Skip to main content
Erschienen in: Archive of Applied Mechanics 12/2022

Open Access 08.10.2022 | Original

Exact series solutions of composite beams with rotationally restrained boundary conditions: static analysis

verfasst von: Yuan Zhang, Lei Zhang, Sigong Zhang

Erschienen in: Archive of Applied Mechanics | Ausgabe 12/2022

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

The structural performance of composite beams is sensitive to load distribution as well as actual boundary conditions. Although the composite beam theory has been solidly established and exact solutions have been readily developed for various loading and boundary conditions, almost all of them are limited to classical boundary conditions (free, pinned and clamped) and there has been little discussion about the actual support conditions in real structures. The general representation of actual boundary conditions can be defined as rotationally restrained edges. In this research, an analytical model with exact series-type solutions was developed for composite beams with rotationally restrained edges. The model and displacement solutions were validated by other analytical methods and numerical results. The influence of rotational and end-slip restraints was investigated. It was found that the deflection and interface slip of composite beams are highly affected by both restraints introduced by the actual boundary conditions. The current model can be used as a benchmark for future design methods considering the realistic boundary conditions of composite beams.
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Composite structures combining dissimilar materials through various shear connectors are extensively implemented in civil engineering applications as floor/wall components and bridge decks. The most common form of composite elements in construction is steel–concrete composites, which has attracted tremendous research interests over last decades (e.g. [13]). In parallel, other types of composites including steel–timber and timber–concrete composites (TCC) have been increasingly developed as alternative solutions to traditional reinforced concrete and steel–concrete composites or refurbishment techniques for existing structures. More recently, with rapidly emerging movement towards sustainable construction and significant advances in engineered wood products and mass timber panels, TCC systems have been increasingly used in new structures and present numerous advantages over timber-only or reinforced concrete floors [4, 5]. Such a trend substantially fuels the need to improve the existing analysis procedure and design methods for composite beams.
The behaviour of composite members is complex as the shear connectors generally permit relative slip between individual components of the member and only partial composite action can be achieved [6]. Since 1940s, a considerable amount of literature has been published on the composite beam theory from linear elastic models [711] to nonlinear plastic methods [1215]. A detailed background to the composite beam theory and the literature in the field can be found in [16]. Based on earlier theoretical models, simplified design methods for composite beams with partial interaction have been developed. In timber engineering, simplified design approaches such as Gamma method in Eurocode 5 (Annex B) [17] and Newmark–McCutcheon formula [8, 18, 19] have been commonly used in practical designs. The Gamma method, which was recommended by Ceccotti [4] for linear-elastic analysis, was developed based on the simple linear elastic model of Möhler [9] for timber–timber composite beams and derived from the exact solution for a simply supported composite beam subjected to a sinusoidal distributed load. A closely related deduction can be found in [20]. According to this approach, the “Effective Stiffness” concept is introduced and a “Connection Efficiency Factor” (\(\gamma \)) is used to account for the composite efficiency of shear connectors, with \(\gamma =1\) representing fully composite action, and \(\gamma =0\) no connection at all. It also provides an elegant solution with acceptable accuracy for simply supported beams under a uniformly distributed load or concentrated load. In addition, the coefficient of \(\gamma \) renders a favourable illustration of the nature of the problem [21]. In parallel, Newmark et al. [8] developed similar elastic composite beam model for steel–concrete composite beams and derived the exact solution for simply supported composite beam under a concentrated load. The expression was simplified by McCutcheon [18, 19] for timber components, which is denoted as Newmark–McCutcheon formula herein. Surprisingly, these two independent methods (Gamma method and Newmark–McCutcheon formula) produced almost the same result. The Newmark–McCutcheon formula has been widely used in North America for calculating effective composite bending stiffness such as ATC Design Guide 1 [22].
However, increasing concerns have been raised about the validity and accuracy of the well-known simplified methods for arbitrary boundary conditions other than the simple support. Exact solutions and simplified formula have been readily developed on various boundary conditions for uniform composite beams [6, 11, 16, 2325]. It has been found that the error of the Gamma method in the clamped–clamped case is around 20% [25], even up to almost 30% [16]. For extensions of the Gamma method to beams with other boundary conditions, Eurocode 5 [17] recommends 0.8 of the relevant span for continuous beams and twice the length for cantilever beams. Furthermore, a general and more accurate way of determining the effective beam length was proposed by Girhammber [16] for arbitrary boundary and loading conditions.
The newly proposed design methods/formulas for so-called arbitrary boundary conditions [16, 23] merely cover an arbitrary combination of free, pinned (simply supported) and clamped conditions. These conditions are idealized extreme conditions that do not reflect actual boundary conditions in structures. In fact, it is prohibitively difficult, if not impossible, to achieve clamped conditions, such that all displacement and rotation components vanish, at supports [26], especially for timber since it is relatively ‘compressible’ in the across grain direction [27]. In most simply supported cases, some amount of rotational stiffness can be introduced by the boundary conditions. For instance, platform-frame construction in timber engineering often result in large restraints at end supports due to weight of upper stories [28]. In reality, the boundary conditions are normally semi-rigid, which can be defined as restrained conditions against rotation. Nevertheless, very little attention has been paid to the actual support conditions for composite beams.
The current research aims to fill this gap by developing an analytical model for static analysis of composite beams with rotationally restrained edges and deriving an exact series solution. Series type solutions (e.g. Navier solution and Levy solution) have been extensively used for simple beam and plate structures from the beginning of the last century [29] and have also been applied to composite beams [30], stressed-skin panels [31] and stiffened plates [32]. The benefit of series-type solutions is that they can be easily generalized for various loading scenarios and boundary conditions. While the application of series type solutions is straightforward to various loading cases, it turns out to be rather tedious to satisfy complex boundary conditions. Accordingly, rigorous series-type solutions are widely applied for simply supported conditions. Although laborious superposition techniques can be used [33] for complex boundary conditions, it is common practice to resort to approximate techniques based on Ritz–Galerkin methodology or direct numerical analysis [34]. It is the use of Stokes’ transformation that releases those complex boundary conditions [35] leading to a simple, unified and systematic procedure dealing with beams and plates with arbitrary boundary conditions [3638].
In the current study, the series-type solutions with the employment of Stokes’ transformation were extended to composite beams with rotationally restrained edges. The fourth-order governing differential equations for the displacement of partially composite beams with rotationally restrained edges were first derived. Then, by applying Stokes’ transformation, exact series solutions were deducted. The validity and accuracy of the proposed solution were assessed by comparing predicted deflections and slips with other analytical models and numerical results. Finally, the parametric studies were conducted to examine the influence of different boundary restraints.

