Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

In this chapter both the classical Hamilton-Kirchhoff Principle and a convolutional variational principle of Gurtin’s type that describes completely a solution to an initial-boundary value problem of elastodynamics are used to solve a number of typical problems of elastodynamics.

1 The Hamilton-Kirchhoff Principle

To formulate H-K principle we introduce a notion of kinematically admissible process, and by this we mean an admissible process that satisfies the strain-displacement relation, the stress-strain relation, and the displacement boundary condition.

(H-K) The Hamilton-Kirchhoff Principle. Let \(P\) denote the set of all kinematically admissible processes \({p}=[\mathbf{u,E,S}]\) on \({\overline{\mathrm{{B}}}}\times [0,\infty )\) satisfying the conditions

$$\begin{aligned} \mathbf{{u}}(\mathbf{{x}},\mathrm{{t}}_1 )=\mathbf{{u}}_1 (\mathbf{{x}}),\quad \mathbf{{u}}(\mathbf{{x}},\mathrm{{t}}_2 )=\mathbf{{u}}_2 (\mathbf{{x}})\quad \mathrm{{on}}\quad {\overline{\mathrm{{B}}}} \end{aligned}$$
(5.1)

where \(t_1\) and \(t_2\) are two arbitrary points on the \(t\)-axis such that \(0\le {t}_1 <{t}_2 <\infty \), and \(\mathbf{u}_1 (\mathbf{x})\) and \(\mathbf{u}_2 (\mathbf{x})\) are prescribed fields on \(\overline{\mathrm{B}}\). Let \(\mathsf{K }=\mathsf{K }\{{p}\}\) be the functional on \(P\) defined by

$$\begin{aligned} \mathsf{{K}}\{{p}\}=\int \limits _{{t}_1 }^{{t}_2 } {[\mathrm{F}({t})-\mathrm{K}({t})]\,{dt}} \end{aligned}$$
(5.2)

where

$$\begin{aligned} {{F}(t)}=\mathrm{U}_\mathrm{C} \{\mathbf{E}\}-\int \limits _\mathrm{B} {\mathbf{b}\cdot \mathbf{u}} \,{ dv}-\int \limits _{\partial \mathrm{B}_2} {\hat{\mathbf{s}}\cdot \mathbf{u}\,{da}} \end{aligned}$$
(5.3)

and

$$\begin{aligned} {{K}(t)}=\frac{1}{2}\int \limits _\mathrm{B} {\rho \,\dot{\mathbf{u}}^2{dv}} \end{aligned}$$
(5.4)

for every \({p}=[\mathbf{u, E, S}]\in {P}\). Then

$$\begin{aligned} \delta \,\mathsf{K }\{{p}\}=0 \end{aligned}$$
(5.5)

if and only if p satisfies the equation of motion and the traction boundary condition.

Clearly, in the (H-K) principle a displacement vector \(\mathbf{u}=\mathbf{u}(\mathbf{x},\mathrm{t})\) needs to be prescribed at two points \(t_1\) and \(t_2\) of the time axis. If \(t_1 =0\), then \(\mathbf{u}(\mathbf{x},0)\) may be identified with the initial value of the displacement vector in the formulation of an initial-boundary value problem, however, the value \(\mathbf{u}(\mathbf{x},\mathrm{t}_2 )\) is not available in this formulation. This is the reason why the (H-K) principle can not be used to describe the initial-boundary value problem. A full variational characterization of an initial-boundary value problem of elastodynamics is due to Gurtin, and it has the form of a convolutional variational principle. The idea of a convolutional variational principle of elastodynamics is now explained using a traction initial-boundary value problem of incompatible elastodynamics. In such a problem we are to find a symmetric second-order tensor field \(\mathbf{S}=\mathbf{S}(\mathbf{x},{t})\;\mathrm{on}\;\overline{\mathrm{B}}\times [0,\infty )\) that satisfies the field equation

$$\begin{aligned} \hat{\nabla }[\rho ^{-1}(\mathrm{div}\,\mathbf{S})]-{\mathbf{\mathsf{{K}} }}[\ddot{\mathbf{S}}]=-\mathbf{B}\quad \mathrm{on}\quad \mathrm{B}\times [0,\infty ) \end{aligned}$$
(5.6)

subject to the initial conditions

$$\begin{aligned} \mathbf{S}(\mathbf{x},0)=\mathbf{S}_0 (\mathbf{x}),\quad \dot{\mathbf{S}}(\mathbf{x},0)=\dot{\mathbf{S}}_0 (\mathbf{x})\quad \mathrm{for}\quad \mathbf{x}\in \mathrm{B} \end{aligned}$$
(5.7)

and the boundary condition

$$\begin{aligned} \mathbf{s}=\mathbf{Sn}=\hat{\mathbf{s}}\quad \mathrm{on}\quad \partial \mathrm{B}\times [0,\infty ) \end{aligned}$$
(5.8)

Here \(\mathbf{S}_0 \) and \(\dot{\mathbf{S}}_0 \) are arbitrary symmetric tensor fields on B, and B is a prescribed symmetric second-order tensor field on \(\overline{\mathrm{B}}\times [0,\infty )\). Moreover, \(\rho ,{\mathbf{\mathsf{{K}} }},\) and \(\hat{\mathbf{s}}\) have the same meaning as in classical elastodynamics.

First, we note that the problem is equivalent to the following one. Find a symmetric second-order tensor field on \(\overline{\mathrm{B}}\times [0,\infty )\) that satisfies the integro-differential equation

$$\begin{aligned} \hat{\nabla }[\rho ^{-1}{ t}*(\mathrm{div}\,\mathbf{S})]-{\mathbf{\mathsf{{K}} }}[\mathbf{S}]=-\hat{\mathbf{B}}\quad \mathrm{on}\quad \mathrm{B}\times [0,\infty ) \end{aligned}$$
(5.9)

subject to the boundary condition

$$\begin{aligned} \mathbf{s}=\mathbf{Sn}=\hat{\mathbf{s}}\quad \mathrm{on}\quad \partial \mathrm{B}\times [0,\infty ) \end{aligned}$$
(5.10)

where

$$\begin{aligned} \hat{\mathbf{B}}={t}*\mathbf{B}+{\mathbf{\mathsf{{K}} }}[\mathbf{S}_0 +\mathrm{t}\,\dot{\mathbf{S}}_0 ] \end{aligned}$$
(5.11)

and \(*\) stands for the convolution product, that is, for any two scalar functions \({a}={a}(\mathbf{x},{t})\;\mathrm{and}\;{b}={b}(\mathbf{x},{t})\)

$$\begin{aligned} ({a}*{b})(\mathbf{x},{t})=\int \limits _0^{t} {{a}(\mathbf{x},\tau ){b}(\mathbf{x},{t}-\tau )} \,{d}\tau \end{aligned}$$
(5.12)

Next, the convolutional variational principle is formulated for the problem described by Eqs. (5.9)–(5.10).

Principle of Incompatible Elastodynamics. Let \(N\) denote the set of all symmetric second-order tensor fields S on \(\overline{\mathrm{B}}\times [0,\infty )\) that satisfy the traction boundary condition (5.8) \(\equiv \) (5.10). Let \(\mathrm{I}_{t} \{\mathbf{S}\}\) be the functional on \(N\) defined by

$$\begin{aligned} \mathrm{I}_{t} \{\mathbf{S}\}=\frac{1}{2}\int \limits _\mathrm{B} {\{\,\rho ^{-1}{t}*(\mathrm{div}\,\mathbf{S})\,*}\, (\mathrm{div}\,\mathbf{S})+\mathbf{S}*{\mathbf{\mathsf{{K}} }}[\mathbf{S}]-2\,\mathbf{S}*\hat{\mathbf{B}}\,\}\,{dv} \end{aligned}$$
(5.13)

Then

$$\begin{aligned} \delta \,\mathrm{I}_\mathrm{t} \{\mathbf{S}\}=0 \end{aligned}$$
(5.14)

at a particular \(\mathbf{S}\in \mathrm{N}\) if and only if S is a solution to the traction problem described by Eqs.  (5.6)–(5.8).

Note. When the fields B, \(\mathbf{S}_0 \), and \(\dot{\mathbf{S}}_0 \) are arbitrarily prescribed, the principle of incompatible elastodynamics may be useful in a study of elastic waves in bodies with various types of defects.

2 Problems and Solutions Related to Variational Principles of Elastodynamics

Problem 5.1.

