Namespaces
Variants
Actions

Difference between revisions of "Oberbeck-Boussinesq equations"

From Encyclopedia of Mathematics
Jump to: navigation, search
m (AUTOMATIC EDIT (latexlist): Replaced 39 formulas out of 39 by TEX code with an average confidence of 2.0 and a minimal confidence of 2.0.)
 
Line 1: Line 1:
Equations giving an approximate description of the thermo-mechanical response of linearly viscous fluids (Navier–Stokes fluids or Newtonian fluids) that can only sustain volume-preserving motions (isochoric motions) in isothermal processes, but can undergo motions that are not volume-preserving during non-isothermal processes. Such a restriction on the response of the fluid requires that the determinant of the deformation gradient <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200101.png" /> be a function of the temperature <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200102.png" />, i.e.,
+
<!--This article has been texified automatically. Since there was no Nroff source code for this article,  
 +
the semi-automatic procedure described at https://encyclopediaofmath.org/wiki/User:Maximilian_Janisch/latexlist
 +
was used.
 +
If the TeX and formula formatting is correct, please remove this message and the {{TEX|semi-auto}} category.
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200103.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a1)</td></tr></table>
+
Out of 39 formulas, 39 were replaced by TEX code.-->
 +
 
 +
{{TEX|semi-auto}}{{TEX|done}}
 +
Equations giving an approximate description of the thermo-mechanical response of linearly viscous fluids (Navier–Stokes fluids or Newtonian fluids) that can only sustain volume-preserving motions (isochoric motions) in isothermal processes, but can undergo motions that are not volume-preserving during non-isothermal processes. Such a restriction on the response of the fluid requires that the determinant of the deformation gradient $\mathcal{F}$ be a function of the temperature $\theta$, i.e.,
 +
 
 +
\begin{equation} \tag{a1} \operatorname { det } \mathcal{F} = f ( \theta ). \end{equation}
  
 
An approximation of the balance of mass, momentum and energy within the context of the above constraint was first discussed by A. Oberbeck [[#References|[a1]]] and later by J. Boussinesq [[#References|[a2]]]. Such an approximate system has relevance to a plethora of problems in astrophysics, geophysics and oceanography.
 
An approximation of the balance of mass, momentum and energy within the context of the above constraint was first discussed by A. Oberbeck [[#References|[a1]]] and later by J. Boussinesq [[#References|[a2]]]. Such an approximate system has relevance to a plethora of problems in astrophysics, geophysics and oceanography.
Line 11: Line 19:
 
The constraint (a1) implies that
 
The constraint (a1) implies that
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200104.png" /></td> </tr></table>
+
\begin{equation*} \operatorname { div } \mathbf{v} = \frac { f ^ { \prime } ( \theta ) } { f ( \theta ) } \left( \frac { \partial \theta } { \partial t } + \nabla \theta \cdot \mathbf{v} \right) = \alpha ( \theta ) \left( \frac { \partial \theta } { \partial t } + \nabla \theta \cdot \mathbf{v} \right), \end{equation*}
  
where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200105.png" /> is the velocity field, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200106.png" /> the temperature, and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200107.png" /> the coefficient of thermal expansion.
+
where $\mathbf{v}$ is the velocity field, $\theta$ the temperature, and $\alpha$ the coefficient of thermal expansion.
  
 
Thus, while in an incompressible fluid
 
Thus, while in an incompressible fluid
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200108.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a2)</td></tr></table>
+
\begin{equation} \tag{a2} \operatorname { div } \mathbf{v} = 0, \end{equation}
  
 
this is not necessarily so for fluids obeying (a1). However, in the Oberbeck–Boussinesq equations the constraint (a2) holds to within the order of approximation.
 
this is not necessarily so for fluids obeying (a1). However, in the Oberbeck–Boussinesq equations the constraint (a2) holds to within the order of approximation.
  