2 Formulation

2.1 Governing equations

The present model is based on Euler–Bernoulli beam theory (for each element of the composite beam). All materials (including shear connectors) are assumed to remain within the linear-elastic range until elements fail. The shear connectors between two elements are regarded as continuous and uniformly distributed along the longitudinal direction. Elements are subjects to bending and no axial loads are applied.
A typical composite beam with two sub-elements of different materials, as shown in Fig. 1, is studied. \(E_i\), \(A_i\) and \(I_i\) (\(i=1,\ 2\)) are Young’s modulus, cross-sectional area and moment of inertia of two elements. L and H denote the length and height of the composite beam, respectively. The symbol \(h_1\) denotes the distance from the centroid of the upper element to the interface between the elements. Similarly, the distance from the centroid of the lower element to the interface is represented by \(h_2\). The subscripts “1” and “2” will refer to the upper and lower elements, respectively. The x-axis of the coordinate system is located in the centroid of the composite section. The coordinate origin is placed at the left support, and the positive direction of z-axis is downward. The displacements in the x- and z-directions are denoted as u and w, respectively.
The kinematical relation can be determined first. In the case that partial composite action is developed between two elements, the interface slippage presents. Based on the above assumptions, the geometrical relation of the slippage \(u_\mathrm{s}\), the deflection w and the axial displacements of two components, \(u_1\) and \(u_2\), is illustrated in Fig. 2. Thus, the relation between the slippage \(u_{\mathrm{s}}\) and the deflection w is
$$\begin{aligned} u_{\mathrm{s}}=u_2-u_1+w' h \end{aligned}$$
(1)
Then, equilibrium of forces can be established. Fig. 3 shows a free-body diagram of an infinitesimal element of length \(\mathrm{d}x\) with internal and external forces upon it. Bending moment, shear force and axial force are denoted by M, Q and N, respectively. Equilibrium equations for the element are
$$\begin{aligned} Q'= & {} -q \end{aligned}$$
(2a)
$$\begin{aligned} N_1+N_2= & {} 0 \end{aligned}$$
(2b)
$$\begin{aligned} M'= & {} Q \end{aligned}$$
(2c)
Considering the equivalence of the internal and external actions on the left-hand side of the free-body diagram in Fig. 3, it can be obtained
$$\begin{aligned} Q= & {} Q_1+Q_2 \end{aligned}$$
(3a)
$$\begin{aligned} N_1'= & {} -N_2'=-q_\mathrm{s} \end{aligned}$$
(3b)
$$\begin{aligned} M= & {} M_1+M_2-N_1 h \end{aligned}$$
(3c)
Assuming both layers have equal curvature, it yields
$$\begin{aligned} \frac{M_1}{E_1I_1}=\frac{M_2}{E_2I_2}=-w'' \end{aligned}$$
(4)
Thus, the internal axial forces can be expressed with the first derivative of displacements as:
$$\begin{aligned} N_1=E_1A_1u_1', \quad N_2=E_2A_2u_2' \end{aligned}$$
(5)
The first governing equation can be obtained by combining Eqs. (2b) and (5) as:
$$\begin{aligned} E_1A_1 u_1'+E_2A_2 u_2'=0 \end{aligned}$$
(6)
The relation of the interface slippage and the shear force of the shear connector is
$$\begin{aligned} q_6=ku_{\mathrm{s}} \end{aligned}$$
(7)
where k denotes the shear stiffness of the shear connector per unit length (unit: \(\mathrm{N/m}^2\)). Applying Eq. (3b), it gives
$$\begin{aligned} N'_1=-N'_2=-ku_{\mathrm{s}} \end{aligned}$$
(8)
By considering Eqs. (1), (5) and (8), the second governing equation can be derived as:
$$\begin{aligned} E_1A_1u_1''+k(u_2-u_1+hw')=0 \quad \mathrm{or} \quad E_2A_2u_2''-k(u_2-u_1+hw')=0 \end{aligned}$$
(9)
Finally, substituting Eqs. (4) and (5) in Eq. (3c) results in
$$\begin{aligned} M=-EI_ow''-E_1A_1 h u_1' \end{aligned}$$
(10)
where \(EI_o=E_1I_1+E_2I_2\). Then, substituting Eqs. (2c), (10) in (2a), it gives
$$\begin{aligned} EI_o w''''+E_1A_1 hu_1'''=q \end{aligned}$$
(11)
which is the third governing differential equation.