A symmetrical elastic beam of flexural rigidity \(EI\), density \(\rho \), and length \(L\), is acted upon by: (i) the transverse force \({F}={F}({x_{1}}, {t})\), (ii) the end shear forces \(V_0 \;\mathrm{and}\;{V}_{L}\), and (iii) the end bending moments \({M}_0 \;\mathrm{and}\;{M}_{L}\) shown in Fig. 5.1. The strain energy of the beam is

$$\begin{aligned} {F}({t})=\frac{1}{2}\int \limits _0^{L} {{E\,I}({u}^{\prime \prime }_2 )^2\,{dx}_1 } \end{aligned}$$
(5.15)

the kinetic energy of the beam is

$$\begin{aligned} {K(t)}=\frac{1}{2}\int \limits _0^{L} {\rho \,(\dot{u}_2 )^2\,{dx}_1 } \end{aligned}$$
(5.16)
Fig. 5.1
figure 1figure 1

The symmetrical beam

and the energy of external forces is

$$\begin{aligned} V(t)&=-\int \limits _0^L {F(x_1 ,t)\,u_2 (x_1 ,t)\,dx_1 +V_0 } u_2 (0,t)\nonumber \\&\quad \;+\,M_0 {u}^{\prime }_2 (0,t)-V_L u_2 (L,t)-M_L {u}^{\prime }_2 (L,t) \end{aligned}$$
(5.17)

where the prime denotes differentiation with respect to \({x}_1\). Let \(U\) be the set of functions \({u}_2 ={u}_2 ({x}_1, {t})\) that satisfies the conditions

$$\begin{aligned} {u_2 (x_1, t_1 )=u(x_1 ),\quad u_2 (x_1, t_2 )=v(x_1 )} \end{aligned}$$
(5.18)

where \({t}_1 \;\mathrm{and}\;{t}_2 \) are two arbitrary points on the \(t\)-axis such that \(0\le {t}_1 <{t}_2 <\infty \), and \({u(x}_1 )\) and \({v(x}_1 )\) are prescribed fields on \([0,{L}]\). Define a functional \(\hat{K}\{{u}_2 \}\) on \(U\) by

$$\begin{aligned} \hat{K}\{{u}_2 \}=\int \limits _{{t}_1 }^{{t}_2 } {[F(t)+V(t)-K(t)]\,dt} \end{aligned}$$
(5.19)

Show that

$$\begin{aligned} \delta \hat{K}\{{u}_2 \}=0 \end{aligned}$$
(5.20)

if and only if \({u}_2 \) satisfies the equation of motion

$$\begin{aligned} ({EI{u}}^{\prime \prime }_2 {)}^{\prime \prime }+\rho \,\ddot{u}_2 ={F}\quad \mathrm{on}\quad [0,{L}]\times [0,\infty ) \end{aligned}$$
(5.21)

and the boundary conditions

$$\begin{aligned}{}[({EI\,{u}}^{\prime \prime }_2 {)}^{\prime }]\,(0,{t})&=-{V}_0 \quad \; \mathrm{on}\quad [0,\infty ) \end{aligned}$$
(5.22)
$$\begin{aligned} {[}({EI\,{u}}^{\prime \prime }_2 )]\,(0,{t})&= {M}_0 \qquad \mathrm{on}\quad [0,\infty ) \end{aligned}$$
(5.23)
$$\begin{aligned} {[}({EI\,{u}}^{\prime \prime }_2 {)}^{\prime }]\,({L,t})&= -{V}_{L} \quad \; \mathrm{on}\quad [0,\infty ) \end{aligned}$$
(5.24)
$$\begin{aligned} {[}({EI\,{u}}^{\prime \prime }_2 )]\,({L,t})&= {M}_{L} \qquad \mathrm{on}\quad [0,\infty ) \end{aligned}$$
(5.25)

The field equaton (5.21) and the boundary conditions (5.22) through (5.25) describe flexural waves in the beam.

Solution.

Introduce the notation

$$\begin{aligned} { u}_{2}(x_{1},{ t})\equiv { u}({ x,t}) \end{aligned}$$
(5.26)

Then the functional \(\hat{ K}\{{ u}_{2}\}\) takes the form

$$\begin{aligned} \hat{ K}\{{ u}\}&=\frac{1}{2}\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{0}^{ L}{ dx}\ [{ EI}({ u}^{\prime \prime })^{2}-\rho (\dot{ u})^{2}] \nonumber \\&\quad +\int \limits _{{ t}_{1}}^{{ t}_{2}}\left\{ -\int \limits _{0}^{ L}\!{ Fu \, dx }+{ V}_{0}{ u}(0,{ t})+{ M}_{0} { u}^{\prime }(0,{ t})-{ V}_{ L}{ u}({ L,t})-{ M}_{ L} { u}^{\prime }({ L,t})\!\right\} \!{ dt} \end{aligned}$$
(5.27)

Let \({ u}\in { U}\) and \({ u}+\omega \tilde{ u}\in { U}\). Then

$$\begin{aligned} \tilde{ u}({ x},{ t}_{1})=\tilde{ u}({ x},{ t}_{2})=0\qquad { x}\in [0,{ L}] \end{aligned}$$
(5.28)

Computing \(\delta \hat{ K}\{{ u}\}\) we obtain

$$\begin{aligned} \delta \hat{ K}\{{ u}\}&= \left. \frac{{ d}}{{ d}\omega }\ \hat{ K}\{{ u}+\omega \tilde{ u}\}\right| _{\omega =0} =\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{0}^{ L}{ dx}\ [EI{u}^{\prime \prime }\,{\tilde{u}}{\,}^{\prime \prime }-\rho \,\dot{u}\,\dot{\tilde{u}}]\nonumber \\&\;\quad +\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\left\{ -\int \limits _{0}^{ L}{ F}\tilde{ u}\,{ dx}+{ V}_0\tilde{ u}(0, { t})+{ M}_0\tilde{ u}^{\prime }(0,{ t})\right. \nonumber \\&\;\quad \left. -{ V}_{ L}\tilde{ u}({ L,t})-{ M}_{ L}\tilde{ u}^\prime ({ L, t})\right\} \end{aligned}$$
(5.29)

Next, note that integrating by parts we obtain

$$\begin{aligned} \left. \int \limits _{0}^{ L}{ dx}({ EIu}^{\prime \prime }\tilde{ u}{\,}^{\prime \prime })\right.&=\left. ({ EIu}^{\prime \prime })\tilde{ u}{\,}^{\prime }\right| _{{ x}\,=\,0}^{{ x}\,=\,{ L}}-\int \limits _{0}^{ L}{ dx}({ EIu}^{\prime \prime })^{\prime }\tilde{ u}{\,}^{\prime }\nonumber \\&=\left. ({ EIu}^{\prime \prime })\tilde{ u}{\,}^{\prime }\right| _{{ x}\,=\,0}^{{ x}\,=\,{ L}}\left. -({ EIu}^{\prime \prime })^{\prime }\tilde{ u}\right| _{{ x}\,=\,0}^{{ x}\,=\,{ L}}+\int \limits _{0}^{ L}{ dx}({ EIu}^{\prime \prime })^{\prime \prime }\tilde{ u} \end{aligned}$$
(5.30)

and

$$\begin{aligned} -\int \limits _{{ t}_{1}}^{{ t}_{2}}\rho \dot{ u}\dot{\tilde{u}}\,{ dt}=-\rho {\dot{ u}\tilde{ u}}\bigg |_{{ t}\,=\,{ t}_{1}}^{{ t}\,=\,{ t}_{2}}+\int \limits _{{ t}_{1}}^{{ t}_{2}}\rho \ddot{ u}\tilde{ u}\,{ dt} \end{aligned}$$
(5.31)

Hence, using the homogeneous conditions (5.28) we reduce (5.29) to the form

$$\begin{aligned} \delta \hat{ K}\{{ u}\}&=\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{0}^{{ L}}{ dx}\ [({ EIu}^{\prime \prime })^{\prime \prime }+\rho \ddot{ u}-{ F}]\tilde{ u}({ x, t})\nonumber \\&\quad +\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\{[{ V}_{0}+({ EIu}^{\prime \prime })^{\prime }(0,{ t})]\tilde{ u}(0,{ t}) -[{ V_{L}}+({ EIu}^{\prime \prime })^{\prime }({ L,t})]\tilde{ u}({ L,t})\nonumber \\&\quad +[{ M}_{0}-({ EIu}^{\prime \prime })(0,{ t})]\tilde{ u}{\,}^{\prime }(0,{ t}) -[{ M}_{ L}-({ EIu}^{\prime \prime })({ L,t})]\tilde{ u}{\,}^{\prime }({ L,t})\} \end{aligned}$$
(5.32)

Now, if \({ u}={ u}({ x,t})\) satisfies (5.21)–(5.25) then \(\delta \hat{ K}\{{ u}\}=0\). Conversely, if \({\delta \hat{ K}\{{ u}\}=0}\) then selecting \(\tilde{ u}=\tilde{ u}({ x,t})\) in such a way that \(\tilde{ u}=\tilde{ u}({ x,t})\) is an arbitrary smooth function on \([{ 0,L}]\times [{ t}_{1},{ t}_{2}]\) and such that \(\tilde{ u}({ 0,t})=\tilde{ u}({ L,t})=0\) on \([{ t_{1},t_{2}}]\) and \(\tilde{ u}^{\prime }({ 0,t})=\tilde{ u}{\,}^{\prime }({ L,t})=0\) on \([{ t_{1},t_{2}}]\), from Eq. (5.32) we obtain

$$\begin{aligned} \int \limits _{{ t}_{1}}^{{ t}_{2}}\int \limits _{0}^{ L}[({ EIu}^{\prime \prime })^{\prime \prime }+\rho {\ddot{u}}-{ F}]\tilde{ u}\ { dtdx}=0 \end{aligned}$$
(5.33)

and by the Fundamental Lemma of the calculus of variations we obtain

$$\begin{aligned} ({ EIu}^{\prime \prime })^{\prime \prime }+\rho {\ddot{ u}}={ F} \end{aligned}$$
(5.34)

