This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Variational setting for cracked beams and shallow arches

Semion Gutman, Junhong Ha and Sudeok Shon 1 Department of Mathematics, University of Oklahoma, Norman, Oklahoma 73019, USA, e-mail: [email protected] 1School of Liberal Arts, Korea University of Technology and Education, Cheonan 31253, South Korea, e-mail: [email protected] 3Department of Architectural Engineering, Korea University of Technology and Education, Cheonan 31253, South Korea, e-mail: [email protected]
Abstract.

We develop a rigorous mathematical framework for the weak formulation of cracked beams and shallow arches problems. First, we discuss the crack modeling by means of massless rotational springs. Then we introduce Hilbert spaces, which are sufficiently wide to accommodate such representations. Our main result is the introduction of a specially designed linear operator that ”absorbs” the boundary conditions at the cracks.

We also provide mathematical justification and derivation of the Modified Shifrin’s method for an efficient computation of the eigenvalues and the eigenfunctions for cracked beams.

Key words and phrases:
Shallow arch, beam, cracks, eigenvalues and eigenfunctions
2010 Mathematics Subject Classification:
47J35, 35Q74, 35D30, 70G75

1. Introduction

Detection of cracks is an important engineering problem. It requires a rigorous mathematical framework for modeling the motion of cracked beams and shallow arches. The main goal of this paper is to develop a variational setting for such a framework. We also present an elegant and efficient method (a modification of the Shifrin’s method) for the computation of the eigenvalues and the eigenfunctions for the cracked elements.

For a theory of cracked Bernoulli-Euler beams see [5]. A significant effort has been directed at the vibration analysis of cracked beams. Representation of a crack by a rotational spring has been proven to be accurate, and it is often used, see [2, 3] and the extensive bibliography there. Determination of the beam natural frequencies is discussed in [11, 13, 14]. S. Caddemi and his colleagues have further developed the theory using energy functions in [3]. However, no full variational setting has been presented so far, making it difficult to study evolution problems. As we have already mentioned, our work closes a gap in this development.

The theory of uniform beams and shallow arches is well developed. An early exposition can be found in [1]. More general models in the multidimensional setting, and a literature survey are presented in [6]. A review for vibrating beams is given in [10]. Motion of uniform arches and a related parameter estimation problem are studied in [8]. These results are extended to point loads in [7]. The existence of a compact, uniform attractor is established in [9].

The transverse motion of a beam or an arch is described by the function y(x,t),x[0,π],t0y(x,t),\,x\in[0,\pi],\,t\geq 0, which represents the deformation of the beam/arch measured from the xx-axis. For definiteness, the boundary conditions are of the hinged type

(1.1) y(0,t)=y′′(0,t)=0,y(π,t)=y′′(π,t)=0,t(0,T).y(0,t)=y^{\prime\prime}(0,t)=0,\quad y(\pi,t)=y^{\prime\prime}(\pi,t)=0,\quad t\in(0,T).

Other types of boundary conditions, can be treated similarly.

Crack modeling is considered in Section 2. Suppose that there are mm cracks located at 0<x1<x2<<xm<π0<x_{1}<x_{2}<\dots<x_{m}<\pi. A crack at x=xix=x_{i} is represented by a rotational spring with the flexibility θi\theta_{i}, i=1,,mi=1,\dots,m. This is expressed as

(1.2) y(xi+,t)y(xi,t)=θiy′′(xi,t),t>0,i=1,,m.y^{\prime}(x_{i}^{+},t)-y^{\prime}(x_{i}^{-},t)=\theta_{i}y^{\prime\prime}(x_{i},t),\quad t>0,\ i=1,\dots,m.

In Section 3 we introduce special Hilbert spaces VV and HH satisfying

(1.3) VHV,V\subset H\subset V^{\prime},

with continuous and dense embeddings. These spaces are broad enough to contain continuous functions with discontinuous derivatives at the joint points.

Section 4 contains our main result. We introduce the operator 𝒜:VV\mathcal{A}:V\to V^{\prime}, by

(1.4) 𝒜u,vV=i=1m+1(u′′,v′′)i+i=1m1θiJ[u](xi)J[v](xi),\langle\mathcal{A}u,v\rangle_{V}=\sum_{i=1}^{m+1}(u^{\prime\prime},v^{\prime\prime})_{i}+\sum_{i=1}^{m}\frac{1}{\theta}_{i}J[u^{\prime}](x_{i})J[v^{\prime}](x_{i}),

for any u,vVu,v\in V, where J[u](x)=u(x+)u(x)J[u^{\prime}](x)=u^{\prime}(x^{+})-u^{\prime}(x^{-}).

Then we show that the solution uu of the equation 𝒜u=f\mathcal{A}u=f in HH satisfies the joint conditions, including (1.2). Thus the operator 𝒜\mathcal{A} ”absorbs” the boundary conditions, as expected of the weak formulation of the steady state problem.

This result allows us to prove the existence of the eigenvalues and the eigenfunctions of 𝒜\mathcal{A}. An efficient Modified Shifrin’s Method for their computation is presented in Section 5.

The results in this paper form the basis for a comprehensive study of dynamic behavior of cracked beams and arches. It will be presented elsewhere.

2. Crack modeling

A crack is a disruption in the material, that has a negligible extent in the direction of the beam/arch axis, but of a non-negligible depth. It is fully described by its position along the axis, and the crack depth ratio μ^\hat{\mu}, as shown in Figure 1.

Refer to caption
Figure 1. Crack parameters.

According to [4], a crack is modeled by a massless rotational spring. The spring flexibility θ=θ(μ^)\theta=\theta(\hat{\mu}) depends on the crack depth ratio μ^\hat{\mu}, and on whether the crack is one-sided or two-sided, open or closed, and so on. The flexibility θ\theta is equal to 0 if there is no crack, and it increases with the crack depth. Explicit expressions for the functions θ(μ^)\theta(\hat{\mu}) are provided in Section 5.

Remark 2.1.

The following discussion is applicable to both arches and beams, but to avoid repetitions we will refer just to arches.

Suppose that there are mm cracks along the length of the arch, located at 0<x1<<xm<π0<x_{1}<\dots<x_{m}<\pi. For convenience, we denote x0=0x_{0}=0, and xm+1=πx_{m+1}=\pi. Consequently, the cracked arch is modeled as a collection of m+1m+1 uniform arches over the intervals li=(xi1,xi),i=1,,m+1l_{i}=(x_{i-1},x_{i}),\,i=1,\dots,m+1, as shown in Figure 2(b).

We consider only the transverse motion of the arch, so its position can be described by the function y=y(x,t)y=y(x,t), 0xπ0\leq x\leq\pi, t0t\geq 0. The boundary conditions at the cracks enforce the continuity of the displacement field yy, the bending moment y′′y^{\prime\prime}, and the shear force y′′′y^{\prime\prime\prime}. Condition y(xi+,t)y(xi,t)=θiy′′(xi+,t)y^{\prime}(x_{i}^{+},t)-y^{\prime}(x_{i}^{-},t)=\theta_{i}y^{\prime\prime}(x_{i}^{+},t) expresses the discontinuity of the arch slope at the ii-th crack, where θi=θ(μ^i)\theta_{i}=\theta(\hat{\mu}_{i}), see Figure 2(b).

Refer to caption
Figure 2. Beam or shallow arch: (a) uniform, (b) with two cracks.

To simplify the statement of the boundary conditions at the cracks, we introduce the notion of the jump J[u](x)J[u](x) of a function u=u(x)u=u(x) at any x(0,π)x\in(0,\pi), as follows

(2.1) J[u](x)=u(x+)u(x).J[u](x)=u(x^{+})-u(x^{-}).

