavelength of the gust, as these perturbations are most likely to penetrate a boundary layer and generate laminar streaks (Matsubara & Alfredsson Reference Matsubara and Alfredsson2001). As the convected gust is transported passively by the uniform base flow, the pressure fluctuations are negligible at leading order in $k_x \ll 1$ (refer to Goldstein (Reference Goldstein1978, (2.3)) and Leib et al. (Reference Leib, Wundrow and Goldstein1999, (3.2))). Equation (2.1) satisfies the continuity equation as the term emerging from $\partial u / \partial x$ is $O(k_x) \ll 1$, i.e.
Without losing generality, $m>0$ is taken. The continuity equation also determines the $\bar {r}^{-1}$ dependence for the radial velocity and the dependence on the derivative of the Bessel function and the coefficient $-\textrm {i}/\xi _{mn}$ for the the azimuthal velocity.
The Reynolds number is $Re_\lambda =U_{\infty }^*\lambda ^*/\nu ^* \gg 1$. Inside the pipe, the disturbances evolve downstream on a length scale that is comparable with the streamwise gust wavelength. A distinguished scaling is thus $k_x=O(Re_\lambda ^{-1})$ or $\bar {x}=k_x x=2{\rm \pi} x^*/\lambda _x^*=O(1)$. The disparity between the streamwise and azimuthal scales implies that disturbances of amplitude $O(\varepsilon )$ may generate streamwise velocity perturbations with an amplitude of $O(\varepsilon /k_x)$ in the viscous layers. These amplitudes are assumed to be much smaller than $U^*_\infty$, which translates to $\varepsilon /k_x \ll 1$ or $\varepsilon Re_\lambda \ll 1$ and implies that the Navier–Stokes equations may be linearized.
Vortical disturbances analogous to those defined in (2.1) have been found in other cylindrical flows, i.e. the instability modes of a rotating pipe flow also have the streamwise velocity component proportional to $J_m$ and the radial velocity proportional to $J_m/r$ (Pedley Reference Pedley1969). A similar choice for the oncoming disturbance was employed for the flat-plate boundary-layer cases studied by Ricco, Luo & Wu (Reference Ricco, Luo and Wu2011) and Marensi, Ricco & Wu (Reference Marensi, Ricco and Wu2017), who prescribed the vorticity fluctuations as a pair of gusts with equal and opposite spanwise wavenumbers. Disturbances (2.1) are physically realistic because they may be generated, for example, by a grid at the entry of the pipe, as a circular analogue to the vibrating wires used in the receptivity studies by Dietz (Reference Dietz1999) and Borodulin et al. (Reference Borodulin, Ivanov, Kachanov and Roschektayev2021a,Reference Borodulin, Ivanov, Kachanov and Roschektayevb) for two- and three-dimensional free stream vortical disturbances, respectively.
As shown in figure 1, the domain is divided into six asymptotic regions. The effect of the base-flow streamwise pressure gradient is negligible in regions I, II and III, mild in region IV, and fully influential in region V. In region VI, the base flow is fully developed. In region I, at $x=O(1)$, the base and the perturbation flows are inviscid and rapid distortion theory applies (Goldstein Reference Goldstein1978). A velocity potential is used and the wall is taken into account through the no-penetration condition on the radial velocity component. In the viscous region II, $\delta ^*/\lambda ^* \ll 1$, where $\delta ^*$ is the boundary-layer thickness. Regions I and II are studied in § 3.2.1. Region III occurs where $\delta ^*/\lambda ^*=O(1)$ and $\bar {x}=O(1)$ and it is governed by the linearized unsteady boundary-region equations (Leib et al. Reference Leib, Wundrow and Goldstein1999). In region IV, which is in the pipe core and surrounded by region III, the base flow is inviscid and the perturbation flow is viscous. Regions III and IV are discussed in § 3.2.2. In region V the base flow is not self-similar because of the unknown pressure gradient. The streamwise base-flow velocity at the pipe axis increases with $x$ and the base flow is non-parallel as the radial velocity plays a key role. The latter vanishes farther downstream in region VI, where the parabolic Poiseuille profile occurs.
2.2. Governing equations and boundary conditions
The flow field is governed by the continuity and Navier–Stokes equations,
where $\boldsymbol {u}=\{ u_x,u_r,u_\theta \}$. The velocity $\boldsymbol {u}$ and the pressure $p$ are expressed as the superposition of the base flow and the perturbation flow, i.e. $\{\boldsymbol {u},p\}=\{\bar {\boldsymbol {U}}(\bar {x},r), \bar {P}(\bar {x})\} +\{\boldsymbol {u}',p'\}$, where
and
The notation in (2.6)–(2.7) is as close as possible to the flat-plate case by Leib et al. (Reference Leib, Wundrow and Goldstein1999) for ease of understanding. The main difference between $\bar {u}_r$ and $\bar {u}_r^{(0)}$ and the corresponding components in (4.1) on page 176 of Leib et al. (Reference Leib, Wundrow and Goldstein1999) is the missing factor $(2 \bar {x})^{1/2}$ in (2.6), which arises in Leib et al. (Reference Leib, Wundrow and Goldstein1999)'s case because of the Blasius-similarity scaling.
By substituting (2.5)–(2.6) into (2.3)–(2.4) and collecting the terms of $O(1)$, one obtains the base-flow boundary-layer equations (Hornbeck Reference Hornbeck1964),
for $k_x, Re_\lambda ^{-1} \ll 1$ and $\mathcal {F} = k_x Re_\lambda = O(1)$. As $P$ is unknown, (2.8)–(2.9) are solved together with the mass conservation law, $\int _0^R U r \,\rm {d}r = R^2/2$, found by averaging across a pipe cross-section. Equations (2.8) and (2.9) are subject to the no-slip and no-penetration conditions at the pipe wall, $U=V=0$ at $r=R$, and to the symmetry conditions at the axis, $\partial U/\partial r=V=0$ at $r=0$. As $p'$ at the wall is unknown, we eliminate the pressure from (2.3)–(2.4) (Kim, Moin & Moser Reference Kim, Moin and Moser1987) to obtain the perturbation-flow equations,
The coefficients of (2.11) are
The boxed terms in (2.13)–(2.15) pertain to the unsteadiness and are asymptotically smaller than the other terms in the low-frequency limit $\mathcal {F} \ll 1$. The procedure to obtain (2.11)–(2.12) is given in the supplementary material S2. Equations (2.11) and (2.12) are also satisfied by $\{\bar {u}_x^{(0)},\bar {u}_r^{(0)}\}$. At the start of the analysis, (2.12) is the radial vorticity equation, but it reduces to the $x$-momentum equation because the leading-order contribution to the radial vorticity is given by $\bar {u}_x$ due to $k_x \ll 1$. Equations (2.11) and (2.12) are subject to
at $r=R$ because of the no-slip condition at the pipe wall and the continuity equation (S2.5). The boundary conditions at $r=0$ are
where the prime here indicates the derivative with respect to $r$. For $m=1$, (2.11)–(2.12) are solved together with the continuity equation (S2.5) because $\bar {u}_\theta$ appears in (2.28). For $m=2$, this complication is avoided by using $\bar {u}_r''=0$ instead of the last condition in (2.29). The behaviour of the perturbation flow near the pipe axis and the derivation of the boundary conditions is further discussed in the supplementary material S3. The same boundary conditions are found for $\bar {u}^{(0)}_x,\bar {u}^{(0)}_r,\bar {u}^{(0)}_\theta$. The azimuthal velocity component $\bar {u}_\theta$ and the pressure $\bar {p}$ are computed a posteriori from the continuity equation (S2.5) and the azimuthal momentum equation (S2.8).
