3.org/TR/xhtml1/DTD/xhtml1-transitional.dtd"> Unified proof to oscillation property of discrete beam
    Appl. Math. Mech. -Engl. Ed.     2014, Vol. 35 Issue (5) : 621–636     PDF       
http: //dx. doi. org/10.1007/s10483-014-1817-6
The Chinese Meteorological Society
0

Article Information

Zi-jun ZHENG ,Pu CHEN ,Da-jun WANG 2014.
Unified proof to oscillation property of discrete beam
Appl. Math. Mech. -Engl. Ed., 35 (5) : 621–636
http: //dx. doi. org/10.1007/s10483-014-1817-6

Article History

Received 2013-03-01;
in final form 2013-10-30
Unified proof to oscillation property of discrete beam
Zi-jun ZHENG, Pu CHEN , Da-jun WANG       
Department of Mechanics and Engineering Science, State Key Laboratory for Turbulence and Complex Systems, College of Engineering, Peking University, Beijing 100871, P. R. China
ABSTRACT:The oscillation property (OP) is a fundamental and important qualitative property for the vibrations of single span one-dimensional continuums such as strings, bars, torsion bars, and Euler beams. Any properly discretized continuum model should keep the OP. In literatures, the OP of discrete beam models is discussed essentially by means of matrix factorization. The discussion is model-specific and boundary-conditionspecific. Besides, matrix factorization is difficult in handling finite element (FE) models of beams. In this paper, according to a sufficient condition for the OP, a new approach to discuss the property is proposed. The local criteria on discrete displacements rather than global matrix factorizations are given to verify the OP. Based on the proposed approach, known results such as the OP for the 2-node FE beams via the HeilingerReissener principle (HR-FE beams) as well as the 5-point finite difference (FD) beams are verified. New results on the OP for the 2-node PE-FE beams and the FE Timoshenko beams with small slenderness are given. Through a simple manipulation, the qualitative property of discrete multibearing beams can also be discussed by the proposed approach.
Keywordsoscillation property (OP)        finite element method (FEM)        finite difference method (FDM)        Euler beam        Timoshenko beam       

1 Introduction

Studies on the natural modes of structures play a very important role in vibration theory and its applications. Besides the quantitative extractions of natural modes of different mechanical systems,qualitative analyses of modes are useful in checking the results of experiments and numerical calculations in structure designs and the solutions of inverse vibration problems[1, 2, 3, 4, 5, 6, 7, 8, 9] .

For the modes of the single span one-dimensional continuums such as strings,bars,torsion bars,and Euler beams (the beams mentioned hereafter are single span Euler beams if not specified),Krein and Gantmacher[1] summarized the following qualitative properties:

P1 There is no repetitive natural frequency (except for a beam with two free ends whose first two frequencies are both zero);

P2 The 1st mode shape has no unconstrained zero point;

P3 The ith mode shape has exactly (i − 1) nodes;

P4 The nodes of two contiguous mode shapes alternate;

P5 Any superposition of the ith,(i + 1)th,· · · ,jth mode shapes reverses its sign no less than (i − 1) times and no more than (j − 1) times.

The ensemble of P1-P5 is known as the oscillation property (OP),from which many other qualitative properties can be obtained[2, 3, 4, 5, 6] . The pioneering work on the OP can be traced back to Sturm,a French mathematician in the 19th century,who established the OP of analytic bars. The unified proof of the OP for analytic strings,bars,and beams was given by Gantmacher and Krein with their excellent oscillatory kernel/matrix theory[1] . It has been proved that some constrained one-dimensional continuums possess the OP as well[7, 8] .

Vibration modes of various discrete models of one-dimensional continuums are supposed to reflect the OP. Otherwise,the discrete models are not reasonable. The discrete form of the OP is given by replacing the mode shapes in P2-P5 with u-lines,which are defined as the polylines connecting the successive deflected nodes of the discrete models (see the dot-dash line in Fig. 1)[1, 6, 7] .

Fig. 1. Simple supported beam loaded by three concentrated forces.