With this notation the conditions at the cracks (joint conditions) are

(2.2) J[y](xi,t)=0,J[y′′](xi,t)=0,J[y′′′](xi,t)=0,J[y](x_{i},t)=0,\quad J[y^{\prime\prime}](x_{i},t)=0,\quad J[y^{\prime\prime\prime}](x_{i},t)=0,

and

(2.3) J[y](xi,t)=θiy′′(xi+,t),J[y^{\prime}](x_{i},t)=\theta_{i}y^{\prime\prime}(x_{i}^{+},t),

where θi=θ(μ^i)\theta_{i}=\theta(\hat{\mu}_{i}), i=1,2,,mi=1,2,\dots,m and t0t\geq 0. Note that y′′(xi+,t)=y′′(xi,t)y^{\prime\prime}(x_{i}^{+},t)=y^{\prime\prime}(x_{i}^{-},t) by (2.2).

3. Hilbert spaces

We introduce Hilbert spaces HH and VV suitable for working with cracked elements. Suppose that an arch has mm cracks at the joint points 0<x1<<xm<π0<x_{1}<\dots<x_{m}<\pi. This partition of the interval [0,π][0,\pi] is associated with m+1m+1 subintervals li=(xi1,xi),m=1,,m+1l_{i}=(x_{i-1},x_{i}),\,m=1,\dots,m+1.

Let HH be the Hilbert space

(3.1) H=i=1m+1L2(li).H=\bigoplus_{i=1}^{m+1}L^{2}(l_{i}).

Let the inner product and the norm in L2(li)L^{2}(l_{i}) be denoted by (,)i(\cdot,\cdot)_{i} and ||i|\cdot|_{i} correspondingly. The inner product and the norm in HH are defined by

(3.2) (u,v)H=i=1m+1(u,v)i,|u|H2=i=1m+1|u|i2.(u,v)_{H}=\sum_{i=1}^{m+1}(u,v)_{i},\quad|u|^{2}_{H}=\sum_{i=1}^{m+1}|u|^{2}_{i}.

Consider the Sobolev space H2(a,b)H^{2}(a,b) on a bounded interval (a,b)(a,b)\subset{\mathbb{R}}, and let uH2(a,b)u\in H^{2}(a,b). Then u,uu,u^{\prime} are continuous functions on [a,b][a,b], up to a set of measure zero, and u′′L2(a,b)u^{\prime\prime}\in L^{2}(a,b). Therefore, for such uu, we will always assume that u,uC[a,b]u,u^{\prime}\in C[a,b].

Define the linear space

(3.3) V={ui=1m+1H2(li):u(0)=u(π)=0,J[u](xi)=0,i=1,,m}.V=\left\{u\in\bigoplus_{i=1}^{m+1}H^{2}(l_{i})\,:\,u(0)=u(\pi)=0,\ J[u](x_{i})=0,\ i=1,\dots,m\right\}.

We interpret uVu\in V as a continuous function on [0,π][0,\pi], such that u(0)=u(π)=0u(0)=u(\pi)=0, with uL2(0,π)u^{\prime}\in L^{2}(0,\pi), i.e. uH01(0,π)u\in H_{0}^{1}(0,\pi). Furthermore, u|liH2(li)u|_{l_{i}}\in H^{2}(l_{i}), and u|liC[xi1,xi]u^{\prime}|_{l_{i}}\in C[x_{i-1},x_{i}] for i=1,2,,m+1i=1,2,\dots,m+1.

Define the inner product on VV by

(3.4) ((u,v))V=i=1m+1(u′′,v′′)i+i=1mJ[u](xi)J[v](xi),for anyu,vV,((u,v))_{V}=\sum_{i=1}^{m+1}(u^{\prime\prime},v^{\prime\prime})_{i}+\sum_{i=1}^{m}J[u^{\prime}](x_{i})J[v^{\prime}](x_{i}),\quad\text{for any}\ u,v\in V,

where (u′′,v′′)i=liu′′(x)v′′(x)𝑑x(u^{\prime\prime},v^{\prime\prime})_{i}=\int_{l_{i}}u^{\prime\prime}(x)v^{\prime\prime}(x)\,dx.

It is clear that ((,))V((\cdot,\cdot))_{V} is a symmetric, bilinear form on VV. To see that ((u,u))V=0((u,u))_{V}=0 implies u=0u=0, notice that any function uu with ((u,u))V=0((u,u))_{V}=0 is piecewise linear and continuous on [0,π][0,\pi]. Furthermore, J[u](xi)=0J[u^{\prime}](x_{i})=0 for any i=1,2,,mi=1,2,\dots,m. Therefore uu^{\prime} is continuous on [0,π][0,\pi]. In fact, it is a constant there, since u′′=0u^{\prime\prime}=0 a.e. on [0,π][0,\pi]. Thus uu is a linear function on [0,π][0,\pi] satisfying the zero boundary conditions at the ends of the interval. Therefore u=0u=0 on [0,π][0,\pi], and ((,))V((\cdot,\cdot))_{V} is a well-defined inner product on VV. The corresponding norm in VV is

(3.5) uV2=i=1m+1|u′′|i2+i=1m|J[u](xi)|2,for anyuV,\|u\|_{V}^{2}=\sum_{i=1}^{m+1}|u^{\prime\prime}|^{2}_{i}+\sum_{i=1}^{m}|J[u^{\prime}](x_{i})|^{2},\quad\text{for any}\ u\in V,

where ||i|\cdot|_{i} is the norm in L2(li)L^{2}(l_{i}). We will show in Lemma 3.2 that VV is a Hilbert space.

Let uVu\in V. We define the derivatives of uu component-wise in the spaces H2(li)H^{2}(l_{i}), that is u(x)=(u|li)(x),u′′(x)=(u|li)′′(x)u^{\prime}(x)=(u|_{l_{i}})^{\prime}(x),u^{\prime\prime}(x)=(u|_{l_{i}})^{\prime\prime}(x), and so on, for xlix\in l_{i}, i=1,,m+1i=1,\dots,m+1. For definiteness, we will assume that the derivative uu^{\prime} is continuous from the right on [0,π][0,\pi].

Some useful properties of functions in VV are established in the following lemma.

Lemma 3.1.

Let c0c\geq 0 denote various constants independent of uVu\in V. Then

  1. (i)

    The second derivative u′′u^{\prime\prime} is bounded in HH, and

    (3.6) |u′′|HuV.|u^{\prime\prime}|_{H}\leq\|u\|_{V}.
  2. (ii)

    The derivative uu^{\prime} is bounded on [0,π][0,\pi],

    (3.7) sup{|u(x)|,x[0,π]}c(|u|H+|u′′|H).\sup\{|u^{\prime}(x)|,\ x\in[0,\pi]\}\leq c\left(|u^{\prime}|_{H}+|u^{\prime\prime}|_{H}\right).

    Moreover,

    (3.8) |u|HcuV,andsup{|u(x)|,x[0,π]}uV.|u^{\prime}|_{H}\leq c\|u\|_{V},\quad\text{and}\quad\sup\{|u^{\prime}(x)|,\ x\in[0,\pi]\}\leq\|u\|_{V}.
  3. (iii)

    Function uu is Lipschitz continuous, with the Lipschitz constant cuVc\|u\|_{V}. Also, uu is bounded on [0,π][0,\pi],

    (3.9) u=max{|u(x)|,x[0,π]}cuV,\|u\|_{\infty}=\max\{|u(x)|,\ x\in[0,\pi]\}\leq c\|u\|_{V},

    and

    (3.10) |u|HcuV.|u|_{H}\leq c\|u\|_{V}.
Proof.

We only show (ii). Let uVu\in V. Then its derivative uu^{\prime} is continuous on any interval [xi1,xi][x_{i-1},x_{i}], i=1,,m+1i=1,\dots,m+1. By the Mean Value Theorem for Integrals, there exists ci[xi1,xi]c_{i}\in[x_{i-1},x_{i}], such that

u(ci)=1|li|liu(s)𝑑s.u^{\prime}(c_{i})=\frac{1}{|l_{i}|}\int_{l_{i}}u^{\prime}(s)\,ds.