The base-flow equations (2.8)–(2.9) and the perturbation (2.11)–(2.12) are parabolic and thus the specification of the appropriate initial conditions is of crucial importance. We have devoted great attention to the formulation of physically meaningful initial conditions for both the base flow and the perturbation flow. The initial base-flow velocity profile has usually been assumed uniform (Hornbeck Reference Hornbeck1964). We instead take into account the interaction between the oncoming flow and the pipe wall by deriving an asymptotic solution composed of the Blasius flow near the walls (inner solution) and the inviscid flow in the pipe core, distorted by the developing boundary layers (outer solution). The supplementary material S4 discusses the conditions under which the curvature effects can be neglected near the wall. The small-$x$ base-flow asymptotic initial condition used to solve the base-flow equations (2.8)–(2.9) is described in § 3.1.
Perturbation inflow conditions for open boundary-layer computations are often specified as the continuous spectrum of the Orr–Sommerfeld equations (Jacobs & Durbin Reference Jacobs and Durbin2001; Brandt, Schlatter & Henningson Reference Brandt, Schlatter and Henningson2004) or by selecting optimal perturbations, as those obtained for the flat-plate case by Andersson, Berggren & Henningson (Reference Andersson, Berggren and Henningson1999) and Luchini (Reference Luchini2000). As inflow conditions, Buffat et al. (Reference Buffat, Le Penven, Cadiou and Montagnier2014) imposed optimal disturbances within one of the boundary layers near the entrance of a channel, while perturbations were absent in the inviscid core.
We prescribe initial conditions that are superposed on the inviscid core flow and match the oncoming vortical disturbances at $\bar {x}=0$, given in (2.1). Like the base flow, these initial conditions are expressed as asymptotic composite solutions of a viscous inner solution within the boundary layer and an outer solution where the base flow is inviscid in the pipe core because we are interested in their entrainment of vortical disturbances that occupy the entire cross-plane near the pipe mouth. The initial condition is not imposed at $\bar {x}=0$ because the radial base-flow velocity is singular there and because the flow field is governed by the full Navier–Stokes equations in the immediate surroundings of the inlet. Therefore, the initial conditions are imposed at a location $0 < \bar {x}_0 \ll 1$.
Section 3.2 presents results for the development of the perturbation flow at small-$\bar {x}$ locations where the base-flow pressure gradient is negligible. These flow fields are small-$\bar {x}$ asymptotic solutions of the flow in region V and can therefore be used as initial conditions to solve equations (2.11)–(2.12) along the pipe, where the base-flow streamwise pressure gradient is fully influential.
2.3. Critical streamwise location for linear stability
As we are interested in the spatial transient growth of vortical disturbances in the entrance region, a first step is to report, as a function of the Reynolds number, the streamwise locations where TS waves appear. Figure 2 shows the neutral stability locations $x_c/R$ as a function of the Reynolds number $Re_R=R Re_\lambda =U^*_\infty R^*/\nu ^*$ computed by Gupta & Garg (Reference Gupta and Garg1981) (blue line) and da Silva & Moss (Reference da Silva and Moss1994) (red circles). The entrance flow is linearly stable for $Re_R<11\,000$ and the unstable region has a finite extent along the $x$ direction as the flow is always linearly stable when it reaches the fully developed condition. We also show the experimental neutral points by Sarpkaya (Reference Sarpkaya1975), measured by introducing disturbances at fixed locations on the pipe wall by electromagnetic excitation. da Silva & Moss (Reference da Silva and Moss1994) claimed that the disagreement may be due to the finite amplitude of Sarpkaya (Reference Sarpkaya1975)'s perturbations, which cannot be captured by the linear stability theory. We focus on Reynolds numbers in the range $1000< Re_R<10\,000$ (marked by the light grey area in figure 2), i.e. small enough for TS waves not to appear and large enough for algebraically growing perturbations or transition to turbulence to have been observed in experimental studies.
3. Initial flow development
In this section we discuss the theoretical and numerical results of the base and perturbation flows near the pipe entrance, where the effect of the pipe confinement is important, but the influence of the base-flow pressure gradient is negligible at leading order. This mathematical analysis is relevant because, through the asymptotic formulation, the physical flow features near the entrance are revealed and the flow evolution farther downstream inside the pipe can be computed. As explained at the end of § 2, the results obtained in this section will indeed specify the appropriate initial conditions for the computation of the base and the perturbation flows in region V, where the dynamics is fully influenced by the streamwise pressure gradient.
3.1. Initial base flow in regions I and II
The initial base-flow solution is constructed according to the method of matched asymptotic expansions. The outer solution is the inviscid base flow in the core of the pipe, while the inner solution is the viscous flow attached to the pipe wall. The composite solution reads
where the subscripts in, out and com stand for inner, outer and common, respectively, and $\bar {\boldsymbol {U}}$ is defined in (2.5). The common part is defined as
where
is the scaled radial coordinate of the inner solution and in (3.2) the plus sign applies to the streamwise velocity and the minus sign applies to the radial velocity because the coordinates $r$ and $\eta$ point in opposite directions.
For the flow in the entrance region of a channel, the inner solution near the inlet corresponds to the Blasius flow as the Reynolds number is large (Wilson Reference Wilson1970; Rubin, Khosla & Saari Reference Rubin, Khosla and Saari1977; Duck Reference Duck2005; Buffat et al. Reference Buffat, Le Penven, Cadiou and Montagnier2014). We also benefit from the Blasius solution in the pipe-flow case as we have proved in the supplementary material S4 that the curvature effects are negligible near the wall. The inner flow therefore satisfies the self-similar Blasius equation
where the prime indicates differentiation with respect to $\eta$. The boundary conditions for (3.4) are $F(0)=0$, $F'(0)=0$ and $F'\rightarrow 1$ as $\eta \rightarrow \infty$. The inner base flow reads
As $Re_\lambda \gg 1$, the base-flow viscous effects are negligible in the pipe core near the entrance, which is consistent with the use of the boundary-layer approximation of the Navier–Stokes equations. In the pipe core, the axisymmetric outer flow is described by the inviscid Stokes stream function $\psi$,
i.e. $\bar {U}_{out}=r^{-1}\partial \psi /\partial r$ and $\bar {V}_{out}=-r^{-1}\partial \psi /\partial x$. The leading-order term $r^2/2$ in (3.6) represents the uniform streamwise flow. The second-order stream function $\psi _2$, which defines the flow due to the pipe confinement and to the Blasius boundary layer developing on the pipe wall, satisfies
subject to
where $\beta =\lim _{\eta \rightarrow \infty } (\eta - F) =1.217 \ldots\, $. The inviscid base flow in the pipe core is irrotational and therefore the governing equation (3.7) is obtained by setting the azimuthal vorticity component to zero. Equation (3.7) is not the Laplace equation because of the negative sign of the last term (Panton Reference Panton2013). The boundary condition (3.8a) is obtained by asymptotic matching, i.e. the radial component of the outer velocity must match the outer limit of the base-flow wall-normal velocity of the Blasius boundary layer,
where the minus sign in front of the $\eta$-limit is due to the inner and outer velocity components pointing towards opposite directions. It follows that
at $r=R$ for $x>0$. The boundary condition (3.8b) is obtained as follows. The base flow is uniform and streamwise only as $x \rightarrow -\infty$ and, as it is not influenced by the presence of the pipe, no wall-normal base-flow velocity occurs as $x \rightarrow -\infty$. Also, the horizontal line $r=R$ for $x<0$ is a streamline with stagnation point at $x=0^-,r=R$ and no wall-normal base-flow velocity occurs along it because the base flow approaches the pipe wall parallel to it. Therefore, $\bar {V}_{out}=0$ at $r=R$ for $x<0$, i.e. $\partial \psi / \partial x=\partial \psi _2 / \partial x=0$. As we choose $\psi _2=0$ as $x \rightarrow -\infty$, by integrating $\partial \psi _2 / \partial x=0$ from $x \rightarrow -\infty$ for $r=R$, it follows that $\psi _2=0$ along $r=R$ for $x<0$. The boundary condition (3.8c) follows from $\bar {V}_{out}=0$ along the pipe axis at any $x$ due to the flow axial symmetry.
