\documentclass[reqno]{amsart} \usepackage{hyperref} \usepackage{graphicx} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2012 (2012), No. 181, pp. 1--13.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2012 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2012/181\hfil Optimal design of a bar] {Optimal design of a bar with an attached mass for maximizing the heat transfer} \author[B. P. Belinskiy, J. W. Hiestand, M. L. McCarthy \hfil EJDE-2012/181\hfilneg] {Boris P. Belinskiy, James W. Hiestand, Maeve L. McCarthy } % in alphabetical order \address{Boris P. Belinskiy \newline Department of Mathematics\\ University of Tennessee at Chattanooga\\ 615 McCallie Avenue\\ Chattanooga, TN 37403-2598, USA} \email{Boris-Belinskiy@utc.edu} \address{James W. Hiestand \newline College of Engineering\\ University of Tennessee at Chattanooga\\ 615 McCallie Avenue\\ Chattanooga, TN 37403-2598, USA} \email{James-Hiestand@utc.edu} \address{Maeve L. McCarthy \newline Department of Mathematics \& Statistics\\ Murray State University\\ 6C Faculty Hall\\ Murray, KY 42071-334, USA} \email{mmccarthy@murraystate.edu} \thanks{Submitted August 15, 2012. Published October 19, 2012.} \subjclass[2000]{74A15, 74P10, 49K15, 34B24} \keywords{Optimal design; heat transfer; heat equation; least eigenvalue; \hfill\break\indent Sturm-Liouville problem; Helly's principle; calculus of variations} \begin{abstract} We maximize, with respect to the cross sectional area, the rate of heat transfer through a bar of given mass. The bar serves as an extended surface to enhance the heat transfer surface of a larger heated known mass to which the bar is attached. In this paper we neglect heat transfer from the sides of the bar and consider only conduction through its length. The rate of cooling is defined by the first eigenvalue of the corresponding Sturm-Liouville problem. We establish existence of an optimal design via rearrangement techniques. The necessary conditions of optimality admit a unique optimal design. We compare the rate of heat transfer for that bar with the rate for the bar of the same mass but of a constant cross-section area. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{definition}[theorem]{Definition} \newtheorem{example}[theorem]{Example} \newtheorem{remark}[theorem]{Remark} \allowdisplaybreaks \def \pd#1#2{\frac{\partial#1}{\partial#2}} \section{Introduction}\label{intro_sect} Materials are often cooled by convection to a surrounding ambient medium such as the atmosphere. For example, the heat generated in an automobile engine is transferred first to the cooling water that circulates through the engine and then to the atmosphere through the radiator. Convective heat transfer is described by the equation \begin{equation} \dot Q=h A_s (T-T_\infty) \label{1.1} \end{equation} where $\dot Q$ is the heat transfer rate, $h$ is an empirical heat transfer coefficient, $A_s$ is the surface area, $T$ is the temperature of the surface and $T_\infty$ is the temperature of the surrounding medium. We want to maximize the surface area since the convective heat transfer rate is proportional to this area. On the other hand, we want to minimize the volume of the heat transfer region, in order to keep its weight and hence material cost as low as possible. Thus we seek to maximize the surface to volume ratio of the heat transfer surface. Extended surfaces attached to a given base mass, $M_0$, are frequently used in commercial applications to increase the heat transfer surface area without significantly increasing the associated mass and hence the material cost of the device. The additional surface might be in the form of thin donuts around a central pipe, parallel plates attached to the surface as in small engines or automobile radiators, or fins extending outward like hairs from a surface. The mass of the added extended surface is small compared to the base mass $M_0$. For this reason a high surface-to-volume ratio for the extended surface is sought. Such mass additions to enhance heat transfer are referred to as fins. Nature also ``designs" according to this criterion. The ears of an elephant