Thus |u(ci)|c|u|H|u^{\prime}(c_{i})|\leq c|u^{\prime}|_{H}. Also, for any x[xi1,xi]x\in[x_{i-1},x_{i}],

|u(x)u(ci)|li|u′′(s)|𝑑sc|u′′|H.|u^{\prime}(x)-u^{\prime}(c_{i})|\leq\int_{l_{i}}|u^{\prime\prime}(s)|\,ds\leq c|u^{\prime\prime}|_{H}.

Therefore |u(x)|c(|u|H+|u′′|H)|u^{\prime}(x)|\leq c\left(|u^{\prime}|_{H}+|u^{\prime\prime}|_{H}\right) for any x[0,π]x\in[0,\pi], giving (3.7). This inequality implies |J[u](x)|2c(|u|H+|u′′|H)|J[u^{\prime}](x)|\leq 2c\left(|u^{\prime}|_{H}+|u^{\prime\prime}|_{H}\right).

We have

(3.11) abu′′(x)𝑑x\displaystyle\int_{a}^{b}u^{\prime\prime}(x)\,dx =u|ab+a<xib(u(xi)u(xi+))\displaystyle=u^{\prime}|_{a}^{b}+\sum_{a<x_{i}\leq b}\left(u^{\prime}(x_{i}^{-})-u^{\prime}(x_{i}^{+})\right)
=u(b)u(a)a<xibJ[u](xi).\displaystyle=u^{\prime}(b)-u^{\prime}(a)-\sum_{a<x_{i}\leq b}J[u^{\prime}](x_{i}).

Therefore

(3.12) |u(b)u(a)|0L|u′′(x)|𝑑x+i=1m|J[u](xi)|cuV.|u^{\prime}(b)-u^{\prime}(a)|\leq\int_{0}^{L}|u^{\prime\prime}(x)|\,dx+\sum_{i=1}^{m}|J[u^{\prime}](x_{i})|\leq c\|u\|_{V}.

First, choose a[0,π]a\in[0,\pi] be such that u(a)0u^{\prime}(a)\leq 0, which is always possible, since u(0)=u(L)=0u(0)=u(L)=0. Then, by (3.12), for any b[0,π]b\in[0,\pi] we have u(b)cuVu^{\prime}(b)\leq c\|u\|_{V}. This establishes the upper bound for u(x),x[0,π]u^{\prime}(x),\,x\in[0,\pi]. Similarly, choosing aa such that u(a)0u^{\prime}(a)\geq 0, we establish the lower bound for u(x),x[0,π]u^{\prime}(x),\,x\in[0,\pi]. Inequalities in (3.8) follow. ∎

Lemma 3.2.

Let uVu\in V,

N12(u)=|u|H2+|u|H2+|u′′|H2,N^{2}_{1}(u)=|u|^{2}_{H}+|u^{\prime}|^{2}_{H}+|u^{\prime\prime}|^{2}_{H},

and

N22(u)=|u|H2+|u′′|H2.N^{2}_{2}(u)=|u^{\prime}|^{2}_{H}+|u^{\prime\prime}|^{2}_{H}.

Then

  1. (i)

    The norms N1,N2N_{1},N_{2} and V\|\cdot\|_{V} are equivalent on VV.

  2. (ii)

    i=1m+1H2(li)\bigoplus_{i=1}^{m+1}H^{2}(l_{i}) is a Hilbert space, and VV is its closed subspace of
    co-dimension m+2m+2.

  3. (iii)

    VV is a Hilbert space.

Proof.

(i) This follows from Lemma 3.1, and the observation that |J[u](x)|2c(|u|H+|u′′|H)|J[u^{\prime}](x)|\leq 2c\left(|u^{\prime}|_{H}+|u^{\prime\prime}|_{H}\right).

(ii) Let X=i=1m+1H2(li)X=\bigoplus_{i=1}^{m+1}H^{2}(l_{i}). Then XX is a Hilbert space with the norm N1()N_{1}(\cdot). By the Trace Theorem [12, Theorem 3.2], functionals g0(u)=u(0+),g_{0}(u)=u(0^{+}), and gπ(u)=u(π)g_{\pi}(u)=u(\pi^{-}), uXu\in X are continuous linear functional on XX. Therefore {uX:g0(u)=0,gπ(u)=0}\{u\in X:g_{0}(u)=0,\ g_{\pi}(u)=0\} is a closed subspace of XX of co-dimension 22 in XX.

Similarly, the functionals J[u](xi)=u(xi+)u(xi)J[u](x_{i})=u(x_{i}^{+})-u(x_{i}^{-}), uXu\in X, i=1,,mi=1,\dots,m, are linear and continuous on XX. By the definition (3.3) of VV, we conclude that (V,N1())(V,N_{1}(\cdot)) is a closed subspace of XX of co-dimension m+2m+2.

It remains to show that on VV the norm N1N_{1} is equivalent to the norm V\|\cdot\|_{V}, defined in (3.5), but this was established in (i).

(iii) Since the space X=i=1m+1H2(li)X=\bigoplus_{i=1}^{m+1}H^{2}(l_{i}) is complete, then so is its closed subspace VV. ∎

Lemma 3.3.

The identity embedding i:VHi:V\to H is linear, continuous, with a dense range in H01H_{0}^{1}. Furthermore, it is compact.

Proof.

The embedding is linear. By (3.10), |u|HcuV|u|_{H}\leq c\|u\|_{V} for uVu\in V. Thus the embedding is continuous.

Let BVB\subset V be the unit ball of VV. By Lemma 3.1, functions uBu\in B are equicontinuous and equibounded on [0,π][0,\pi]. Hence they form a precompact set in C[0,π]C[0,\pi], and in L2(0,π)L^{2}(0,\pi). Similarly, functions {u}uB\{u^{\prime}\}_{u\in B} are precompact in C[xi1,xi]C[x_{i-1},x_{i}], for any i=1,,m+1i=1,\dots,m+1, hence they are precompact in L2(0,π)L^{2}(0,\pi). The compactness of the embedding follows.

For the density of the embedding, note that C0(0,π)VHC_{0}^{\infty}(0,\pi)\subset V\subset H, and C0(0,π)C_{0}^{\infty}(0,\pi) is dense in HH. Therefore VV is dense in HH. ∎

Lemma 3.3 allows us to define the Gelfand triple VHVV\subset H\subset V^{\prime}, with the dense embeddings. Furthermore, the embedding VHV\subset H is compact. The pairing ,V\langle\cdot,\cdot\rangle_{V} between VV and VV^{\prime} extending the inner product in HH. This means that given fH=HVf\in H=H^{\prime}\subset V^{\prime}, and vVv\in V, we have f,vV=(f,v)H\langle f,v\rangle_{V}=(f,v)_{H}.

4. Variational setting and operator 𝒜\mathcal{A}

We introduce the operator 𝒜:VV\mathcal{A}:V\to V^{\prime} that ”absorbs” the junction boundary conditions. This operator is central to the variational setting of problems for cracked beams and arches. The existence of its eigenvalues and the eigenfunctions is established as well.

Definition 4.1.

Define the operator 𝒜\mathcal{A} on VV by

(4.1) 𝒜u,vV=i=1m+1(u′′,v′′)i+i=1m1θiJ[u](xi)J[v](xi),\langle\mathcal{A}u,v\rangle_{V}=\sum_{i=1}^{m+1}(u^{\prime\prime},v^{\prime\prime})_{i}+\sum_{i=1}^{m}\frac{1}{\theta}_{i}J[u^{\prime}](x_{i})J[v^{\prime}](x_{i}),

for any u,vVu,v\in V. We will also write 𝒜u,v\langle\mathcal{A}u,v\rangle for 𝒜u,vV\langle\mathcal{A}u,v\rangle_{V}, if it does not cause a confusion.