In fact,commonly used discrete models possess the OP. Some finite difference (FD) bars and finite element (FE) bars are proved to possess the property for any arbitrary mesh[6, 10]by stiffness matrix factorization. The stiffness matrices of 3-point FD bars and 2-node FE bars are both tri-diagonal,and thus can be uniformly factorized into the simple matrix as follows:

where L and are diagonal matrices,whose elements are grid sizes and nodal stiffness coeffi- cients,respectively,and E is known as the differential matrix expressed as

The factors of K can be easily identified as sign-oscillatory matrices. Therefore,the 3-point FD and the 2-node FE bars possess the OP unconditionally[6, 10]. However,for discrete beams,the patterns of stiffness matrices differ from model to model. Therefore,the discussion based on matrix-factorization is model-specific and boundary-condition- specific. The OP of 5-point FD beams was proved[6, 7, 8, 9] by factoring the stiffness matrix as follows:

in which the differential matrix E depends on the boundary conditions. The OP of 2-node FE beams via the potential energy principle (PE-FE beams) was first studied by Gladwell[11]. For cantilever conditions,he provided

where

For other boundary conditions,the factor matrices are more complicated than the above. Even for cantilever beams,the factors (5) and (6) are too complicated to implement further discus- sion. Assuming that the bending stiffness (flexural rigidity) EI was element-wise constant, Gladwell[12] simplified the matrices and established the OP of PE-FE beams.

In fact,if EI is element-wise constant,then the mode shapes of PE-FE beams and the analytic beams are both piecewise cubic polynomial between masses. According to the charac- teristics of the FE method (FEM),these PE-FE beams possess identical mode shapes at nodes to the analytic beams with concentrated masses only,and thus possess the OP[12]. This idea also delivers that arbitrarily meshed 2-node HR-FE beams with the variable bending stiffness EI possess the OP[12]. But if EI is not element-wise constant,the OP may fail for PE-FE beams,as shown in Ref. [12].

However,since the above two approaches,i.e.,the stiffness matrix factorization and the ana- log to analytic beam,are closely coupled to the discrete models,when more complicated discrete beam models are put into consideration,the existing proofs cannot be easily transplanted.

In this paper,a new approach is proposed to discuss the OP of beams,which makes unified discussion on the properties of more complicated discrete beam models possible. We start our discussion with a proposition about a static property of one-dimensional continuums closely related to the OP,which is hence named as the statics oscillation property (SOP).

Proposition 1 (A) If a single span one-dimensional continuum is under the action of one concentrated force (nodal force for discrete systems), then the deflection of any unconstrained point (unconstrained node for discrete models) is non-zero and has the same sign.

(B) If a single span one-dimensional continuum is under the action of k (k = 2,3,• • • ) arbitrary concentrated forces (nodal forces for discrete systems),then its deflection line (u-line for discrete systems) reverses the sign no more than (k − 1) times.

This proposition agrees with our common experiences. See Fig. 1 for example,when a simply supported beam is under the action of three concentrated forces,the deflection line reverses its sign no more than twice. Gantmacher and Krein[1] proved that possessing the SOP is a sufficient condition for a one-dimensional continuum to possess the OP. Zheng et al.[12, 13] extended the conclusion to discrete models with nodal concentrated masses only,and proved the equivalence between the OP and the SOP in discrete cases.

In this paper,the SOP is directly studied instead of the OP. The SOP of one-dimensional continuums can be recognized as a corollary of Rolle’s Theorem[1].Similarly,the SOP of discrete strings,bars,and torsion bars can be proved by analyzing the monotonicity of the u-lines. Unfortunately,the proof of the SOP for discrete bar models cannot be directly used for discrete beam models. Up to now,direct study on the SOP of discrete beam models has not been found in literatures.

Based on the local criteria of the balance,the deformation,and the positivity of the strain energy obtained in an analytic beam model,the SOP of discrete beam models is proved con- sistently in this paper. The result covers the known conclusions that the SOP/OP exists in 5-point FD beam models[6, 7, 8] and 2-node HR-FE beam models[12]. A sufficient condition for the SOP/OP of PE-FE beam models is also delivered from the result,which includes the condition that the bending stiffness is element-wise constant,as discussed in Refs. [11, 12],and supports the conjecture in Ref. [11] that the OP exists in the PE-FE beams with element-wise linear bending stiffness. Besides,the SOP/OP of FE Timoshenko beams with a small slenderness ratio is guaranteed by the approach.