have large surface area compared to their volume which allows the blood passing through them to be efficiently cooled. Likewise members of species like deer that live near the equator must be able to dissipate heat efficiently. They achieve this desirable ratio by being smaller than their counterparts that live towards the poles (think of a sphere where the surface to volume is inversely proportional to $r$). Engineering heat transfer texts sometimes consider fins of variable cross-section \cite[pp. 124-126]{Incropera} but the cross-section is assumed to be for a regular shape, e.g. a cylinder with variable radius. Or the optimization of a fin with a given cross-section (e.g. a rectangle) is optimized with respect to the length and thickness \cite[pp. 74]{Schneider}. However, a general variation of shape to maximize the heat transfer from the fin is not considered. Heat transfer within the surface is by conduction and the rate is given by the equation \begin{equation} \dot Q=-{{kA \Delta T}\over l} \end{equation} where $k$ is the thermal conductivity of the material, $A$ is the cross-sectional area, and $l$ is the length of the material. Here $\Delta T$ is the temperature difference between the end points of the heat transfer. The equations above, along with the corresponding physical background, may be found in \cite[pp. 110-114]{Incropera}. If an energy balance is performed for the region of the bar between $x$ and $x+\Delta x$, energy enters by conduction at $x$ and leaves by conduction at $x+\Delta x$ and also from the side by convection (see Equation \eqref{1.1}). The difference is the rate of change of the energy content of that region of the bar $$ \text{rate of change of energy}=\text{energy in}-\text{energy out}\quad \text{or} $$ \begin{equation} \rho c A \Delta x {{\Delta T}\over {\Delta t}}=-kA{{\Delta T}\over {\Delta x}} \Big \vert_x+kA{{\Delta T}\over {\Delta x}}\Big \vert_{x+\Delta x}-hA_s(T-T_{\infty}). \label{1.3} \end{equation} Here $T(x,t)$ is the temperature distribution. The surface area is $A_s=P\Delta x$, where $P(x)$ is the perimeter of the cross-section at the point $x$. The ratio ${{\Delta T}\over {\Delta t}}$ is the rate of change of temperature with time and ${{\Delta T}\over {\Delta x}}$ is the local temperature gradient. The following bar material parameters are introduced, the density $\rho$, the specific heat capacity $c$, the thermal conductivity $k$, and the convective heat transfer coefficient $h$. It is assumed that $\rho,c,k,h$ are positive constants. Dividing by $\Delta x$, and taking the limit as $\Delta x$ and $\Delta t$ approach zero, yields the partial differential equation \begin{equation} A{\pd {T}{t}}={k\over {\rho c}}{\pd {}{x}}\Big(A{\pd {T}{x}}\Big) -{{hP}\over {\rho c}}(T-T_{\infty}), (x,t)\in (0,l)\times (0,\infty). \label{1.4} \end{equation} We discuss a particular case of this general equation when convective heat transfer from the side of the bar is neglected, i.e., the limiting case $h\to 0$ is considered. It is the purpose of this paper to find the optimal distribution of the cross-section area, $A$, of a surface of revolution of a given mass such that the heat transfer rate is a maximum. This will produce a maximum cooling per unit mass and may be considered the optimum. Detailed discussion of techniques and results in structural optimization can be found in \cite{Dym,Atanackovic} and the references therein. We mention in particular the maximization of a column's buckling load \cite{Taylor,TaylorLiu}, the minimization of the mass of an oscillating bar \cite{Taylor2,Turner} or a rotating rod \cite{Atanackovic2001,Atanackovic2004}, the maximization of a column's height \cite{KellerNiordson} and the minimization of the moment of inertia of an oscillating turbine \cite{BelinskiyMcCarthy}. If the design variable tapers too rapidly, the eigenvalue being optimized is not isolated from the remainder of the spectrum \cite{KellerNiordson,CoxMcCarthy,Atanackovic2001}. In these cases optimality conditions can be derived using non-smooth analysis in conjunction with the more classical Calculus of Variations techniques \cite{McCarthy,McCarthy2}. The complexity of the Sturm-Liouville problem also increases if the boundary conditions contain an eigenparameter. This is due to the fact the the Sturm-Liouville operator is not self-adjoint with respect to the usual $L^2(0,l)$ inner product. The spectral properties of the Sturm-Liouville problems that arise from diverse mechanical models and contain the spectral parameter in the boundary condition(s) have been studied in \cite{Walter,Fulton,Hinton2,BBS,Belinskiy}. Numerical schemes for the inverse problem were developed by \cite{McCarthyRundell}. Design problems of this type have been considered by \cite{Turner,BelinskiyMcCarthy}. In the latter, existence of an optimal design was treated seriously, as it will be here. In the design problem considered here, we encounter a spectral parameter in a boundary condition. In Section \ref{sl_sect} we give the mathematical description of the model, apply separation of variables and formulate the spectral properties of the corresponding Sturm-Liouville problem. We also give the solution for an elementary case of the problem when the cross-section area is constant, which we need later for comparison with the solution in case of variable cross-section area. In Section \ref{cov_sect} we derive the necessary conditions of optimality and hence find an optimal form of the bar. In Section \ref{rearr_sect} we use a rearrangement technique to prove that the optimal design is increasing and maximizes the first eigenvalue of the Sturm-Liouville problem. In Section \ref{num_sect} we give the numerical comparison of cooling properties for the bar of optimal shape and a bar having the same mass but with the constant cross-section area. In the Appendix, we prove that the rate of cooling for the bar with the optimal cross-section area is greater than for a bar of the same mass but with constant cross-section area. \section{Heat transfer of a bar of a variable cross-section area: separation of variables and the Sturm-Liouville problem} \label{sl_sect} We consider the heat transfer in a bar $\{0z \label{6.20} \end{equation} holds for any positive ratio $M/M_0$. This is confirmed by our numerical results in Section \ref{num_sect}. \section{The Optimal Design is Increasing} \label{rearr_sect} We prove here that the heat transfer rate $\lambda_1(A)$ can be increased through the use of increasing rearrangements of the cross-sectional area $A(x)$. This is achieved through the use of an alternative characterization of $\lambda_1(A)$. We begin by defining decreasing and increasing rearrangements, and stating some of their relevant properties. \begin{definition} \rm The decreasing rearrangement of a nonnegative function, $f$, on $(a,b)$ is simply $$ f^*(x) \equiv \sup \{ t > 0 : \mu_f(t) > x \}, $$ where $\mu_f$ is the distribution function of $f$, $$\mu_f(t) = |\{x\in (a,b):f(x)>t\}| \quad t \ge 0. $$ The increasing rearrangement of $f$ is $f_*(x) \equiv f^*(b-x)$. \end{definition} If $g$ and $h$ are nonnegative functions on $(a,b)$, with $g$ increasing and $h$ decreasing, then \begin{equation} \int_a^b f dx = \int_a^b f^* dx = \int_a^b f_* dx , \label{5.7} \end{equation} By \eqref{5.7}, if we replace a particular design $A\in ad$ by either its increasing or decreasing rearrangements $A_*$ or $A^*$ then the new design has the same integral. Furthermore, \begin{equation} \int_a^b f^* g dx \leq \int_a^b fg dx , \qquad \int_a^b f_* h dx \leq \int_a^b fh dx. \label{5.8} \end{equation} These results are a special case of those established in \cite[pp. 153]{PolyaSzego}. \begin{theorem} \label{rearr_thm} For any cross-sectional area $A$ satisfying $$ 00$. Since $\lambda_1(A_*), M_0, \rho, u_1(x), A(x) >0$, this implies that $u_1$ is decreasing. The second part of \eqref{5.8} implies that $\int_0^{x} Au_1 dy \geq \int_0^{x} A_* u_1 dy $ and so $$ {{1}\over{\lambda_1(A)}} \geq {{M_0}\over{\rho}} \int_0^l \left( \int_0^x A_*(y) u_1(y) dy \right) {{dx}\over{A(x)}} +\int_0^l \left( \int_0^x A_*(y) u_1(y) dy \right)^2 {{dx}\over{A(x)}}. $$ Clearly, the functions $( \int_0^x A_* u_1 dy)$ and $( \int_0^x A_*u_1 dy)^2$ are nonnegative increasing