See Section 2 for the setup for the junction (crack) points xix_{i}, and the flexibilities θi\theta_{i}. Recall, that a linear operator A:VVA:V\to V^{\prime} is called coercive, if there exists c>0c>0, such that Au,ucuV2\langle Au,u\rangle\geq c\|u\|^{2}_{V} for any uVu\in V.

Lemma 4.2.

Let 𝒜\mathcal{A} be defined by (4.1). Then 𝒜\mathcal{A} is a symmetric, continuous, linear, and coercive operator from VV onto VV^{\prime}.

Proof.

Clearly, 𝒜\mathcal{A} is a symmetric linear operator. Since all θi>0\theta_{i}>0, we conclude that there exists a constant C>0C>0, such that |𝒜u,v|CuVvV|\langle\mathcal{A}u,v\rangle|\leq C\|u\|_{V}\|v\|_{V}. Therefore 𝒜\mathcal{A} is defined on all of VV, and it is bounded.

Similarly,

|𝒜u,u|=i=1m+1|u′′|2+i=1m1θi|J[u](xi)|2cuV2.|\langle\mathcal{A}u,u\rangle|=\sum_{i=1}^{m+1}|u^{\prime\prime}|^{2}+\sum_{i=1}^{m}\frac{1}{\theta}_{i}|J[u^{\prime}](x_{i})|^{2}\geq c\|u\|_{V}^{2}.

Therefore 𝒜\mathcal{A} is coercive on VV, and its range is VV^{\prime}, see [15, Theorem 2.2.1]. ∎

As was mentioned in Section 2, functions u=u(x)u=u(x) modeling an arch with cracks are expected to satisfy certain boundary conditions. For convenience, we restate them here:

(4.2) u(0)=u(π)=0,u′′(0)=u′′(π)=0,u(0)=u(\pi)=0,\quad u^{\prime\prime}(0)=u^{\prime\prime}(\pi)=0,

and

(4.3) J[u](xi)=0,J[u′′](xi)=0,J[u′′′](xi)=0,J[u](xi)=θiu′′(xi+),J[u](x_{i})=0,\quad J[u^{\prime\prime}](x_{i})=0,\quad J[u^{\prime\prime\prime}](x_{i})=0,\quad J[u^{\prime}](x_{i})=\theta_{i}u^{\prime\prime}(x_{i}^{+}),

for i=1,,mi=1,\dots,m.

The next theorem is the main result of this paper.

Theorem 4.3.

Let the domain of 𝒜\mathcal{A} be D(𝒜)={vV:𝒜vH}D(\mathcal{A})=\{v\in V:\mathcal{A}v\in H\}.

  1. (i)

    If uD(𝒜)u\in D(\mathcal{A}), then u|liH4(li)u|_{l_{i}}\in H^{4}(l_{i}), 𝒜u=u′′′′\mathcal{A}u=u^{\prime\prime\prime\prime} a.e. on lil_{i}, i=1,,m+1i=1,\dots,m+1, and uu satisfies conditions (4.2)–(4.3).

  2. (ii)

    Let fHf\in H, then equation 𝒜u=f\mathcal{A}u=f in VV^{\prime} has a unique solution uD(𝒜)u\in D(\mathcal{A}).

Proof.

By Lemma 4.2, the operator 𝒜\mathcal{A} is coercive, and its range is VV^{\prime}. Since H=HVH=H^{\prime}\subset V^{\prime}, condition fHf\in H implies that fVf\in V^{\prime}. Therefore equation 𝒜u=f\mathcal{A}u=f in VV^{\prime} has a solution uD(𝒜)u\in D(\mathcal{A}), which is unique since 𝒜\mathcal{A} is coercive.

To investigate the properties of functions in D(𝒜)D(\mathcal{A}), recall that l1=(x0,x1)l_{1}=(x_{0},x_{1}). Notice that C0(l1)VC_{0}^{\infty}(l_{1})\subset V, where it is assumed that the functions from C0(l1)C_{0}^{\infty}(l_{1}) are extended by zero outside of l1l_{1}. Thus v(x)=v(x)=0v(x)=v^{\prime}(x)=0, for x=0x=0 and any xx1x\geq x_{1}, vC0(l1)v\in C_{0}^{\infty}(l_{1}).

Let uD(𝒜)u\in D(\mathcal{A}), so 𝒜u=f\mathcal{A}u=f for some fHf\in H. By the definition of VV, we have u|l1H2(l1)u|_{l_{1}}\in H^{2}(l_{1}). For any vC0(l1)v\in C_{0}^{\infty}(l_{1}), by the definition of 𝒜\mathcal{A}, we have

𝒜u,v=l1u′′(x)v′′(x)𝑑x.\langle\mathcal{A}u,v\rangle=\int_{l_{1}}u^{\prime\prime}(x)v^{\prime\prime}(x)\,dx.

Integration by parts gives

l1u′′(x)v′′(x)𝑑x=l1u(x)v′′′′(x)𝑑x=(f,v)H=l1f(x)v(x)𝑑x.\int_{l_{1}}u^{\prime\prime}(x)v^{\prime\prime}(x)\,dx=\int_{l_{1}}u(x)v^{\prime\prime\prime\prime}(x)\,dx=(f,v)_{H}=\int_{l_{1}}f(x)v(x)\,dx.

Therefore D(4)u=fD^{(4)}u=f in the sense of the weak derivatives on l1l_{1}. Thus u|l1H4(l1)u|_{l_{1}}\in H^{4}(l_{1}), and u′′′′=fu^{\prime\prime\prime\prime}=f a.e. on l1l_{1}. Repeating this argument for other intervals lil_{i}, we conclude that u|liH4(li)u|_{l_{i}}\in H^{4}(l_{i}), and u′′′′=fu^{\prime\prime\prime\prime}=f a.e. on lil_{i}, i=1,,m+1i=1,\dots,m+1.

It remains to show the satisfaction of the conditions (4.2)–(4.3). So, let 𝒜u=fH\mathcal{A}u=f\in H. Since we have already established that u|liH4(li)u|_{l_{i}}\in H^{4}(l_{i}), i=1,,m+1i=1,\dots,m+1, we can do the Integration by Parts on every interval lil_{i}, to obtain that for any vVv\in V

(f,v)H\displaystyle(f,v)_{H} =𝒜u,v=i=1m+1(u′′,v′′)i+i=1m1θiJ[u](xi)J[v](xi)\displaystyle=\langle\mathcal{A}u,v\rangle=\sum_{i=1}^{m+1}(u^{\prime\prime},v^{\prime\prime})_{i}+\sum_{i=1}^{m}\frac{1}{\theta}_{i}J[u^{\prime}](x_{i})J[v^{\prime}](x_{i})
=i=1m+1(u′′′′,v)iu′′′v|0π+i=1mu′′′v|xixi++u′′v|0πi=1mu′′v|xixi+\displaystyle=\sum_{i=1}^{m+1}(u^{\prime\prime\prime\prime},v)_{i}-u^{\prime\prime\prime}v|_{0}^{\pi}+\sum_{i=1}^{m}u^{\prime\prime\prime}v|_{x_{i}^{-}}^{x_{i}^{+}}+u^{\prime\prime}v^{\prime}|_{0}^{\pi}-\sum_{i=1}^{m}u^{\prime\prime}v^{\prime}|_{x_{i}^{-}}^{x_{i}^{+}}
+i=1m1θiJ[u](xi)J[v](xi).\displaystyle+\sum_{i=1}^{m}\frac{1}{\theta}_{i}J[u^{\prime}](x_{i})J[v^{\prime}](x_{i}).