The proposed approach in this paper can also be transplanted to discuss the qualitative property of vibrating multibearing beams. The natural vibration of multibearing beams shows a qualitative property closely related to P1-P5,which is known as the OP of multibearing beams[9]. The corresponding property of FD multibearing beams has also been proved by matrix factorization with an assumption that the sizes of the two grids clamping any inner support do not vary acutely[9]. However,through the present approach,the OP of a few kinds of discrete multibearing beam models,such as FD and FE beams,can be obtained by performing a simple transformation,and the mesh size assumption is no longer necessary.

The entire discussion of the OP for discrete beam models with lumped masses will be given as follows. In Section 2,the preliminary knowledge is given. For easy understanding,a proof to the SOP of analytic beams presented by Gantmacher and Krein[1] is shortly reviewed. Based on their concept,a unified sufficient condition for the SOP of discrete beams is proposed in Section 3. In Section 4,many practical models are shown to satisfy the proposed condition, such as 5-point FD beams,2-node HR-FE beams,most 2-node PE-FE beams,and most 2-node Timoshenko FE beams. In Section 5,the OP of multibearing beams is briefly introduced and discussed by the proposed approach,and other discrete multibearing beam models are found to possess the property besides the FD ones discussed in literatures. In Section 6,the main conclusions of this paper are reviewed. 2 Preliminary

Before discussion,we first review a few lemmas and the establishment of the SOP/OP of analytic beams in Ref. [1].

Lemma 1[1, 10] One-dimensional continuums possess the OP for any mass distribution (continuous and/or lumped) possessing the SOP. The discrete models of one-dimensional con- tinuums with any lumped mass possessing the SOP possess the OP.

For analytic beams,the SOP can be derived by analyzing the governing equation with the following result.

Lemma 2 If a continuous function f (x) has k roots in an interval (a,b),then its integral

has at most (k + 1) roots in [a,b] for any arbitrary constant C.

Take a simply supported beam for example,whose governing equation is

where q(x) is the distributed force,and EI(x) is the bending stiffness. Integrating (7) twice yields

in which M (x) is the bending moment distribution in physics.

Suppose that the beam is under the action of k concentrated forces. Because the shear force varies only at the place where there is a concentrated force,the bending moment M (x) is a polyline with (k + 1) segments,and thus has at most (k + 1) roots. Since two of these roots are at the ends according to the boundary condition,there are (k − 1) roots in (0,L). According to Lemma 2,there are at most k roots of w '(x) in [0,L],and there are (k + 1) roots of w(x) in the close interval. Since two of the roots of w(x) are at the ends because of the boundary condition,w(x) reverses its sign no more than (k − 1) times. Figure 2 shows the case k = 3,in which the sign reversion numbers of M (x),w '(x),and w(x) are 2,3,and 2,respectively.

Fig. 2. Beam under action of three concentrated forces.

The beams under other boundary conditions can also be discussed similarly. As shown in Table 1,for the beams with the last two boundary conditions,the conclusions are obtained by analyzing their conjugated beams[7, 8] whose corresponding boundary conditions are given as fixed-fixed and fixed-hinged,respectively.

Table 1. Maximum number of isolated roots under k concentrated forces

Although the analytic beams possess the SOP,it is not obvious that their discrete coun- terparts reflect the property,especially for the FE models which involve a weighted residual procedure. The deduction for analytic beams does not work directly for discrete beams,which is not governed by differential equations. A counter example is given to demonstrate that in Ref. [12]. 3 Sufficient condition for SOP/OP of discrete beams

Before the main theorem of this paper is given,let us review a few local qualitative properties of a beam segment [a,b] subject to the concentrated moments Ma and Mb and the concentrated forces fa and fb at its two ends,as shown in Fig. 3. The equilibrium delivers

Fig. 3. Static balanced beam segment under action of moments and forces.

If Ma> 0 and Mb> 0,then M (x) > 0 in [a,b]. Integrating (8) yields

in which θa and θbare the slopes at a and b,respectively. This inequality indicates that if Ma and Mb are both positive,the average slope defined as satisfies

Clearly,if Ma and Mb are both negative in turn,the inequality (11) reverses its direction.