Next, by selecting \(\tilde{ u}=\tilde{ u}({ x,t})\) in such a way that \(\tilde{u}\) is an arbitrary smooth function on \([{ 0,L}]\times [{ t_{1},t_{2}}]\) that complies with the conditions \(\tilde{ u}({ 0,t})\ne 0\) on \([{ t_{1},t_{2}}],\ \tilde{ u}({ L,t})=0,\tilde{ u}{\,}^{\prime }({ 0,t})=\tilde{ u}{\,}^{\prime }({ L,t})=0\) on \([{ t_{1},t_{2}}]\), and by using (5.32) and (5.34), we obtain

$$\begin{aligned} \int \limits _{{ t}_{1}}^{{ t}_{2}}[{ V}_{0}+({ EIu}^{\prime \prime })^{\prime }(0,{ t})]\tilde{ u} (0,{ t})=0 \end{aligned}$$
(5.35)

This together with the Fundamental Lemma of calculus of variations yields

$$\begin{aligned} ({ EIu}^{\prime \prime })^{\prime }(0,{ t})=-{ V}_{0} \end{aligned}$$
(5.36)

Next, by selecting \(\tilde{ u}\) to be an arbitrary smooth function on \([{ 0,L}]\times [{ t_{1},t_{2}}]\) that satisfies the conditions \(\tilde{ u}({ L,t})\ne 0\) on \([{ t_{1},t_{2}}],\ \tilde{ u}{\,}^{\prime }({ 0,t})=0\), and \(\tilde{ u}{\,}^{\prime }({ L,t})=0\) on \([{ t_{1},t_{2}}]\), we find from Eqs. (5.34), (5.36), and (5.32) that

$$\begin{aligned} \int \limits _{{ t}_{1}}^{{ t}_{2}}[{ V}_{ L}+({ EIu}^{\prime \prime })^{\prime }({ L,t})]\tilde{ u}({ L,t}){ dt}=0 \end{aligned}$$
(5.37)

Equation (5.37) together with the Fundamental Lemma of calculus of variations imply that

$$\begin{aligned} ({ EIu}^{\prime \prime })^{\prime }({ L,t})=-{ V}_{ L} \end{aligned}$$
(5.38)

Next, by selecting \(\tilde{u}\) to be an arbitrary smooth function on \([{ 0,L}]\times [{ t_{1},t_{2}}]\) that meets the conditions \(\tilde{ u}{\,}^{\prime }({ 0,t})\ne 0\) on \([{ t_{1},t_{2}}]\), and \(\tilde{ u}{\,}^{\prime }({ L,t})=0\) on \([{ t_{1},t_{2}}]\), by virtue of Eqs. (5.34), (5.36), (5.38), and (5.32), we obtain

$$\begin{aligned} \int \limits _{{ t}_{1}}^{{ t}_{2}}[{ M}_{0}-({ EIu}^{\prime \prime })(0,{ t})]\tilde{ u}{\,}^{\prime }(0,{ t}){ dt}=0 \end{aligned}$$
(5.39)

This together with the Fundamental Lemma of calculus of variations yields

$$\begin{aligned} ({ EIu}^{\prime \prime })(0,{ t})={ M}_{0} \end{aligned}$$
(5.40)

Finally, by letting \(\tilde{u}\) to be an arbitrary smooth function on \([{ 0,L}]\times [{ t_{1},t_{2}}]\) and such that \(\tilde{ u}{\,}^{\prime }({ L,t})\ne 0\), from Eqs. (5.34), (5.36), (5.38), (5.40), and (5.32) we obtain

$$\begin{aligned} \int \limits _{{ t}_{1}}^{{ t}_{2}}[{ M_{L}}-({ EIu}^{\prime \prime })({ L,t})]\tilde{ u}{\,}^{\prime }({ L,t})=0 \end{aligned}$$
(5.41)

Equation (5.41) together with the Fundamental Lemma of calculus of variations yields

$$\begin{aligned} ({ EIu}^{\prime \prime })({ L,t})={ M_{L}} \end{aligned}$$
(5.42)

This completes a solution to Problem 5.1.

Problem 5.2.

A thin elastic membrane of uniform area density \(\hat{\rho }\) is stretched to a uniform tension \(\hat{T}\) over a region \({C}_0 \) of the \({x}_1, {x}_2\) plane. The membrane is subject to a vertical load \({f}={f}(\mathbf{x},{t})\;\mathrm{on}\;{C}_0 \times [0,\infty )\) and the initial conditions

$$\begin{aligned} {u}(\mathbf{x},0)={u}_0 (\mathbf{x}),\quad \dot{u}(\mathbf{x},0)=\dot{u}_0 (\mathbf{x})\quad \mathrm{for}\quad \mathbf{x}\in {C}_0 \end{aligned}$$

where \({u}={u}(\mathbf{x},{t})\) is a vertical deflection of the membrane on \(\overline{C}_0 \times [0,\infty )\), and \({u}_0 (\mathbf{x})\) and \(\dot{u}_0 (\mathbf{x})\) are prescribed functions on \({C}_0 \). Also, \({u}={u}(\mathbf{x},{t})\) on \(\partial {C}_0 \times [0,\infty )\) is represented by a given function \({g}={g}(\mathbf{x},{t})\). The strain energy of the membrane is

$$\begin{aligned} {F(t)}=\frac{\hat{T}}{2}\int \limits _{\mathrm{C}_0 } {{u}_{,\alpha }} {u}_{,\alpha } \,{da} \end{aligned}$$
(5.43)

The kinetic energy of the membrane is

$$\begin{aligned} {K(t)}=\frac{\hat{\rho }}{2}\int \limits _{\mathrm{C}_0 } {(\dot{u})^2} \,{da} \end{aligned}$$
(5.44)

The external load energy is

$$\begin{aligned} {V(t)}=-\int \limits _{\mathrm{C}_0} {f\,u\,da} \end{aligned}$$
(5.45)

Let \(U\) be the set of functions \({u}={u}(\mathbf{x},{t})\) on \({C}_0 \times [0,\infty )\) that satisfy the conditions

$$\begin{aligned} {u}(\mathbf{x},{t}_1)={a}(\mathbf{x}),\quad {u}(\mathbf{x},{t}_2 )={b}(\mathbf{x})\quad \mathrm{for}\quad \mathbf{x}\in {C}_0 \end{aligned}$$
(5.46)

and

$$\begin{aligned} {u}(\mathbf{x},{t})={g}(\mathbf{x},{t})\quad \mathrm{on}\quad \partial {C}_0 \times [0,\infty ) \end{aligned}$$
(5.47)

where \({t}_1 \;\mathrm{and}\;{t}_2\) have the same meaning as in Problem 5.1, and \(a\)(x) and \(b\)(x) are prescribed functions on \(C_0\). Define a functional \(\hat{K}\{.\}\) on \(U\) by

$$\begin{aligned} \hat{K}\{{u}\}=\int \limits _{{t}_1 }^{{t}_2}{[F(t)+V(t)-K(t)]\,dt} \end{aligned}$$
(5.48)

Show that the condition

$$\begin{aligned} \delta \,\hat{K}\{{u}\}=0\quad \mathrm{on}\quad {U} \end{aligned}$$
(5.49)

implies the wave equation

$$\begin{aligned} \left( {\nabla ^2-\frac{1}{{c}^2}\frac{\partial ^2}{\partial {t}^2}} \right) \,{u}=-\frac{{f}}{\hat{T}}\quad \mathrm{on}\quad {C}_0 \times [0,\infty ) \end{aligned}$$
(5.50)

where

$$\begin{aligned} {c}=\sqrt{\frac{\hat{T}}{\hat{\rho }}} \end{aligned}$$
(5.51)

Note that \([\hat{T}]=[\mathrm{Force}\times {L}^{-1}],\;[\hat{\rho }]=[\mathrm{Density}\times {L}],\;[{c}]=[{LT}^{-1}],\) where \(L\) and \(T\) are the length and time units, respectively.

Solution.