Since vVv\in V, we have v(0)=v(π)=0v(0)=v(\pi)=0, and vv is continuous on [0,π][0,\pi]. Therefore the above equality can be rewritten as

(4.4) i=1m+1(u′′′′f,v)i+i=1mJ[u′′′](xi)v(xi)+u′′v|0πi=1mu′′v|xixi+\displaystyle\sum_{i=1}^{m+1}(u^{\prime\prime\prime\prime}-f,v)_{i}+\sum_{i=1}^{m}J[u^{\prime\prime\prime}](x_{i})v(x_{i})+u^{\prime\prime}v^{\prime}|_{0}^{\pi}-\sum_{i=1}^{m}u^{\prime\prime}v^{\prime}|_{x_{i}^{-}}^{x_{i}^{+}}
+i=1m1θiJ[u](xi)J[v](xi)=0.\displaystyle+\sum_{i=1}^{m}\frac{1}{\theta}_{i}J[u^{\prime}](x_{i})J[v^{\prime}](x_{i})=0.

The first sum is zero, since u′′′′=fu^{\prime\prime\prime\prime}=f a.e. on lil_{i}, i=1,,m+1i=1,\dots,m+1. Next, choose a continuously differentiable vVv\in V, which is not zero only in a small neighborhood of x=0x=0, and v(0)0v^{\prime}(0)\not=0. Conclude that u′′(0)=0u^{\prime\prime}(0)=0. Similarly, u′′(π)=0u^{\prime\prime}(\pi)=0.

Choose a continuously differentiable vVv\in V, such that v(xi)=0v^{\prime}(x_{i})=0, and v(xi)=0v(x_{i})=0 for all i=2,,mi=2,\dots,m, but v(x1)0v(x_{1})\not=0, v(x1)=0v^{\prime}(x_{1})=0. Conclude that J[u′′′](x1)=0J[u^{\prime\prime\prime}](x_{1})=0. Repeat this procedure for other points xix_{i}, one at a time. Thus J[u′′′](xi)=0J[u^{\prime\prime\prime}](x_{i})=0, i=1,,mi=1,\dots,m. We are left with

(4.5) i=1m[1θiJ[u](xi)J[v](xi)u′′v|xixi+]=0.\sum_{i=1}^{m}\left[\frac{1}{\theta}_{i}J[u^{\prime}](x_{i})J[v^{\prime}](x_{i})-u^{\prime\prime}v^{\prime}|_{x_{i}^{-}}^{x_{i}^{+}}\right]=0.

Choose a continuously differentiable vVv\in V, which is not zero only in a small neighborhood of x1x_{1}, and such that v(x1)0v^{\prime}(x_{1})\not=0. This implies J[u′′](x1)v(x1)=0J[u^{\prime\prime}](x_{1})v^{\prime}(x_{1})=0. Therefore J[u′′](x1)=0J[u^{\prime\prime}](x_{1})=0. Repeat for other points xix_{i}. Thus u′′(xi+)=u′′(xi)u^{\prime\prime}(x_{i}^{+})=u^{\prime\prime}(x_{i}^{-}) for i=1,,mi=1,\dots,m. Now we can rewrite (4.5) as

i=1m[1θiJ[u](xi)u′′(xi+)]J[v](xi)=0.\sum_{i=1}^{m}\left[\frac{1}{\theta}_{i}J[u^{\prime}](x_{i})-u^{\prime\prime}(x_{i}^{+})\right]J[v^{\prime}](x_{i})=0.

Choose a continuous, piecewise linear vVv\in V, such that v(x1)=1v(x_{1})=1, vv is linear on [0,x1][0,x_{1}], and on [x1,π][x_{1},\pi]. Note that J[v](xi)=0J[v^{\prime}](x_{i})=0 for i=2,,mi=2,\dots,m, and J[v](x1)0J[v^{\prime}](x_{1})\not=0. Conclude that J[u](x1)=θ1u′′(x1+)J[u^{\prime}](x_{1})=\theta_{1}u^{\prime\prime}(x_{1}^{+}). Repeat for other points xix_{i}, i=2,,mi=2,\dots,m. Thus uu satisfies all the conditions (4.2)–(4.3). ∎

Remark. The fact that u′′′′=fu^{\prime\prime\prime\prime}=f a.e. on (0,π)(0,\pi) in Theorem 4.3 does not imply that uH4(0,π)u\in H^{4}(0,\pi). This is similar to the fact that the strong derivative pp^{\prime} of a step function pp on (0,π)(0,\pi) is zero a.e. on (0,π)(0,\pi). However, pH1(0,π)p\not\in H^{1}(0,\pi).

Finally in this section we discuss the eigenvalues and the eigenfunctions of the operator 𝒜\mathcal{A}. It was shown in Lemma 4.2 that 𝒜\mathcal{A} is a continuous, linear, symmetric, and coercive operator from VV onto VV^{\prime}. Following [15, Section 2.2.1], 𝒜\mathcal{A} can also be considered as an unbounded operator in HH. By Lemma 3.3, the embedding VHV\subset H is compact. Therefore the standard spectral theory for Sturm-Liouville boundary value problems is applicable. The eigenfunctions belong to HH. Therefore, by Theorem 4.3, they are in the domain D(𝒜)VD(\mathcal{A})\subset V, thus continuous on [0,π][0,\pi], and satisfy conditions (4.2)–(4.3).

We summarize these results in the following lemma.

Lemma 4.4.

Let 𝒜\mathcal{A} be the operator defined in (4.1). Then

  1. (i)

    There exists an increasing sequence of its real positive eigenvalues
    λ14,λ24,\lambda^{4}_{1},\lambda^{4}_{2},\dots, with limkλk4=\lim_{k\to\infty}\lambda^{4}_{k}=\infty.

  2. (ii)

    The corresponding eigenfunctions φkD(𝒜)V\varphi_{k}\in D(\mathcal{A})\subset V, k1k\geq 1, and they satisfy the junction conditions (4.2)–(4.3).

  3. (iii)

    The eigenfunctions φk\varphi_{k} satisfy 𝒜φk=λk4φk\mathcal{A}\varphi_{k}=\lambda^{4}_{k}\varphi_{k} in HH, k1k\geq 1. That is, φk′′′′(x)=λk4φk(x)\varphi_{k}^{\prime\prime\prime\prime}(x)=\lambda^{4}_{k}\varphi_{k}(x) a.e. on every interval lil_{i}, i=1,,m+1i=1,\dots,m+1.

  4. (iv)

    The set {φk}k=1\{\varphi_{k}\}_{k=1}^{\infty} is a complete orthonormal basis in HH.

Algorithms for a computational determination of the eigenvalues and the eigenfunctions of 𝒜\mathcal{A} are discussed in Section 5.

Remark. If the arch is uniform, i.e. it has no cracks, then the results presented in this section are simplified. Specifically, the spaces V,HV,H, and the operator 𝒜\mathcal{A} take the following forms

(4.6) V=H2(0,π)H01(0,π),H=L2(0,π),𝒜u,vV=(u′′,v′′)H,V=H^{2}(0,\pi)\cap H_{0}^{1}(0,\pi),\quad H=L^{2}(0,\pi),\quad\langle\mathcal{A}u,v\rangle_{V}=(u^{\prime\prime},v^{\prime\prime})_{H},

for any u,vVu,v\in V. See [8] for an investigation of this case.

5. Eigenvalues and eigenfunctions

In this section we present the Modified Shifrin’s method for the computation of the eigenfunctions φk,k1\varphi_{k},\,k\geq 1 and the corresponding eigenvalues λk4\lambda^{4}_{k}, of the operator 𝒜\mathcal{A}. The existence of the eigenvalues and the eigenfunctions has been established in Lemma 4.4.

Transition matrices method. This is a common method for the determination of the eigenfunctions and the eigenvalues, so we just briefly mention it for completeness, see [11] for details.