Besides,the semi-positivity of the strain energy reserved in the segment leads to

The three conclusions based on the beam theory,i.e.,the balance (9),the deformation (11),

Theorem 1 Suppose a discrete beam with N nodes {xi}N1(0 = x1< x2< • • • < xN= L) is under the action of any given set of nodal forces. The discrete beam possesses the SOP if the corresponding nodal sequences of the deflection {wi}N1, the rotation angle {θi}N1, and the bending moment {Mi}N1 satisfy the following three local static properties:

(i) If the bending moments at any two successive nodes Mi and Mi+1 are both non-negative, then

where both equalities in (13) hold simultaneously only when Mi and Mi+1 both vanish; if Mi and Mi+1 are both non-positive,the inequality signs in (13) reverse (see Fig. 4 for example); for FD models,in which the nodal rotation θi is not given explicitly,the nodal rotation can be defined as

Fig. 4. Deformation of beam segment under action of two identical-sign nodal moments.

(ii) The strain energy of any beam segment [xi,xi+1] is non-negative,i.e.,

(iii) The internal shear forces Q(x+i ),Q(xi+1 ) and bending moments Mi,Mi+1 are statically balanced in (x+i ,xi+1 ).

Furthermore,if the concentrated mass matrix is employed,it possesses the OP.

As shown before,all assumptions are derived from an analytic beam segment. The proof generally follows the concept in Section 2,where the monotonicity of w(x) is indicated by the slope θ(x). However,the slope sequence {θi}N1is no longer directly related to the deflection sequence {wi}N1for discrete beams. Instead of {θi}N1,the average slope sequence {θi,i+1}}N−11 plays the role of θ(x) in turn. Therefore,a bridge between {θi}N1and {θi,i+1}}N−11 is the key of the proof.

Proof For convenience,simply supported beams are considered again. Beams with other boundary conditions can be discussed similarly.

Assume that a simply supported discrete beam is under the action of k concentrated forces. Because of the equilibrium condition (iii) in Theorem 1,there are at most (k+1) monotonic sub- sequences of {Mi}N1. Except the two ending ones,in each monotonic sub-sequence of {Mi}N1 , there is at most one sign reversion. Thus,the sequence {Mi}N1can be divided into p (p≤k) sub-sequences,i.e.,{Mi}nrnr+1−1(r = 1,2,3,• • • ,p),such that each sub-sequence {Mi}nrnr+1−1is sign-definite (see Fig. 2(b),where the three sign-definite sub-sequences are separated by vertical dash lines).

If a sub-sequence {Mi}nrnr+1−1 is non-negative,Assumption (i) in Theorem 1 yields θnr≤ θnr+1 ≤• • • ≤θnr+1−1, i.e.,the sub-sequence {θi}nrnr+1−1 increases. Similarly,if the sequence {Mi}nrnr+1−1 is non-positive,the sub-sequence {θi}nrnr+1−1 decreases. Particularly,if Mj−1Mj<0 and MjMj+1<0 appear simultaneously for a node j,we say that Mj forms a sign-definite sub-sequence of {Mi}N1 . In that case,we define θj as a monotonic sub-sequence of {θi}N1 (see Fig. 5(b)). Anyhow,the sequence {θi}N1 can be partitioned into exactly p monotonic sub- sequences corresponding to the p sign-definite sub-sequences of {Mi}N1 ,and the monotonicity of the sub-sequences alternates (see Figs. 2(b) and 2(c),in which the monotonic sub-sequences of {θi}N1 are partitioned by vertical dash lines).

Extending these monotonic sub-sequences of {θi}N1 such that the boundary terms of any two successive sub-sequences overlay,we can still obtain p monotonic sub-sequences since the monotonicity of the original sub-sequences alternates. Now,in each of these sub-sequences, {θi}N1 reverses its sign at most once.

However,if there is a non-negative local minimum term θj,which is defined as a non- negative ending term of a decreasing sequence and a starting term of an increasing sequence, or a non-positive local maximum term,which can be defined similarly,then the two successive monotonic sub-sequences θjdo not reverse their signs,as shown in Fig. 5(a). Figure 5(b) shows a degenerated condition in which θjincreases according to its previous definition. Though the series decreases in the illustrated part,θjis a positive local minimum term. If there are totally c1non-negative local minima and non-positive local maxima under the k forces,then the sequence {θi}N1 reverses its sign no more than (p − 2c1) ( k − 2c1) times.