The functional \(\hat{ K}=\hat{ K}\{{ u}\}\) takes the form

$$\begin{aligned} \hat{ K}\{{ u}\}=\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{{ C}_{0}}\left( \frac{\widehat{ T}}{2}{ u}, _{\alpha }{ u}, _{\alpha }-\frac{\hat{\rho }}{2}\dot{ u}^{2}-{ fu}\right) { da}\nonumber \\ \qquad \qquad \qquad \qquad \qquad \mathrm{for\ every}\qquad u\in {U} \end{aligned}$$
(5.52)

Let \({ u}\in { U}\) and \({ u}+{ \omega }\tilde{ u}\in { U}\). Then

$$\begin{aligned} \tilde{ u}(\mathbf x ,{ t}_{1})=\tilde{ u}(\mathbf x ,{ t}_{2})=0\quad \mathrm{for}\ \mathbf x \in { C}_{0} \end{aligned}$$
(5.53)

and

$$\begin{aligned} \tilde{ u}(\mathbf x ,{ t})=0\quad \mathrm{on}\ \partial { C}_{0}\times [0,\infty ) \end{aligned}$$
(5.54)

Computing \(\delta \widehat{ K}\{{ u}\}\) we obtain

$$\begin{aligned} \delta \widehat{ K}\{{ u}\}&=\frac{ d}{{ d}\omega }\widehat{ K}\{{ u}+\omega \tilde{ u}\}|_{\omega \,=\,0} \nonumber \\&=\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{{ C}_{0}}(\widehat{ T}{ u}, _{\alpha }\tilde{ u}, _{\alpha }-\hat{\rho }\dot{ u}\dot{\tilde{ u}}-{ f}\tilde{ u}){ da} \end{aligned}$$
(5.55)

Since

$$\begin{aligned} { u}, _{\alpha }\tilde{ u}, _{\alpha }=({ u}, _{\alpha }\tilde{ u}), _{\alpha }-{ u}, _{\alpha \alpha }\,\tilde{ u} \end{aligned}$$
(5.56)

and

$$\begin{aligned} \dot{ u}\dot{\tilde{ u}}=(\dot{ u}\tilde{ u})^{.}-\ddot{ u}\tilde{ u} \end{aligned}$$
(5.57)

therefore, using the divergence theorem and the homogeneous conditions (5.53) and (5.54), we reduce (5.55) into the form

$$\begin{aligned} \delta \widehat{ K}\{{ u}\}=\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{{ C}_{0}}(-\widehat{ T}{ u}, _{\alpha \alpha }+\hat{\rho }\ddot{ u}-{ f})\tilde{ u}\ { da} \end{aligned}$$
(5.58)

Hence, the condition

$$\begin{aligned} \delta \widehat{ K}\{{ u}\}=0\quad \mathrm{on}\ U \end{aligned}$$
(5.59)

together with the Fundamental Lemma of calculus of variations imply Eq. (5.50). This completes a solution to Problem 5.2.

Problem 5.3.

Transverse waves propagating in a thin elastic membrane are described by the field equation (see Problem 5.2.)

$$\begin{aligned} \left( {\nabla ^2-\frac{1}{{c}^2}\frac{\partial ^2}{\partial {t}^2}} \right) \,{u}=-\frac{{f}}{\hat{T}}\quad \mathrm{on}\quad {C}_0 \times [0,\infty ) \end{aligned}$$
(5.60)

the initial conditions

$$\begin{aligned} {u}(\mathbf{x},0)={u}_0 (\mathbf{x}),\quad \dot{u}(\mathbf{x},0)=\dot{u}_0 (\mathbf{x})\quad \mathrm{for}\quad \mathbf{x}\in {C}_0 \end{aligned}$$
(5.61)

and the boundary condition

$$\begin{aligned} {u}(\mathbf{x},{t})={g}(\mathbf{x},{t})\quad \mathrm{on}\quad \partial {C}_0 \times [0,\infty ) \end{aligned}$$
(5.62)

Let \(\hat{{U}}\) be a set of functions \({u}={u}(\mathbf{x},{t})\) on \({C}_0 \times [0,\infty )\) that satisfy the boundary condition (5.62). Define a functional \(\mathcal F _{ t} \{.\}\;\mathrm{on}\;\hat{{U}}\) in such a way that

$$\begin{aligned} \delta {F}_{t} \{{u}\}=0 \end{aligned}$$
(5.63)

if and only if \({u}={u}(\mathbf{x},{t})\) is a solution to the initial-boundary value problem (5.60) through (5.62).

Solution.

By transforming the initial-boundary value problem (5.60)–(5.62) to an equivalent integro-differential boundary-value problem in a way similar to that of the Principle of Incompatible Elastodynamics [see Eqs. (5.6)–(5.12)] we find that the functional \(\mathcal F _{ t}\{{ u}\}\) on \(\hat{ U}\) takes the form

$$\begin{aligned} \mathcal F _{ t}\{{ u}\}=\frac{1}{2}\int \limits _{{ C}_{0}}({ i}*{ u}, _{\alpha }*\,{u}, _{\alpha }+\frac{1}{{ c}^{2}}{ u}*{ u}-2{ g}*{ u}){ da} \end{aligned}$$
(5.64)

where

$$\begin{aligned} { i}={ i(t)}={ t} \end{aligned}$$
(5.65)

and

$$\begin{aligned} { g}={ i}*\frac{ f}{\widehat{ T}}+\frac{1}{{ c}^{2}}({ u}_{0}+{ t}\dot{ u}_{0}) \end{aligned}$$
(5.66)

The associated variational principle reads:

$$\begin{aligned} \delta \mathcal F _{ t}\{{ u}\}=0\quad \mathrm{on}\ \hat{ U} \end{aligned}$$
(5.67)

if and only if \(u\) is a solution to the initial-boundary value problem (5.60)–(5.62). This completes a solution to Problem 5.3.

Problem 5.4.

A homogeneous isotropic thin elastic plate defined over a region \({C}_0 \) of the \({x}_1, {x}_2 \) plane, and clamped on its boundary \(\partial C_0 \), is subject to a transverse load \({p}={p}(\mathbf{x},{t})\;\mathrm{on}\;{C}_0 \times [0,\infty )\). The strain energy of the plate is

$$\begin{aligned} {F(t)}=\frac{D}{2}\int \limits _{\mathrm{C}_0}{(\nabla ^2{w})^2} \,{da} \end{aligned}$$
(5.68)

The kinetic energy of the plate is

$$\begin{aligned} {K(t)}=\frac{\hat{\rho }}{2}\int \limits _{\mathrm{C}_0} {(\dot{w})^2} \,{da} \end{aligned}$$
(5.69)

The external energy is

$$\begin{aligned} {V(t)}=-\int \limits _{\mathrm{C}_0} {p\,w\,da} \end{aligned}$$
(5.70)

Here, \({w}={w}(\mathbf{x},{t})\) is a transverse deflection of the plate on \({C}_0 \times [0,\infty )\), \(D\) is the bending rigidity of the plate (\({[D]} = \mathrm{[Force}\times \mathrm{Length}]\)), and \(\hat{\rho }\) is the area density of the plate ([\(\hat{\rho }\)] \(=\) [Density \(\times \) Length]).

Let \(W\) be the set of functions \({w}={w}(\mathbf{x},{t})\) on \({C}_0 \times [0,\infty )\) that satisfy the conditions

$$\begin{aligned} {w}(\mathbf{x},{t}_1 )={a}(\mathbf{x}),\quad {w}(\mathbf{x},{t}_2 )={b}(\mathbf{x})\quad \mathrm{for}\quad \mathbf{x}\in {C}_0 \end{aligned}$$
(5.71)

and

$$\begin{aligned} {w}=0,\quad \frac{\partial {w}}{\partial {n}}=0\quad \mathrm{on}\quad \partial {C}_0 \times [0,\infty ) \end{aligned}$$
(5.72)

where \({t}_1, \;{t}_2 \), \(a\)(x) and \(b\)(x) have the same meaning as in Problem 5.2, and \(\partial /\partial {n}\) is the normal derivative on \(\partial {C}_0 \). Define a functional \(\hat{K}\{.\}\) on \(W\) by

$$\begin{aligned} \hat{K}\{{w}\}=\int \limits _{{t}_1 }^{{t}_2 } {[F(t)+V(t)-K(t)]\,dt} \end{aligned}$$
(5.73)

Show that

$$\begin{aligned} \delta \hat{K}\{{w}\}=0\quad \mathrm{on}\quad {W} \end{aligned}$$
(5.74)

if and only if \({w}={w}(\mathbf{x},{t})\) satisfies the differential equation

$$\begin{aligned} \nabla ^2\nabla ^2{w}+\frac{\hat{\rho }}{{D}}\frac{\partial ^2{w}}{\partial {t}^2}=\frac{{p}}{{D}}\quad \mathrm{on}\quad {C}_0 \times [0,\infty ) \end{aligned}$$
(5.75)

and the boundary conditions

$$\begin{aligned} {w}=0,\quad \frac{\partial {w}}{\partial {n}}=0\quad \mathrm{on}\quad \partial {C}_0 \times [0,\infty ) \end{aligned}$$
(5.76)

Solution.

