1. Introduction
Mathematical models of thermal convection in horizontal fluid layers are studied both as examples of complexity in nonlinear systems and as idealizations of convention in astrophysical, geophysical, and engineering applications. Convection in a layer can be driven by internal heating or cooling, by the boundary conditions, or both. Rayleigh–Bénard (RB) convection (Siggia Reference Siggia1994; Getling Reference Getling1998; Ahlers, Grossmann & Lohse Reference Ahlers, Grossmann and Lohse2009), which has enjoyed the most attention, is driven solely by the boundary conditions. Internally heated (IH) convection, which is no less fundamental, is driven in its simplest models by constant and uniform volumetric heating. The IH configuration most commonly studied is a fluid layer bounded below by a perfect insulator and above by a perfect conductor. Here, we study a model of IH convection that also is bounded below by a perfect insulator but is bounded above by a poor conductor: a configuration considered in very few previous works (Hewitt, McKenzie & Weiss Reference Hewitt, McKenzie and Weiss1980; Ishiwatari, Takehiro & Hayashi Reference Ishiwatari, Takehiro and Hayashi1994). This model also describes the dynamics of internally cooled convection with the top and bottom boundary conditions exchanged, though here we speak only in terms of internal heating.
The model studied is of interest for several reasons. First, convection that is wholly or partly driven by internal heating or cooling occurs in the Earth’s mantle (Schubert, Turcotte & Olson Reference Schubert, Turcotte and Olson2001) and atmosphere (Berlengiero et al. Reference Berlengiero, Emanuel, von Hardenberg, Provenzale and Spiegel2012), other planetary atmospheres (Ingersoll & Porco Reference Ingersoll and Porco1978; Kaspi, Flierl & Showman Reference Kaspi, Flierl and Showman2009), the cores of large main-sequence stars (Kippenhahn & Weigert Reference Kippenhahn and Weigert1994) and engineered systems involving exothermic chemical or nuclear reactions, including nuclear accident scenarios (Asfia & Dhir Reference Asfia and Dhir1996; Nourgaliev, Dinh & Sehgal Reference Nourgaliev, Dinh and Sehgal1997; Grötzbach & Wörner Reference Grötzbach and Wörner1999). In particular, in the mantle and certain nuclear accidents, the upper boundary may be closer to a poor conductor than to the perfect conductor adopted in many models. Second, the convective configuration studied here is among the simplest possible in the sense that, when it is modelled using the Boussinesq equations, only two dimensionless parameters enter the dynamics (aside from any parameters used in describing the geometry). There are six configurations with this property (Goluskin Reference Goluskin2015), three instances of RB convection and three of IH convection, and the present model is by far the least studied of the six. Finally, the model makes for an unusually tractable ‘textbook example’ of fluid stability analysis; the linear and nonlinear stability thresholds are close but not identical, and both can be computed analytically for any boundary conditions on the velocity.
We are aware of only two studies of the present configuration (Hewitt et al. Reference Hewitt, McKenzie and Weiss1980; Ishiwatari et al. Reference Ishiwatari, Takehiro and Hayashi1994). Both examined scale selection using two-dimensional simulations, and for free-slip boundaries Ishiwatari et al. (Reference Ishiwatari, Takehiro and Hayashi1994) used long-wavelength asymptotics to find the linear instability threshold of the static state and derive an asymptotic equation for the dynamics near onset. Beyond those studies, our results can be compared with work on RB convection between poorly conducting boundaries (e.g. Sparrow, Goldstein & Jonsson Reference Sparrow, Goldstein and Jonsson1963; Hurle, Jakeman & Pike Reference Hurle, Jakeman and Pike1967; Otero et al. Reference Otero, Wittenberg, Worthing and Doering2002; Johnston & Doering Reference Johnston and Doering2009) and work on IH convection with a top that conducts perfectly, rather than poorly. The latter configuration was studied early on by Roberts (Reference Roberts1967), Tritton & Zarraga (Reference Tritton and Zarraga1967) and Thirlby (Reference Thirlby1970), in many subsequent works reviewed by Kulacki & Richards (Reference Kulacki and Richards1985), and more recently in simulations both at finite Prandtl numbers (Ichikawa et al. Reference Ichikawa, Kurita, Yamagishi and Yanagisawa2006; Cartland Glover & Generalis Reference Cartland Glover and Generalis2009; Takahashi et al. Reference Takahashi, Tasaka, Murai, Takeda and Yanagisawa2010; Cartland Glover, Fujimura & Generalis Reference Cartland Glover, Fujimura and Generalis2013) and in the infinite limit (Houseman Reference Houseman1988; Schubert, Glatzmaier & Travis Reference Schubert, Glatzmaier and Travis1993; Parmentier, Sotin & Travis Reference Parmentier, Sotin and Travis1994).
Our mathematical model and its basic features are laid out in § 2. For various boundary conditions on the velocity, linear and nonlinear stability thresholds of the static state are found in § 3. Integral quantities important to heat transport are addressed in § 4, much of which is devoted to proving a lower bound on the mean temperature, and § 5 gives concluding remarks.
2. The model
In dimensional terms, we are considering a fluid of thermal diffusivity
${\it\kappa}$
in a layer of height
$d$
, heated internally at rate
$H$
. The quantity
$H$
has units of temperature per time and is equal to the specific rate of heating, normalized by density and specific heat. Figure 1 shows a schematic of this configuration. The perfectly insulating bottom boundary is enforced by a vanishing temperature flux, and the poorly conducting top boundary is modelled by a fixed heat flux, enforced by fixing the vertical temperature gradient to
$-{\it\Gamma}$
there. The better the fluid transports heat, relative to the top boundary, the more accurate it is to model that boundary with a fixed heat flux (Hurle et al.
Reference Hurle, Jakeman and Pike1967). Our model thus describes a situation where the fluid transports heat much better than the top boundary, which in turn transports heat much better than the bottom boundary. Any layer whose top boundary is much more conductive than its bottom one should be well described by our model whenever convection is sufficiently strong.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_fig1g.gif?pub-status=live)
Figure 1. Schematic of the convective configuration studied in the present work. Quantities shown are dimensional. The internal heat source (
$H$
) and gravitational acceleration (
$g$
) are constant and uniform.