The solution to (3.7) together with the boundary conditions (3.8) is found by use of the complex Fourier transform along $x$. The solution reads
where $I_1$ is the modified Bessel function of the first kind and $\gamma \in \mathbb {R}<0$. The supplementary material S5 presents the derivation of (3.11) (Dettman Reference Dettman1965). The composite solution of the streamwise velocity $U$ is
The composite solution of the radial velocity $V$, defined in (2.5), is
Figure 3 shows the inner and outer solutions, their common parts, and the composite solutions for the base flow for $Re_\lambda =500$ and $2000$ at $x=0.04$. The composite base-flow streamwise velocity, depicted in figure 3(a,b), agrees well with the inner Blasius solution across the pipe radius as the acceleration in the pipe core is still small at this $x$ location. The increase of inviscid streamwise velocity balances the decrease within the viscous region to conserve the mass flow rate. The displacement effect of the Blasius boundary layers, given by (3.9), causes a small streamwise pressure gradient related to the dependence of the displacement stream function $\psi _2$ on the $x$ coordinate. This small pressure gradient is negligible at leading order in the boundary layer. The inviscid streamwise velocity is larger than unity near the wall because it accelerates along $x$ to conserve the mass flow rate as the radial velocity decreases from its boundary-layer blowing value as the pipe axis is approached. This acceleration gives rise to the local near-wall peak in the $\bar {U}$ profile, also reported in Durst et al. (Reference Durst, Ray, Ünsal and Bayoumi2005) for entry-pipe flow and in Sparrow, Lin & Lundgren (Reference Sparrow, Lin and Lundgren1964), Panton (Reference Panton2013) and Alizard et al. (Reference Alizard, Cadiou, Le Penven, Di Pierro and Buffat2018) for entry-channel flows. The peak occurs because the inviscid streamwise velocity is larger than the viscous streamwise velocity deficit of the boundary layer.
The base-flow radial velocity, shown in figure 3(c,d), agrees well with the viscous solution only in the proximity of the wall, while the two fail to overlap in the inviscid core, where the composite profile coincides with the outer solution. The composite solution shows a distinct near-wall peak that moves closer to the wall and decreases in intensity as the Reynolds number increases. The velocity is not exactly zero at the wall. This small slip velocity decreases as $O(Re_\lambda ^{-1/2})$ and induces a viscous layer at the next order, which is not computed.
3.2. Initial perturbation flow
The precise flow specification at the pipe entrance is relevant as our interest lies in the mathematical description of the inlet profiles to start the computations and to understand how the vortical disturbances are entrained in the pipe mouth and how they evolve in the entrance region. The initial perturbation flow, fully influenced by the confinement of the pipe and only mildly by the base-flow streamwise pressure, is studied in this section via matched asymptotic expansions. Asymptotic composite solutions are derived for regions I and II in § 3.2.1. The flow in regions III and IV is discussed in § 3.2.2.
3.2.1. Perturbation flow in regions I and II
In region I the base flow is uniform and inviscid, the perturbation flow is inviscid, and $\bar {x}\ll 1$ with $\mathcal {F}=O(1)$. The flow field can thus be adequately described by rapid distortion theory (Goldstein Reference Goldstein1978). The velocity is expressed as
The velocity $\boldsymbol {u}^{(1)}$ is due to the interaction of the free stream gust (2.1) with the pipe wall, while $\boldsymbol {u}^{(1)}_1$ is generated by the wall-normal boundary-layer perturbation velocity via $\boldsymbol {\nabla } \phi _1$. The perturbation potential $\phi$ satisfies the Laplace equation
subject to
where $u_{\infty r}(r=R)= [2 \hat u^\infty _{r,mn} J_m(\xi _{mn}/2)/\xi _{mn}] \exp [\textrm {i} k_x (x-t) + \textrm {i} m \theta ]$ is the gust radial velocity found from (2.1) as the pipe wall is approached. Expression (3.16b) is the no-penetration boundary condition at the wall for the radial velocity. To solve (3.15) in the infinite strip $-\infty < x<\infty$, $0\leq r\leq R$, one needs a boundary condition at $r=R$ for $x<0$, which is unknown. However, the interest here is in the solution at $x \gg 1$, which can be found by separation of variables. In this limit, it is found that $\phi = \hat \phi (r) \exp [\textrm {i} k_x (x-t) + \textrm {i} m \theta ]$, where $\hat \phi (r)$ satisfies the modified Bessel equation
In the limit $k_x\ll 1$, the solution is
which is not singular when $m=0$ because $\hat u^\infty _{r,0n}=0$. Details of the derivation to obtain (3.18) are found in the supplementary material S6. The velocity components of $\boldsymbol {u}^{(1)}=\{u_x^{(1)},u_r^{(1)},u_\theta ^{(1)}\}$ in (3.14) are
Using (2.6) and (2.7), the leading-order streamwise and radial velocities in region I are
The leading-order azimuthal velocity $\bar {u}_{\theta,out}=u_\theta ^{(1)}$ is given by (3.21) because of (2.7b). As the pipe wall is approached, $\bar {u}_{\theta,out}$ drives the inner flow in the viscous region II, where the azimuthal viscous diffusion effects do not play a leading role. As proved by the scaling in the supplementary material S4, the flow in region II is described by the unsteady boundary-layer equations in Cartesian coordinates because the curvature effects are negligible near the wall. The inner solution for the radial velocity is $\bar {u}_{\theta,in}=\bar {u}_{\theta,com} F'(\eta )$, found by use of (4.13) in Leib et al. (Reference Leib, Wundrow and Goldstein1999), where the common part $\bar {u}_{\theta,com}$ is
The composite azimuthal velocity $\bar {u}_{\theta }$ is
The solution (3.24) is not needed for the numerical computation in region V because $\bar {u}_{\theta }$ is absent from (2.11) and (2.12). However, the computation of $\bar {u}_{\theta,com}$ is essential because it determines the amplitude of the inner streamwise and radial velocities. The streamwise velocity component in region II is $\bar {u}_{x,in}=\bar {u}_{\theta,com}\bar {x} \eta F''/(2R)$. The $\bar {x} \eta F''/2$ dependence is found using (4.13) in Leib et al. (Reference Leib, Wundrow and Goldstein1999), and the amplitude $\bar {u}_{\theta,com}$ is obtained by converting the continuity equation from cylindrical to Cartesian coordinates and by taking the limit $r \rightarrow R$. As $\bar {u}_{x,out}=0$, the composite streamwise velocity $\bar {u}_x$ is
The asymptotic approach is clearly an invaluable tool as the amplitude of the streamwise perturbation velocity is uniquely linked to the oncoming free stream flow characteristics through $\bar {u}_{\theta,com}$, given in (3.23). Had the perturbation been prescribed within the boundary layer without relating to the oncoming fluctuations, the streak amplitude should have been assigned arbitrarily.