The functional \(\hat{ K}=\hat{ K}\{{ w}\}\) on \(W\) takes the form

$$\begin{aligned} \widehat{ K}\{{ w}\}=\frac{1}{2}\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{{ C}_{0}}[{ D}(\nabla ^{2}{ w})^{2}-\hat{\rho }\dot{ w}^{2}-2{ pw}]{ da} \end{aligned}$$
(5.77)

Let \({ w}\in { W,w}+\omega \tilde{ w}\in { W}\). Then

$$\begin{aligned} \tilde{ w}(\mathbf x ,{ t}_{1})=\tilde{ w}(\mathbf x ,{ t}_{2})=0\quad \mathrm{for}\ \mathbf x \in { C}_{0} \end{aligned}$$
(5.78)

and

$$\begin{aligned} \tilde{ w}=0,\quad \frac{\partial \tilde{ w}}{\partial { n}}=0\quad \mathrm{on}\ \partial { C}_{0}\times [0,\infty ) \end{aligned}$$
(5.79)

Hence, we obtain

$$\begin{aligned} \delta \widehat{ K}\{{ w}\}&=\frac{ d}{{ d}\omega }\widehat{ K}\{{ w}+\omega \tilde{ w}\}|_{\omega \,=\,0} \nonumber \\&=\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{{ C}_{0}}[{ D}(\nabla ^{2}{ w})(\nabla ^{2}\tilde{ w})-\hat{\rho }\dot{ w}\dot{\tilde{ w}}-{ p}\tilde{ w}]{ da} \end{aligned}$$
(5.80)

Since

$$\begin{aligned} (\nabla ^{2}{ w})(\nabla ^{2}\tilde{ w})&={ w}, _{\alpha \alpha }\ \tilde{ w}, _{\beta \beta }=({ w}, _{\alpha \alpha }\ \tilde{ w}, _{\beta }), _{\beta } \nonumber \\ -{ w}, _{\alpha \alpha \beta }\ \tilde{ w}, _{\beta }&=({ w}, _{\alpha \alpha }\tilde{ w}, _{\beta }-{ w}, _{\alpha \alpha \beta }\tilde{ w}), _{\beta }+{ w}, _{\alpha \alpha \beta \beta }\tilde{ w} \end{aligned}$$
(5.81)

and

$$\begin{aligned} \dot{ w}\dot{\tilde{ w}}=(\dot{ w}\tilde{ w})^{.}-\ddot{ w}\tilde{ w} \end{aligned}$$
(5.82)

therefore, using the divergence theorem as well as the homogeneous conditions (5.78) and (5.79), we reduce (5.80) to the form

$$\begin{aligned} \delta \widehat{ K}\{{ w}\}=\int \limits _{{ t}_{1}}^{{ t}_{2}}{ dt}\int \limits _{{ C}_{0}}({ D}\nabla ^{4}{ w}+\hat{\rho }\ddot{ w}-{ p})\tilde{ w}{ da} \end{aligned}$$
(5.83)

Hence, by virtue of the Fundamental Lemma of calculus of variations

$$\begin{aligned} \delta \widehat{ K}\{{ w}\}=0\quad \mathrm{on}\ { W} \end{aligned}$$
(5.84)

if and only if \(w\) satisfies the differential equation

$$\begin{aligned} \nabla ^{2}\nabla ^{2}{ w}+\frac{\hat{\rho }}{ D}\frac{\partial ^{2}{ w}}{\partial { t}^{2}}=\frac{ p}{ D}\quad \mathrm{on}\ { C}_{0}\times [0,\infty ) \end{aligned}$$
(5.85)

and the boundary conditions

$$\begin{aligned} { w}=\frac{\partial { w}}{\partial { n}}=0\quad \mathrm{on}\ \partial { C}_{0}\times [0,\infty ) \end{aligned}$$
(5.86)

This completes a solution to Problem 5.4.

Problem 5.5.

Transverse waves propagating in a clamped thin elastic plate are described by the equations (see Problem 5.4)

$$\begin{aligned} \nabla ^2\nabla ^2{w}+\frac{\hat{\rho }}{{D}}\frac{\partial ^2{w}}{\partial {t}^2}=\frac{{p}}{{D}}\quad \mathrm{on}\quad {C}_0 \times [0,\infty ) \end{aligned}$$
(5.87)
$$\begin{aligned} {w}(\mathbf{x},0)={w}_0 (\mathbf{x}),\quad \dot{{w}}(\mathbf{x},0)=\dot{{w}}_0 (\mathbf{x})\quad \mathrm{for}\quad \mathbf{x}\in { C}_0 \end{aligned}$$
(5.88)

and

$$\begin{aligned} {w}=0,\quad \frac{\partial {w}}{\partial {n}}=0\quad \mathrm{on}\quad \partial {C}_0 \times [0,\infty ) \end{aligned}$$
(5.89)

where \({w}_0 (\mathbf{x})\) and \(\dot{{w}}_0 (\mathbf{x})\) are prescribed functions on \({C}_0\). Let \({W}^*\) denote the set of functions \({w}={w}(\mathbf{x},{t})\) that satisfy the homogeneous boundary conditions (5.89). Find a functional \(\hat{\mathcal{F }}_{t} \{.\}\;\mathrm{on}\;{W}^*\) with the property that

$$\begin{aligned} \delta \,\hat{\mathcal{F }}_{t} \{{w}\}=0\quad \mathrm{on}\quad {W}^*\end{aligned}$$
(5.90)

if and only if \(w\) is a solution to the initial-boundary value problem (5.87) through (5.89).

Solution.

First, we note that the initial-boundary value problem (5.87)–(5.89) is equivalent to the following boundary-value problem. Find \({ w}={ w}(\mathbf x ,{ t})\) on \({ C}_{0}\times [0,\infty )\) that satisfies the integro-differential equation.

$$\begin{aligned} { i}*\nabla ^{4}{ w}+\frac{1}{{ c}^{2}}{ w}={ h}\quad \mathrm{{on}}\quad { C}_{0}\times [0,\infty ) \end{aligned}$$
(5.91)

subject to the boundary conditions

$$\begin{aligned} {w}=\frac{\partial { w}}{\partial {n}}=0\quad \mathrm{{on}}\quad \partial { C}_{0}\times [0,\infty ) \end{aligned}$$
(5.92)

Here,

$$\begin{aligned}&{ i}={ i(t)}={ t},\quad { h}(\mathbf x ,{ t})={ i}*\frac{ p}{ D}+\frac{1}{{ c}^{2}}({ w}_{0}+{ t}\dot{ w}_{0}),\nonumber \\&\mathrm{{and}}\quad \frac{1}{{ c}^{2}}=\frac{\hat{\rho }}{ D} \end{aligned}$$
(5.93)

Next, we define a functional \({\hat{\mathcal{F }}}_{ t}\ \{{ w}\}\) on \({ W}^{*}\) by

$$\begin{aligned} {\hat{\mathcal{F }}}_{ t}\{{ w}\}=\frac{1}{2}\int \limits _{{ C}_{0}}\left( { i}*\nabla ^{2}{ w}*\nabla ^{2}{ w}+\frac{1}{{ c}^{2}}{ w}*{ w}-2{ h}*{ w}\right) { da} \end{aligned}$$
(5.94)

By computing \(\delta \mathcal{\hat{F} }_{ t}\ \{{ w}\}\), we obtain

$$\begin{aligned} \delta {\hat{\mathcal{F }}}_{ t}\{{ w}\}=\int \limits _{{ C}_{0}}\left( { i}*\nabla ^{4}{ w}+\frac{1}{{ c}^{2}}{ w}-{ h}\right) *\tilde{ w}{ da} \end{aligned}$$
(5.95)

where \(\tilde{ w}\) is an arbitrary smooth function on \({ C}_{0}\) such that

$$\begin{aligned} \tilde{ w}=\frac{\partial \tilde{ w}}{\partial {n}}=0\quad \mathrm{{on}}\ \partial { C}_{0}\times [0,\infty ) \end{aligned}$$
(5.96)

Therefore, using the Fundamental Lemma of calculus of variations, it follows from Eq. (5.95) that the condition

$$\begin{aligned} \delta {\hat{\mathcal{F }}}_{ t}\{{ w}\}=0\quad \mathrm{{on}}\ W^{*} \end{aligned}$$
(5.97)

holds true if and only if \(w\) is a solution to the initial-boundary value problem (5.87)–(5.89). This completes a solution to Problem 5.5.

Problem 5.6.

Free longitudinal vibrations of an elastic bar are defined as solutions of the form

$$\begin{aligned} {u(x,t)}=\phi ({x})\,\sin (\omega \,{t}+\gamma ) \end{aligned}$$
(5.98)

to the homogeneous wave equation

$$\begin{aligned} \frac{\partial }{\partial {x}}\left( {{E}\frac{\partial {u}}{\partial {x}}} \right) -\rho \frac{\partial ^2{u}}{\partial {t}^2}=0\quad \mathrm{{on}}\quad [0,{L}]\times [0,\infty ) \end{aligned}$$
(5.99)

subject to the homogeneous boundary conditions

$$\begin{aligned} {u(0,t)}={u(L,t)}=0\quad \mathrm{{on}}\quad [0,\infty ) \end{aligned}$$
(5.100)

or

$$\begin{aligned} \frac{\partial {u}}{\partial {x}}({0,t})=\frac{\partial {u}}{\partial {x}}({L,t})=0\quad \mathrm{{on}}\quad [0,\infty ) \end{aligned}$$
(5.101)