Statistically steady convection is possible only when the heat flux across the top boundary balances the internal heat production, hence we require
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn1.gif?pub-status=live)
The natural temperature scale in this system is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn2.gif?pub-status=live)
The quantity
$d^{2}H/{\it\kappa}$
is the usual temperature scale of IH convection, and it agrees in this configuration with
$d{\it\Gamma}$
, the temperature scale of fixed-flux RB convection. Modelling the dynamics using the Boussinesq equations (Spiegel & Veronis Reference Spiegel and Veronis1960; Chandrasekhar Reference Chandrasekhar1981), we non-dimensionalize lengths by
$d$
, temperatures by
${\it\Delta}$
and times by the characteristic timescale of thermal diffusion,
$d^{2}/{\it\kappa}$
. The dimensionless dynamics are then governed by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn3.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn4.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn5.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline21.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline22.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline23.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn6.gif?pub-status=live)
where
$g$
is the constant gravitational acceleration acting in the
$-\hat{\boldsymbol{z}}$
direction,
${\it\alpha}$
is the linear coefficient of thermal expansion,
${\it\Delta}$
is the temperature scale defined by (2.2) and
${\it\nu}$
is the kinematic viscosity. The above definition of
$R$
agrees with the usual definitions of Rayleigh numbers as control parameters in IH convection (Kulacki & Richards Reference Kulacki and Richards1985) and fixed-flux RB convection (Sparrow et al.
Reference Sparrow, Goldstein and Jonsson1963). As in RB convection, this
$R$
is roughly the ratio of inertial forces to viscous forces and can be thought of as the strength with which the flow is driven.
The spatial domain of our model has a dimensionless vertical extent of
$0\leqslant z\leqslant 1$
and is infinite or periodic in its one or two horizontal directions. The dimensionless thermal boundary conditions are
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn7.gif?pub-status=live)
For the velocity conditions at the top and bottom, we consider all four permutations of no-slip and free-slip boundaries, which are enforced by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn8.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn9.gif?pub-status=live)
When the fluid is static, the unique steady temperature field is horizontally uniform and has the parabolic vertical profile
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn10.gif?pub-status=live)
as shown in figure 2(a). By contrast, figure 2(b) shows the sort of temperature profile that we expect at large
$R$
, where experience with similar systems suggests that strong convective mixing will render the fluid roughly isothermal outside of an upper thermal boundary layer. Integral quantities related to mean temperature profiles are discussed in § 4.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary-alt:20170721122512-18740-mediumThumb-S0022112015001408_fig2g.jpg?pub-status=live)
Figure 2. Schematics of mean vertical temperature profiles (a) in the static state and (b) as expected in strong convection. The dimensionless thermal boundary conditions are shown.
To provide a concrete example of strong convection in our model, we carried out a two-dimensional simulation using nek5000 (Fisher, Lottes & Kerkemeier Reference Fisher, Lottes and Kerkemeier2013). Figure 3 shows a typical temperature field from that simulation. As expected, cold plumes descend from an upper thermal boundary layer. We have not collected quantitative data on heat transport in this model, nor to the best of the author’s knowledge has anyone else.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary-alt:20170721122512-04722-mediumThumb-S0022112015001408_fig3g.jpg?pub-status=live)
Figure 3. Instantaneous temperature field from a two-dimensional simulation of our model with
$R=1.44\times 10^{8}$
,
$\mathit{Pr}=1$
, a horizontal period of 6 and no-slip boundaries. The hottest fluid (white) is 0.06 dimensionless degrees warmer than the coldest fluid (black).
3. Stability of the static state
To determine the stability of the static state, wherein
$\boldsymbol{u}=\mathbf{0}$
and
$T=T_{st}$
, we decompose the temperature field as
$T(\boldsymbol{x},t)=T_{st}(z)+{\it\theta}(\boldsymbol{x},t)$
, where
${\it\theta}$
is called the temperature fluctuation. Under the Boussinesq equations (2.3)–(2.5), fluctuations evolve according to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn11.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn12.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn13.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline39.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline40.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline41.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline42.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline43.gif?pub-status=live)
3.1. Linear stability eigenproblem
To find a threshold for the linear instability of infinitesimal perturbations, we neglect the nonlinear terms in the fluctuation equations (3.1)–(3.3). The first half of the procedure for finding
$R_{L}$
closely follows the classic calculation for RB convection (Lord Rayleigh 1916; Chandrasekhar Reference Chandrasekhar1981). The linearizations of (3.3) and
$\hat{\boldsymbol{z}}\boldsymbol{\cdot }\boldsymbol{{\rm\nabla}}\times \boldsymbol{{\rm\nabla}}\times (3.2)$
form a closed pair of evolution equations for
$w$
and
${\it\theta}$
that have the same linear stability threshold as the full equations:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn14.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn15.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline48.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn16.gif?pub-status=live)
The final two conditions on
$w$
depend on whether the boundaries are no-slip or free-slip and are derived from the definitions (2.8)–(2.9) with the help of the incompressibility condition (2.3). We consider all four combinations here:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn17.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn18.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn19.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn20.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline50.gif?pub-status=live)
The right-hand side of (3.4)–(3.5) can be regarded as a linear operator acting on
$[{\rm\nabla}^{2}w\,{\it\theta}]^{\text{T}}$
. At the stability threshold we seek, the spectrum of the operator is marginally stable, meaning at least one eigenvalue has a vanishing real part. Here we look only for marginally stable states that are stationary, as opposed to time-dependent. Such time-independent states obey the linear eigenproblem
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn21.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn22.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline52.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline53.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn23.gif?pub-status=live)
The
$R$
for which non-zero solutions exist are generalized eigenvalues; at such an
$R$
there is a zero in the spectrum of an operator taking the form
$\mathscr{A}-R\mathscr{B}$
, where
$\mathscr{A}$
and
$\mathscr{B}$
are linear differential operators.