2.2 General boundary conditions

In order to simulate different boundary conditions, two types of springs were defined to characterize all possible intermediate end supports of a single-span composite beam from simply supported to fully clamped as shown in Fig. 4. Two external springs represent the rotational restraints, \(R_0\) and \(R_\mathrm{L}\), and two internal springs restrain the slip at the beam ends, \(K_0\) and \(K_\mathrm{L}\), denoted as end slip restraints herein (or end shear restraints in [4042]). The rigid end-slip restraint was found to have a significant effect on the behaviour of composite beams, especially when the intermediate shear connector stiffness between two members of a composite beam is small [42]. However, very little attention has been paid to the role of the end slip restraint in the existing composite beam theories.
Rigid shear restraints were defined in [40] to suppress the shear deformation at beam ends leading to plane section at composite beam ends as shown in Fig. 5b, which can be realized by a special support design (although quite rare in current practice). Numerical results [41] have shown that rigid shear restraints exert a stiffening action to the beam and significantly reduce the displacement [42]. In this study, a more general end-slip restraint was defined to characterize a broad range of slip conditions at the beam ends. For zero end-slip restraints, each member of a composite beam freely elongates leading to larger slip at beam ends. Similarly, infinite end-slip restraints imposes the end sections to remain plane (no slip), which was defined as rigid shear restraints in [40]. In addition, although the end-slip restraints are introduced together with the rotational restraints, they can be considered separately. That means even for simply supported beam (with free rotations at the beam ends), end-slip restraints can be added to improve the composite actions.
Six boundary conditions are required to solve the system of governing equations. The general boundary conditions (Fig. 4) can be expressed as:
$$\begin{aligned}&w(0)=w(L)=0 \end{aligned}$$
(12a)
$$\begin{aligned}&M(0)=R_0w'(0), \quad M(L)=-R_\mathrm{L}w'(L) \end{aligned}$$
(12b)
$$\begin{aligned}&u_\mathrm{s}(0)=-N_1(0)/K_0, \quad u_\mathrm{s}(L)=N_1(L)/K_\mathrm{L} \end{aligned}$$
(12c)
It should be noted that \(u_\mathrm{s}(0)\) is negative because the element 1 moves in the negative x-direction and undergoes negative displacement. Substituting Eq. (8) in Eq. (12c) gives
$$\begin{aligned} N_1'(0)/k=N_1(0)/K_0,\quad -N_1'(L)/k=N_1(L)/K_\mathrm{L} \end{aligned}$$
(13)
Substituting Eqs. (10) and (5) in Eqs. (12b) and (13), respectively, the boundary conditions can be expressed in the form of displacement, rotation, and slip as:
$$\begin{aligned}&\displaystyle w(0)=w(L)=0 \end{aligned}$$
(14a)
$$\begin{aligned}&\displaystyle EI_ow''(0)+E_1A_1hu_1'(0)=R_0w'(0),\quad EI_ow''(L)+E_1A_1hu_1'(L)=-R_\mathrm{L}w'(L) \end{aligned}$$
(14b)
$$\begin{aligned}&\displaystyle u'_1(0)= \frac{K_0}{k}u''_1(0),\quad u'_1(L)= -\frac{K_\mathrm{L}}{k}u''_1(L) \end{aligned}$$
(14c)