Here, \(\omega \) is a circular frequency of vibrations, \(\gamma \) is a dimensionless constant, and \(\phi =\phi ({x})\) is an unknown function that complies with Eqs. (5.99) and (5.100), or Eqs. (5.99) and (5.101). Substituting \({u}={u(\mathbf x ,t)}\) from Eq. (5.98) into (5.99) through (5.101) we obtain

$$\begin{aligned} \frac{{d}}{{dx}}\left( {{E}\frac{{d}\phi }{{dx}}} \right) +\lambda \phi =0\quad \mathrm{on}\quad [{0,L}] \end{aligned}$$
(5.102)
$$\begin{aligned} \phi (0)=\phi ({L})=0 \end{aligned}$$
(5.103)

or

$$\begin{aligned} {\phi }^{\prime }(0)={\phi }^{\prime }({L})=0 \end{aligned}$$
(5.104)

where the prime stands for derivative with respect to \(x\), and

$$\begin{aligned} \lambda =\rho \omega ^2 \end{aligned}$$
(5.105)

Therefore, introduction of (5.98) into (5.99) through (5.101) results in an eigenproblem in which an eigenfunction \(\phi =\phi ({x})\) corresponding to an eigenvalue \(\lambda \) is to be found. An eigenproblem that covers both boundary conditions (5.100) and (5.101) can be written as

$$\begin{aligned} \frac{{d}}{{dx}}\left( {{E}\frac{{d}\phi }{{dx}}} \right) +\lambda \phi =0\quad \mathrm{on}\quad [0,{L}] \end{aligned}$$
(5.106)
$$\begin{aligned} {\phi }^{\prime }(0)-\alpha \,\phi (0)=0,\quad {\phi }^{\prime }({L})+\beta \,\phi ({L})=0 \end{aligned}$$
(5.107)

where \(\left| \alpha \right| +\left| \beta \right| >0.\) Let \(U\) be the set of functions \(\phi =\phi ({x})\) on [0, \(L\)] that satisfy the boundary conditions (5.107). Define a functional \(\pi \{.\}\;\mathrm{on}\;{U}\) by

$$\begin{aligned} \pi \{\phi \}=\frac{1}{2}\int \limits _0^{L} {\left[ {{E}\left( {\frac{{d}\phi }{{dx}}} \right) ^2-\lambda \phi ^2} \right] \,{dx}} +\frac{1}{2}\alpha \,{E}(0)\,\,[\phi (0)]^2+\frac{1}{2}\beta \,{E}({L})\,\,[\phi ({L})]^2 \end{aligned}$$
(5.108)

Show that

$$\begin{aligned} \delta \,\pi \{\phi \}=0\quad \mathrm{over}\quad {U} \end{aligned}$$
(5.109)

if and only if \(\phi =\phi ({x})\) is an eigenfunction corresponding to an eigenvalue \(\lambda \) in the eigenproblem (5.106) and (5.107).

Solution.

Let \(\phi \in { U}\) and \(\phi +\omega \ \tilde{\phi }\in { U}\). Then

$$\begin{aligned} \tilde{\phi }^{\prime }(0)-\alpha \tilde{\phi }(0)=0,\quad \tilde{\phi }^{\prime }({ L})+\beta \tilde{\phi }({ L})=0 \end{aligned}$$
(5.110)

and

$$\begin{aligned} \pi \{\phi +\omega \tilde{\phi }\}&=\frac{1}{2}\int \limits _{0}^{ L}[{ E}(\phi ^{\prime }+\omega \tilde{\phi }^{\prime })^{2}-\lambda (\phi +\omega \tilde{\phi })^{2}]{ dx}\nonumber \\&\quad +\frac{1}{2}\alpha { E}(0)[\phi (0)+\omega \tilde{\phi }(0)]^{2}+\frac{1}{2}\beta { E(L)}[\phi ({ L})+\omega \tilde{\phi }({ L})]^{2}\quad \end{aligned}$$
(5.111)

Hence, we obtain

$$\begin{aligned} \delta \pi \{\phi \}&= \left. \frac{ d}{{ d}\omega }\pi \{\phi +\omega \tilde{\phi }\}\right| _{\omega \,=\,0} \nonumber \\&=\int \limits _{0}^{ L}[{ E}\phi ^{\prime }\tilde{\phi }^{\prime }-\lambda \phi \tilde{\phi }]\ { dx} +\alpha { E}(0)\,\phi (0)\,\tilde{\phi }(0)+\beta { E}({ L})\,\phi ({ L})\,\tilde{\phi }({ L}) \end{aligned}$$
(5.112)

Since

$$\begin{aligned} \int \limits _{0}^{ L}{ E}\phi ^{\prime }\tilde{\phi }^{\prime }{ dx}=\left. { E}\phi ^{\prime }\tilde{\phi }\right| _{{x}\,=\,0}^{{x}\,=\,{ L}}-\int \limits _{0}^{ L}({ E}\phi ^{\prime })^{\prime }\tilde{\phi }{ dx} \end{aligned}$$
(5.113)

therefore, Eq. (5.112) takes the form

$$\begin{aligned} \delta \pi \{\phi \}&=-\int \limits _{0}^{ L}[({ E}\phi ^{\prime })^{\prime }+\lambda \phi ]\tilde{\phi }{ dx} -{ E}(0)[\phi ^{\prime }(0)-\alpha \phi (0)]\ \tilde{\phi }(0)\nonumber \\&\quad +{ E(L)}[\phi ^{\prime }({ L})+\beta \phi ({ L})]\tilde{\phi }({ L}) \end{aligned}$$
(5.114)

Now, if \(\phi =\phi ({ x})\) is an eigenfunction corresponding to an eigenvalue \(\lambda \) in the problem (5.106)–(5.107), then by virtue of (5.114) \(\delta \pi \{\phi \}=0\) over \(U\). Conversely, if \(\delta \pi \{\phi \}=0\) then selecting \(\tilde{\phi }=\tilde{\phi }({ x})\) to be a smooth function on [0, \(L\)] such that \(\tilde{\phi }(0)=\tilde{\phi }({ L})=0\), and using the Fundamental Lemma of calculus of variations, we obtain

$$\begin{aligned} ({ E}\phi ^{\prime })^{\prime }+\lambda \phi =0\quad \mathrm{{on}}\quad [0,{ L}] \end{aligned}$$
(5.115)

Next, if \(\delta \pi \{\phi \}=0\) then selecting \(\tilde{\phi }=\tilde{\phi }({ x})\) to be a smooth function on [0, \(L\)] and such that \(\tilde{\phi }({ L})=0\), and \(\tilde{\phi }(0)\ne 0\), by virtue of (5.115), we obtain

$$\begin{aligned} {E}(0)[\phi ^{\prime }(0)-\alpha \phi (0)]\tilde{\phi }(0)=0 \end{aligned}$$
(5.116)

Since

$$\begin{aligned} { E}(0)>0 \end{aligned}$$
(5.117)

Equation (5.116) implies that \(\phi =\phi ({ x})\) satisfies the boundary condition

$$\begin{aligned} \phi ^{\prime }(0)-\alpha \phi (0)=0 \end{aligned}$$
(5.118)

Finally, if \(\delta \pi \{\phi \}=0\) then selecting \(\tilde{\phi }\) to be a smooth function on [0, \(L\)] and such that \(\tilde{\phi }({ L})\ne 0\), by virtue of (5.115) and (5.118), we obtain

$$\begin{aligned} {E(L)} [\phi ^{\prime }({ L})+\beta \phi ({ L})]\tilde{\phi }({ L})=0 \end{aligned}$$
(5.119)

Since \({ E(L)}>0\), Eq. (5.119) implies that

$$\begin{aligned} \phi ^{\prime }({ L})+\beta \phi ({ L})=0 \end{aligned}$$
(5.120)

This shows that if Eq. (5.110) holds true then \((\phi ,\lambda )\) is an eigenpair for the problem (5.106)–(5.107). This completes a solution to Problem 5.6.

Problem 5.7.

Free lateral vibrations of an elastic bar clamped at the end \({x}=0\) and supported by a spring of stiffness \(k\) at the end \({x}={L}\) are defined as solutions of the form

$$\begin{aligned} {u(x,t)}=\phi ({x})\,\sin (\omega {t}+\gamma ) \end{aligned}$$
(5.121)

to the equation [see Problem 5.1, Eq. (5.127) in which \({u}_2 ={u}\), and \({F}=0\)]

$$\begin{aligned} \frac{\partial ^2}{\partial {x}^2}\left( {{EI}\frac{\partial ^2{u}}{\partial {x}^2}} \right) + \ \rho \frac{\partial ^2{u}}{\partial {t}^2}=0\quad \mathrm{on}\quad [0,{L}]\times [0,\infty ) \end{aligned}$$
(5.122)

subject to the boundary conditions

$$\begin{aligned} {u(0,t)={u}^{\prime }(0,t)=0}\quad \mathrm{on}\quad [0,\infty ) \end{aligned}$$
(5.123)
$$\begin{aligned} {u}^{\prime \prime }({L,t})=0,\quad ({EI}\,{u}^{\prime \prime }{)}^{\prime }({L,t})-{k\,u(L,t)}=0\quad \mathrm{on}\quad [0,\infty ) \end{aligned}$$
(5.124)

Let \(\rho =\mathrm{const}\), and \(\lambda =\rho \,\omega ^2\). Then the associated eigenproblem reads

$$\begin{aligned} ({EI}\,{\phi }^{\prime \prime }{)}^{\prime \prime }-\lambda \phi =0\quad \mathrm{on}\quad [0,{L}] \end{aligned}$$
(5.125)
$$\begin{aligned} \phi (0)={\phi }^{\prime }(0)=0 \end{aligned}$$
(5.126)
$$\begin{aligned} {\phi }^{\prime \prime }({L})=0,\quad ({EI}{\phi }^{\prime \prime }{)}^{\prime }({L})-{k}\phi ({L})=0 \end{aligned}$$
(5.127)

Let \(V\) denote the set of functions \(\phi =\phi ({x})\) on [0, \(L\)] that satisfy the boundary conditions (5.126) and (5.127). Define a functional \(\pi \{.\}\;\mathrm{on}\;{V}\) by

$$\begin{aligned} \pi \{\phi \}=\frac{1}{2}\int \limits _0^{L} {{EI}\,({\phi }^{\prime \prime })^2{dx}+\frac{1}{2}{k}[\phi ({ L})]^2-\frac{\lambda }{2}\int \limits _0^{L} {\phi ^2{dx}} } \end{aligned}$$
(5.128)

Show that

$$\begin{aligned} \delta \,\pi \{\phi \}=0\quad \mathrm{over}\quad {V} \end{aligned}$$
(5.129)

if and only if \((\lambda , \phi )\) is a solution to the eigenproblem (5.125) through (5.127).