Let <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o1200109.png" /> denote the acceleration due to gravity, let <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001010.png" /> be a typical length scale (usually the thickness of the layer of the fluid), <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001011.png" /> a representative density, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001012.png" /> the viscosity, and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001013.png" /> a characteristic temperature difference. Assume that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001014.png" /> is a constant. On introducing a characteristic velocity <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001015.png" /> and characteristic time <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001016.png" /> for the problem through (see [[#References|[a4]]])
+
Let $g$ denote the acceleration due to gravity, let $L$ be a typical length scale (usually the thickness of the layer of the fluid), $\rho$ a representative density, $\mu$ the viscosity, and $\delta \theta _ { 0 }$ a characteristic temperature difference. Assume that $\alpha$ is a constant. On introducing a characteristic velocity $U$ and characteristic time $t$ for the problem through (see [[#References|[a4]]])
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001017.png" /></td> </tr></table>
+
\begin{equation*} U = \sqrt { g L \alpha \delta \theta _ { 0 } } , \quad t = \frac { U } { L }, \end{equation*}
  
 
one can set the relevant non-dimensional parameters to be
 
one can set the relevant non-dimensional parameters to be
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001018.png" /></td> </tr></table>
+
\begin{equation*} Re = \frac { \rho L U } { \mu } , \quad \varepsilon = U ( \frac { \rho } { g \mu } ) ^ { 1 / 3 }, \end{equation*}
  
where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001019.png" /> is called the [[Reynolds number|Reynolds number]].
+
where $\operatorname{Re}$ is called the [[Reynolds number|Reynolds number]].
  
Assuming that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001020.png" /> and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001021.png" /> is a constant, and expressing the (non-dimensional) velocity <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001022.png" />, the temperature <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001023.png" /> and the pressure <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001024.png" /> through
+
Assuming that $\varepsilon \ll 1$ and $\alpha$ is a constant, and expressing the (non-dimensional) velocity $\mathbf{v}$, the temperature $\theta$ and the pressure $p$ through
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001025.png" /></td> </tr></table>
+
\begin{equation*} \left( \begin{array} { l } { \mathbf{v} } \\ { \theta } \\ { p } \end{array} \right) = \sum _ { n = 0 } ^ { \infty } \varepsilon ^ { n } \left( \begin{array} { c } { \mathbf{v} _ { n } } \\ { \theta _ { n } } \\ { p _ { n } } \end{array} \right), \end{equation*}
  
and substituting the above into the governing equations for mass, momentum and energy balance for a Navier–Stokes fluid leads to an hierarchy of equations at different orders of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001026.png" />. Let <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001027.png" /> denote the specific heat at some reference temperature <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001028.png" /> and set <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001029.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001030.png" /> being the thermal conductivity. Let <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001031.png" /> denote the field of the body forces. Then the equations at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001032.png" /> read
+
and substituting the above into the governing equations for mass, momentum and energy balance for a Navier–Stokes fluid leads to an hierarchy of equations at different orders of $\varepsilon$. Let $C ( \theta _ { r } )$ denote the specific heat at some reference temperature $\theta _ { r }$ and set $\Lambda _ { 1 } = U C ( \theta _ { r } ) L / \kappa$, $\kappa$ being the thermal conductivity. Let $\mathbf{b}$ denote the field of the body forces. Then the equations at $O ( 1 )$ read
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001033.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a3)</td></tr></table>
+
\begin{equation} \tag{a3} \left\{ \begin{array} { l } { \operatorname { Re } ( \nabla p _ { 0 } + \mathbf{b} ) = 0, } \\ { \Lambda _ { 1 } C ( \theta _ { r } ) \left( \frac { \partial \theta _ { 0 } } { \partial t } + \nabla \theta _ { 0 } \cdot  \mathbf{v} _ { 0 } \right) = \Delta \theta _ { 0 }, } \\ { \operatorname { div } \mathbf{v} _ { 0 } = 0. } \end{array} \right. \end{equation}
  
Notice that the above set of equations is not adequate to determine the variables. When the equations for the balance of linear momentum at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001034.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001035.png" /> and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001036.png" /> are appended to (a3), i.e.,
+
Notice that the above set of equations is not adequate to determine the variables. When the equations for the balance of linear momentum at $O ( \varepsilon )$, $O ( \varepsilon ^ { 2 } ).$ and $O ( \varepsilon ^ { 3 } )$ are appended to (a3), i.e.,
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001037.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a4)</td></tr></table>
+
\begin{equation} \tag{a4} \left\{ \begin{array} { l } { \nabla p _ { 1 } = \nabla p _ { 2 } = 0, } \\ { \frac { \partial \mathbf{v} _ { 0 } } { \partial t } + [ \nabla \mathbf{v} _ { 0 } ] \mathbf{v} _ { 0 } = \frac { 1 } { Re } \Delta \mathbf{v} _ { 0 } + \operatorname { Re } \nabla p _ { 3 } + \theta _ { 0 } \mathbf{b}. } \end{array} \right. \end{equation}
  