Fig. 5. Non-negative local minimum term of {θi }N1 (hollow points) and corresponding moment under normal and degenerated conditions.

Now,we claim that if

then either θi or θi+1 is a non-negative local minimum or a non-positive local maximum. In fact, the given conditions imply that θi+1 − {θi,i+1} θi− {θi,i+1} > 0. On the other hand,we have MiMi+1<0 (see Fig. 6). Otherwise,{θi,i+1} should be somewhere between θiand θi+1 according to Assumption (i) in Theorem 1. Taking the strain energy inequality (15) into account,we have

Fig. 6. Deflection of beam segment under action of two opposite-sign nodal moments.

Without loss of generality,suppose {θi,i+1}>0, θi≤0, and θi+1≤0. Then,(16) leads to Mi> 0 and Mi+1<0,i.e.,the series {θi}N1reverses its monotonicity from increasing to decreasing either at θior θi+1. Since θi≤0 and θi+1≤0,according to our definition,one of them is a non-positive local maximum (see Fig. 7).

Fig. 7. Sketch of nodal slopes and moments when θi < 0,θi+1 < 0,and {θi,i+1} > 0 (not degenerated condition).

Consider a sign-definite sub-sequence {θi}nj . If

holds for i = j,j + 1,• • • ,n − 1,then the average rotation sub-sequence {θi,i+1}n−1j is also sign-definite. But if {θi,i+1} ≠ 0 and {θi,i+1}(θi+ θi+1) ≤0 at some node {xi},the sequence {θi,i+1}n−1j reverses its sign at most twice for each such i. Since there are (k − 2c1) sign-definite sub-sequences of {θi}N1 ,the sequence {θi,i+1}N−11 reverses its sign no more than (k − 2c1+ 2c2) times,in which c2denotes the number of i such that {θi,i+1} ≠ 0 and {θi,i+1}(θi+ θi+1)≤0 with the condition θiθi+1≤0. As mentioned above,this happens only if θior θi+1 is a non-positive local maximum or a non-negative local minimum. Therefore,c2≤c1.

Up to now,we prove that the sequence {θi,i+1}N−11=reverses its sign at most k ( k − 2c1+ 2c2) times. Because of Lemma 3,the u-line as the integral of the average slope reverses its sign no more than (k − 1) times.

Now,it is explained that if k = 1,then wi≠ 0 (i = 2,3,• • • ,N − 1). In this condition, without loss of generality,assume that M1= MN= 0 and Mi> 0 (i = 2,3,• • • ,N − 1). Assumption (i) in Theorem 1 indicates that the equality in (13) never holds simultaneously. Therefore,

If w2≥0,then (17) leads to w3> 0,and further w4> 0,• • • ,wN> 0,which is against wN= 0. Thus,w2< 0. In the same way,wN-1 < 0. On the other hand,the series {θi,i+1}reverses its sign at most one time. Thus,the series {wi} first decreases and then increases. This together with w2< 0 and wN-1 < 0 leads to wi< 0 (i = 2,3,• • • ,N − 1).

According to the definition,the discrete beam possesses the SOP.

The SOP of the discrete beams under the first four boundary conditions in Table 1 can be established similarly,while that of the discrete beams with the last two boundary conditions can be discussed by employing a conjugated transformation first. 4 SOP/OP of discrete beam models

In this section,Theorem 1 is employed to establish the SOP/OP of different beam models by verifying that they satisfy the assumptions within Theorem 1. 4.1 SOP/OP of analytic beams only with concentrated masses

One of the discrete beam models is the analytic beam only with concentrated masses. Taking the points with concentrated masses as nodes,the assumptions of Theorem 1 are satisfied for this model. Therefore,the SOP/OP holds. 4.2 SOP/OP of FD beams

It has been proved that a 5-point FD beam is equivalent to a spring-mass-bar system (see Fig. 8)[14].

Fig. 8. Equivalent model to FD beam.