3 Exact series solution

The displacement functions can be expressed as Fourier series of the form:
$$\begin{aligned}&w(x)=\sum _{m=1}^{\infty } A_m \sin \alpha _m x, \quad 0 \leqslant x \leqslant L \end{aligned}$$
(15)
$$\begin{aligned}&u_1(x)=B_0+\sum _{m=1}^{\infty } B_m \cos \alpha _m x, \quad 0 \leqslant x \leqslant L \end{aligned}$$
(16)
$$\begin{aligned}&u_2(x)=C_0+\sum _{m=1}^{\infty } C_m \cos \alpha _m x, \quad 0 \leqslant x \leqslant L \end{aligned}$$
(17)
where \(\alpha _m=m\pi /L\) and \(A_m\), \(B_m\) and \(C_m\) are the Fourier coefficients. In general, the applied load can be given by
$$\begin{aligned} q(x)=\sum _{m=1}^{\infty } Q_m \sin \alpha _m x, \quad 0 \leqslant x \leqslant L \end{aligned}$$
(18)
where
$$\begin{aligned} Q_m=\frac{2}{L}\int _{0}^{L } q(x)\sin \alpha _m x \mathrm{d}x \end{aligned}$$
(19)
With a given external force q(x), these Fourier coefficients can be determined by substituting the displacement functions [Eqs. (15), (16) and (17)] in the governing differential equations of Eqs. (6), (9) and (11). The derivation process is presented as follows.
In particular, the derivative of the displacement functions must be carefully handled because term-by-term differentiation may not be valid. A special differentiation technique called Stokes’ transformation can be applied. Mathematical discussions and proofs on this topic can be found on pages 137–139 in [43] and on pages 375–377 in [44]. The first to fourth derivatives of Fourier sine series and cosine series can be found in [36, 45].
Based on Stokes’ transformation, the first derivative of w(x) can be expressed as:
$$\begin{aligned} w'(x)=\sum _{m=1}^{\infty }\alpha _mA_m \cos \alpha _m x, \quad 0 \leqslant x \leqslant L \end{aligned}$$
(20)
However, with regard to the first derivative of \(u_1\) and \(u_2\), term-by-term differentiation is only valid on open interval of \(0<x<L\). They can be expressed as:
$$\begin{aligned}&u'_1=-\sum _{m=1}^{\infty }\alpha _mB_m\sin \alpha _m x \end{aligned}$$
(21)
$$\begin{aligned}&u'_2=-\sum _{m=1}^{\infty }\alpha _mC_m\sin \alpha _m x \end{aligned}$$
(22)
Substituting Eqs. (21) and (22) in Eq. (6), the first governing equation can be expressed as:
$$\begin{aligned} \sum _{m=1}^{\infty }\alpha _m(E_1A_1B_m+E_2A_2C_m)\sin \alpha _m x=0 \end{aligned}$$
(23)
Since Eq. (23) is valid for any x (or Cantor’s theorem [46]), it can be concluded that
$$\begin{aligned} E_1A_1B_m+E_2A_2C_m=0 \end{aligned}$$
(24)
The second derivative of \(u_1\) on interval of \(0 \leqslant x \leqslant L\) can be obtained as:
$$\begin{aligned} u_1''=\frac{u'_1(L)-u'_1(0)}{L}+\sum _{m=1}^{\infty }\left\{ \frac{2}{L}\big [(-1)^m u'_1(L)-u'_1(0)\big ]-\alpha _m^2B_m\right\} \cos \alpha _m x \end{aligned}$$
(25)
Substituting Eqs. (16), (17), (20) and (25) in the second governing equation of Eq. (9), it gives
$$\begin{aligned} \begin{aligned}&\frac{E_1A_1}{k}\big [u'_1(L)-u'_1(0)\big ]+k(C_0-B_0) +\sum _{m=1}^{\infty }\Big \{\frac{2 E_1A_1}{L}\big [(-1)^mu'_1(L)-u'_1(0)\big ]-E_1A_1\alpha _m^2B_m\\&\quad +\,k(C_m-B_m)+kh\alpha _mA_m\Big \}\cos \alpha _m x=0 \end{aligned} \end{aligned}$$
(26)
Based on Cantor’s theorem [46], coefficients of the cosine series in Eq. (26) equal zero. Applying Eq. (24), it arrives
$$\begin{aligned} \frac{k}{E_1A_1}h\alpha _m A_m+\frac{2\big [(-1)^mu'_1(L)-u'_1(0)\big ]}{L}-\left( \alpha _m^2+\frac{k}{EA}\right) B_m=0 \end{aligned}$$
(27)
in which \(\frac{1}{EA}=\frac{1}{E_1A_1}+\frac{1}{E_2A_2}\) . Further simplification can be made by substituting the boundary conditions of Eq. (14c), which results in
$$\begin{aligned} \frac{k}{E_1A_1}h\alpha _m A_m+2\sum _{i=1}^{M}D_{\mathrm{im}}\alpha _i^2B_i-\left( \alpha _m^2+\frac{k}{EA}\right) B_m=0 \end{aligned}$$
(28)
The derivation of \(D_{\mathrm{im}}\) is shown in 6.
At last, the fourth derivative of w and the third derivative of \(u_1\) are given by
$$\begin{aligned}&w''''(x)=-\sum _{m=1}^{\infty }\alpha _m\left\{ \frac{2}{L}\big [(-1)^mw''(L)-w''(0)\big ]-\alpha _m^3A_m\right\} \sin \alpha _m x \end{aligned}$$
(29)
$$\begin{aligned}&u_1'''(x)=-\sum _{m=1}^{\infty }\alpha _m\left\{ \frac{2}{L}\big [(-1)^m u_1'(L)-u_1'(0)\big ]-\alpha _m^2B_m\right\} \sin \alpha _m x \end{aligned}$$
(30)
in which \(0<x<L\). After substituting the load function of Eq. (18) and displacement derivatives of Eqs. (29) and (30) in Eq. (11), the first governing equation can be expressed as:
$$\begin{aligned} \begin{aligned}&\sum _{m=1}^{\infty }\Big \{-EI_o\alpha _m\frac{2}{L}\big [(-1)^mw''(L)-w''(0)\big ] -E_1A_1h\alpha _m\frac{2}{L}\big [(-1)^m u_1'(L)-u_1'(0)\big ]\\&\quad +\,EI_o\alpha _m^4A_m+E_1A_1h\alpha _m^3B_m-Q_m\Big \}\sin \alpha _m x=0 \end{aligned} \end{aligned}$$
(31)
Based on Cantor’s theorem [46], coefficients of the trigonometric series in Eq. (31) equal zero, which arrives
$$\begin{aligned} \begin{aligned}&-EI_o\alpha _m\frac{2}{L}\big [(-1)^mw''(L)-w''(0)\big ] -E_1A_1h\alpha _m\frac{2}{L}\big [(-1)^m u_1'(L)-u_1'(0)\big ] \\&\quad +EI_o\alpha _m^4A_m+E_1A_1h\alpha _m^3B_m=Q_m \end{aligned} \end{aligned}$$
(32)
Applying Eq. (14b), Eq. (32) can be rearranged as:
$$\begin{aligned} \alpha _m\frac{2}{L}\big [(-1)^{m}R_\mathrm{L}w'(L)+R_0w'(0)\big ]+EI_o\alpha _m^4A_m+E_1A_1h\alpha _m^3B_m=Q_m \end{aligned}$$
(33)
Using Eq. (20), it can be obtained that
$$\begin{aligned} w'(L)=\sum _{m=1}^{\infty }(-1)^m\alpha _mA_m, \quad w'(0)=\sum _{m=1}^{\infty }\alpha _mA_m \end{aligned}$$
(34)
Substituting Eq. (34) into Eq. (33), it gives
$$\begin{aligned} EI_o\alpha _m^4A_m+\frac{2}{L}\alpha _m\sum _{i=1}^{\infty }\left[ (-1)^{i+m}R_\mathrm{L}+R_0\right] \alpha _iA_i+E_1A_1h\alpha _m^3B_m=Q_m \end{aligned}$$
(35)
For convenience, a non-dimensional parameter, r, was introduced herein to replace the rotational restraint stiffness, R, as
$$\begin{aligned} r=\frac{1}{1+\frac{3EI_o}{RL}} \end{aligned}$$
(36)
in which r is so-called end fixity factors, which is widely used in semi-rigid steel frame connections [38, 47]. In particular, \(r=0\) represents the simply supported condition and \(r=1\) is used for the fully clamped. Accordingly, Eq. (35) can be expressed as:
$$\begin{aligned} EI_o\alpha _m^4A_m+\frac{6EI_o\alpha _m}{L^2}\sum _{i=1}^{\infty }\left[ (-1)^{i+m}\frac{r_\mathrm{L}}{1-r_\mathrm{L}}+\frac{r_0}{1-r_0}\right] \alpha _iA_i+E_1A_1h\alpha _m^3B_m=Q_m \end{aligned}$$
(37)
Equations (28) and (37) are two infinite systems of linear equations with respect to coefficients \(A_m\) and \(B_m\). (Here, the subscripts were set as the same for convenience.) The infinite number of equations can be truncated to finite terms, say M, for each coefficients. (The truncated terms for \(A_m\) and \(B_m\) can be different as well.) Then, it will produce 2M equations with 2M unknown coefficients, which can be easily solved. Once coefficients \(B_m\) have been determined, \(C_m\) can also be calculated based on Eq. (24). As a result, all exact displacements, w(x), \(u_1(x)\) and \(u_2(x)\), can be determined. In addition, the slip can be obtained by using Eq. (1). With all these displacements known, other moments and inner forces can easily be derived, which are not presented here.
The results are theoretically exact when the number of terms is infinite. Convergent solutions with sufficient accuracy can be obtained by a finite number of terms. The larger the number of terms used in a particular analysis, the more accurate the solution will be. No generalizations can be made concerning the choice of an adequate number of terms, as this will vary depending upon the load distribution and boundary restraints. Zhang and Xu [38] have shown that this type of solution converges rapidly for orthotropic plates with rotationally restrained edges. A similar fast convergence can also be observed for the composite beam in this study. For all results presented in this study, M is taken to be 1000.