one obtains a determinate system of equations (a3)–(a4), referred to as the Oberbeck–Boussinesq equations. Thus, the Oberbeck–Boussinesq equations do not follow from retaining the perturbances of the same order in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001038.png" />. This perturbation procedure, discussed in [[#References|[a3]]] in detail, also provides the corrections to the Oberbeck–Boussinesq equations at higher order of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/o/o120/o120010/o12001039.png" />.
+
one obtains a determinate system of equations (a3)–(a4), referred to as the Oberbeck–Boussinesq equations. Thus, the Oberbeck–Boussinesq equations do not follow from retaining the perturbances of the same order in $\varepsilon$. This perturbation procedure, discussed in [[#References|[a3]]] in detail, also provides the corrections to the Oberbeck–Boussinesq equations at higher order of $\varepsilon$.
  
 
A similar heuristic approach has been developed for the thermo-mechanical response of non-Newtonian fluids [[#References|[a5]]].
 
A similar heuristic approach has been developed for the thermo-mechanical response of non-Newtonian fluids [[#References|[a5]]].
Line 50: Line 58:
  
 
====References====
 
====References====
<table><TR><TD valign="top">[a1]</TD> <TD valign="top">  A. Oberbeck,  "Ueber die Wärmeleitung der Flüssigkeiten bei Berücksichtigung der Strömungen infolge von Temperaturdifferenzen"  ''Ann. Phys. Chem.'' , '''VII'''  (1879)  pp. 271–292</TD></TR><TR><TD valign="top">[a2]</TD> <TD valign="top">  J. Boussinesq,  "Théorie analytique de la chaleur" , Gauthier-Villars  (1903)</TD></TR><TR><TD valign="top">[a3]</TD> <TD valign="top">  K.R. Rajagopal,  M. Ruzicka,  A.R. Srinivasa,  "On the Oberbeck–Boussinesq approximation"  ''Math. Meth. Appl. Sci.'' , '''6'''  (1996)  pp. 1157–1167</TD></TR><TR><TD valign="top">[a4]</TD> <TD valign="top">  S. Chandrasekar,  "Hydrodynamic and hydromagnetic stability" , Oxford Univ. Press  (1961)</TD></TR><TR><TD valign="top">[a5]</TD> <TD valign="top">  A.V. Shenoy,  R.A. Mashelkar,  "Thermal convenction in non-Newtonian fluids"  J.P. Hartnett (ed.)  T.F. Irvine (ed.) , ''Advances in Heat Transfer'' , '''15''' , Acad. Press  (1982)  pp. 143–225</TD></TR><TR><TD valign="top">[a6]</TD> <TD valign="top">  C. Foias,  O. Manley,  R. Temam,  "Attractors for the Bénard problem: existence and physical bounds on their fractal dimension"  ''Nonlinear Anal. Theory Methods Appl.'' , '''11'''  (1987)  pp. 939–967</TD></TR><TR><TD valign="top">[a7]</TD> <TD valign="top">  R. Temam,  "Infinite-dimensional dynamical systems in mechanics and physics" , Springer  (1997)  (Edition: Second)</TD></TR></table>
+
<table><tr><td valign="top">[a1]</td> <td valign="top">  A. Oberbeck,  "Ueber die Wärmeleitung der Flüssigkeiten bei Berücksichtigung der Strömungen infolge von Temperaturdifferenzen"  ''Ann. Phys. Chem.'' , '''VII'''  (1879)  pp. 271–292</td></tr><tr><td valign="top">[a2]</td> <td valign="top">  J. Boussinesq,  "Théorie analytique de la chaleur" , Gauthier-Villars  (1903)</td></tr><tr><td valign="top">[a3]</td> <td valign="top">  K.R. Rajagopal,  M. Ruzicka,  A.R. Srinivasa,  "On the Oberbeck–Boussinesq approximation"  ''Math. Meth. Appl. Sci.'' , '''6'''  (1996)  pp. 1157–1167</td></tr><tr><td valign="top">[a4]</td> <td valign="top">  S. Chandrasekar,  "Hydrodynamic and hydromagnetic stability" , Oxford Univ. Press  (1961)</td></tr><tr><td valign="top">[a5]</td> <td valign="top">  A.V. Shenoy,  R.A. Mashelkar,  "Thermal convenction in non-Newtonian fluids"  J.P. Hartnett (ed.)  T.F. Irvine (ed.) , ''Advances in Heat Transfer'' , '''15''' , Acad. Press  (1982)  pp. 143–225</td></tr><tr><td valign="top">[a6]</td> <td valign="top">  C. Foias,  O. Manley,  R. Temam,  "Attractors for the Bénard problem: existence and physical bounds on their fractal dimension"  ''Nonlinear Anal. Theory Methods Appl.'' , '''11'''  (1987)  pp. 939–967</td></tr><tr><td valign="top">[a7]</td> <td valign="top">  R. Temam,  "Infinite-dimensional dynamical systems in mechanics and physics" , Springer  (1997)  (Edition: Second)</td></tr></table>