In the spring-mass-bar system,the nodal slope {θi } can be defined as (14). Substituting (14) into Hooks’ law,one obtains

It implies that Assumption (i) of Theorem 1 holds. Substituting (14) and Hooks’ law into (15),it follows that

Assumption (ii) of Theorem 1 holds. At last,the system in Fig. 8 is a physical system,and Assumption (iii) of balance holds,too.

Corollary 1 5-point FD beams (with lumped mass matrices) possess the SOP (and hence the OP) for arbitrary sectional parameters and meshes. 4.3 SOP/OP of HR-FE beams

Because of the characteristics of the FEM,Assumptions (ii) and (iii) of Theorem 1 are always satisfied. For the FE interpolations,the rotational angles are usually employed as nodal variables. Thus,there is no need to define additional nodal slopes {θi}N1.

Let I and J be the two ends of a single element. For most practical FE beam elements, the strain energy is determined by the nodal moments MI and MJ at the two ends if there is no load inside the element. Then,the relative rotational angles,i.e.,ΔθI= θIθI,J and ΔθJ= θJθI,J,which are work-conjugate with MIand MJ,are determined by MIand MJ, too. In other words,ΔθIand ΔθJdo not depend on wI and wJ. In that condition,Assumption (i) of Theorem 1 can be verified by examining a simply supported beam element and checking whether the corresponding flexibility matrix,i.e.,

is positive definite and satisfies

or equivalently,the stiffness matrix of the beam element is positive definite and satisfies

For a simply supported FE beam segment,the equilibrium equation derived by the Heilinger- Reissener (HR) principle can be expressed as

where EI (x) is the bending stiffness and is positive over the segment, is the length of the segment,and NF = (1 − ξ,−ξ) is the shape function of M (x). The flexibility matrix in (23) can be further expressed as

Obviously,every component of the matrix is sign-definite,and (21) is always satisfied for any positive bending stiffness EI (x).

Corollary 2 2-node HR-FE beams (with lumped mass matrices) possess the SOP (and hence the OP) for arbitrary sectional parameters and meshes. 4.4 SOP/OP of PE-FE beams

The equilibrium equation of a simply supported FE beam segment derived by the potential energy principle can be expressed as

Here,

The stiffness matrix in (25) can be further expressed as

Obviously,for the positive bending stiffness EI(ξ),(22) holds if and only if

Then,Corollary 2 leads to Corollary 3.

Corollary 3 2-node PE-FE beam models (with lumped mass matrices) possess the SOP (and hence the OP) if (27) holds for all of its elements.

Note that Corollary 3 is a sufficient condition,not a necessary one. But it includes most practical conditions since EI(ξ) can be linear,quadratic,or monotonic in each element. In fact, by illustrating the weight function of (27),i.e.,,in Fig. 9,one can easily find that the condition (27) fails only if EI (ξ) is averagely much larger in (1/3,2/3) than in [0,1/3] ∪ [2/3,1]. Obviously,this seldom happens in practical conditions.

Fig. 9. Weighting function in (27).
4.5 SOP/OP of FE Timoshenko beams

For analytic beams,only those based on the Euler beam theory show the SOP/OP. How- ever,in our proof to the SOP/OP of discrete beam models,the Euler-Bernoulli formula is not employed. Thus,the SOP/OP may exist for discrete beam models based on other theories. In this subsection,the SOP/OP of a few typical 2-node FE Timoshenko beam models is discussed by the proposed approach.

The stiffness matrix of a simply supported Timoshenko beam element can be derived via the HR principle or potential energy (PE) principle. For the former,the flexibility matrix is

in which GA(ξ) and k stand for the shear stiffness and the shear coefficient,respectively. Obviously,the inequality (21) is satisfied if and only if

Thus,we obtain Corollary 4.

Corollary 4 2-node HR-FE Timoshenko beams (with lumped mass matrices) possess the SOP (and hence the OP) if (29) holds for all of its elements.

For the Timoshenko element derived via the PE principle,we assume that EI and GA are element-wise constant for simplicity. If the element displacement w(ξ) is interpolated by cubic functions,while the cross sectional rotation θ(ξ) is interpolated by quadratic functions,the stiffness matrix of a simply supported element after static condensation is[15]

It can be found that (22) holds if and only if

Thus,for such a PE-FE Timoshenko beam,if (31) is satisfied for all of its elements,the SOP/OP holds.