The composite radial velocity $\bar {u}_r$ is found by first summing the outer solution (3.22) and the inner solution obtained by multiplying the region-II wall-normal velocity (4.13) in Leib et al. (Reference Leib, Wundrow and Goldstein1999) by the amplitude $\bar {u}_{\theta,com}$ and by $(2 \bar {x})^{1/2}$ on using (4.1) in Leib et al. (Reference Leib, Wundrow and Goldstein1999). The common part, i.e. the large-$\eta$ limit of the region-II solution, is then subtracted. The common part also emerges from the second-order term of the Taylor expansion of the outer solution (3.22) as $r \rightarrow R$, while the leading-order term vanishes because of the no-penetration condition (3.16b). The composite radial velocity $\bar {u}_r$ is
The initial conditions for $\{ \bar {u}_x^{(0)}, \bar {u}_r^{(0)}, \bar {u}_\theta ^{(0)}\}$, defined in (2.7), are also found through a composite solution. The solution for the streamwise velocity $\bar {u}_x^{(0)}$ is found by combining the outer solution (3.19) and the inner solution given by (4.13) in Leib et al. (Reference Leib, Wundrow and Goldstein1999), valid for $\mathcal {F} = O(1)$ and $\bar {x}\ll 1$. The composite streamwise velocity $\bar {u}_x^{(0)}$ is
The first term in (3.19) can be neglected because it is $O(k_x)$. The inner solution for the radial velocity $\bar {u}_r^{(0)}$ is found by (4.13) in Leib et al. (Reference Leib, Wundrow and Goldstein1999), which is also valid for $\mathcal {F} = O(1)$ and $\bar {x}\ll 1$. It reads
Physically this perturbation velocity represents a radial blowing/suction effect, an unsteady analogue to the one experienced by the base-flow Blasius boundary layer. In the Blasius case, the base wall-normal velocity is generated by the free stream uniform $U^*_\infty$, while $\bar {u}^{(0)}_{r,out}$ is driven by $\bar {u}^{(0)}_{x,out}$. The transpiration velocity induces the inviscid velocity field $\boldsymbol {u}^{(1)}_1$ in (3.14) and the associated radial pressure gradient. The velocity potential $\phi _1$ satisfies
subject to
Similar to the problem of solving (3.15), (3.29) cannot be solved in the infinite strip $-\infty < x<\infty$, 0$\leq r\leq R$ because the boundary condition at $r=R$ for $x<0$ is unknown. In our region of interest, $x\gg 1$, the second derivative of $\phi _1$ with respect to $x$ is asymptotically smaller than the other derivatives as long as $k_x \ll \bar {x} \ll 1$ because of the $(2\bar {x})^{-1/2}$ behaviour of the boundary conditions (3.30b). By separating the variables and expressing $\phi _1=\hat \phi _1(r)\exp (\textrm {i} m \theta -\textrm {i} k_x t)/(2 \bar {x})^{1/2}$, equation (3.29) simplifies to the Euler differential equation. The solution is
The composite radial velocity $\bar {u}_r^{(0)}$ reads
It is verified that $k_x\bar {u}_r^{(0)}$ is asymptotically smaller than $\textrm {i} m \bar {u}_r$ and $\bar {u}_x^{(0)}$ where $k_x/Re_\lambda \ll \bar {x}\ll 1$, as required by the expansion (2.7). Differently from the leading-order components where $\bar {u}_{\theta,out}$ is needed to determine the amplitude of $\{\bar {u}_{x,in},\bar {u}_{r,in}\}$, it is not necessary to compute $\bar {u}^{(0)}_{\theta,out}$ as we have already found $\{\bar {u}_x^{(0)},\bar {u}_r^{(0)}\}$ to start the integration of the boundary-region equations (2.11) and (2.11).
Figures 4–7 show velocity profiles for different $m$ and $n$ values. The flow parameters have been chosen as representative of the most significant cases, i.e. azimuthal wavelengths at $r=R$ and characteristic radial length scale comparable with the radius, Reynolds number with a middle value in the range of interest given in figure 2, and sufficiently small $\bar {x}$ for the asymptotic solutions to be valid.
In figures 4 and 5, increasing $n$ causes the velocity profiles that are directly driven by the free stream forcing (2.1), i.e. $\bar {u}_r$, $\bar {u}_\theta$ and $\bar {u}_x^{(0)}$, to have a more intense modulation along the radial coordinate because the Bessel-function zeros $\xi _{mn}$, which appear in the argument of the Bessel functions in (2.1), increase in amplitude. The zeros $\xi _{mn}$ can thus be interpreted as analogous to wall-normal wavenumbers in the Cartesian geometry. Figures 4(c,d) and 6(c,d) show how the inner profiles (region II) and the outer profiles (region I) of the azimuthal velocities combine to create the composite profiles valid at any radial location. Figure 8 shows contour plots of the composite velocity components and the pressure at a fixed time instant in a cross-plane near the pipe mouth.
The composite solutions (3.25)–(3.27) and (3.32) for $\bar {u}_x,\bar {u}_r,\bar {u}_x^{(0)}$ and $\bar {u}_r^{(0)}$ in regions I and II are used as initial conditions for the computation in regions V and VI.
3.2.2. Perturbation flow in regions III and IV
The outer expressions (3.22) do not take into account the viscous decay as they are only valid in the inviscid region I, i.e. for $\bar {x} \ll 1$. Viscous effects become important as the outer flow evolves downstream through region IV. The flow in regions III and IV can be found by a composite solution. The expression for the outer velocity field in region IV can be written as
where the stream function $\psi (x,r)$ is given in (3.6) and (3.11). As depicted in figure 1, the base-flow pressure gradient only has a second-order effect through the $x$-dependence of $\psi (x,r)$. In the Cartesian geometry the outer velocity $\boldsymbol {u}_{out}$ can be obtained analytically by solving the parabolic region-IV momentum equation (5.9) on page 181 in Leib et al. (Reference Leib, Wundrow and Goldstein1999). In the cylindrical geometry, the radial and the azimuthal momentum equations are instead coupled and the system has to be solved numerically. The inner solution must also found numerically by solving the boundary-region equations (5.2)–(5.5) on page 180 of Leib et al. (Reference Leib, Wundrow and Goldstein1999), complemented by mixed-type boundary conditions obtained by asymptotically matching with the outer solution (3.33). These numerical calculations are not pursued herein because our region-V solution includes the solutions of regions III and IV and covers the whole streamwise flow evolution as the base-flow pressure gradient is accounted for at leading order.