The condition
$R>R_{L}$
is sufficient for linear instability, but, because we have assumed stationarity, it may not be necessary. Showing that it is necessary would require proving that all marginal states are indeed stationary. This is fairly easy in RB convection (Pellew & Southwell Reference Pellew and Southwell1940), but the analogous proof fails here because of the non-constant coefficient in (3.5). A functional analytical approach has been used to prove stationarity in certain IH configurations (Herron Reference Herron2001, Reference Herron2003) but apparently not for fixed-flux thermal boundary conditions.
Because the eigenproblem (3.11)–(3.12) is linear and lacks horizontal boundaries, we can Fourier transform it in
$x$
and
$y$
(or, equivalently, apply a normal mode substitution). This yields a separate eigenproblem for each magnitude,
$k$
, of the horizontal wavevector:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn24.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn25.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline63.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline64.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline65.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline66.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline67.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline68.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn26.gif?pub-status=live)
where
$k^{2}$
cannot be zero because horizontally uniform
$w$
would violateincompressibility.
In similar models of IH convection, the value of
$R_{L}$
must be found by solving the eigenproblem (3.14)–(3.15) numerically for various fixed
$k$
(Roberts Reference Roberts1967; Kulacki & Goldstein Reference Kulacki and Goldstein1975). In the present case, we have carried out such numerics only to confirm that, for all four pairs of velocity conditions, the generalized eigenvalue
$R$
decreases monotonically as
$k^{2}\rightarrow 0$
. The infimum of expression (3.16) can thus be replaced by the long-wavelength limit,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn27.gif?pub-status=live)
and an exact analytical expression for
$R_{L}$
can be found by asymptotically expanding the eigenproblem in
$k^{2}$
. This has been done for free-slip boundaries by Ishiwatari et al. (Reference Ishiwatari, Takehiro and Hayashi1994) and is carried out for other velocity conditions in § 3.3.
Monotonic decrease of the generalized eigenvalue
$R$
as
$k^{2}\rightarrow 0$
has been found in various other convective systems where heat fluxes are fixed on both boundaries (Sparrow et al.
Reference Sparrow, Goldstein and Jonsson1963; Chapman, Childress & Proctor Reference Chapman, Childress and Proctor1980; Depassier & Spiegel Reference Depassier and Spiegel1982). We do not know of any analytical proofs of this feature, although it seems to be a fairly general consequence of such boundary conditions.
3.2. Energy stability eigenproblem
Returning to the nonlinear fluctuation equations (3.1)–(3.3), we now seek a Rayleigh number,
$R_{E}$
, below which the static state is globally stable. As in most studies of fluid stability, we prove such a threshold using the energy method (Serrin Reference Serrin1959; Joseph Reference Joseph1976; Straughan Reference Straughan2004). In particular, we follow Joseph (Reference Joseph1965) in considering (generalized) energies of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn28.gif?pub-status=live)
where
${\it\gamma}>0$
is a coupling parameter to be chosen later, and where
$\unicode[STIX]{x2A0F}$
denotes an average over the volume. The stability of the static state follows if the energy is a Lyapunov functional; that is, if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn29.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn30.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline82.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline83.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline84.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline85.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline86.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline87.gif?pub-status=live)
We prove stability up to the largest threshold we can by choosing the value of
${\it\gamma}$
that maximizes
$R_{E_{{\it\gamma}}}$
. An even larger threshold might be proven by optimizing over a broader family of Lyapunov functions than the ansatz (3.18). However, the condition (3.20) is generally very hard to check for a candidate functional. Like most authors, with a few exceptions (Kaiser, Tilgner & Von Wahl Reference Kaiser, Tilgner and Von Wahl2005; Chernyshenko et al.
Reference Chernyshenko, Goulart, Huang and Papachristodoulou2014), we have avoided this difficulty by the energy method, which entails restricting ourselves to Lyapunov functions that (i) are quadratic in the fluctuation variables, and (ii) are conserved by the nonlinear terms of the evolution equations (3.1)–(3.3). For such energies, a condition on the spectrum of a linear eigenproblem, introduced below, suffices to guarantee that
$(\text{d}/\text{d}t)E_{{\it\gamma}}\leqslant 0$
.
The spectral condition that is sufficient for Lyapunov stability has been derived and solved numerically in similar IH configurations (Kulacki & Goldstein Reference Kulacki and Goldstein1975; Straughan Reference Straughan1990). We can see how the eigenproblem arises by adding the volume averages of
$(1/(\mathit{Pr}\,R))\boldsymbol{u}\boldsymbol{\cdot }(3.2)$
and
${\it\gamma}{\it\theta}\times (3.3)$
, and then integrating by parts to find
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn31.gif?pub-status=live)
Relaxing the dynamical constraints on
$\boldsymbol{u}$
and
${\it\theta}$
in the above expression gives
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn32.gif?pub-status=live)
where the infimum of the functional is over sufficiently smooth
$\boldsymbol{u}$
and
${\it\theta}$
that are subject only to incompressibility and the dynamical boundary conditions. The Euler–Lagrange equations of this functional, like the linear stability equations, can be reduced to a closed system for
$w$
and
${\it\theta}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn33.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn34.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline99.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline100.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline101.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline102.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline103.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline104.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn35.gif?pub-status=live)
As in the linear stability analysis of § 3.1, we can Fourier transform in the horizontal directions to get an ordinary differential equation eigenproblem for each horizontal wavevector magnitude,
$k$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn36.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn37.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn38.gif?pub-status=live)
Since
$E_{{\it\gamma}}$
is only certain to be a valid Lyapunov functional when
$R<R_{E_{{\it\gamma}}}$
, we get the strongest result by choosing
${\it\gamma}$
to maximize
$R_{E_{{\it\gamma}}}$
. This optimized threshold is what we call
$R_{E}$
. That is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn39.gif?pub-status=live)
As in the linear stability analysis, the infimum in the above expression occurs as
$k^{2}\rightarrow 0$
. We have confirmed this statement, at least for the optimal values of
${\it\gamma}$
that we eventually choose, by numerically solving the eigenproblem (3.26)–(3.27) for various
$k$
. Again we can replace the infimum with the
$k^{2}\rightarrow 0$
limit,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn40.gif?pub-status=live)
and we can solve the energy stability eigenproblem by expanding asymptotically in
$k^{2}$
.