functions of $x$. Once again \eqref{5.8} yields \begin{align*} {{1}\over{\lambda_1(A)}} & \geq {{M_0}\over{\rho}} \int_0^l \Big( \int_0^x A_*(y) u_1(y) dy \Big) \Big( {{1}\over{A(x)}} \Big)^* dx \\ &\quad +\int_0^l \Big( \int_0^x A_*(y) u_1(y) dy \Big)^2 \Big( {{1}\over{A(x)}} \Big)^* dx. \end{align*} If $f$ is decreasing on the range of $g$ then the composition $(f\circ g)^*=f\circ g_*$, see \cite{Cox}, which implies that \begin{align*} {{1}\over{\lambda_1(A)}} &\geq {{M_0}\over{\rho}} \int_0^l \Big( \int_0^x A_*(y) u_1(y) dy \Big) {{dx}\over{A_*(x)}} + \int_0^l \Big( \int_0^x A_*(y) u_1(y) dy \Big)^2 {{dx}\over{A_*(x)}}\\ & = {{1}\over{\lambda_1(A_*)}}. \qedhere \end{align*} \end{proof} \begin{theorem} \label{thm4.3} The design satisfying the first order optimality conditions \eqref{XXX}-\eqref{var-A} $$ A(x)=\frac{M\sqrt{\lambda_1}\coth(\sqrt{\lambda_1}l)}{\rho \cosh^2 (\sqrt{\lambda_1}(x-l))} $$ maximizes the functional $A\mapsto\lambda_1(A)$ on the set $$ ad = \big\{ A : 01$ in the graph are shown for illustrative purposes only. \begin{figure}[ht] \begin{center} \includegraphics[width=0.7\textwidth]{fig1} \end{center} \caption{Comparision of optimal design and constant area design} \label{bhm} \end{figure} Numerical results show that the advantage of the optimum cross-section over the constant cross-section is small and becomes less so as the base mass, $M_0$, increases. This is physically reasonable. Indeed, recall that convective heat transfer from the side of the area has been neglected. Hence addition of the extended surface does little but move the boundary condition at $x=l$ that distance from $M_0$. Furthermore, as $M$ becomes small compared to $M_0$, its very presence becomes negligible and hence its shape does not matter. In each case $z^{\rm opt}\ge z$, as shown in the Appendix. This numerical observation is certainly in agreement with the general results of Sections \ref{cov_sect} and \ref{rearr_sect}. However, the effect is not large because of the physical reasons explained above. Moreover, for $M/M_0 \to 0$ the optimum and constant area results merge. This numerical observation is in agreement with the asymptotic formula \eqref{B.5}. \section{Conclusion} We have found the optimal distribution of the cross-section area of a bar in the form of a surface of revolution of a given total mass with a point mass attached at the end such that the heat transfer rate is a maximum. That rate is defined by the least eigenvalue of the corresponding Sturm-Liouville problem. This is of independent interest because the spectral parameter appears not only in the differential equation but also in the boundary condition. The bar will produce the maximum cooling per unit mass and may be considered the optimum. The optimal distribution coincides with one found by Taylor \cite{Taylor} and M.J. Turner \cite{Turner} for the design of a bar having a maximum lowest eigenfrequency with the given mass. Numerical results show that the advantage of the optimal design over the constant cross-section is small and decreases as the base mass increases. We believe this to be a result of the fact that our model neglects heat transfer from the side of the bar. We should emphasize that we have considered a special case of the heat transfer assuming that convective heat transfer from the side of the bar is neglected and only conduction through the length of the bar is considered (see Section \ref{intro_sect}). We expect that the solution of the optimal design problem for the more general problem will show a more noticeable difference between the optimal design and the constant case. If we were to include the heat transfer phenomenon from the sides of the bar, we would have to consider the partial differential equation \begin{equation} a^2(x){\pd {T}{t}}={k\over {\rho c}}{\pd {}{x}}\Big(a^2(x){\pd {T}{x}}\Big) -{{ha(x)\sqrt{1+(a'(x)^2}}\over {\rho c}}(T-T_{\infty}), \end{equation} where $ (x,t)\in (0,l)\times (0,\infty)$, and $a(x)$ would be the radius of the body of revolution that represents the bar. The corresponding Sturm-Liouville problem has a discrete spectrum and a complete set