4 Validations with other models

In this section, the proposed series-type solution was validated and compared with other existing composite beam models by considering specific supporting conditions and loading patterns.

4.1 Reduce to the simply supported condition

The preceding sections are devoted to composite beams with general boundary conditions which include simply supported and fully clamped as well as their intermediate cases. Such general expressions can be reduced to the simply supported case by assuming zero rotational (\(R_0=R_\mathrm{L}=0\)) and end-slip (\(K_0=K_\mathrm{L}=0\)) restraints in Eqs. (35) and (28). Then, Eqs. (35) and (28) can be reduced to the following form:
$$\begin{aligned} EI_o\alpha _m^4A_m+E_1A_1h\alpha _m^3B_m=Q_m \end{aligned}$$
(38a)
$$\begin{aligned} \frac{kh}{E_1A_1}\alpha _mA_m-\left( \alpha _m^2+\frac{k}{EA}\right) B_m=0 \end{aligned}$$
(38b)
Fourier coefficients \(A_m\) and \(B_m\) can easily be solved from the above set of equations. Then, the exact displacement of composite beam under arbitrary transverse loads is given by
$$\begin{aligned} w(x)=\sum _{m=1}^{\infty } \frac{1}{EI_o+\frac{kh^2}{\alpha _m^2+k/EA}}\frac{Q_m}{\alpha _m^4}\sin \alpha _m x \end{aligned}$$
(39)
The value of constant \(Q_m\) in Eq. (39) was defined in Eq. (19), which is dependent upon the type of load distribution. The most common types of load distribution are uniform distributed loads and concentrated loads. For a uniform distributed load, q, along the span,
$$\begin{aligned} Q_m=\frac{2q}{m\pi }(1-(-1)^m) \end{aligned}$$
(40)
For a concentrated load, P, as shown in Fig. 6
$$\begin{aligned} Q_m=\frac{2P}{L}\sin \alpha _m e \end{aligned}$$
(41)
where e is the distance of the load from left-hand support.
Smith [30] derived exact series-type solutions for deflections of built-up composite beams by expressing the moment and sectional normal forces due to applied load as two Fourier series. The present solution in Eq. (39) is equivalent to those results by Smith [30] but in a more concise form.