3.3. Analytical solution of the stability eigenproblems
To evaluate expression (3.17) for
$R_{L}$
and expression (3.30) for
$R_{E}$
, we expand the eigenproblems (3.14)–(3.15) and (3.26)–(3.27), respectively, in the small quantity
$k^{2}$
. Long-wavelength expansions have been applied previously to convective models with fixed-flux thermal boundary conditions, both to find
$R_{L}$
and to capture the nonlinear dynamics near onset (Chapman et al.
Reference Chapman, Childress and Proctor1980; Chapman & Proctor Reference Chapman and Proctor1980; Ishiwatari et al.
Reference Ishiwatari, Takehiro and Hayashi1994; Childress & Spiegel Reference Childress, Spiegel, Givoli, Grote and Papanicolaou2004), although we are not aware of their use in finding
$R_{E}$
.
Anticipating that
$O({\hat{w}})=k^{2}O(\hat{{\it\theta}})$
in the asymptotic solutions of both eigenproblems, we apply the expansions
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn41.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn42.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn43.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline122.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline123.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline124.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline125.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline126.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline127.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline128.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn44.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn45.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline129.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline130.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline131.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline132.gif?pub-status=live)
In both the linear and energy analyses, the
${\it\theta}_{0}$
equations and their boundary conditions require that
${\it\theta}_{0}$
be constant. The non-zero constants arbitrarily fix the magnitudes of the eigenfunctions, so we take
${\it\theta}_{0}\equiv 1$
for convenience. The
$W_{0}$
equations give
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn46.gif?pub-status=live)
where
$P(z)$
and
$Q_{{\it\gamma}}(z)$
are the unique polynomials of orders 4 and 5, respectively, that satisfy the
$w$
boundary conditions and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn47.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn48.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline140.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline141.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline142.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline143.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline144.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline145.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline146.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn49.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn50.gif?pub-status=live)
Values of
$R_{L}$
for various boundary conditions on the velocity are given in table 1. These result from evaluating the integral (3.39) with the
$P(z)$
given in the appendix A. Since no-slip boundaries exert stresses that slow the fluid, it is unsurprising that the Rayleigh number needed to induce convection is smallest when the velocity conditions are both free-slip, larger when the conditions are mixed, and larger still when the conditions are both no-slip. When the velocity boundary conditions are mixed,
$R_{L}$
is smaller when the top boundary is the free-slip one. This is reasonable because it is the unstable temperature gradient near the top boundary that drives the flow, and we expect from related studies of IH convection (Kulacki & Richards Reference Kulacki and Richards1985) that mean velocities will be larger in the top half of the layer. A free-slip top thus encourages motion more than a free-slip bottom does.
Table 1. Rayleigh numbers above which the static state is linearly unstable (
$R_{L}$
) and below which the static state is Lyapunov stable (
$R_{E}$
), along with the percentage of
$R_{L}$
by which
$R_{E}$
falls short of
$R_{L}$
. Exact expressions for
$R_{E}$
are given in table 2. The finding
$R_{L}=240$
for free-slip boundaries agrees with Ishiwatari et al. (Reference Ishiwatari, Takehiro and Hayashi1994).
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_tab1.gif?pub-status=live)
Table 2 gives the expressions for
$R_{E_{{\it\gamma}}}$
that are found by evaluating the integral (3.40) with the
$Q_{{\it\gamma}}(z)$
given in the appendix A. Also shown are the optimal coupling parameters,
${\it\gamma}^{\ast }$
, that maximize these
$R_{E_{{\it\gamma}}}$
, as well as the maximized values,
$R_{E}$
.
Table 2. The threshold (
$R_{E_{{\it\gamma}}}$
) below which the energy
$E_{{\it\gamma}}$
is proven to be a Lyapunov functional for the static state, the optimal coupling parameter (
${\it\gamma}^{\ast }$
) that maximizes this threshold, and the maximized threshold (
$R_{E}$
). Numerical approximations of
$R_{E}$
are given in table 1.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary-alt:20170721122512-04406-mediumThumb-S0022112015001408_tab2.jpg?pub-status=live)
The approximate numerical value of each
$R_{E}$
is given alongside the corresponding
$R_{L}$
in table 1. For each pair of velocity conditions,
$R_{E}$
falls short of
$R_{L}$
by less than 1 %. It is unknown whether subcritical convection can occur in the small gap between
$R_{E}$
and
$R_{L}$
, although it can indeed occur when the top boundary is a perfect conductor, rather than a poor one (Tveitereid & Palm Reference Tveitereid and Palm1976; Busse Reference Busse2014).
4. Heat transport
The stability analysis of the static state in § 3 determines whether convection occurs at a given
$R$
, except in the narrow range between
$R_{E}$
and
$R_{L}$
. At
$R$
large enough for convection to occur, we would like to predict quantitative features of the flow, especially quantities related to vertical heat transport.
Net heat transport is fixed in our model, being equal to both the flux at the top boundary and the rate of internal heating. The relative contributions of convection and conduction to the net transport, on the other hand, are dynamically determined. Several integral quantities are useful in characterizing these contributions. One such quantity, the mean fluid temperature, we bound from below in § 4.1. Other useful quantities are discussed in § 4.2.