4. Downstream flow development
In § 3 the base and perturbation flows are influenced by the confinement of the pipe, but the streamwise location is sufficiently upstream for the base-flow streamwise pressure gradient not to play a leading-order role. In the present section, we consider region V, i.e. streamwise locations where the base-flow pressure gradient instead plays a leading-order role on the base flow and therefore on the perturbation flow. In region VI, the pressure gradient adjusts downstream to a constant value as the base flow develops to the fully developed Poiseuille flow. The base-flow profiles are obtained by solving (2.8)–(2.9) and the perturbation profiles are found by solving (2.11)–(2.12).
4.1. Base flow
The base-flow streamwise and wall-normal velocity profiles across the pipe are shown in figure 9 at different streamwise positions. The small-$x$ asymptotic profiles (3.1) are also shown (dashed lines). The flow field approaches the fully developed regime as it develops downstream: the base-flow streamwise velocity evolves to the parabolic Poiseuille profile, while the wall-normal velocity decreases to zero.
The base-flow streamwise velocity is shown in figure 10 at various $r/R$ locations as a function of the streamwise position. There is excellent agreement with the numerical data by Hornbeck (Reference Hornbeck1964) (red circles) and very good agreement with the experimental data by Reshotko (Reference Reshotko1958) (white circles), except at the two locations closest to the pipe wall, where the uncertainty of the measurements may have been influential.
The downstream adjustment of the pressure gradient can be monitored through the correction pressure function
which measures the deviation of the base-flow pressure, defined in (2.5), from the fully developed Poiseuille value. Our numerical data, shown in figure 11, agree well with the numerical data based on the series solution by Sparrow et al. (Reference Sparrow, Lin and Lundgren1964) (red circles) and satisfactorily with the experimental data by Mohanty & Asthana (Reference Mohanty and Asthana1978). Our fully developed computed value, $\mathcal {K}_\infty =\lim _{x \rightarrow \infty }\mathcal {K}(x)=0.63$, matches the one by Sparrow et al. (Reference Sparrow, Lin and Lundgren1964) and is slightly lower than the experimental values by Shapiro, Siegel & Kline (Reference Shapiro, Siegel and Kline1954) (square) and Knibbs (reported by Sparrow et al. (Reference Sparrow, Lin and Lundgren1964), triangle).
We can define boundary-layer thicknesses to quantify the diffusion of the viscous effects as the flow develops downstream. They are defined as
where $k=R$ identifies the boundary thickness that matches the pipe radius in the fully developed downstream limit ($\varDelta _R=3$ is obtained by substituting $\delta _R=R$, $\lim _{x \rightarrow \infty } \bar {U}(x,r)=2[1-(r/R)^2]$ and $\lim _{x \rightarrow \infty } \bar {U}_{cen}=2$ into (4.2)) and $k=LWG,PIPE$ denotes the boundary thickness that matches the one employed by Leib et al. (Reference Leib, Wundrow and Goldstein1999) as $x \rightarrow 0$, i.e., $\delta _{LWG}=(2 x/Re_\lambda )^{1/2}$ ($\varDelta _{LWG,PIPE}=\beta ^{-1} = 0.822$ is obtained by substituting (3.12) into (4.2)). Figure 12(a) shows the boundary-layer thicknesses as functions of the streamwise coordinate. The thickness $\delta _{LWG,PIPE}$ in our pipe-flow flow case is thinner than the corresponding Blasius-flow $\delta _{LWG}$ because of the accelerating core caused by the favourable pressure gradient.
We also quantify the entry length, i.e. the distance from the pipe mouth where region V ends and the fully developed region VI starts. The entry length is typically defined by the streamwise location where the axis base velocity $\bar {U}_{cen}$ reaches $99\,\%$ of its fully developed value. We can first use (23) in Durst et al. (Reference Durst, Ray, Ünsal and Bayoumi2005), i.e. $x_{e,u}=2 R[0.4642 +(2 C_{e,u} R Re_\lambda )^{1.6} ]^{1/1.6}$, where $C_{e,u}=0.0567$. As we operate under the assumption $Re_\lambda \gg 1$, Durst et al. (Reference Durst, Ray, Ünsal and Bayoumi2005)'s equation reduces to $x_{e,u}=4 C_{e,u} R^2 Re_\lambda$, which is consistent with the scaling adopted in figure 10. We compute $C_{e,u}=0.057$, which is within the uncertainty range provided by Durst et al. (Reference Durst, Ray, Ünsal and Bayoumi2005). We also quantify the entrance region as $x_{e,pres}=4 C_{e,pres} R^2 Re_\lambda$, i.e. the streamwise distance from the pipe mouth where $\mathcal {K}=0.99\mathcal {K}_\infty$, that is, where the pressure gradient has reached its fully developed constant value. We compute $C_{e,pres}=0.055$.
Crabtree, Küchemann & Sowerby on page 440 of Rosenhead (Reference Rosenhead1963) remark that in a pipe entrance flow: ‘… the whole of the fluid across a section becomes influenced by viscosity before the parabolic profile is reached’. We can examine this statement, although they do not specify how the diffusion of viscous effects is defined mathematically. The development of the flow to the Poiseuille parabolic profile is already quantified by the entry length $x_{e,u}$, based on the downstream evolution of $\bar {U}_{cen}$, but we also further monitor it by an adjustment length $x_{e,pois}=4 C_{e,pois} R^2 Re_\lambda$, defined as the streamwise location where the average difference between the streamwise velocity and the Poiseuille velocity, i.e. $\mathscr {E}(x)=R^{-1}\int _0^R |\bar {U}(x,r)-2[1-(r/R)^2]|\,\rm {d}r$ (shown by the dashed line in figure 12b), has decayed to 1 % of $\lim _{x\rightarrow 0}\mathscr {E}=2/3$. We find $C_{e,pois}=0.057$, i.e. the same as $C_{e,u}$. The diffusion of viscous effects can be quantified by two adjustment lengths. We first obtain $x_{e,u2}=4 C_{e,u2} R^2 Re_\lambda$, i.e. the downstream distance from the pipe mouth where the second derivative of the streamwise velocity with respect to the radial coordinate at the pipe axis, $\partial ^2\bar {U} / \partial r^2 |_{r=0}$ (shown by the solid line in figure 12b), is $99\,\%$ of its fully developed value. We choose this quantity because it represents radial viscous effects and the axis is the last radial location where the viscous diffusion from the wall is felt. We compute $C_{e,u2}=0.085$. We then find $x_{e,\delta }=4 C_{e,\delta } R^2 Re_\lambda$, i.e. the downstream distance from the entrance where $\delta _R=0.99$$R$, and obtain $C_{e,\delta }=0.059$. Table 1 summarizes the computed entrance lengths.
We therefore find that $C_{e,u2}$, $C_{e,\delta }>C_{e,u}, C_{e,pois}$, i.e. the flow becomes viscous for the whole wall-normal extent of the pipe slightly downstream from where the flow can be considered in good agreement with the Poiseuille profile. Therefore, there does not exist a distinct streamwise region along which viscous diffusion affects the whole wall-normal extent of the pipe and the velocity profile has not yet developed to the parabolic profile, which appears to be in contradiction with the statement by Crabtree, Küchemann & Sowerby. The adjustment length $x_{e,u2}$, based on the second wall-normal derivative of the streamwise velocity at the pipe axis, is the most conservative amongst the four lengths, as also visually evident in figure 12(b).