4.2 Deduction of Gamma method

Van der Linden [20] presented a closed-form solution for a simply supported composite beam with a sinusoidal load along its length and deduced equivalent expressions of the Gamma method, which can also be derived by the present series-type solution. The sinusoidal load can be defined as:
$$\begin{aligned} q(x)=q_0\sin \alpha _1 x \end{aligned}$$
(42)
where \(\alpha _1=\pi /L\), which is the value of \(\alpha _m\) in Eq. (18) when \(m=1\). In this manner, the displacement function can be also expressed in a sinusoidal shape as:
$$\begin{aligned} w(x)=A\sin \alpha _1 x \end{aligned}$$
(43)
From Eq. (39), the constant A can be derived as:
$$\begin{aligned} A=\frac{q_0L^4}{\pi ^4EI_{\mathrm{eff}}} \end{aligned}$$
(44)
in which \(EI_{\mathrm{eff}}\) is the effective flexural stiffness and is expressed as:
$$\begin{aligned} EI_{\mathrm{eff}}=EI_o+\frac{kh^2}{\left( \frac{\pi }{L}\right) ^2+\frac{k}{EA}} \end{aligned}$$
(45)
Equations (45) differ from the way Van der Linden presented in [20], and the results are nevertheless the same. Further rearrangements can be carried out leading to the prominent expression of Gamma method proposed in the Annex B of the Eurocode 5 [17]
$$\begin{aligned} EI_{\mathrm{eff}}=EI_o+\gamma E_1A_1 (h-a)^2+E_2A_2a^2 \end{aligned}$$
(46)
in which \(\gamma =[1+\pi ^2E_1A_1/(kL^2)]^{-1}\) is the shear connection reduction factor and a is the distance from the centroid of the bottom element to the neutral axis of the composite section, which is given by
$$\begin{aligned} a=\frac{\gamma E_1A_1}{\gamma E_1A_1+E_2A_2}h \end{aligned}$$
(47)

5 Numerical examples and parametric studies

In this section, a 4 m single-span timber–concrete composite beam reported by Girhammer in [16] was studied. The cross section and dimensions of the beam are shown in Fig. 7. The material properties are \(E_1=12\, \mathrm{GPa}\) and \(E_2=8\, \mathrm{GPa}\). The slip modulus (i.e. stiffness of shear connectors) is \(k=50\,\mathrm{N/mm/mm}\). For this beam, the influence of end slip restraints was first discussed, followed by deflection predictions and comparisons with other models.

5.1 Influence of end slip restraints

Based on the expression of \(D_{\mathrm{im}}\) in Eq. (A.6), a non-dimensional parameter, \(\lambda \), was defined herein for end-slip restraints as:
$$\begin{aligned} \lambda _0=\frac{K_0}{kL},\quad \lambda _\mathrm{L}=\frac{K_\mathrm{L}}{kL} \end{aligned}$$
(48)
where k is the slip modulus, \(K_0\) and \(K_\mathrm{L}\) are stiffness of end-slip restraints at two ends, respectively, and the associated parameters are \(\lambda _0\) and \(\lambda _\mathrm{L}\). For simplicity, \(\lambda _0=\lambda _\mathrm{L}=\lambda \) in the following analysis. \(\lambda \) represents the relative importance of the end-slip restraints compared to the shear connectors. A very large variation in \(\lambda \) is possible: \(\lambda \) equal to zero corresponds to no end-slip restraints, and a higher \(\lambda \) results in smaller slip at beam ends.
First, consider the beam in Fig. 7 simply supported and loaded by a uniformly distributed load, 1 kN/m. For this beam, four different end-slip restraints were studied and the associated values of \(\lambda \) were 0, 0.1, 1 and 10. The beam displacements and slips were determined for the four cases as illustrated in Fig. 8. It can be found that both displacements and slips were notably affected by the end-slip restraints. More importantly, the slips at beam ends were close to zero when \(\lambda \geqslant 1\), extremely close for \(\lambda =10\), as shown in Fig. 8b. Thus, it can be imagined that when the non-dimensional end-slip parameter \(\lambda \) is larger than 10, the relative slip at beam ends can be considered to be fully suppressed and the slip at other beam positions will almost be the same for different \(\lambda \).
Table 1
Maximum deflection of a single-span composite beam
Boundary conditions
Load distributions
Maximum deflections (mm)
  
