Vibration of plates

The vibration of plates is a special case of the more general problem of mechanical vibrations. The equations governing the motion of plates are simpler than those for general three-dimensional objects because one of the dimensions of a plate is much smaller than the other two. This suggests that a two-dimensional plate theory will give an excellent approximation to the actual three-dimensional motion of a plate-like object, and indeed that is found to be true.[1]

There are several theories that have been developed to describe the motion of plates. The most commonly used are the Kirchhoff-Love theory[2] and the Mindlin-Reissner theory. Solutions to the governing equations predicted by these theories can give us insight into the behavior of plate-like objects both under free and forced conditions. This includes the propagation of waves and the study of standing waves and vibration modes in plates.

Contents

Kirchhoff-Love plates

The governing equations for the dynamics of a Kirchhoff-Love plate are


   \begin{align}
     N_{\alpha\beta,\beta} & = J_1~\ddot{u}_\alpha \\
     M_{\alpha\beta,\alpha\beta} - q(x,t) & = J_1~\ddot{w} - J_3~\ddot{w}_{,\alpha\alpha}
   \end{align}

where u_\alpha are the in-plane displacements of the mid-surface of the plate, w is the transverse (out-of-plane) displacement of the mid-surface of the plate, q is an applied transverse load, and the resultant forces and moments are defined as


   N_{\alpha\beta}�:= \int_{-h}^h \sigma_{\alpha\beta}~dx_3 \quad \text{and} \quad
   M_{\alpha\beta}�:= \int_{-h}^h x_3~\sigma_{\alpha\beta}~dx_3 \,.

Note that the thickness of the plate is 2h and that the resultants are defined as weighted averages of the in-plane stresses \sigma_{\alpha\beta}. The derivatives in the governing equations are defined as


   \dot{u}_i�:= \frac{\partial u_i}{\partial t} ~;~~ \ddot{u}_i�:= \frac{\partial^2 u_i}{\partial t^2} ~;~~
   u_{i,\alpha}�:= \frac{\partial u_i}{\partial x_\alpha} ~;~~ u_{i,\alpha\beta}�:= \frac{\partial^2 u_i}{\partial x_\alpha \partial x_\beta}

where the Latin indices go from 1 to 3 while the Greek indices go from 1 to 2. Summation over repeated indices is implied. The x_3 coordinates is out-of-plane while the coordinates x_1 and x_2 are in plane. For a uniformly thick plate of thickness 2h and homogeneous mass density \rho


   J_1�:= \int_{-h}^h \rho~dx_3 = 2\rho h \quad \text{and} \quad
   J_3�:= \int_{-h}^h x_3^2~\rho~dx_3 = \frac{2}{3}\rho h^3 \,.

Isotropic plates

For an isotropic and homogeneous plate, the stress-strain relations are


   \begin{bmatrix}\sigma_{11} \\ \sigma_{22} \\ \sigma_{12} \end{bmatrix}
   = \cfrac{E}{1-\nu^2}
   \begin{bmatrix} 1 & \nu & 0 \\
                   \nu & 1 & 0 \\
                   0 & 0 & 1-\nu \end{bmatrix}
    \begin{bmatrix}\varepsilon_{11} \\ \varepsilon_{22} \\ \varepsilon_{12} \end{bmatrix} \,.

where \varepsilon_{\alpha\beta} are the in-plane strains. The strain-displacement relations for Kirchhoff-Love plates are


  \varepsilon_{\alpha\beta} = \frac{1}{2}(u_{\alpha,\beta}%2Bu_{\beta,\alpha})
      - x_3\,w_{,\alpha\beta} \,.

Therefore, the resultant moments corresponding to these stresses are


   \begin{bmatrix}M_{11} \\ M_{22} \\ M_{12} \end{bmatrix} = 
   -\cfrac{2h^3E}{3(1-\nu^2)}~\begin{bmatrix} 1 & \nu & 0 \\
                   \nu & 1 & 0 \\
                   0 & 0 & 1-\nu \end{bmatrix}
   \begin{bmatrix} w_{,11} \\ w_{,22} \\ w_{,12} \end{bmatrix}

If we ignore the in-plane displacements u_{\alpha\beta}, the governing equations reduce to


   D\nabla^2\nabla^2 w = -q(x,t) - 2\rho h\ddot{w} \,.