4.2. Perturbation flow
The downstream evolution of the perturbation flow is discussed in this section. The scaling adopted in § 2, which leads to $\boldsymbol {\bar {u}_{x,0}}=\boldsymbol {\bar {u}_{x,0}}(\bar {x},r;Re_\lambda, k_x, n,m,R,\hat u_x^\infty,\hat u_r^\infty )$, is useful for the asymptotic analysis and it relates directly to the open-boundary-layer case of Leib et al. (Reference Leib, Wundrow and Goldstein1999). However, it is does not convey an immediate physical meaning as, for example, the scaled frequency $k_x$ appears as an independent variable and in the scaling of $\bar {x}$. We therefore express the solution as $\boldsymbol {\bar {u}_{x,0}}=\boldsymbol {\bar {u}_{x,0}}(x_R,r_R;Re_R,\lambda _{xR},n,\lambda _R,\hat u_x^\infty,\hat u_r^\infty )$, where the subscript $R$ indicates scaling by the pipe radius $R^*$, i.e. $\lambda _R=\lambda ^*/R^*$. In the cylindrical geometry, varying the index $n$ allows studying the role of radial length scale, i.e. the larger $n$, the smaller the radial length scale.
In the analysis, we fix $\hat u_{x,mn}^\infty =1$, and the kinetic energy of the gust at the pipe mouth, defined as
(where $r_R=r^*/R^*$), is kept equal to $\mathcal {E}_{gust,11}$, the energy for $m=n=1$ and $\hat u^\infty _{r,11}=1$ ($u_{g,x},u_{g,r},u_{g,\theta }$ are the amplitudes of the free stream gust velocity components, defined in (2.1)). Expression (4.3) is written explicitly as
where $\bar {r}=\xi _{mn}r_R/2$ and $\hat u^\infty _{\theta,mn}$ has been eliminated by using (2.2). Setting $\mathcal {E}_{gust}=\mathcal {E}_{gust,11}$ allows computing $\hat u^\infty _{r,mn}$ for $m,n$ of choice and use of (2.2) gives $\hat u^\infty _{\theta,mn}$.
We first study the velocity and pressure profiles of a reference case that is representative of realistic water and air pipe-flow experiments. Table 2 presents the flow parameters of these cases. We then study the effect of the inflow parameters on the perturbation field.
4.2.1. Initial reference-case flow
The region-I-II asymptotic initial conditions require a smooth matching with the numerical solutions of the boundary-region equations (2.11) and (2.12) at small $x_R$ locations. This matching is monitored by comparison in figure 13. The asymptotic profiles (dashed lines) are consistent with the numerical solutions (solid lines) at small $x_R$. The $\bar {u}_x$ profiles resemble the profiles of the Klebanoff modes appearing in free stream open boundary layers due to free stream vortical disturbances, shown in figure 3 on page 184 of Leib et al. (Reference Leib, Wundrow and Goldstein1999). This profile shape is expected because at short distances from the pipe entrance the flow confinement and the base pressure gradient do not play a leading role, a valid representation being given by the free stream boundary-layer region-II profile (3.25). Nevertheless, the amplitude and growth rate of $\bar {u}_x$ in our case are given by $\bar {u}_{\theta,com}$ in (3.23) and are thus influenced by the presence of the pipe wall. At these locations the dominant part of the total streamwise disturbance velocity $\bar {u}_{x,0}$ is due to $\bar {u}_x^{(0)}$, which is given in (3.27) and is caused by the direct free stream forcing action of the inviscid streamwise velocity in (3.19). The part due to $\bar {u}_x$ is smaller, which means that the streaks have not formed yet and do not dominate the boundary layers over the pipe wall.
4.2.2. Downstream evolution of the reference-case flow
The downstream evolution of the disturbance field for the reference case of table 2 is investigated ($n=3,m=4$), assigned as perturbations at the pipe entrance. The inflow perturbations are streamwise-stretched vortices with comparable radial and azimuthal length scales in cross-sectional $r-\theta$ planes and a much longer streamwise wavelength. As discussed in § 2.3, at this bulk Reynolds number, $Re_R=1500$, no TS waves exist, but nevertheless an intense transient growth is detected. This growth is monitored by the energy of the perturbation, defined as
where $\bar {u}_{x,0}, \bar {u}_{r,0}, \bar {u}_{\theta,0}$ are defined in (2.7). Figure 14(a) shows that, while at the entrance the three velocity components have comparable intensity, the streamwise velocity component is the major contributor to the perturbation dynamics as the total energy increases from the pipe inlet, while the other two velocity components decay at comparable rate, as shown in the inset of figure 14(a). Downstream of $x_R=20$, the total energy is almost entirely due to $\bar {u}_{x,0}$. It reaches its peak at $x_R=28$ and then decays monotonically because of viscous dissipation. Almost no perturbation energy is computed downstream of $x_R=120$. The transient growth is confined to region V, upstream of the fully developed region VI, because the entry region, estimated through the centreline base velocity, terminates at $x_R=339$. The energy growth and decay is qualitatively similar the transient results in fully developed pipe flows by Reshotko & Tumin (Reference Reshotko and Tumin2001) in the spatial case and by O'Sullivan & Breuer (Reference O'Sullivan and Breuer1994) and Trefethen et al. (Reference Trefethen, Trefethen and Schmid1999) in the temporal case if the time scale is converted to the axial coordinate by use of the streamwise velocity at the pipe axis for qualitative comparison.
As shown in figure 14(b), the maximum of the perturbation energy near the pipe mouth is confined near the pipe wall, within the base-flow viscous layer. As the flow evolves downstream, the perturbation spreads towards the core of the pipe, although its peak remains in the outer quarter radial region during the entire flow evolution (radial location $r_{\mathcal {E}}>0.6$). The inset of figure 14(b) depicts the downstream evolution of the magnitude of the pressure at the wall, i.e.
Relation (4.6) is obtained from the $\theta$-momentum equation (S2.8). The wall pressure decays up to $x_R=0.5$, grows to a maximum at $x_R=4$, and then decays monotonically downstream.
The radial profiles of $|\bar {u}_{x,0}|$ are shown in figure 15(a), for growing disturbances, and figure 15(b), for decaying disturbances. The base-flow streamwise gradients enhance the perturbation near the wall and the maximum of $|\bar {u}_{x,0}|$ moves towards the centreline at all streamwise locations. During the growth phase, the wall-shear stress induced by the perturbation remains constant, while it decays when the perturbation energy decreases.
The azimuthal velocity $|\bar {u}_{\theta,0}|$ in figure 16(a) decays at a faster rate near the centreline and so does the radial velocity $|\bar {u}_{r,0}|$ (not shown). Figure 16(b) confirms that the pressure $|\bar {p}_0|$ is closely related to $|\bar {u}_{\theta,0}|$ as the near-wall maximum of $|\bar {u}_{\theta,0}|$ is located at the same radial position of the local maximum of $|\bar {p}_0|$, although the overall maximum pressure disturbance is found at the pipe wall.