Soution.

Let \(\phi \in { V}\) and \(\phi +\omega \tilde{\phi }\in { V}\). Then

$$\begin{aligned} \tilde{\phi }(0)=0,\quad \tilde{\phi }^{\prime }(0)=0 \end{aligned}$$
(5.130)

Computing the first variation of the functional \(\pi \{\phi \}\) given by (5.128), we obtain

$$\begin{aligned} \delta \pi \{\phi \}&=\frac{ d}{{ d}\omega }\pi \left. \{\phi +\omega \tilde{\phi }\}\right| _{\omega \,=\,0} \nonumber \\&=\int \limits _{0}^{ L}({ EI}\ \phi ^{\prime \prime }\tilde{\phi }^{\prime \prime }-\lambda \phi \tilde{\phi }){ dx} +{ k}\phi ({ L})\ \tilde{\phi }({ L}) \end{aligned}$$
(5.131)

Since

$$\begin{aligned} \int \limits _{0}^{ L}{ EI}\ \phi ^{\prime \prime }\tilde{\phi }^{\prime \prime }{ dx}=({ EI}\ \phi ^{\prime \prime })\tilde{\phi }^{\prime }\bigg |_{{ x}\,=\,0}^{{ x}\,=\,{ L}} -({ EI}\ \phi ^{\prime \prime })^{\prime }\tilde{\phi }\bigg |_{{ x}\,=\,0}^{{ x}\,=\,{ L}}+\int \limits _{0}^{ L}({ EI}\ \phi ^{\prime \prime })^{\prime \prime }\ \tilde{\phi }{ dx} \end{aligned}$$
(5.132)

therefore, using (5.130) we reduce (5.131) into the form

$$\begin{aligned} \delta \pi \{\phi \}=\int \limits _{0}^{ L}[({ EI}\ \phi ^{\prime \prime })^{\prime \prime }-\lambda \phi ]\tilde{\phi }{ dx} +({ EI}\ \phi ^{\prime \prime })({ L})\tilde{\phi }^{\prime }({ L}) -[({ EI}\ \phi ^{\prime \prime })^{\prime }({ L})-{ k}\phi ({ L})]\tilde{\phi }({ L}) \end{aligned}$$
(5.133)

Now, if \((\lambda ,\phi )\) is a solution to the eigenproblem (5.125)–(5.127), then \(\delta \pi \{\phi \}=0\). Conversely, if \(\delta \pi \{\phi \}=0\) over \(V\), then selecting \(\tilde{\phi }\) to be an arbitrary smooth function on [0, \(L\)] such that \(\tilde{\phi }({ x})\not \equiv 0\) for \({ x}\in (0,{ L}),\tilde{\phi }^{\prime }({ L})=0,\tilde{\phi }({ L})=0\), we obtain

$$\begin{aligned} \int \limits _{0}^{ L}[({ EI}\ \phi ^{\prime \prime })^{\prime \prime }-\lambda \phi ]\ \tilde{\phi }\,{ dx}=0 \end{aligned}$$
(5.134)

Equation (5.134) together with the Fundamental Lemma of calculus of variations implies

$$\begin{aligned} ({ EI}\ \phi ^{\prime \prime })^{\prime \prime }-\lambda \phi =0\quad \mathrm{on}\quad [0,{ L}] \end{aligned}$$
(5.135)

Next, by selecting \(\tilde{\phi }\) on [0, \(L\)] in such a way that

$$\begin{aligned} \tilde{\phi }^{\prime }({ L})\ne 0,\quad \tilde{\phi }({ L})=0 \end{aligned}$$
(5.136)

we find that the condition \(\delta \pi \{\phi \}=0\) and Eq. (5.135) imply that

$$\begin{aligned} ({ EI}\ \phi ^{\prime \prime })({ L})=0 \end{aligned}$$
(5.137)

Since

$$\begin{aligned} { E(L)}>0,\quad { I(L)}>0 \end{aligned}$$
(5.138)

we obtain

$$\begin{aligned} \phi ^{\prime \prime }({ L})=0 \end{aligned}$$
(5.139)

Finally, by selecting \(\tilde{\phi }\) on [0, \(L\)] in such a way that

$$\begin{aligned} \tilde{\phi }({ L})\ne 0 \end{aligned}$$
(5.140)

we conclude that the condition \(\delta \pi \{\phi \}=0\) together with Eqs. (5.135), and (5.139) lead to the boundary condition

$$\begin{aligned} ({ EI}\ \phi ^{\prime \prime })^{\prime }({ L})-{ k}\,\phi ({ L})=0 \end{aligned}$$
(5.141)

This completes a solution to Problem 5.7.

Problem 5.8.

Show that the eigenvalues \(\lambda _\mathrm{i} \) and the eigenfunctions \(\phi _{i} =\phi _{i} ({x})\) for the longitudinal vibrations of a uniform elastic bar having one end clamped and the other end free are given by the relations

$$\begin{aligned} \omega _\mathrm{i} =\sqrt{\frac{\lambda _\mathrm{i} }{\rho }} =\frac{(2{i}-1)}{2{L}}\sqrt{\frac{{E}}{\rho }} \end{aligned}$$
$$\begin{aligned} \phi _\mathrm{i} ({x})=\sin \frac{(2{i}-1)\pi \,{x}}{2{L}},\quad {i}=1,2,3,\ldots , 0\le {x}\le {L} \end{aligned}$$

(see Problem 5.6).

Solution.

For an elastic bar that is clamped at \({ x}=0\) and free at \({ x}={ L}\) the eigenproblem reads

$$\begin{aligned} { E}\phi ^{\prime \prime }({ x})+\lambda \phi ({ x})=0\qquad x\in [0,{ L}] \end{aligned}$$
(5.142)
$$\begin{aligned} \phi (0)=0,\quad \phi ^{\prime }({ L})=0 \end{aligned}$$
(5.143)

where

$$\begin{aligned} \lambda =\omega ^{2}\rho \end{aligned}$$
(5.144)

There is an infinite sequence of eigensolutions \((\lambda _{ i},\phi _{ i})\) to the problem (5.142)–(5.143) of the form

$$\begin{aligned} \lambda _{ i}&=\frac{(2{ i}-1)^{2}\pi ^{2}}{4{ L}^{2}}{ E} \end{aligned}$$
(5.145)
$$\begin{aligned} \phi _{ i}({ x})&=\sin \frac{(2{ i}-1)\pi { x}}{2{ L}},\quad { i}=1,2,3,\ldots \end{aligned}$$
(5.146)

This can be shown by substituting (5.145) and (5.146) into (5.142), and by showing that \(\phi _{ i}({ x})\) satisfies (5.143). By combining (5.144) and (5.145) we obtain

$$\begin{aligned} \omega _{ i}\equiv \sqrt{\frac{\lambda _{ i}}{\rho }}=\frac{(2{ i}-1)\pi }{2{ L}}\sqrt{\frac{ E}{\rho }} \end{aligned}$$
(5.147)

This completes a solution to Problem 5.8.

Problem 5.9.

Show that the eigenvalues \(\lambda _{i}\) and the eigenfunctions \(\phi _{i} =\phi _{i} ({x})\) for the lateral vibrations of a uniform, simply supported elastic beam are given by the relations

$$\begin{aligned} \omega _{i}&=\sqrt{\frac{\lambda _{i} }{\rho }} =\frac{\pi ^2{i}^2}{{L}^2}\sqrt{\frac{{EI}}{\rho }}\\ \phi _{i} ({x})&=\sin \frac{{i}\,\pi \,{x}}{{L}},\quad {i}=1,2,3\ldots ,0\le {x}\le {L} \end{aligned}$$

(see Problem 5.1).

Solution.