When EI and GA are element-wise constant,(29) is equivalent to (31). Physically,the left-hand side of (31), ,is the slenderness ratio of the element. In practice,it is supposed to be small such that the beam theory is applicable. Thus,in practice,(31) is usually satisfied. 5 SOP/OP of discrete multibearing beams

The proposed approach can also be employed to discuss relevant qualitative properties of discrete multibearing beams. In fact,the qualitative property of the multibearing FD beams proposed in literature[9] also holds for any discrete multibearing beam satisfying the assumptions of Theorem 1. In this section,we shall take a beam with one overhang for example (see Fig. 10(a)) to demonstrate that.

Fig. 10. ue-line of overhang beam.

Let (ui,uo) be the displacement vector of the overhang beam,in which ui denotes the displacements of the nodes between the two supports,while uo is the displacement of the overhang nodes. If we employ generalized coordinates in the overhang ,which are the opposites of the nodal displacements (see the hollow points in Figs. 10(b) and 10(c)),and employ the original displacement coordinates between the two supports,then the vibration equation can be expressed as

where is the flexibility matrix with respect to ,and can be expressed as

Denote -line as the polyline connecting successive deflected free nodes (xi,i) (see Figs. 10(b) and 10(c)). It is found that for the FD models of such an overhang beam, is oscillatory. Thus,if the mode shapes in P1-P5 are replaced with -lines,P1-P5 also hold[9].

More generally,for a multibearing FD beam consisting of s spans, can be defined as

in which ui denotes the displacement vector in the ith span. Then,the -lines possess the same properties. This conclusion is known as the OP of multibearing beams due to its relation with P1-P5. By employing the -line,the concept of the SOP of multibearing beams is also obtained.

Now,the SOP/OP of the discrete overhang beam is established again under the assumptions of Theorem 1 by the proposed approach (see Fig. 10). It is sufficient to show that when the beam is under the action of k concentrated forces,the -line reverses its sign no more than (k − 1) times.

First,under the action of the k forces,the bending moments {Mi }N1 can be divided into no more than (k + 2) monotonic sub-sequences because the monotonicity of {Mi }N1 varies only at the (k + 1) points subjected to the forces (including the reacting force of the inner support).

Then,through exact the same progress as in the proof of Theorem 1,we have that {θi,i+1}N1 reverses its sign no more than (k + 1) times. Thus,the u-line can be divided into no more than (k + 1) monotonic sub intervals.

Finally,the sign reversion of the -line can be obtained by studying the sign reversion of the u-line. Under the k forces,the deflection of the overhang beam falls into one of the following categories:

(I) {θi,i+1} N1 does not reverse its sign across the inner support (see Fig. 10(b)). As the u-line can be divided into no more than (k + 1) monotonic sub intervals and vanishes at the outer support,it reverses its sign no more than k times. Because the u-line reverses its sign over the inner support in this condition,the -line does not reverse its sign over the support. The rest roots of the u-line are also roots of the -line,and vice versa. Thus,the -line reverses its sign one time less than the u-line does,i.e.,no more than (k − 1) times.

(II) {θi,i+1}N1reverses its sign over the inner support (see Fig. 10(c)). Then,the u-line does not reverse its sign in its two successive monotonic intervals clamping the inner support. Thus, the u-line reverses its sign no more than (k − 2) times. In this condition,the -line reverses its sign over the inner support. Thus,it reverses one time more than the u-line does,i.e.,at most (k − 1) times.

In a word,the -line reverses its sign no more than (k− 1) times subjected to k concentrated forces with the assumptions of Theorem 1. Therefore,we have the following corollary.

Corollary 5 A simply supported beam model with one overhang (and lumped mass matrix) possesses the SOP (and hence the OP) if the assumptions of Theorem 1 are satisfied.

In the same way,the SOP/OP of other multibearing beams can also be obtained under the assumptions of Theorem 1. 6 Conclusions

In this paper,a new approach is proposed to discuss the OP of discrete beams. It focuses on a static property closely related to the OP instead of studying the OP directly.

A unified proof of the SOP/OP of discrete beam models is given under three local assump- tions,which are obtained from the classical Euler-Bernoulli beam theory. Compared with the known approaches,the proof focuses on the static properties of a local segment instead of the whole models,and thus no stiffness matrix assembling or factorization is required. Benefiting from that,the verification of the SOP/OP is almost straight forward.