Latest revision as of 17:02, 1 July 2020

Equations giving an approximate description of the thermo-mechanical response of linearly viscous fluids (Navier–Stokes fluids or Newtonian fluids) that can only sustain volume-preserving motions (isochoric motions) in isothermal processes, but can undergo motions that are not volume-preserving during non-isothermal processes. Such a restriction on the response of the fluid requires that the determinant of the deformation gradient $\mathcal{F}$ be a function of the temperature $\theta$, i.e.,

\begin{equation} \tag{a1} \operatorname { det } \mathcal{F} = f ( \theta ). \end{equation}

An approximation of the balance of mass, momentum and energy within the context of the above constraint was first discussed by A. Oberbeck [a1] and later by J. Boussinesq [a2]. Such an approximate system has relevance to a plethora of problems in astrophysics, geophysics and oceanography.

Numerous attempts have been made to provide a rigorous justification for the Oberbeck–Boussinesq equations, the details of which can be found in [a3].

While plausible arguments based on physical grounds are advanced to derive the Oberbeck–Boussinesq equations (see [a4]), namely that the effects of the variations in the density with respect to the temperature are more significant in the buoyancy forces than in the inertial effects, no compelling evidence or a rigorous mathematical basis is available.

The constraint (a1) implies that

\begin{equation*} \operatorname { div } \mathbf{v} = \frac { f ^ { \prime } ( \theta ) } { f ( \theta ) } \left( \frac { \partial \theta } { \partial t } + \nabla \theta \cdot \mathbf{v} \right) = \alpha ( \theta ) \left( \frac { \partial \theta } { \partial t } + \nabla \theta \cdot \mathbf{v} \right), \end{equation*}

where $\mathbf{v}$ is the velocity field, $\theta$ the temperature, and $\alpha$ the coefficient of thermal expansion.

Thus, while in an incompressible fluid

\begin{equation} \tag{a2} \operatorname { div } \mathbf{v} = 0, \end{equation}

this is not necessarily so for fluids obeying (a1). However, in the Oberbeck–Boussinesq equations the constraint (a2) holds to within the order of approximation.

Let $g$ denote the acceleration due to gravity, let $L$ be a typical length scale (usually the thickness of the layer of the fluid), $\rho$ a representative density, $\mu$ the viscosity, and $\delta \theta _ { 0 }$ a characteristic temperature difference. Assume that $\alpha$ is a constant. On introducing a characteristic velocity $U$ and characteristic time $t$ for the problem through (see [a4])

\begin{equation*} U = \sqrt { g L \alpha \delta \theta _ { 0 } } , \quad t = \frac { U } { L }, \end{equation*}

one can set the relevant non-dimensional parameters to be

\begin{equation*} Re = \frac { \rho L U } { \mu } , \quad \varepsilon = U ( \frac { \rho } { g \mu } ) ^ { 1 / 3 }, \end{equation*}

where $\operatorname{Re}$ is called the Reynolds number.

Assuming that $\varepsilon \ll 1$ and $\alpha$ is a constant, and expressing the (non-dimensional) velocity $\mathbf{v}$, the temperature $\theta$ and the pressure $p$ through

\begin{equation*} \left( \begin{array} { l } { \mathbf{v} } \\ { \theta } \\ { p } \end{array} \right) = \sum _ { n = 0 } ^ { \infty } \varepsilon ^ { n } \left( \begin{array} { c } { \mathbf{v} _ { n } } \\ { \theta _ { n } } \\ { p _ { n } } \end{array} \right), \end{equation*}