For our notation, we let an overline denote an average over the horizontal directions and infinite time, while angle brackets denote an average over volume and infinite time. Assuming periodicity on a horizontal domain of
$[0,L_{x}]\times [0,L_{y}]$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn51.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn52.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline195.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline196.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline197.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline198.gif?pub-status=live)
4.1. Lower bound on mean temperature
We now prove a lower bound on
$\langle T-\overline{T}_{T}\rangle$
: the mean temperature of the fluid, relative to the mean temperature of the top boundary,
$\overline{T}_{T}$
. The comparison with
$\overline{T}_{T}$
is crucial; the volume average of
$T$
alone cannot change from its initial value and so says nothing about the flow. The quantity
$\langle T-\overline{T}_{T}\rangle$
satisfies the uniform bounds
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn53.gif?pub-status=live)
The lower bound follows from expression (4.7) below. To derive the upper bound, we integrate
$z^{2}$
against the
$T$
(2.5) to find
$1/3-\langle T-\overline{T}_{T}\rangle =\langle zwT\rangle$
. Incompressibility then gives
$\langle zwT\rangle =\langle zw{\it\theta}\rangle$
, multiplying
${\it\theta}$
against the temperature fluctuation (3.3) gives
$\langle zw{\it\theta}\rangle =\langle |\boldsymbol{{\rm\nabla}}{\it\theta}|^{2}\rangle \geqslant 0$
, and combining these relations gives
$\langle T-\overline{T}_{T}\rangle \leqslant 1/3$
.
The upper bound on
$\langle T-\overline{T}_{T}\rangle$
is saturated only by the static state, in which heat is transported up and out of the layer by conduction alone. When
$R$
is raised, and some of the net heat transport is taken over by convection,
$\langle T-\overline{T}_{T}\rangle$
must fall. Since
$R$
is proportional to the dimensional heating rate,
$H$
, this decrease in
$\langle T-\overline{T}_{T}\rangle$
might seem counterintuitive until we recall that temperature has essentially been normalized by its value in the static state. If
$\langle T-\overline{T}_{T}\rangle$
decreases as
$R$
is raised, this means only that the dimensional mean temperature,
${\it\Delta}\langle T-\overline{T}_{T}\rangle$
, grows sublinearly with
$H$
.
Experience with other convective systems strongly suggests that the flow will become ever more energetic and complicated as
$R\rightarrow \infty$
, and meanwhile
$\langle T-\overline{T}_{T}\rangle \rightarrow 0$
. The belief that
$\langle T-\overline{T}_{T}\rangle$
vanishes is analogous to the belief that the Nusselt number grows unboundedly in RB convection, and we are not aware of any method for proving such claims. What we can prove, beyond the uniform bounds (4.3), is a lower bound on how quickly
$\langle T-\overline{T}_{T}\rangle$
decreases toward zero. The proof occupies the remainder of this subsection, but its result at leading order in
$R$
is simply
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn54.gif?pub-status=live)
In dimensional terms, the above bound says that the mean temperature, relative to that of the top boundary, grows with the heating rate no slower than
$H^{2/3}$
. We cannot judge the tightness of this bound since we are unaware of any experimental studies of
$\langle T-\overline{T}_{T}\rangle$
.
4.1.1. Background decomposition
We derive a lower bound on
$\langle T-\overline{T}_{T}\rangle$
using the background method (Doering & Constantin Reference Doering and Constantin1992; Constantin & Doering Reference Constantin and Doering1996). The result is proven only for no-slip boundaries, much like the related bound that has been proven for fixed-flux RB convection (Otero et al.
Reference Otero, Wittenberg, Worthing and Doering2002). The background method entails decomposing the temperature field into a chosen background profile,
${\it\tau}(z)$
, and remaining part,
${\it\Theta}(\boldsymbol{x},t)$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn55.gif?pub-status=live)
The bound we obtain depends on our choice of
${\it\tau}$
. This
${\it\tau}$
does not generally solve the governing equations, in which case
${\it\Theta}$
does not obey the fluctuation equations (3.1)–(3.3).
The background profile
${\it\tau}$
must satisfy three conditions. First, it must be continuous. Second, it must obey the same fixed-flux boundary conditions (2.7) as
$T$
, so that
${\it\Theta}$
satisfies the homogenous conditions
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn56.gif?pub-status=live)
As explained shortly, the boundary conditions on
${\it\tau}$
do not actually constrain our choice of background profile since they can be met by vanishingly thin boundary layers that do not affect the resulting bound. Finally, we must choose a
${\it\tau}$
for which we can show that a particular functional
$\mathscr{Q}$
, introduced below, is non-negative.
We will bound
$\langle T-\overline{T}_{T}\rangle$
subject not to its full dynamical constraints but only to incompressibility, boundary conditions, and three integral relations found by taking
$\langle T\times (2.5)\rangle$
,
$\langle {\it\tau}\times (2.5)\rangle$
and
$\langle \boldsymbol{u}\boldsymbol{\cdot }(2.4)\rangle$
. After integration by parts, these relations are
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn57.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn58.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn59.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline244.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline245.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline246.gif?pub-status=live)
The quantity we seek to bound appears in relation (4.7), which can be expanded as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn60.gif?pub-status=live)
As done by Lu, Doering & Busse (Reference Lu, Doering and Busse2004), we apply relations (4.8)–(4.9) to (4.10) and find
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn61.gif?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn62.gif?pub-status=live)
and
$a>0$
is to be fixed later. We must choose an admissible
${\it\tau}$
for which we can show that
$\mathscr{Q}\geqslant 0$
, where
$\mathscr{Q}$
is treated as a functional of any
${\it\Theta}$
and incompressible
$\boldsymbol{u}$
that are sufficiently smooth and satisfy the dynamical boundary conditions. When
$\mathscr{Q}\geqslant 0$
, expression (4.11) gives the bound
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn63.gif?pub-status=live)
Having already relaxed the full dynamical constraints, we further limit our scope to
${\it\tau}$
profiles that are piecewise linear. These simplifications lead to a suboptimal bound but let us reach it analytically, which is necessary for the bound to apply at arbitrarily large
$R$
.