Figure 17 shows contour plots of the velocity components and pressure at a fixed time instant in a cross-plane where the total perturbation energy reaches its maximum. The spanwise and streamwise velocity components appear twisted about the pipe axis with respect to the initial symmetrical distribution observed near the pipe mount, shown in figure 8.
4.2.3. Effect of Reynolds number
Figure 18 shows the evolution of the total energy $\mathcal {E}$ at different bulk Reynolds numbers $Re_R$. As $Re_R$ increases up to approximately $Re_R=6000$, the perturbation energy increases, with the peak occurring farther downstream. The initial growth rate is independent of $Re_R$. For $6000< Re_R<10\,000$, the evolution of the energy up to the maximum is Reynolds number independent, while downstream of the peak the energy decays at a slower rate as $Re_R$ increases. This dependence on the Reynolds number also occurs in free stream boundary layers exposed to convected gusts, where the spanwise wavenumber $\kappa =k_z/\mathcal {F}^{1/2}$ is the governing parameter (Leib et al. Reference Leib, Wundrow and Goldstein1999). In figures 5 and 6 of Leib et al. (Reference Leib, Wundrow and Goldstein1999), as $\kappa$ decreases from 1 to 0.1 the the global perturbation energy, dominated by the streamwise velocity, increases and its maximum occurs farther downstream. This behaviour is thus consistent with figure 18, where $\kappa =m/(k_{x,R} Re_R)^{1/2}$ is in the range $0.12<\kappa <0.3$ for $1500< Re_R<10\,000$. The radial locations of the energy maxima move closer to the wall at larger $Re_R$ as the boundary layers becomes thinner (not shown).
4.2.4. Effect of streamwise wavelength
Studying the influence of the streamwise wavelength $\lambda _{xR}$ of the oncoming disturbance is equivalent to studying that of the frequency because the two quantities are inversely proportional to each other and related through $U_\infty ^*$ for small-amplitude free stream gusts. Figure 19 shows that the initial amplitude of the perturbation energy is independent of $\lambda _{xR}$. Up to $x_R=10$, $\lambda _{xR}$ has no influence on the disturbance growth as the profiles overlap on one another. Farther downstream, longer wavelengths (i.e. smaller frequencies) lead to perturbations that grow more intensely, with their maxima occurring at larger $x_R$ locations.
The trends at the two highest $\lambda _{xR}$ overlap (red line and thin black line), showing the independence of the dynamics in the limit of low frequency. The asymptotic scaling at large $\lambda _{xR}$ can be studied for $\epsilon = \mathcal {F} \ll 1$. In this limit, we introduce $\hat U = U$, $\hat V = \epsilon V$, $\hat x = \bar {x}/\epsilon$ and $\hat {u}_x = \bar {u}_x/\epsilon$. All the terms in the base-flow equations (2.8)–(2.9) are retained and the coefficient in front of the parenthesis in (2.9) changes to unity. The boundary-region equations (2.11)–(2.12) simplify as all the terms in (2.13)–(2.26) that are proportional to $i$, which are due to the unsteadiness of the oncoming disturbance (2.1), are asymptotically smaller, and $\mathcal {F}$ changes to unity. The low-frequency disturbance dynamics is thus steady at leading order. The analogous scaling for open free stream boundary layers was first found by Leib et al. (Reference Leib, Wundrow and Goldstein1999) on pages 183–185. In figure 19, the behaviour is consistent with the asymptotic analysis for $\epsilon \ll 1$: $Re_R$ is constant for those cases and thus it does not affect the scaling $\hat x$. At leading order the total energy is
which is independent of $\lambda _{xR}$. The prediction that the optimum conditions leading to the maximum transient growth are steady is fully consistent with the optimal-growth results of vortical perturbations that are present inside a Blasius boundary layer and evolve downstream from locations near the leading edge (Andersson et al. Reference Andersson, Berggren and Henningson1999; Luchini Reference Luchini2000). Figure 19 shows that oncoming perturbations with measurable unsteadiness, i.e. with frequency $f^*=1$ Hz ($\lambda _{xR}=200$) for the air pipe in table 2, are equally likely to lead to the flow breakdown as they achieve a growth which is very close to the optimal one.
4.2.5. Effect of radial characteristic scale
In the expansion based on Fourier–Bessel coefficients (2.1), the characteristic radial scale is defined by the zeros of the Bessel function $\xi _{mn}$. For fixed azimuthal index $m$, the larger the radial coefficient $n$, the smaller the radial length scale, as shown in figure 20(a) by the profiles of the real part of $\bar {u}_r$ at small $x_R$ locations. The transient growth decreases as the radial length scale of the free stream disturbance becomes smaller, as shown in figure 20(b). This effect is due to the more intense viscous dissipation caused by the large velocity gradients related to the small radial scale. The location of the maximum energy is computed closer to the pipe inlet as the radial scale decreases. For a Fourier–Bessel coefficient $n=9$, the perturbation decreases upstream of the expected transient growth and subsequent viscous dissipation. For $n=12$, the perturbation decreases at any streamwise location.
The influence of the radial scale on the disturbance evolution cannot be investigated through the non-normality of the eigenfunctions of the developing boundary layers on the pipe wall. An eigenvalue-based approach would model the near-entrance disturbance as completely confined within the base-flow boundary layer, whereas the radial scale only pertains to the free stream gust flow and does not enter the classical stability problem of the boundary layer. The adopted boundary-region approach instead takes into account all the characteristics of the gust flow because, being an initial-value problem, the specification of the pipe-entrance flow is vital for the evolution of the perturbation inside the pipe.
4.2.6. Effect of azimuthal wavelength
Figure 21(a) shows the real part of $\bar {u}_\theta$ for $n=6$ near the pipe mouth, at $x_R=0.18$, for different $\lambda _{zR}=2{\rm \pi} /m$. For this $n$ value, varying $m$ only impacts on the azimuthal modulation of the disturbance, while the characteristic radial scale is maintained. The influence of the azimuthal wavelength $\lambda _{zR}$ on the perturbation energy inside the pipe is shown in figure 21(b). There exists an optimal $\lambda _{zR}$ (the reference $\lambda _{zR}=2.09$, $m=3$) that leads to the most energetic growth. This azimuthal coefficient was also found to lead to the most energetic spatial transient disturbance by Mayer & Reshotko (Reference Mayer and Reshotko1997) in a fully developed laminar flow. Disturbances with smaller $\lambda _{zR}$ are influenced by viscous dissipation that hampers their growth. Oncoming disturbances with a larger azimuthal wavelength, $\lambda _{zR}=3.14$ ($m=2$) are almost two-dimensional and do not lead to the maximum growth because the dynamics of the streaks is a strictly three-dimensional phenomenon (Leib et al. Reference Leib, Wundrow and Goldstein1999), unlike TS waves, the amplification of which is maximum in the two-dimensional case (Schmid & Henningson Reference Schmid and Henningson2001). The perturbations with $\lambda _{zR}=3.14$ persist downstream to larger distances than in the optimal case because they are less influenced by viscous dissipation due to azimuthal shear effects.