The general solution of the equation w′′′′=λ4ww^{\prime\prime\prime\prime}=\lambda^{4}w on li=(xi1,xi)l_{i}=(x_{i-1},x_{i}), for i=1,,m+1i=1,\dots,m+1 is

(5.1) wiλ(x)\displaystyle w_{i}^{\lambda}(x) =Aisinλ(xxi1)+Bicosλ(xxi1)\displaystyle=A_{i}\sin\lambda(x-x_{i-1})+B_{i}\cos\lambda(x-x_{i-1})
+Cisinhλ(xxi1)+Dicoshλ(xxi1).\displaystyle+C_{i}\sinh\lambda(x-x_{i-1})+D_{i}\cosh\lambda(x-x_{i-1}).

Let vector Ai=[Ai,Bi,Ci,Di]T\vec{A}_{i}=[A_{i},B_{i},C_{i},D_{i}]^{T} be composed of the coefficients of the expansion in (5.1), on interval lil_{i}.

Suppose that vector A1\vec{A}_{1} is known. Then it defines function w1λw_{1}^{\lambda} on l1l_{1}, and the boundary conditions for w1λw_{1}^{\lambda} at x1x_{1}^{-}. The junction conditions (4.2)–(4.3) define the boundary conditions for w2λw_{2}^{\lambda} at x1+x_{1}^{+}. Then the initial value problem (w2λ)′′′′=λ4w2λ(w_{2}^{\lambda})^{\prime\prime\prime\prime}=\lambda^{4}w_{2}^{\lambda} on l2=(x1,x2)l_{2}=(x_{1},x_{2}), uniquely defines the expansion coefficients A2\vec{A}_{2} on l2l_{2}. It is readily seen that the transformation from A1\vec{A}_{1} to A2\vec{A}_{2} is linear, and it is given by a 4×44\times 4 matrix T(1)T^{(1)}, i.e. A2=T(1)A1\vec{A}_{2}=T^{(1)}\vec{A}_{1}.

Extending this process to all the subintervals li,i=1,,m+1l_{i},i=1,\dots,m+1, we get

(5.2) Am+1=T(m)T(1)A1.\vec{A}_{m+1}=T^{(m)}\cdots T^{(1)}\vec{A}_{1}.

Note that all the matrices T(i)T^{(i)} are λ\lambda-dependent in a non-linear way.

To satisfy the hinged boundary conditions at x=0x=0, we require A1=[A1,0,C1,0]T\vec{A}_{1}=[A_{1},0,C_{1},0]^{T}.

Let the 2×42\times 4 matrix B(m+1)B^{(m+1)} transform the solution wm+1λw_{m+1}^{\lambda} determined by the vector Am+1\vec{A}_{m+1} into the boundary conditions for wm+1λw_{m+1}^{\lambda} and (wm+1λ)′′(w_{m+1}^{\lambda})^{\prime\prime} at x=πx=\pi.

To satisfy the hinged boundary conditions at x=πx=\pi, we have to solve the matrix equation

(5.3) [0,0]T=B(m+1)T(m)T(1)A1.[0,0]^{T}=B^{(m+1)}T^{(m)}\cdots T^{(1)}\vec{A}_{1}.

This matrix equation has a non-trivial solution, if the corresponding λ\lambda-dependent 2×22\times 2 determinant is equal to zero. This amounts to finding an eigenvalue λ4\lambda^{4} of the problem. Numerically, the highly nonlinear equation is solved by a Newton type method. The computations can be quite expensive, so the applicability of this method is usually restricted to a small number of cracks, see [11].

Modified Shifrin’s method. The original method is described in [14]. We modify it by placing it within the framework of this paper. Also, notice that in our study we have proved the existence of the eigenvalues and the eigenfunctions for the operator 𝒜\mathcal{A}, which provides the theoretical justification for the method.

Let VlVV_{l}\subset V be the linear space of continuous piecewise linear functions on [0,π][0,\pi], which are linear on every interval li=(xi1,xi)l_{i}=(x_{i-1},x_{i}), i=1,m+1i=1,\dots m+1. Note that VlV_{l} is an mm-dimensional space.

The goal of the next result is to show that any function uVu\in V can be uniquely represented as u=us+ulu=u_{s}+u_{l}, where usu_{s} is smooth, and ulVlu_{l}\in V_{l}. Thus ulu_{l} absorbs all the jumps of the derivative uu^{\prime} on (0,π)(0,\pi).

Lemma 5.1.

Let uVu\in V. Then there exists a unique decomposition

(5.4) u=us+ul,u=u_{s}+u_{l},

where usH01(0,π)H2(0,π)u_{s}\in H_{0}^{1}(0,\pi)\cap H^{2}(0,\pi), and ulVlu_{l}\in V_{l}.

Proof.

Let vv be defined by

v(x)=0x[0ξu′′(τ)𝑑τ]𝑑ξ,x,ξ[0,π].v(x)=\int_{0}^{x}\left[\int_{0}^{\xi}u^{\prime\prime}(\tau)\,d\tau\right]d\xi,\quad x,\xi\in[0,\pi].

Then vH2(0,π)v\in H^{2}(0,\pi), v(0)=v(0)=0v(0)=v^{\prime}(0)=0, and v′′=u′′v^{\prime\prime}=u^{\prime\prime}. Let us(x)=v(x)v(π)x/πu_{s}(x)=v(x)-v(\pi)x/\pi, x[0,π]x\in[0,\pi]. Then us(0)=us(π)=0u_{s}(0)=u_{s}(\pi)=0. Therefore usH01(0,π)H2(0,π)Vu_{s}\in H_{0}^{1}(0,\pi)\cap H^{2}(0,\pi)\subset V. Note that J[us](xi)=J[v](xi)=0J[u_{s}^{\prime}](x_{i})=J[v^{\prime}](x_{i})=0, and us′′=u′′u_{s}^{\prime\prime}=u^{\prime\prime}. If u~s\tilde{u}_{s} is another such function, then u~susV=0\|\tilde{u}_{s}-u_{s}\|_{V}=0, and u~s=us\tilde{u}_{s}=u_{s}.

Now we just let ul=uusu_{l}=u-u_{s}. Then ul′′=0u_{l}^{\prime\prime}=0 on every interval lil_{i}, and ulVlu_{l}\in V_{l}. The function ulu_{l} is unique, since uulu-u_{l} is smooth, which is already shown to be unique. ∎

Let φ=φk\varphi=\varphi_{k} be an eigenfunction of 𝒜\mathcal{A}. The index kk will be suppressed for notational simplicity.

By Lemma 4.4(iii),

(5.5) φ′′′′(x)=λ4φ(x)\varphi^{\prime\prime\prime\prime}(x)=\lambda^{4}\varphi(x)

where the equality is satisfied a.e. on every interval lil_{i}. By Lemma 5.1 we can represent φ\varphi as

(5.6) φ=φs+φl,\varphi=\varphi_{s}+\varphi_{l},

where φs\varphi_{s} is smooth, and φlVl\varphi_{l}\in V_{l}. Since φl′′=0\varphi_{l}^{\prime\prime}=0 on every lil_{i}, equation (5.5) becomes

(5.7) φs′′′′=λ4φs+λ4φl.\varphi_{s}^{\prime\prime\prime\prime}=\lambda^{4}\varphi_{s}+\lambda^{4}\varphi_{l}.

Let {wi}i=1m\{w_{i}\}_{i=1}^{m} be the basis in VlV_{l}, defined by

(5.8) wi(x)={xiππx,0xxixiπ(xπ),xi<xπ.w_{i}(x)=\begin{cases}\frac{x_{i}-\pi}{\pi}x,&0\leq x\leq x_{i}\\ \frac{x_{i}}{\pi}(x-\pi),&x_{i}<x\leq\pi.\end{cases}

Note that wi(x)<0w_{i}(x)<0, for 0<x<π0<x<\pi, J[wi](xi)=1J[w_{i}^{\prime}](x_{i})=1, and xix_{i} is the only discontinuity point of wiw_{i}^{\prime} on (0,π)(0,\pi).