For a uniform, simply supported beam with the lateral vibrations, the eigenproblem takes the form

$$\begin{aligned} {EI}\phi ^{(4)}-\lambda \phi&=0\quad \mathrm{on}\quad [0, { L}] \end{aligned}$$
(5.148)
$$\begin{aligned} \phi (0)=\phi ^{\prime \prime }(0)&=0,\quad \phi ({ L})=\phi ^{\prime \prime }({ L})=0 \end{aligned}$$
(5.149)

where

$$\begin{aligned} \lambda =\omega ^{2}{ \rho } \end{aligned}$$
(5.150)

There is an infinite sequence of eigensolutions \((\lambda _{ i},\phi _{ i})\) to the problem (5.148)–(5.149) of the form

$$\begin{aligned} \lambda _{ i}&={ EI}\left( \frac{{ i}\pi }{{ L}}\right) ^{4} \end{aligned}$$
(5.151)
$$\begin{aligned} \phi _{ i}({ x})&=\sin \frac{{ i}\pi { x}}{{ L}}\quad { i}=1,2,3,\ldots \end{aligned}$$
(5.152)

To prove that \((\lambda _{ i},\phi _{ i})\) given by (5.151)–(5.152) satisfies Eqs. (5.148)–(5.149), we note that

$$\begin{aligned} \phi _{ i}^{\prime \prime }({ x})=-\left( \frac{{ i}\pi }{{ L}}\right) ^{2}\phi _{ i}({ x}) \end{aligned}$$
(5.153)

and

$$\begin{aligned} \phi _{ i}^{(4)}({ x})=\left( \frac{{ i}\pi }{{ L}}\right) ^{4}\phi _{ i}({ x}) \end{aligned}$$
(5.154)

Substituting (5.151) and (5.152) into (5.148) and using (5.154) we find that \(\phi _{ i}=\phi _{ i}({ x})\) satisfies Eq. (5.148) on [0, \(L\)]. Also, it follows from Eqs. (5.152) and (5.153) that the boundary conditions (5.149) are satisfied; and Eqs. (5.150) and (5.151) imply that

$$\begin{aligned} \omega _{ i}=\sqrt{\frac{\lambda { i}}{\rho }}=\frac{\pi ^{2}{ i}^{2}}{{ L}^{2}}\sqrt{\frac{{ EI}}{\rho }} \end{aligned}$$
(5.155)

These steps complete a solution to Problem 5.9.

Problem 5.10.

Show that the eigenvalues \(\lambda _{mn} \) and the eigenfunctions \(\phi _{mn} =\phi _{mn} ({x})\) for the transversal vibrations of a rectangular elastic membrane: \(0\le {x}_1 \le {a}_1, \;0\le {x}_2 \le {a}_2 \), that is clamped on its boundary, are given by

$$\begin{aligned} \omega _{mn} =\sqrt{\frac{\lambda _{mn}}{\hat{\rho }}} =\pi \sqrt{\frac{\hat{T}}{\hat{\rho }}\left( {\frac{{m}^2}{{a}_1^2 }+\frac{{n}^2}{{a}_2^2 }} \right) } \end{aligned}$$
$$\begin{aligned} \begin{array}{c} \phi _{mn} ({x}_1, {x}_2 )=\sin \,\dfrac{{m}\,\pi \,{x}_1 }{{a}_1 }\,\sin \dfrac{{n}\,\pi \,{x}_2 }{{a}_2 }, \\ {m,n}=1,2,3,\ldots ,0\le {x}_1 \le {a}_1, \,\,0\le {x}_2 \le {a}_2 \\ \end{array} \end{aligned}$$

(See Problem 5.2).

Solution.

Let \({ C}_{0}\) denote the rectangular region

$$\begin{aligned} 0<{ x}_{1}<{ a}_{1},\quad 0<{ x}_{2}<{ a}_{2} \end{aligned}$$
(5.156)

and let \(\partial { C}_{0}\) be its boundary. Then the associated eigenproblem reads. Find an eigenpair \((\lambda , \phi )\) such that

$$\begin{aligned} \hat{{ T}}\ \nabla ^{2}\phi +\lambda \phi =0\quad \mathrm{on} \ { C}_{0} \end{aligned}$$
(5.157)

and

$$\begin{aligned} \phi =0\quad \mathrm{on}\ \partial { C}_{0} \end{aligned}$$
(5.158)

where

$$\begin{aligned} \lambda =\omega ^{2}\hat{\rho } \end{aligned}$$
(5.159)

There is an infinite number of eigenpairs \((\lambda _{ mn}, \phi _{ mn}),\ { m, n}=1, 2, 3, \ldots \) that satisfy Eqs. (5.157) and (5.158), and they are given by Equation

$$\begin{aligned} \lambda _{ mn}&=\pi ^{2}\hat{{ T}}\left( \frac{{ m}^{2}}{{ a}_{1}^{2}}+\frac{{ n}^{2}}{{ a}_{2}^{2}}\right) \end{aligned}$$
(5.160)
$$\begin{aligned} \phi _{ mn}({ x_{1},x}_{2})&=\sin \left( \frac{{ m}\pi { x}_{1}}{{ a}_{1}}\right) \sin \left( \frac{{ n}\pi { x}_{2}}{{ a}_{2}}\right) \end{aligned}$$
(5.161)

This can be proved by substituting (5.152) and (5.153) into (5.149) and (5.150).

Also, the eigenvalues \(\lambda _{ mn}\) generate the eigenfrequencies \(\omega _{ mn}\) by the formulas

$$\begin{aligned} \omega _{ mn}=\sqrt{\frac{\lambda _{ mn}}{\hat{\rho }}}=\pi \sqrt{\frac{\hat{{ T}}}{\hat{\rho }}}\sqrt{\left( \frac{{ m}^{2}}{{ a}_{1}^{2}}\right) +\left( \frac{{ n}^{2}}{{ a}_{2}^{2}}\right) } \end{aligned}$$
(5.162)

This completes a solution to Problem 5.10.

Problem 5.11.

Show that the eigenvalues \(\lambda _{mn} \) and the eigenfunctions \(\phi _{mn} =\phi _{mn} ({x}_1, {x}_2 )\) for the transversal vibrations of a thin elastic rectangular plate: \({0\le {x}_1 \le {a}_1, \;0\le {x}_2 \le {a}_2}\), that is simply supported on its boundary are given by the relations

$$\begin{aligned} \omega _{mn} =\sqrt{\frac{\lambda _{mn} }{\hat{\rho }}} =\pi ^2\left( {\frac{{m}^2}{{a}_1^2 }+\frac{{n}^2}{{a}_2^2 }} \right) \sqrt{\frac{{D}}{\hat{\rho }}} \end{aligned}$$
$$\begin{aligned} \begin{array}{c} \phi _{{mn}} ({x}_1, {x}_2 )=\sin \,\dfrac{{m}\,\pi \,{x}_1}{{a}_1 }\,\sin \dfrac{{n}\,\pi \,{x}_2 }{{a}_2 }, \\ {m,n}=1,2,3,\ldots ,0\le {x}_1 \le {a}_1, \,\,0\le {x}_2 \le {a}_2 \\ \end{array} \end{aligned}$$

(See Problem 5.4).

Solution.

The eigenproblem associated with the transversal vibrations of a thin elastic rectangular plate that is simply supported on its boundary, reads [see Eq. (5.85) of Problem 5.4]

$$\begin{aligned} { D}\ \nabla ^{2}\nabla ^{2}\ \phi -\lambda \phi =0\quad \mathrm{on}\ { C}_{0} \end{aligned}$$
(5.163)
$$\begin{aligned} \phi =\nabla ^{2}\phi =0\quad \mathrm{on}\ \partial { C}_{0} \end{aligned}$$
(5.164)

where

$$\begin{aligned} \lambda =\omega ^{2}\hat{\rho } \end{aligned}$$
(5.165)

and \({ C}_{0}\) and \(\partial { C}_{0}\) are the same as in Problem 5.10.

There are an infinite number of eigenpairs \((\lambda _{ mn},\phi _{ mn})\) that satisfy Eqs. (5.163) and (5.164), and the eigenpairs are given by

$$\begin{aligned} \lambda _{ mn}&=\pi ^{4}{ D}\left( \frac{{ m}^{2}}{{ a}_{1}^{2}}+\frac{{ n}^{2}}{{ a}_{2}^{2}}\right) ^{2} \end{aligned}$$
(5.166)
$$\begin{aligned} \phi _{ mn}({ x_{1},x}_{2})&=\sin \left( \frac{{ m}\pi { x}_{1}}{{ a}_{1}}\right) \sin \left( \frac{{ n}\pi { x}_{2}}{{ a}_{2}}\right) \quad { m,n}=1,2,3,\ldots \end{aligned}$$
(5.167)

This is proved by substituting (5.166) and (5.167) into (5.163) and (5.164).

Also, by using (5.165) the eigenfrequencies \(\omega _{mn}\) are obtained

$$\begin{aligned} \omega _{mn}=\sqrt{\frac{\lambda _{mn}}{\hat{\rho }}}=\pi ^{2}\left( \frac{{ m}^{2}}{{ a}_{1}^{2}}+\frac{{ n}^{2}}{{ a}_{2}^{2}}\right) \sqrt{\frac{{ D}}{\hat{\rho }}} \end{aligned}$$
(5.168)

This completes a solution to Problem 5.11.