4.1.2. Piecewise linear background profile
Figure 4 shows the family of
${\it\tau}$
that we consider here. In principle, the boundary layers of thickness
${\it\epsilon}$
are needed to satisfy the fixed-flux thermal boundary conditions, so that
${\it\Theta}$
obeys the corresponding homogenous conditions. In practice, however, carrying
${\it\epsilon}$
through the analysis and then taking
${\it\epsilon}\rightarrow 0$
yields the same bound as setting
${\it\epsilon}\equiv 0$
at the start, so we simply do the latter. Our calculation thus excludes the
$O({\it\epsilon})$
terms that would make it fully rigorous but arrives at the same result. The fact that the thermal boundary conditions on
${\it\tau}$
effectively can be ignored relies on the fixed-flux conditions. When a boundary layer is used to meet a fixed-temperature condition, its effect on the resulting bound does not generally vanish as its thickness goes to zero because its slope approaches infinity.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary-alt:20170721122512-63774-mediumThumb-S0022112015001408_fig4g.jpg?pub-status=live)
Figure 4. Schematic of the background profile,
${\it\tau}(z)$
, that we consider. The parameters
${\it\delta}$
,
$a$
and
$b$
are optimized to maximize the lower bound (4.13) while maintaining the non-negativity of
$\mathscr{Q}$
. We can neglect the two layers of thickness
${\it\epsilon}$
in our analysis (see the text).
With
${\it\epsilon}\equiv 0$
, our ansatz for
${\it\tau}$
consists of only two linear pieces:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn64.gif?pub-status=live)
where figure 4 shows the geometric meanings of
${\it\delta}$
,
$a$
and
$b$
. The top temperature is fixed as
${\it\tau}_{T}=0$
for convenience since adding a constant to
${\it\tau}$
does not affect the bound (4.13). The upper piece of
${\it\tau}$
turns out to be a boundary layer because we must choose an expression for its thickness,
${\it\delta}$
, that vanishes as
$R\rightarrow \infty$
. The lower piece of
${\it\tau}$
is chosen to have a slope of
$a/2$
, whatever the value of
$a$
we fix later: a known trick for making the sign-indefinite term of
$\mathscr{Q}$
vanish outside the boundary layer (Constantin & Doering Reference Constantin and Doering1996; Lu et al.
Reference Lu, Doering and Busse2004). With the ansatz (4.14) chosen for
${\it\tau}$
, the lower bound (4.13) becomes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn65.gif?pub-status=live)
4.1.3. Optimal parameter choices
We seek the optimal parameters,
${\it\delta}^{\ast }$
,
$a^{\ast }$
and
$b^{\ast }$
, that maximize the lower bound (4.13) while still letting us show
$\mathscr{Q}\geqslant 0$
. There is one optimality condition that is unaffected by the requirement that
$\mathscr{Q}\geqslant 0$
; the lower bound (4.15) is maximized when its partial derivative with respect to
$a$
vanishes, and this requires that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn66.gif?pub-status=live)
Making this optimal choice, we eliminate
$b$
from the bound to find
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn67.gif?pub-status=live)
It remains to choose
${\it\delta}$
and
$a$
optimally, but first we must find conditions on these parameters that ensure
$\mathscr{Q}\geqslant 0$
.
Determining conditions sufficient for
$\mathscr{Q}\geqslant 0$
requires some functional analysis to bound the magnitude of the sign-indefinite term,
$\langle (2{\it\tau}^{\prime }-a)w{\it\Theta}\rangle$
. Following Otero et al. (Reference Otero, Wittenberg, Worthing and Doering2002), we can proceed in spectral space, noting that
$\mathscr{Q}$
is bounded below by an integral over horizontal wavevectors:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn68.gif?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn69.gif?pub-status=live)
and where
${\hat{w}}_{\boldsymbol{k}}(z)$
and
$\hat{{\it\Theta}}_{\boldsymbol{k}}(z)$
are the horizontal Fourier transforms of
$w$
and
${\it\Theta}$
,
$\widetilde{w}_{\boldsymbol{k}}$
is the complex conjugate of
${\hat{w}}_{\boldsymbol{k}}$
, and
$\mathit{Re}$
denotes the real part of a complex quantity. Incompressibility has been used to eliminate horizontal velocity components from
$\mathscr{Q}_{\boldsymbol{k}}$
. The sign-indefinite term of
$\mathscr{Q}_{\boldsymbol{k}}$
is nonzero only in the boundary layer, and its magnitude there is bounded by (Otero et al.
Reference Otero, Wittenberg, Worthing and Doering2002)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn70.gif?pub-status=live)
where we have made use of the fact that
$2{\it\tau}^{\prime }-a=-1$
in the boundary layer for the optimal choice
$b^{\ast }=({\it\delta}-a)/2$
. The above estimate, and no other part of our proof, relies on the assumption of no-slip boundaries. The two previous expressions give
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn71.gif?pub-status=live)
The non-negativity of all four coefficients in the above inequality suffices for the non-negativity of each
$\mathscr{Q}_{\boldsymbol{k}}$
and, in turn, of
$\mathscr{Q}$
. We choose
${\it\alpha}={\it\beta}={\it\delta}^{2}/4\sqrt{2}$
, which is as large as the latter two coefficients allow. The non-negativity of the first coefficient, which also implies that of the second, then requires
$a\geqslant (1/32)R{\it\delta}^{4}$
. Our lower bound will be maximized by letting
${\it\delta}$
be as large as possible while guaranteeing
$\mathscr{Q}\geqslant 0$
, so we choose
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn72.gif?pub-status=live)
after which the bound (4.17) becomes
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn73.gif?pub-status=live)
To avoid both positive terms in the lower bound being subdominant when
$R\rightarrow \infty$
, we must choose
${\it\delta}$
no larger than
$O(R^{-1/3})$
. For such
${\it\delta}$
, the last three terms are subdominant, so
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn74.gif?pub-status=live)
at large
$R$
. The optimal
${\it\delta}^{\ast }$
that maximizes this leading-order expression is proportional to
$R^{-1/3}$
. Finding this
${\it\delta}^{\ast }$
and using it to put expressions (4.22) and (4.16) for
$a^{\ast }$
and
$b^{\ast }$
in terms of
$R$
gives
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn75.gif?pub-status=live)
With these parameter choices,
${\it\tau}^{\prime }=-1/2+O(R^{-1/3})$
in the boundary layer. This is roughly half of what
$\partial T/\partial z$
would be in the thermal boundary layer of an actual flow. In fixed-flux RB convection, on the other hand, the boundary layers of a similarly optimized
${\it\tau}$
profile have the same
$z$
-derivative as the dynamical
$T$
field at the boundaries.