Let Δi=J[φ](xi)\Delta_{i}=J[\varphi^{\prime}](x_{i}). Note that Δi=Δi(λ)\Delta_{i}=\Delta_{i}(\lambda). Since J[φl](xi)=J[φ](xi)J[\varphi_{l}^{\prime}](x_{i})=J[\varphi^{\prime}](x_{i}), and {wi}i=1m\{w_{i}\}_{i=1}^{m} is a basis in VlV_{l}, the function φl\varphi_{l} can be represented as

(5.9) φl(x;λ)=i=1mΔi(λ)wi(x).\varphi_{l}(x;\lambda)=\sum_{i=1}^{m}\Delta_{i}(\lambda)w_{i}(x).

Therefore equation (5.7) can be written as

(5.10) φs′′′′=λ4φs+λ4i=1mΔiwi.\varphi_{s}^{\prime\prime\prime\prime}=\lambda^{4}\varphi_{s}+\lambda^{4}\sum_{i=1}^{m}\Delta_{i}w_{i}.

In [14], equation (6) on p. 412 is used instead of (5.7). Note, that while our approaches are similar, the notations may be defined differently. Thus the formulas in [14] cannot be used in our framework.

Equation (5.10) for φs\varphi_{s} is a linear non-homogeneous, fourth order ODE on (0,π)(0,\pi). Its general solution is the sum of the complementary solution

(5.11) (φs)c(x;λ)=Acos(λx)+Bsin(λx)+Ccosh(λx)+Dsinh(λx),\left(\varphi_{s}\right)_{c}(x;\lambda)=A\cos(\lambda x)+B\sin(\lambda x)+C\cosh(\lambda x)+D\sinh(\lambda x),

and a particular solution (φs)p(x;λ)\left(\varphi_{s}\right)_{p}(x;\lambda). The latter can be found using the Laplace transform. Its expression in a convolution form is

(5.12) (φs)p(x;λ)=λ412λ3(sinh(λx)sin(λx))i=1mΔiwi(x).\left(\varphi_{s}\right)_{p}(x;\lambda)=\lambda^{4}\frac{1}{2\lambda^{3}}\left(\sinh(\lambda x)-\sin(\lambda x)\right)*\sum_{i=1}^{m}\Delta_{i}w_{i}(x).

Let

(5.13) Mi(x;λ)=0x(sinh(λ(xs))sin(λ(xs)))wi(s)𝑑s.M_{i}(x;\lambda)=\int_{0}^{x}\left(\sinh(\lambda(x-s))-\sin(\lambda(x-s))\right)w_{i}(s)\,ds.

Then the general solution φs=(φs)c+(φs)p\varphi_{s}=(\varphi_{s})_{c}+(\varphi_{s})_{p} of (5.10) can be written as

(5.14) φs(x;λ)=(φs)c(x;λ)+λ2i=1mΔiMi(x;λ).\varphi_{s}(x;\lambda)=\left(\varphi_{s}\right)_{c}(x;\lambda)+\frac{\lambda}{2}\sum_{i=1}^{m}\Delta_{i}M_{i}(x;\lambda).

Note that all the expressions Mi(x;λ),i=1,,mM_{i}(x;\lambda),\,i=1,\dots,m can be computed explicitly.

Recall from the junction conditions (4.3), that φ′′(xj)=φ′′(xj+)\varphi^{\prime\prime}(x_{j}^{-})=\varphi^{\prime\prime}(x_{j}^{+}), and Δj=J[φ](xj)=θjφ′′(xj+)\Delta_{j}=J[\varphi^{\prime}](x_{j})=\theta_{j}\varphi^{\prime\prime}(x_{j}^{+}) for j=1,,mj=1,\dots,m. By construction φ′′(x)=φs′′(x)\varphi^{\prime\prime}(x)=\varphi_{s}^{\prime\prime}(x). Thus φs′′(xj)=φs′′(xj+)\varphi^{\prime\prime}_{s}(x_{j}^{-})=\varphi^{\prime\prime}_{s}(x_{j}^{+}), and we can write Δj=θjφs′′(xj)\Delta_{j}=\theta_{j}\varphi^{\prime\prime}_{s}(x_{j}), j=1,,mj=1,\dots,m. Therefore, the last expression becomes

(5.15) Δj=θj[(φs)c′′(xj;λ)+λ2i=1mΔiMi′′(xj;λ)],\Delta_{j}=\theta_{j}\left[\left(\varphi_{s}\right)_{c}^{\prime\prime}(x_{j};\lambda)+\frac{\lambda}{2}\sum_{i=1}^{m}\Delta_{i}M_{i}^{\prime\prime}(x_{j};\lambda)\right],

where the derivatives are in xx, and j=1,,mj=1,\dots,m.

Thus (5.15) gives mm linear equations for m+4m+4 unknowns Δj,j=1,,m\Delta_{j},\,j=1,\dots,m, and A,B,C,DA,B,C,D. Note that these mm equations are valid for any boundary conditions at the ends of the interval (0,π)(0,\pi). The missing four equations are derived from the boundary conditions of the problem.

For example, for the hinged boundary conditions (1.1), we have the following four linear equations: φs(0;λ)=φs′′(0;λ)=0\varphi_{s}(0;\lambda)=\varphi_{s}^{\prime\prime}(0;\lambda)=0, and φs(π;λ)=φs′′(π;λ)=0\varphi_{s}(\pi;\lambda)=\varphi_{s}^{\prime\prime}(\pi;\lambda)=0. More explicitly, using (5.14), equation φs(0;λ)=0\varphi_{s}(0;\lambda)=0 becomes

(5.16) (φs)c(0;λ)+λ2i=1mΔiMi(0;λ)=0,orA+C=0.\left(\varphi_{s}\right)_{c}(0;\lambda)+\frac{\lambda}{2}\sum_{i=1}^{m}\Delta_{i}M_{i}(0;\lambda)=0,\quad\text{or}\quad A+C=0.

Similarly, φs′′(0;λ)=0\varphi_{s}^{\prime\prime}(0;\lambda)=0 becomes

(5.17) λ2A+λ2C+λ2i=1mΔiMi′′(0;λ)=0.-\lambda^{2}A+\lambda^{2}C+\frac{\lambda}{2}\sum_{i=1}^{m}\Delta_{i}M_{i}^{\prime\prime}(0;\lambda)=0.

Equations φs(π;λ)=φs′′(π;λ)=0\varphi_{s}(\pi;\lambda)=\varphi_{s}^{\prime\prime}(\pi;\lambda)=0 are

(5.18) (φs)c(π;λ)+λ2i=1mΔiMi(π;λ)=0,\left(\varphi_{s}\right)_{c}(\pi;\lambda)+\frac{\lambda}{2}\sum_{i=1}^{m}\Delta_{i}M_{i}(\pi;\lambda)=0,

and

(5.19) (φs)c′′(π;λ)+λ2i=1mΔiMi′′(π;λ)=0.\left(\varphi_{s}\right)_{c}^{\prime\prime}(\pi;\lambda)+\frac{\lambda}{2}\sum_{i=1}^{m}\Delta_{i}M_{i}^{\prime\prime}(\pi;\lambda)=0.

Writing the linear system (5.15)–(5.19) in the matrix form 𝐔(λ)𝐱¯=𝟎¯\bf U(\lambda)\bar{x}=\bar{0} shows that it has non-trivial solutions only if

(5.20) det(𝐔(λ))=0.\det({\bf U(\lambda)})=0.

This non-linear equation has infinitely many solutions λk,k1\lambda_{k},\,k\geq 1, corresponding to the eigenvalues λk4,k1\lambda_{k}^{4},\,k\geq 1.