Girhammar [16]
Present
Difference (%)
Pinned-pinned
Distributed load (1 kN/m)
7.56
7.526
− 0.4
Point load (1 kN)
3.08
3.061
− 0.6
Pinned-clamped
Distributed load (1 kN/m)
4.01
3.995
− 0.4
Point load (1 kN)
1.76
1.754
− 0.3
Clamped-clamped
Distributed load (1 kN/m)
2.32
2.323
0.1
Point load (1 kN)
1.16
1.159
− 0.1
Similarly, parametric studies were conducted for the same simply supported composite beam with different slip moduli. Four slip moduli were considered, namely \(k=1, 5, 25\) and \(50\,N/mm/mm\). For each slip modulus, two end-slip restraint scenarios were investigated: \(\lambda =0\) and \(\lambda =1\). In order to illustrate the influence of end-slip restraints varying with different slip moduli, both numerical results of two different \(\lambda \) were presented in the same figure (left-hand side is for \(\lambda =0\) and right hand side \(\lambda =1\)). In this manner, Fig. 9 demonstrates the beam displacements and slips for the composite beam with different slip moduli and end-slip restraints. It is clear that the effect of end-slip restraints is more evident in the case with smaller slip moduli (see \(k=1\,\mathrm{N/mm/mm}\)).

5.2 Influence of end support restraints

Not only the simply supported boundary (pinned-pinned) conditions, but also pinned–clamped and clamped–clamped conditions were analysed and compared with results reported in [16]. For the simply supported case, r and \(\lambda \) were both set as zero. The clamped cases were simulated by setting \(r=0.999\) [\(r=1\) results in a zero denominator in Eq. (37)] and \(\lambda =10{,}000\) (far enough for end-slip restraints). Two transverse loading scenarios were examined: uniformly distributed loading and mid-span concentrated loading. The results and comparisons are presented in Table 1. It can be seen that the present results obtained from exact series solutions are in excellent agreement with results in [16]. In addition, the displacement and slip curves for the concentrated loading case are presented in Fig. 10.
Since very little results can be found for composite beams with rotational restrained edges, no comparison was conducted for those restrained cases. For demonstration, the displacements and slips of restrained composite beams (\(r=0.5\)) with different end-slip restraints (\( \lambda =0, 0.1\ \& \ 1\)) are presented in Fig. 11 as well as the simply supported case (\( r=0\ \& \ \lambda =0\)) and the clamped case (\( r=0.999\ \& \ \lambda =10{,}000\)). It can be observed that the deflection and slip of composite beams are highly dependent on boundary restraints. More importantly, the end-slip restraints have a marginal influence on the deflection but a major effect on the slip.

6 Conclusions

The behaviour of composite beams is governed by load distributions and actual boundary conditions. Existing solutions normally deal with classical boundary conditions, namely free, simply supported and clamped, which often deviates from the actual conditions in reality. The realistic supporting conditions are generally semi-rigid, especially for timber structures since it is notoriously difficult to achieve fully clamped conditions for timber elements. The current study first summarized widely used design methods for composite beams both in Europe and North America, and concluded that they are basically the same method. Then, an analytical model for composite beams with rotationally restrained edges was developed, and its exact series solutions were derived by using Stokes’ transformation.
By comparing them to other simple models, it has been shown that the recently proposed model can be reduced to the simple models by setting the end restraints to zero. The parametric studies revealed that both rotational and end-slip restraints significantly influence the deflection and interlayer slip of composite beams. Particularly, the effect of end-slip restraints is more evident when the slip modulus is small, and is notable for slips but marginal for deflections.
In the future, the proposed model can be useful as a benchmark for future analysis such as finite-element modelling. Simple design formulas can also be developed for hand-calculation design solutions.

Declarations

Conflict of interests

The authors declared no potential conflicts of interest with respect to the research, authorship and/or publication of this article.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Anhänge

Appendix A: Derivation of \(D_{\mathrm{im}}\)