Applying the optimal
${\it\delta}^{\ast }$
to the exact expression (4.23), we at last obtain our lower bound on the mean temperature,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn76.gif?pub-status=live)
At large
$R$
, this bound scales like
$R^{-1/3}$
and reduces to expression (4.4). The lower bound on mean temperature proven by Lu et al. (Reference Lu, Doering and Busse2004) for a different IH configuration also scales like
$R^{-1/3}$
.
4.2. Other quantities important to heat transport
We turn now to other integral quantities that, like
$\langle T-\overline{T}_{T}\rangle$
, bear on the relative contributions of conduction and convection. The vertical heat flux,
$J$
, at a point is the sum of the fluxes due to conduction,
$J_{cond}$
, and convection,
$J_{conv}$
. In our non-dimensionalization,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn77.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn78.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline339.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline340.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_inline341.gif?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn79.gif?pub-status=live)
This balance expresses the fact that, because the bottom boundary is insulating, the mean upward flux at height
$z$
is equal to the rate, also
$z$
, at which heat is produced below that height. Integrating expression (4.29) over the vertical extent gives another useful balance,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn80.gif?pub-status=live)
where
${\it\delta}\overline{T}:=\overline{T}_{B}-\overline{T}_{T}$
is the difference between the mean temperatures at the bottom and top boundaries. That is, the mean heat flux over the layer is
$1/2$
and is the sum of the conductive and convective parts,
${\it\delta}\overline{T}$
and
$\langle wT\rangle$
.
4.2.1. Mean temperature difference
The relative contributions of conduction and convection can be characterized in similar but not identical ways by two quantities: the relative mean temperature of the fluid,
$\langle T-\overline{T}_{T}\rangle$
, and the mean temperature difference between the boundaries,
${\it\delta}\overline{T}$
, which is also the mean conductive transport across the layer. (We could equally well speak of
$\langle wT\rangle$
instead of
${\it\delta}\overline{T}$
since we know they sum to
$1/2$
.) For
${\it\delta}\overline{T}$
we have the uniform upper bound
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn81.gif?pub-status=live)
A lower bound of zero seems likely, but we have not proven it. The upper bound follows from expressions (4.9) and (4.30). Much like
$\langle T-\overline{T}_{T}\rangle$
, the quantity
${\it\delta}\overline{T}$
saturates its upper bound of
$1/2$
only in the static state, and we expect but do not know how to prove that
${\it\delta}\overline{T}\rightarrow 0$
as
$R\rightarrow \infty$
. In this limit, the net heat transport would be accomplished solely by convection, rather than conduction.
At large
$R$
, we expect
${\it\delta}\overline{T}$
and
$\langle T-\overline{T}_{T}\rangle$
to be even more similar. Strong convection typically renders
$\overline{T}(z)$
profiles roughly isothermal outside of boundary layers, as in the schematic of figure 2(b), and here this would mean
${\it\delta}\overline{T}\sim \left\langle T-\overline{T}_{T}\right\rangle$
. Even with boundary conditions for which large-scale shear might keep the interior far from isothermal (Goluskin et al.
Reference Goluskin, Johnston, Flierl and Spiegel2014; van der Poel et al.
Reference van der Poel, Ostilla-Mónico, Verzicco and Lohse2014), we still expect
${\it\delta}\overline{T}$
and
$\langle T-\overline{T}_{T}\rangle$
to scale similarly at large
$R$
. Thus, since we have proven in § 4.1 that
$\langle T-\overline{T}_{T}\rangle$
can decay no faster than
$R^{-1/3}$
, it seems likely that the same is true of
${\it\delta}\overline{T}$
.
Proving a parameter-dependent lower bound on
${\it\delta}\overline{T}$
remains an open challenge. The challenge is novel because
${\it\delta}\overline{T}$
is related to
$\langle |\boldsymbol{{\rm\nabla}}\boldsymbol{u}|^{2}\rangle$
, via expressions (4.9) and (4.30), but not to
$\langle |\boldsymbol{{\rm\nabla}}T|^{2}\rangle$
. The background method would thus require decomposing
$\boldsymbol{u}$
, whereas past applications of the method to convection have decomposed
$T$
and bounded quantities related to
$\langle |\boldsymbol{{\rm\nabla}}T|^{2}\rangle$
.
4.2.2. Nusselt numbers and diagnostic Rayleigh numbers
In convective systems, the relative contributions of conduction and convection to net heat transport are often expressed using dimensionless Nusselt numbers,
$N$
. One particular definition of
$N$
, together with a diagnostic Rayleigh number,
$\mathit{Ra}$
, works well to reveal the parallels between RB configurations with different thermal boundary conditions (Otero et al.
Reference Otero, Wittenberg, Worthing and Doering2002; Verzicco & Sreenivasan Reference Verzicco and Sreenivasan2008; Johnston & Doering Reference Johnston and Doering2009; Wittenberg Reference Wittenberg2010). We define
$N$
and
$\mathit{Ra}$
in a way that agrees with these RB studies and extends to IH convection:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn82.gif?pub-status=live)
where
$\langle J\rangle$
and
$\langle J_{cond}\rangle$
refer to the developed flow, while
$\langle J_{cond}\rangle _{st}$
refers to the static state. In our present model,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn83.gif?pub-status=live)
The definition (2.6a
) of the control parameter
$R$
uses the dimensional temperature scale
${\it\Delta}$
, given in expression (2.2), that is proportional to the temperature difference between the boundaries in the static state. The diagnostic parameter
$\mathit{Ra}$
essentially replaces this static temperature difference with that in the developed flow. The two parameters agree in the static state, but
$\mathit{Ra}<R$
in sustained convection.