of eigenfunctions. We could derive a Rayleigh-Ritz ratio for the least eigenvalue similar to expression \eqref{2.14}. But the technique of the Calculus of Variations used in Section \ref{cov_sect} will not lead to an explicit form of the cross-section area and the least eigenvalue. For that problem, we had hoped to use a numerical approach based on the discretization of our bar that would reduce the problem of the optimal design to the problem of optimization for a function of several variables (this idea was developed for the optimal design of mechanical systems in \cite{Cardou}). We have recently learned that an equation with similar appearance of the function $a(x)$ is optimized in \cite{Henrot-Privat}. The techniques used there may be applicable when we consider the more general heat transfer model. This will be will be considered in a future paper. \section{Appendix} Having derived an explicit formula for the optimal rate of cooling and an equation for the rate for the bar with the constant cross-section area, we may compare them directly. We prove below that the optimal rate is greater than the rate for the bar of the same mass but with the constant cross-section area. This inequality clearly is demonstrated in our numerical results above, but is proven here for completeness. \begin{lemma} The inequality \begin{equation} z^{\rm opt} >z \label{B.1} \end{equation} where \begin{equation} z^{\rm opt}=\ln\Big(\mu+\sqrt{\mu^2+1} \Big) \label{B.2} \end{equation} and $z$ is the minimal positive root of the equation \begin{equation} z \tan z= \mu^2 \label{B.3} \end{equation} holds for any positive quantity \begin{equation} \mu\equiv \sqrt{{M\over {M_0}}}. \label{B.4} \end{equation} \end{lemma} \begin{proof} We consider both $z^{\rm opt}$ and $z$ as functions of $\mu> 0$. Note first that $z^{\rm opt}(\mu)\asymp \mu, z(\mu)\asymp \mu$, and hence \begin{equation} \lim_{\mu\to 0} z^{\rm opt}(\mu)=\lim_{\mu\to 0} z(\mu)=0. \label{B.5} \end{equation} Hence the inequality \eqref{B.1} holds if the derivative of $z^{\rm opt}(\mu)$ is greater or equal to the derivative of $z(\mu)$. It is easy to prove that the first positive solution of \eqref{B.3} is a uniquely defined function on $z\in (0,\pi/2)$ with the derivative \begin{equation} z'(\mu)={{2\mu}\over {\tan z+{z\over {\cos^2 z}}}}={{2\mu}\over {{{\mu^2}\over z} +z+{{\mu^4}\over z}}} \label{B.6} \end{equation} where we used \eqref{B.3} to get the final form. The derivative of $z^{\rm opt}$ can be easily found from \eqref{B.2}. We finally come up with the necessity to prove the following inequality \begin{equation} F(\mu)\equiv {1\over {\sqrt{1+\mu^2}}}-{{2\mu z}\over {\mu^2+\mu^4+z^2}}\ge 0. \label{B.7} \end{equation} We find first \begin{equation} \mu^2+\mu^4+z^2\ge 2\sqrt{\mu^2+\mu^4} z=2\mu \sqrt{1+\mu^2} z. \label{B.8} \end{equation} Hence \begin{equation} F(\mu)\ge {1\over {\sqrt{1+\mu^2}}}-{{2\mu z}\over {2\mu \sqrt{1+\mu^2} z}}=0 \label{B.9} \end{equation} which proves \eqref{B.7} and, along with \eqref{B.5}, proves \eqref{B.1}. \end{proof} \subsection*{Acknowledgments} The first author was supported in part by a University of Tennessee at Chattanooga Faculty Research Grant. The third author was supported in part by NSF DMS \# 0209562 and \# 0531865. The authors are grateful to the referee for pointing out the work of Henrot and Privat. \begin{thebibliography}{00} \bibitem{Atanackovic2001} T.~M. Atanackovic. \newblock \protect{On the Optimal Shape of a Rotating Rod}. \newblock {\em Transactions of the ASME}, 68:860--864, 2001. \bibitem{Atanackovic2004} T.~M. Atanackovic. \newblock \protect{On the Optimal Shape of a Compressed Rotating Rod}. \newblock {\em Meccanica}, 39:147--157, 2004. \bibitem{Belinskiy} B.~P. Belinskiy and J.~P. Dauer. \newblock Eigenoscillations of mechanical systems with boundary conditions containing the frequency. \newblock {\em Quarterly of Applied Mathematics}, 56(3):521--541, 1998. \bibitem{BelinskiyMcCarthy} B.~P. Belinskiy, C.~M. McCarthy, and T.~J. Walters. \newblock Optimal design of turbines with an attached mass. \newblock {\em European Series in Applied and Industrial Mathematics: Control, Optimization and Calculus of Variations (ESAIM: COCV)}, 9:217--230, 2003. \bibitem{BBS} P.