4.2.7. Comparison with experimental and numerical data
Wygnanski & Champagne (Reference Wygnanski and Champagne1973) experimentally studied the air flow in the entrance region of a circular pipe in order to understand the formation of turbulent slug and puffs. Just like Reynolds (Reference Reynolds1883), they recognized the impact of the entry disturbances on the flow character farther downstream, especially the intensity of inlet perturbations. They utilized an orifice, a suspended circular disk and a honeycomb to generate the perturbation at the pipe mouth. In the entrance region, i.e. where the flow is stable according to classical stability theory, Wygnanski & Champagne (Reference Wygnanski and Champagne1973) detected intense growth of low-frequency oscillations upstream of the turbulent region and proposed that ‘… the growth of perturbations is associated with the large shear existing within the boundary layer …’. They also observed that the longitudinal fluctuations rose abruptly, much more than the cross-flow velocity components. These experimental results fully agree with our theoretical results.
Figure 22 shows their flow visualization of the pipe entrance flow by dye injection at two downstream locations. It is evident that, upstream of where the flow breaks down to turbulent puffs (right-hand part of figure 22a), the disturbances are elongated in the axial direction and characterized by a well-defined azimuthal length scale, as predicted by the low-frequency theory.
Figure 23 presents a comparison between our calculations and three sets of experimental data by Wygnanski & Champagne (Reference Wygnanski and Champagne1973). In figure 23(a,b), the base-flow comparison is excellent for the case at low level of free stream turbulence (figure 23a; $\vert u'\vert /\bar {U}_{cl}=0.16\,\%$, where $\bar {U}_{cl}$ is the centreline mean velocity), whereas the experimental base-flow data at much higher levels of free stream turbulence (figure 23b; $\vert u'\vert /\bar {U}_{cl}=5.8\,\%$, and $7.8\,\%$) deviate considerably from the base flow computed via (2.8) and (2.9). These results are expected because, when the oncoming disturbances are weak (generated by the upstream honeycomb), the nonlinear effects on the perturbation dynamics are negligible and do not distort the laminar base flow. When the disturbances are more intense (generated by the orifice and the suspended disk), nonlinearity distorts the base flow, even though the flow has not broken down to turbulence. The comparisons of the perturbation data, shown in figure 23(c,d), are consistent with the mean-flow results, i.e. much better agreement is obtained in the left graph, where nonlinear effects are likely to play a marginal role. The results in figure 23(c) are robust because the profile shape varies only slightly when the frequency, and the azimuthal and radial indexes, $m$ and $n$, are varied independently about the reference case $k_{x,R}=0.118$, $n=2$, $m=3$. The flow parameter ranges were chosen based on the description of the experimental conditions given in Wygnanski & Champagne (Reference Wygnanski and Champagne1973), i.e. streamwise wavelength longer than the radius and comparable azimuthal and radial length scales because generated by the honeycomb upstream of the pipe entrance. It would certainly be interesting to carry out further comparisons between better controlled experiments and computations of the nonlinear boundary-region equations and direct numerical simulations.
Wu et al. (Reference Wu, Moin, Adrian and Baltzer2015) studied the pipe-flow entrance problem via direct numerical simulations. In one of their cases at $Re_R=4000$, the flow at the pipe inlet was a superposition of a uniform base flow and vortical disturbances located near the pipe wall with a quite distinct azimuthal modulation. The perturbation did not grow as the base flow evolved along the entry region and the breakdown to turbulence was detected much farther downstream, i.e. at locations where the laminar base flow reached the parabolic profile. This result qualitatively agrees with our theoretical prediction. A visual inspection of their figure 1D leads to an estimate of $m=8$ or larger, that is, sufficient for viscous effects to dominate and attenuate the algebraic growth, as observed in figure 21 ($m=8$ corresponds to $\lambda _{zR}=0.79$).
5. Conclusions
In this paper, we have presented the first theoretical study of the entrainment and growth of unsteady three-dimensional vortical disturbances in the entrance region of an incompressible pipe flow, as a step towards the full comprehension of laminar to turbulent transition in confined flows. This fundamental problem has been central in the fluid mechanics research community since the pioneering work of Reynolds (Reference Reynolds1883). In the other theoretical studies of the pipe-entrance flow, the entrainment of disturbances in the confined space was not studied because velocity or pressure fluctuations were absent in the pipe core flow and were only present in the viscous boundary layer attached to the pipe wall (references in § 1.3).
The mathematical framework takes inspiration from the pioneering work by Leib et al. (Reference Leib, Wundrow and Goldstein1999) on the generation of low-speed streaks in a flat-plate boundary layer. The analysis is based on the method of matched asymptotic expansions and on the assumptions of high Reynolds number and of flow perturbations of low amplitude and frequency. The low-frequency hypothesis has been motivated by the evidence that these disturbances amplify the most in free stream boundary layers to form streamwise-elongated streaks. This approach has allowed for an analytical description of realistic vortical perturbations that can be created in a laboratory set-up at the pipe inlet. The effects of the pipe confinement, the streamwise pressure gradient, the viscous/inviscid interplay and the interactions amongst the velocity components have been revealed. The amplitude of the initial disturbance is related uniquely to that of the oncoming free stream perturbation. The composite velocity profiles of the disturbance flow have been useful for the specification of the appropriate initial conditions for the boundary-region perturbation equations.
The base-flow velocity components and pressure gradient agree with results from direct numerical simulations and experiments. We propose to utilize the second derivative of the base flow at the pipe centreline to ascertain whether the flow has reached the fully developed Poiseuille flow. This method is the most conservative one among those tested for quantifying the entry length because the pipe axis is the last location where the base flow is subject to the diffusing viscous effects. Near the pipe mouth, the base pressure gradient plays a minimal role and the disturbance flow is mostly dominant within the boundary layer at the pipe wall. The perturbations appear in the form of streamwise-elongated streaky structures and eventually evolve towards the pipe core, where the base flow is still mostly inviscid. The disturbance growth is enhanced as the frequency decreases, and the bulk Reynolds number and the characteristic radial length scale increase. The azimuthal wavelength that generates the most intense downstream growth has also been computed.
The significant algebraic growth and the viscous decay only occur in the entrance region, i.e. where the base flow is still streamwise dependent. Nevertheless, as evident from the direct numerical simulations of Wu et al. (Reference Wu, Moin, Adrian and Baltzer2015), vortical disturbances that initially decay from the pipe inlet may trigger nonlinear effects farther downstream and lead the flow to the breakdown to turbulence. Our results compare favourably with the experimental data of Wygnanski & Champagne (Reference Wygnanski and Champagne1973) when the amplitude of the oncoming perturbation is low, thus suggesting that the linearized dynamics is valid. The less satisfactory agreement for more intense perturbations highlights the urgency for the inclusion of the nonlinear effects. We therefore plan to extend our analysis by relaxing the assumption of small amplitude to study the nonlinear dynamics and to explore the link between the perturbations generated at the pipe inlet and the travelling waves discovered by Faisst & Eckhardt (Reference Faisst and Eckhardt2003) and Wedin & Kerswell (Reference Wedin and Kerswell2004). Accurate experimental data are also required in order to further test our results.
Supplementary material
Supplementary materials are available at https://doi.org/10.1017/jfm.2021.1005.
Acknowledgements
The authors would like to thank the A*STAR-Sheffield Research Attachment Programme and the University of Sheffield for funding this research. We also indebted to Professor L. Tuckermann for guidance on the conditions at the pipe axis. We are also thankful to the referees for their reviews which greatly improved the quality of our work.
Declaration of interests
The authors report no conflict of interest.