Free vibrations

For free vibrations, the governing equation of an isotropic plate is


   D\nabla^2\nabla^2 w = - 2\rho h\ddot{w} \,.

Circular plates

For freely vibrating circular plates,  w = w(r,t), and the Laplacian in cylindrical coordinates has the form


  \nabla^2 w \equiv \frac{1}{r}\frac{\partial }{\partial r}\left(r \frac{\partial w}{\partial r}\right) \,.

Therefore, the governing equation for free vibrations of a circular plate of thickness 2h is


   \frac{1}{r}\frac{\partial }{\partial r}\left[r \frac{\partial }{\partial r}\left\{\frac{1}{r}\frac{\partial }{\partial r}\left(r \frac{\partial w}{\partial r}\right)\right\}\right] = -\frac{2\rho h}{D}\frac{\partial^2 w}{\partial t^2}\,.

Expanded out,


  \frac{\partial^4 w}{\partial r^4} %2B \frac{2}{r} \frac{\partial^3 w}{\partial r^3} - \frac{1}{r^2} \frac{\partial^2 w}{\partial r^2} %2B \frac{1}{r^3} \frac{\partial w}{\partial r} = -\frac{2\rho h}{D}\frac{\partial^2 w}{\partial t^2}\,.

To solve this equation we use the idea of separation of variables and assume a solution of the form


   w(r,t) = W(r)F(t) \,.

Plugging this assumed solution into the governing equation gives us


  \frac{1}{\beta W}\left[\frac{d^4 W}{dr^4} %2B \frac{2}{r}\frac{d^3 W}{dr^3} - \frac{1}{r^2}\frac{d^2W}{dr^2}
   %2B \frac{1}{r^3} \frac{d W}{dr}\right] = -\frac{1}{F}\cfrac{d^2 F}{d t^2} = \omega^2

where \omega^2 is a constant and \beta�:= 2\rho h/D. The solution of the right hand equation is


   F(t) = \text{Re}[ A e^{i\omega t} %2B B e^{-i\omega t}] \,.

The left hand side equation can be written as


  \frac{d^4 W}{dr^4} %2B \frac{2}{r}\frac{d^3 W}{dr^3} - \frac{1}{r^2}\frac{d^2W}{dr^2} 
    %2B \frac{1}{r^3} \cfrac{d W}{d r} = \lambda^4 W

where \lambda^4�:= \beta\omega^2. The general solution of this eigenvalue problem that is appropriate for plates has the form


   W(r) = C_1 J_0(\lambda r) %2B C_2 I_0(\lambda r)

where J_0 is the order 0 Bessel function of the first kind and I_0 is the order 0 modified Bessel function of the first kind. The constants C_1 and C_2 are determined from the boundary conditions. For a plate of radius a with a clamped circumference, the boundary conditions are


   W(r) = 0 \quad \text{and} \quad \cfrac{d W}{d r} = 0 \quad \text{at} \quad r = a \,.

From these boundary conditions we find that


   J_0(\lambda a)I_1(\lambda a) %2B I_0(\lambda a)J_1(\lambda a) = 0 \,.

We can solve this equation for \lambda_n (and there are an infinite number of roots) and from that find the modal frequencies \omega_n = \lambda_n^2/\beta. We can also express the displacement in the form


   w(r,t) = \sum_{n=1}^\infty C_n\left[J_0(\lambda_n r) - \frac{J_0(\lambda_n a)}{I_0(\lambda_n a)}I_0(\lambda_n r)\right]
    [A_n e^{i\omega_n t} %2B B_n e^{-i\omega_n t}] \,.

For a given frequency \omega_n the first term inside the sum in the above equation gives the mode shape. We can find the value of C_1 using the appropriate boundary condition at r = 0 and the coefficients A_n and B_n from the initial conditions by taking advantage of the orthogonality of Fourier components.

Rectangular plates

Consider a rectangular plate which has dimensions a\times b in the (x_1,x_2)-plane and thickness 2h in the x_3-direction. We seek to find the free vibration modes of the plate.

Assume a displacement field of the form


   w(x_1,x_2,t) = W(x_1,x_2) F(t) \,.

Then,


   \nabla^2\nabla^2 w = w_{,1111} %2B 2w_{,1212} %2B w_{,2222}
     = \left[\frac{\partial^4 W}{\partial x_1^4} %2B 2\frac{\partial^4 W}{\partial x_1^2 \partial x_2^2} %2B \frac{\partial^4W}{\partial x_2^4}\right] F(t)

and


   \ddot{w} = W(x_1,x_2)\frac{d^2F}{dt^2} \,.