~A. Binding, P.~J. Browne, and K.~Seddighi. \newblock \protect{Sturm-Liouville} problems with eigenparameter dependent boundary condtions. \newblock {\em Proceedings of the Edinburgh Mathematical Society}, 37:57--72, 1993. \bibitem{Cardou} A.~Cardou. \newblock Piecewise uniform optimum design for axial vibration requirement. \newblock {\em AIAA J.}, 11(12):1760--1761, 1973. \bibitem{Cox} S.~J. Cox. \newblock The two phase drum with the deepest base note. \newblock {\em Japan Journal of Industrial and Applied Mathematics}, 8(3):345--355, 1991. \bibitem{CoxMcCarthy} S.~J. Cox and C.~M. McCarthy. \newblock The shape of the tallest column. \newblock {\em SIAM Journal on Mathematical Analysis}, 29(3):1--8, 1998. \bibitem{Dym} C.~L. Dym. \newblock On some recent approaches to structural optimization. \newblock {\em J. Sounds \& Vibration}, 32:49--70, 1974. \bibitem{Fulton} C.~T. Fulton. \newblock Two-point boundary value problems with eigenvalue parameter contained in the boundary conditions. \newblock {\em Proc. Royal Soc. Edin.}, 77A:293--308, 1977. \bibitem{Henrot-Privat} A.~Henrot and Y.~Privat. \newblock Shape minimization of dendritic atttenuation. \newblock {\em Appl. Math. Optim.}, 57:1--16, 2008. \bibitem{Hinton2} D.~Hinton. \newblock Eigenfunction expansions for a singular eigenvalue problem with eigenparameter in the boundary condition. \newblock {\em SIAM J. Math. Anal.}, 12(4):572--584, 1981. \bibitem{Incropera} F.~P. Incropera and D.~P. Dewitt. \newblock {\em Fundamentals of Heat and Mass Transfer}. \newblock John Wiley \& Sons, New York, 4th edition, 1996. \bibitem{KellerNiordson} J.~B. Keller and F.~I. Niordson. \newblock The tallest column. \newblock {\em J. Math. Mech.}, 16:433--446, 1966. \bibitem{Ladyzhenskaia} O.~A. Ladyzhenskaia. \newblock {\em The Boundary Value Problems of Mathematical Physics}. \newblock Springer-Verlag, New York, 1985. \bibitem{McCarthy} C.~M. McCarthy. \newblock The tallest column: Optimalty revisited. \newblock {\em Journal of Computational and Applied Mathematics}, 101(1-2):27--37, 1999. \bibitem{McCarthy2} C.~M. McCarthy. \newblock The optimal design of tubular structures. \newblock {\em Journal of Computational and Applied Mathematics}, 114(1):55--66, 2000. \bibitem{McCarthyRundell} C.~M. McCarthy and W.~Rundell. \newblock \protect{The inverse eigenparameter dependent Sturm-Liouville problem}. \newblock {\em Numerical Functional Analysis and Optimization}, 24(1-2):85--106, 2003. \bibitem{PolyaSzego} G.~Polya and G.~Szego. \newblock {\em Isoperimetric Inequalites in Mathematical Physics}, volume~27 of {\em Ann. of Math. Stud.} \newblock Princeton University Press, Princeton, NJ, 1951. \bibitem{PorterStirling} D.~Porter and D.~Stirling. \newblock {\em Integral Equations}. \newblock Cambridge University Press, Cambridge, UK, 1990. \bibitem{Powers} D.~L. Powers. \newblock {\em Boundary Value Problems}. \newblock Academic Press, New York, 2nd edition, 1979. \bibitem{Schneider} P.~J. Schneider. \newblock {\em Conduction Heat Transfer}. \newblock Addison-Wesley, 1955. \bibitem{Taylor2} J.~E. Taylor. \newblock Minimum mass bar for axial vibrations at specified natural frequency. \newblock {\em AIAA J.}, 5:1911--1913, 1967. \bibitem{Taylor} J.~E. Taylor. \newblock The strongest column: the energy approach. \newblock {\em J. of Appl. Mech.}, 34:486--487, 1967. \bibitem{TaylorLiu} J.~E. Taylor and C.~Y. Liu. \newblock On the optimal design of columns. \newblock {\em AIAA J.}, 6:1497--1502, 1968. \bibitem{Turner} M.~J. Turner. \newblock Design of minimum mass structures with specified natural frequencies. \newblock {\em AIAA J.}, 5:406--412, 1967. \bibitem{Atanackovic} B.D. Vujanovic and T.~M. Atanackovic. \newblock {\em \protect{An Introduction to Modern Variation Techniques in Mechanics and Engineering}}. \newblock Birkh\"{a}user, Boston, 2004. \bibitem{Walter} J.~Walter. \newblock Regular eigenvalue problems with eigenvalue parameter in the boundary condition. \newblock {\em Math. Z.}, 133:301--312, 1973. \end{thebibliography} \end{document}