Through the proposed approach,2-node HR-FE beams with concentrated mass matrices and 5-point FD beams are verified to hold the SOP/OP unconditionally,i.e.,beams with arbitrary section parameters and grids always possess the SOP/OP. That agrees with the literatures[8, 10]. For 2-node PE-FE beams,we greatly widen the cases possessing the SOP/OP by providing a sufficient condition (27),which is satisfied by most practical PE-FE elements.

Besides,the proposed approach also shows that FE beams based on the Timoshenko beam theory possess this property,too,if the slenderness ratio is small in each element. This con- clusion corroborates that the results derived via the Euler beam theory and the Timoshenko beam theory are similar under the condition of small slenderness.

Moreover,the proposed approach is generalized to discuss the SOP/OP of multibearing beams. The property of beams with one overhang is also established uniformly with the same assumptions in Theorem 1.

References
[1] Gant m ach er, F. R . an d K rein , M. G. Oscillation Matrices and Kernels and Small Vibrationsof Mechanical Systems, State Publishing House for Technical-Theoretical Literature,New York,83–244 (1961)
[2] Gladwell, G. M. L. The inverse problem for the vibrating beam. Proceedings of the Royal Societyof London. Series A, Mathematical and Physical Science, 393, 277–295 (1984)
[3] Gladwell, G. M. L. and Gbadeyan, J. On the inverse problem of the vibrating string and rod.Quarterly Journal of Mechanics and Applied Mathematics, 38, 169–174 (1985)
[4] Gladwell, G. M. L. The inverse problem for the Euler-Bernoulli beam. Proceedings of the RoyalSociety of London. Series A, Mathematical and Physical Science, 407, 199–218 (1986)
[5] Gladwell, G. M. L., Willms, N. B., He, B., and Wang, D. How can we recognize an acceptablemode shape for a vibrating beam? Quarterly Journal of Mechanics and Applied Mathematics, 42,303–316 (1989)
[6] Glad well, G. M. L. Inverse Problems in Vibration, 2nd ed., Kluwer Academic Publishers, Dordrecht, 185–331 (2004)
[7] Wang, D. J., Chou, C. S., and Wang, Q. S. Qualitative properties of frequencies and modes ofbeams modeled by discrete systems. Chinese Journal of Mechanics, Series A, 19, 169–175 (2003)
[8] Wang, Q. S. and Wang, D. J. Difference discrete system of Euler beam with arbitrary supportsand sign-oscillatory property of stiffness matrices. Applied Mathematics and Mechanics (EnglishEdition), 19, 393–398 (2006) DOI 10.1007/s10483-006-0316-y
[9] Wang, Q. S., Wang, D. J., He, M., Zhang, L. H., and Qian, H. F. Some qualitative properties ofthe vibration modes of the continuous system of a beam with one or two overhangs. Journal ofEngineering Mechanics, 138, 945–952 (2012)
[10] Gladwell, G. M. L. Qualitative properties of finite element models, I: Sturm-Liouville systems.Quarterly Journal of Mechanics and Applied Mathematics, 44, 249–265 (1991)
[11] Gladwell, G. M. L. Qualitative properties of finite-element models, II: the Euler-Bernoulli beam.Quarterly Journal of Mechanics and Applied Mathematics, 44, 267–284 (1991)
[12] Zheng, Z. J., Chen, P., and Wang, D. J. Oscillation property of modes for FE models of bars andbeams (in Chinese). Journal of Shock and Vibrations, 31, 79–83 (2012)
[13] Zheng, Z. J., Chen, P., and Wang, D. J. Oscillation property of the vibrations for finite elementmodels of an Euler beam. Quarterly Journal of Mechanics and Applied Mathematics, 66, 587–608(2013)
[14] Gladwell, G. M. L. The approximation of uniform beams in transverse vibration by sets of masseselastically connected. Proceedings of the 4th U.S. Congress of Applied Mechanics, American Societyof Mechanical Engineers, New York, 169–172 (1962)
[15] Reddy, J. N. On locking-free shear deformable beam finite elements. Computer Methods in Applied Mechanics and Engineering, 149, 113–132 (1997)