Plugging these into the governing equation gives


   \frac{D}{2\rho h W}\left[\frac{\partial^4 W}{\partial x_1^4} %2B 2\frac{\partial^4 W}{\partial x_1^2 \partial x_2^2} %2B \frac{\partial^4W}{\partial x_2^4}\right]  
   = -\frac{1}{F}\frac{d^2F}{dt^2} = \omega^2

where \omega^2 is a constant because the left hand side is independent of t while the right hand side is independent of x_1,x_2. From the right hand side, we then have


   F(t) = A e^{i\omega t} %2B B e^{-i\omega t} \,.

From the left hand side,


   \frac{\partial^4 W}{\partial x_1^4} %2B 2\frac{\partial^4 W}{\partial x_1^2 \partial x_2^2} %2B \frac{\partial^4W}{\partial x_2^4}
   = \frac{2\rho h \omega^2}{D} W =: \lambda^4 W

where


   \lambda^2 = \omega\sqrt{\frac{2\rho h}{D}} \,.

Since the above equation is a biharmonic eigenvalue problem, we look for Fourier expansion solutions of the form


   W_{mn}(x_1,x_2) = \sin\frac{m\pi x_1}{a}\sin\frac{n\pi x_2}{b} \,.

We can check and see that this solution satisfies the boundary conditions for a freely vibrating rectangular plate with simply supported edges:


  \begin{align}
    w(x_1,x_2,t) = 0 & \quad \text{at}\quad x_1 = 0, a \quad \text{and} \quad x_2 = 0, b \\
    M_{11} = D\left(\frac{\partial^2 w}{\partial x_1^2} %2B \nu\frac{\partial^2 w}{\partial x_2^2}\right) = 0
      & \quad \text{at}\quad x_1 = 0, a \\
    M_{22} = D\left(\frac{\partial^2 w}{\partial x_2^2} %2B \nu\frac{\partial^2 w}{\partial x_1^2}\right) = 0
      & \quad \text{at}\quad x_2 = 0, b \,.
  \end{align}

Plugging the solution into the biharmonic equation gives us


   \lambda^2 = \pi^2\left(\frac{m^2}{a^2} %2B \frac{n^2}{b^2}\right) \,.

Comparison with the previous expression for \lambda^2 indicates that we can have an infinite number of solutions with


  \omega_{mn} = \sqrt{\frac{D\pi^4}{2\rho h}}\left(\frac{m^2}{a^2} %2B \frac{n^2}{b^2}\right) \,.

Therefore the general solution for the plate equation is


  w(x_1,x_2,t) = \sum_{m=1}^\infty \sum_{n=1}^\infty \sin\frac{m\pi x_1}{a}\sin\frac{n\pi x_2}{b}
    \left( A_{mn} e^{i\omega_{mn} t} %2B B_{mn} e^{-i\omega_{mn} t}\right) \,.

To find the values of A_{mn} and B_{mn} we use initial conditions and the orthogonality of Fourier components. For example, if


   w(x_1,x_2,0) = \varphi(x_1,x_2) \quad \text{on} \quad x_1 \in [0,a] \quad \text{and} \quad
   \frac{\partial w}{\partial t}(x_1,x_2,0) = \psi(x_1,x_2)\quad \text{on} \quad x_2 \in [0,b]

we get,


  \begin{align}
    A_{mn} & = \frac{4}{ab}\int_0^a \int_0^b \varphi(x_1,x_2) 
                 \sin\frac{m\pi x_1}{a}\sin\frac{n\pi x_2}{b} dx_1 dx_2 \\
    B_{mn} & = \frac{4}{ab\omega_{mn}}\int_0^a \int_0^b \psi(x_1,x_2) 
                 \sin\frac{m\pi x_1}{a}\sin\frac{n\pi x_2}{b}  dx_1 dx_2\,.
  \end{align}

References

  1. ^ Reddy, J. N., 2007, Theory and analysis of elastic plates and shells, CRC Press, Taylor and Francis.
  2. ^ A. E. H. Love, On the small free vibrations and deformations of elastic shells, Philosophical trans. of the Royal Society (London), 1888, Vol. série A, N° 17 p. 491–549.

See also