Restated in terms of the
$N$
we have defined, the basic features of
${\it\delta}\overline{T}$
described in § 4.2.1 are:
$N=1$
in the static state,
$N>1$
in sustained flows and we expect
$N$
to grow without bound as
$R\rightarrow \infty$
. These same statements apply to the usual Nusselt number of RB convection. We have further argued (without proof) in § 4.2.1 that
${\it\delta}\overline{T}$
can decay no faster than
$R^{-1/3}$
. This would be equivalent to
$N$
growing no faster than
$\mathit{Ra}^{1/2}$
.
Since we have proven an
$R$
-dependent bound on
$\langle T-\overline{T}_{T}\rangle$
, it is natural to ask whether the Nusselt number and diagnostic Rayleigh number could instead be generalized to IH convection in a way that invokes
$\langle T-\overline{T}_{T}\rangle$
, rather than
${\it\delta}\overline{T}$
. This leads us to define quantities that are like
$N$
and
$\mathit{Ra}$
but with averages weighted proportionally to height,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn84.gif?pub-status=live)
Because
$\overline{J}(z)=z$
here, weighting by height is equivalent to weighting by the mean vertical heat flux at each height. For our present boundary conditions,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn85.gif?pub-status=live)
Expressed in these terms, our lower bound (4.4) on
$\langle T-\overline{T}_{T}\rangle$
becomes an upper bound on
$\widetilde{N}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn86.gif?pub-status=live)
The above bound has the same scaling as bounds that have been proven in a variety of RB configurations, so long as the RB bounds are expressed using definition (4.32a,b
) for
$N$
and
$\mathit{Ra}$
(Constantin & Doering Reference Constantin and Doering1996; Otero et al.
Reference Otero, Wittenberg, Worthing and Doering2002; Plasting & Kerswell Reference Plasting and Kerswell2003; Wittenberg Reference Wittenberg2010).
5. Conclusions
This work has addressed IH convection beneath a poor conductor, a basic but largely overlooked configuration. We have found differing linear and nonlinear stability thresholds for the static state. Simple exact expressions exist for both thresholds, which is rare in studies of fluid stability, and they have been found here using long-wavelength asymptotics. Beyond the static state, we have bounded the mean temperature of the convecting fluid, relative to that of the top boundary, assuming no-slip velocity conditions on both boundaries. As the heating rate (
$H$
) is raised, this dimensional mean temperature can grow no slower than
$H^{2/3}$
. In terms of a dimensionless Nusselt number (
$\widetilde{N}$
) and diagnostic Rayleigh number (
$\widetilde{\mathit{Ra}}$
) that we have defined using the mean temperature, our bound takes the same form as upper bounds on Nusselt numbers in various other convective models.
Many fundamental features of the model studied here are yet to be explored. At Rayleigh numbers between our thresholds for global stability and linear instability, it is not known whether subcritical convection can be sustained, although the answer is affirmative when the upper boundary conducts perfectly instead of poorly (Tveitereid & Palm Reference Tveitereid and Palm1976). While we have proved a parameter-dependent lower bound on the mean temperature of sustained flow, only uniform bounds are known for the mean temperature difference between the top and bottom boundaries. The latter quantity is useful in characterizing heat transport in the fluid, and it is much easier to measure experimentally than the mean fluid temperature. A bound on this temperature difference would amount to a bound on viscous dissipation, rather than on thermal dissipation, and a method for constructing it would likely yield novel results in other IH configurations as well (Goluskin Reference Goluskin2015). Beyond analytical results, it seems the only studies of our configuration have been two-dimensional simulations at small-to-moderate Rayleigh numbers (Hewitt et al. Reference Hewitt, McKenzie and Weiss1980; Ishiwatari et al. Reference Ishiwatari, Takehiro and Hayashi1994). The parameter regimes in which convection is strong and complicated are wide open for numerical simulations and laboratory experiments.
Acknowledgements
The author thanks C. R. Doering and the anonymous referees for some very helpful comments on the manuscript. The author was supported during part of this research by the US National Science Foundation (NSF) Mathematical Physics award PHY-1205219.
Appendix A. Polynomial eigenfunctions
The asymptotic solution of the linear stability eigenproblem in § 3.3 makes use of the unique fourth-order polynomials that satisfy
$P^{(4)}(z)=1$
and the velocity boundary conditions (3.7)–(3.10). These polynomials are
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn87.gif?pub-status=live)
The same
$P(z)$
are given by Chapman & Proctor (Reference Chapman and Proctor1980), scaled for their domain of
$-1\leqslant z\leqslant 1$
rather than our domain of
$0\leqslant z\leqslant 1$
. The asymptotic solution of the energy stability eigenproblem requires the unique fifth-order polynomials that satisfy
$Q_{{\it\gamma}}^{(4)}(z)=1/2(1+{\it\gamma}z)$
and the velocity boundary conditions. These polynomials are
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20170721070605003-0108:S0022112015001408:S0022112015001408_eqn88.gif?pub-status=live)
Figure 5 shows that when the energy (3.18) is defined using the optimal coupling parameter
${\it\gamma}^{\ast }$
(cf. table 2), each fifth-order
$Q_{{\it\gamma}}(z)$
closely approximates the corresponding fourth-order
$P(z)$
. In the
$k^{2}\rightarrow 0$
limit, the linear stability eigenmode has vertical velocity proportional to
$P(z)$
, the energy stability eigenmode has vertical velocity proportional to
$Q_{{\it\gamma}}(z)$
, and both eigenmodes have constant temperature. Thus, the closeness of the optimal
$Q_{{\it\gamma}}(z)$
to
$P(z)$
means that the long-wavelength solutions to the linear and energy stability eigenproblems are similar not only in their generalized eigenvalues,
$R_{L}$
and
$R_{E}$
, but also in their eigenfunctions.
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary-alt:20170721122512-48716-mediumThumb-S0022112015001408_fig5g.jpg?pub-status=live)
Figure 5. The polynomials
$P(z)$
(——) and
$Q_{{\it\gamma}}(z)$
(- - - - -), where the latter are evaluated for the optimal coupling parameter
${\it\gamma}^{\ast }$
. Top and bottom boundary conditions on the velocity are (a) both no-slip, (b) free-slip only at the top, (c) free-slip only at the bottom and (d) both free-slip.