Let the corresponding non-trivial solution of the linear system be 𝐱¯={Δ1(λk),,Δm(λk),Ak,Bk,Ck,Dk}{\bf\bar{x}}=\{\Delta_{1}(\lambda_{k}),\dots,\Delta_{m}(\lambda_{k}),A_{k},B_{k},C_{k},D_{k}\}. Then we can compute
φl(x;λk)=i=1mΔi(λk)wi(x)\varphi_{l}(x;\lambda_{k})=\sum_{i=1}^{m}\Delta_{i}(\lambda_{k})w_{i}(x), and φs(x,λk)\varphi_{s}(x,\lambda_{k}) from (5.14). Finally, we get the eigenfunction φk(x)=φs(x,λk)+φl(x;λk)\varphi_{k}(x)=\varphi_{s}(x,\lambda_{k})+\varphi_{l}(x;\lambda_{k}). One may want to normalize φk\varphi_{k} in HH to achieve its uniqueness, up to a sign.

A comparison of the computational efficiency of the methods was conducted in [14]. The first three eigenvalues have been computed by the Transition matrices method, and by the Shifrin’s method. The Shifrin’s method is about twice as fast for the beam with one crack, and about three times as fast for the beam with two cracks.

Natural beam frequencies. For practical applications it is important to express equation (5.5) in physical variables.

Equation of harmonic transverse oscillations v=v(x)v=v(x) of a uniform beam defined on interval (0,L)(0,L) is

(5.21) EIv′′′′(x)=ω2ρAv(x),0<x<L,EIv^{\prime\prime\prime\prime}(x)=\omega^{2}\rho Av(x),\quad 0<x<L,

where EE is the Young’s modulus, AA and II are the cross-sectional area and the area moment of inertia correspondingly, and ω\omega is the natural frequency of the oscillations. For a cracked beam, equation (5.21) is satisfied on every subinterval (xi1,xi)(x_{i-1},x_{i}), i=1,,m+1i=1,\dots,m+1.

To relate this equation to the non-dimensional variables, define the tt-scale ω0\omega_{0}, and the radius of gyration rr by

(5.22) ω0=(πL)2EIρA,r=IA.\omega_{0}=\left(\frac{\pi}{L}\right)^{2}\sqrt{\frac{EI}{\rho A}},\quad r=\sqrt{\frac{I}{A}}.

Then make the change of variables

(5.23) xπxL,vvr,tω0t.x\leftarrow\frac{\pi x}{L},\quad v\leftarrow\frac{v}{r},\quad t\leftarrow\omega_{0}t.

Then equation (5.21) becomes

EIr(πL)4vn′′′′(xn,tn)=ω2ρArvn(xn,tn).EIr\left(\frac{\pi}{L}\right)^{4}v_{n}^{\prime\prime\prime\prime}(x_{n},t_{n})=\omega^{2}\rho Arv_{n}(x_{n},t_{n}).

Thus, in non-dimensional ratios

v′′′′=ω2(Lπ)4ρAEIv.v^{\prime\prime\prime\prime}=\omega^{2}\left(\frac{L}{\pi}\right)^{4}\frac{\rho A}{EI}v.

Comparing this equation with the definition of the eigenvalues and the eigenfunctions φk′′′′=λk4φk\varphi_{k}^{\prime\prime\prime\prime}=\lambda_{k}^{4}\varphi_{k}, we conclude that the natural beam frequencies are given by

(5.24) ωk=λk2(πL)2EIρA,k1.\omega_{k}=\lambda_{k}^{2}\left(\frac{\pi}{L}\right)^{2}\sqrt{\frac{EI}{\rho A}},\quad k\geq 1.

Expressions for flexibilities θi\theta_{i}. The standard approach to modeling a crack is to represent it as a massless rotational spring with the spring constant kk, and the flexibility θ\theta.

The spring constant kk relates the torque to the angle of rotation. In our case this relationship takes the form EIy′′(x)=kJ[v](x)EIy^{\prime\prime}(x)=kJ[v^{\prime}](x), or J[v](x)=θv′′(x)J[v^{\prime}](x)=\theta v^{\prime\prime}(x), where

(5.25) θ=EIk.\theta=\frac{EI}{k}.

If the beam has a rectangular cross-section, as shown in Figure 1, then the area moment of inertia II of the rectangle can be computed explicitly, and (5.25) can be simplified further. If the crack is double-sided, then by [13, Eq. (2.8)-(2.10)], the expression for the flexibility θ\theta becomes

(5.26) θ=6πHμ^2(0.5350.929μ^+3.500μ^23.181μ^3+5.793μ^4),\theta=6\pi H\hat{\mu}^{2}(0.535-0.929\hat{\mu}+3.500\hat{\mu}^{2}-3.181\hat{\mu}^{3}+5.793\hat{\mu}^{4}),

where HH is the half-height of the beam cross-section, and μ^=a/H\hat{\mu}=a/H.

If the crack is single-sided, then by [13, Eq. (2.8)-(2.10)]

(5.27) θ=6πHμ^2(0.63841.035μ^+3.7201μ^25.1773μ^3+7.553μ^47.332μ^5),\theta=6\pi H\hat{\mu}^{2}(0.6384-1.035\hat{\mu}+3.7201\hat{\mu}^{2}-5.1773\hat{\mu}^{3}+7.553\hat{\mu}^{4}-7.332\hat{\mu}^{5}),

where HH is the entire height of the beam cross-section, and μ^=a/H\hat{\mu}=a/H.

References

  • [1] J.M. Ball, Initial-boundary value problems for an extensible beam, J. Math. Anal. Appl, 42 (1973) 61-90.
  • [2] S. Caddemi and I. Calio, Exact closed-form solution for the vibration modes of the Euler–Bernoulli beam with multiple open cracks, Journal of Sound and Vibration, 327(3), 1973, 473 - 489.
  • [3] S. Caddemi and A. Morassi, Multi-cracked Euler-Bernoulli beams: Mathematical modeling and exact solutions, International Journal of Solids and Structures, 50(6), 2013, 944-956.
  • [4] M.N. Cerri and G.C. Ruta, Detection of localised damage in plane circular arches by frequency data, Journal of Sound and Vibration, 270(1), 2004, 39-59.
  • [5] S. Christides and A.D.S. Barr, One-dimensional theory of cracked Bernoulli-Euler beams, International Journal of Mechanical Sciences, 26(11), 1984, 639-648
  • [6] E. Emmrich and M. Thalhammer, A class of integro-differential equations incorporating nonlinear and nonlocal damping with applications in nonlinear elastodynamics: Existence via time discretization, Nonlinearity, 24(9), 2011, 2523-2546.
  • [7] S. Gutman and J. Ha, Shallow arches with weak and strong damping, Journal of the Korean Mathematical Society 54(5), 2017, 945-966.
  • [8] S. Gutman and J. Ha and S. Lee, Parameter identification for weakly damped shallow arches, Journal of Mathematical Analysis and Applications, 403(1), 2013, 297-313.
  • [9] S. Gutman and J. Ha, Uniform attractor of shallow arch motion under moving points load, Journal of Mathematical Analysis and Applications, 464(1), 2018, 557-579.
  • [10] S. M. Han and H. Benaroya and T. Wei, Dynamics of transversely vibrating beams using four engineering theories, Journal of Sound and Vibration, 225(5), 1999, 935-988.
  • [11] H.P. Lin and S.C. Chang and J.D. Wu, Beam vibrations with an arbitrary number of cracks, Journal of Sound and Vibration, 258(5), 2002, 987-999.
  • [12] J.L. Lions and E. Magenes, Nonhomogeneous Boundary Value Problems and Applications, vol 1, 1972, Springer-Verlag, New York.
  • [13] W.M. Ostachowicz and M. Krawczuk, Analysis of the effect of cracks on the natural frequencies of a cantilever beam, Journal of Sound and Vibration, 150(2), 1991, 191-201.
  • [14] E.I. Shifrin and R. Ruotolo, Natural frequencies of a beam with an arbitrary number of cracks, Journal of Sound and Vibration, 222(3), 1999, 409-423.
  • [15] R. Temam, Infinite-Dimensional Dynamical Systems in Mechanics and Physics, Applied Mathematical Sciences, 1997, Springer, New York.