and substituting the above into the governing equations for mass, momentum and energy balance for a Navier–Stokes fluid leads to an hierarchy of equations at different orders of $\varepsilon$. Let $C ( \theta _ { r } )$ denote the specific heat at some reference temperature $\theta _ { r }$ and set $\Lambda _ { 1 } = U C ( \theta _ { r } ) L / \kappa$, $\kappa$ being the thermal conductivity. Let $\mathbf{b}$ denote the field of the body forces. Then the equations at $O ( 1 )$ read

\begin{equation} \tag{a3} \left\{ \begin{array} { l } { \operatorname { Re } ( \nabla p _ { 0 } + \mathbf{b} ) = 0, } \\ { \Lambda _ { 1 } C ( \theta _ { r } ) \left( \frac { \partial \theta _ { 0 } } { \partial t } + \nabla \theta _ { 0 } \cdot \mathbf{v} _ { 0 } \right) = \Delta \theta _ { 0 }, } \\ { \operatorname { div } \mathbf{v} _ { 0 } = 0. } \end{array} \right. \end{equation}

Notice that the above set of equations is not adequate to determine the variables. When the equations for the balance of linear momentum at $O ( \varepsilon )$, $O ( \varepsilon ^ { 2 } ).$ and $O ( \varepsilon ^ { 3 } )$ are appended to (a3), i.e.,

\begin{equation} \tag{a4} \left\{ \begin{array} { l } { \nabla p _ { 1 } = \nabla p _ { 2 } = 0, } \\ { \frac { \partial \mathbf{v} _ { 0 } } { \partial t } + [ \nabla \mathbf{v} _ { 0 } ] \mathbf{v} _ { 0 } = \frac { 1 } { Re } \Delta \mathbf{v} _ { 0 } + \operatorname { Re } \nabla p _ { 3 } + \theta _ { 0 } \mathbf{b}. } \end{array} \right. \end{equation}

one obtains a determinate system of equations (a3)–(a4), referred to as the Oberbeck–Boussinesq equations. Thus, the Oberbeck–Boussinesq equations do not follow from retaining the perturbances of the same order in $\varepsilon$. This perturbation procedure, discussed in [a3] in detail, also provides the corrections to the Oberbeck–Boussinesq equations at higher order of $\varepsilon$.

A similar heuristic approach has been developed for the thermo-mechanical response of non-Newtonian fluids [a5].

Also, in the case of the thermo-mechanical response of solids, the constraint (a1) seems to be applicable. However, the counterparts to the Oberbeck–Boussinesq equations have not yet been established as there is no single model that enjoys the kind of widespread use as the Navier–Stokes equations. From the point of view of mathematical analysis, the Oberbeck–Boussinesq equations are an example of a coupled non-linear system that retain the salient features of the Navier–Stokes equations when temperature effects are included. See [a6] or [a7] for details.

References

[a1] A. Oberbeck, "Ueber die Wärmeleitung der Flüssigkeiten bei Berücksichtigung der Strömungen infolge von Temperaturdifferenzen" Ann. Phys. Chem. , VII (1879) pp. 271–292
[a2] J. Boussinesq, "Théorie analytique de la chaleur" , Gauthier-Villars (1903)
[a3] K.R. Rajagopal, M. Ruzicka, A.R. Srinivasa, "On the Oberbeck–Boussinesq approximation" Math. Meth. Appl. Sci. , 6 (1996) pp. 1157–1167
[a4] S. Chandrasekar, "Hydrodynamic and hydromagnetic stability" , Oxford Univ. Press (1961)
[a5] A.V. Shenoy, R.A. Mashelkar, "Thermal convenction in non-Newtonian fluids" J.P. Hartnett (ed.) T.F. Irvine (ed.) , Advances in Heat Transfer , 15 , Acad. Press (1982) pp. 143–225
[a6] C. Foias, O. Manley, R. Temam, "Attractors for the Bénard problem: existence and physical bounds on their fractal dimension" Nonlinear Anal. Theory Methods Appl. , 11 (1987) pp. 939–967
[a7] R. Temam, "Infinite-dimensional dynamical systems in mechanics and physics" , Springer (1997) (Edition: Second)
How to Cite This Entry:
Oberbeck-Boussinesq equations. Encyclopedia of Mathematics. URL: http://encyclopediaofmath.org/index.php?title=Oberbeck-Boussinesq_equations&oldid=50429
This article was adapted from an original article by Josef MálekK.R. Rajagopal (originator), which appeared in Encyclopedia of Mathematics - ISBN 1402006098. See original article