By using Eqs. (14c) and (25), \(u'_1(0)\) and \(u'_1(L)\) can be expressed as:
$$\begin{aligned}&u'_1(0)= \frac{K_0}{kL}\big [u'_1(L)-u'_1(0)\big ]+\frac{K_0}{k}\sum _{m=1}^{\infty }\left\{ \frac{2}{L}[(-1)^m u'_1(L)-u'_1(0)\big ]-\alpha _m^2B_m\right\} \end{aligned}$$
(A.1)
$$\begin{aligned}&u'_1(L)= -\frac{K_\mathrm{L}}{kL}\big [u'_1(L)-u'_1(0)\big ]-\frac{K_\mathrm{L}}{k}\sum _{m=1}^{\infty }\left\{ \frac{2}{L}[u'_1(L)-(-1)^mu'_1(0)\big ]-(-1)^m\alpha _m^2B_m\right\} \end{aligned}$$
(A.2)
For numerical calculations, the sum of infinite terms in Eqs. (A.1) and (A.2) can be approximately evaluated by the partial sum of finite terms by truncating the infinite number of terms to finite terms, M. Then, a system of functions about \(u'_1(0)\) and \(u'_1(L)\) can be derived by rearranging Eqs. (A.1) and (A.2).
$$\begin{aligned}&\left[ 1+(2M+1)\frac{K_0}{kL}\right] u'_1(0)-(-1)^M\frac{K_0}{kL}u'_1(L)=-\frac{K_0}{k}\sum _{m=1}^{M}\alpha _m^2B_m \end{aligned}$$
(A.3a)
$$\begin{aligned}&(-1)^M\frac{K_\mathrm{L}}{kL}u'_1(0)-\left[ 1+(2M+1)\frac{K_\mathrm{L}}{kL}\right] u'_1(L)=-\frac{K_\mathrm{L}}{k}\sum _{m=1}^{M}(-1)^m\alpha _m^2B_m \end{aligned}$$
(A.3b)
\(u'_1(0)\) and \(u'_1(L)\) can readily be solved and \((-1)^m u'_1(L)-u'_1(0)\) can be expressed as:
$$\begin{aligned} \frac{(-1)^mu'_1(L)-u'_1(0)}{L}=\sum _{i=1}^{M}\left\{ (-1)^i\frac{(-1)^mA-B}{AD-BC}\frac{K_\mathrm{L}}{kL}-\frac{(-1)^mC-D}{AD-BC}\frac{K_0}{kL}\right\} \alpha _i^2B_i \end{aligned}$$
(A.4)
in which
$$\begin{aligned} A=1+(2M+1)\frac{K_0}{kL},\quad B=(-1)^M\frac{K_0}{kL} \end{aligned}$$
(A.5a)
$$\begin{aligned} C=(-1)^M\frac{K_\mathrm{L}}{kL}, \quad D=1+(2M+1)\frac{K_\mathrm{L}}{kL} \end{aligned}$$
(A.5b)
At last, \(D_{\mathrm{im}}\) is defined as:
$$\begin{aligned} D_{\mathrm{im}}=(-1)^i\frac{(-1)^mA-B}{AD-BC}\frac{K_\mathrm{L}}{kL}-\frac{(-1)^mC-D}{AD-BC}\frac{K_0}{kL} \end{aligned}$$
(A.6)
Literatur
1.
Zurück zum Zitat Wang, Y.C.: Deflection of steel–concrete composite beams with partial shear interaction. J. Struct. Eng. 124(10), 1159–1165 (1998)CrossRef Wang, Y.C.: Deflection of steel–concrete composite beams with partial shear interaction. J. Struct. Eng. 124(10), 1159–1165 (1998)CrossRef
2.
Zurück zum Zitat Oehlers, D.J., Bradford, M.A.: Elementary Behaviour of Composite Steel and Concrete Structural Members. Butterworth-Heinemann, Oxford (1999)CrossRef Oehlers, D.J., Bradford, M.A.: Elementary Behaviour of Composite Steel and Concrete Structural Members. Butterworth-Heinemann, Oxford (1999)CrossRef
3.
Zurück zum Zitat Johnson, R.P.: Composite Structures of Steel and Concrete: Beams, Slabs, Columns and Frames for Buildings, 3rd edn. Blackwell Publishing, Hoboken (2004)CrossRef Johnson, R.P.: Composite Structures of Steel and Concrete: Beams, Slabs, Columns and Frames for Buildings, 3rd edn. Blackwell Publishing, Hoboken (2004)CrossRef
4.
Zurück zum Zitat Ceccotti, A.: Composite concrete–timber structures. Prog. Struct. Mat. Eng. 4(3), 264–275 (2002)CrossRef Ceccotti, A.: Composite concrete–timber structures. Prog. Struct. Mat. Eng. 4(3), 264–275 (2002)CrossRef
5.
Zurück zum Zitat Yeoh, D., Fragiacomo, M., De Franceschi, M., Heng Boon, K.: State of the art on timber–concrete composite structures: literature review. J. Struct. Eng. 137(10), 1085–1095 (2011)CrossRef Yeoh, D., Fragiacomo, M., De Franceschi, M., Heng Boon, K.: State of the art on timber–concrete composite structures: literature review. J. Struct. Eng. 137(10), 1085–1095 (2011)CrossRef
6.
Zurück zum Zitat Girhammar, U.A., Pan, D.H.: Exact static analysis of partially composite beams and beam–columns. Int. J. Mech. Sci. 49(2), 239–255 (2007)CrossRef Girhammar, U.A., Pan, D.H.: Exact static analysis of partially composite beams and beam–columns. Int. J. Mech. Sci. 49(2), 239–255 (2007)CrossRef
7.
Zurück zum Zitat Stüssi, F.: Zusammengesetzte vollwandträger. Int. Assoc. Bridge Struct. Eng. (IABSE) 8, 249–269 (1947) Stüssi, F.: Zusammengesetzte vollwandträger. Int. Assoc. Bridge Struct. Eng. (IABSE) 8, 249–269 (1947)
8.
Zurück zum Zitat Newmark, N.M., Siess, C.P., Viest, I.M.: Tests and analysis of composite beams with incomplete interaction. Proc. Soc. Exp. Stress Anal. 9(1), 75–92 (1951) Newmark, N.M., Siess, C.P., Viest, I.M.: Tests and analysis of composite beams with incomplete interaction. Proc. Soc. Exp. Stress Anal. 9(1), 75–92 (1951)
9.
Zurück zum Zitat Möhler, K.: Über das tragverhalten von biegeträgern und druckstäben mit zusammengestezten querschnitten und nachgiebigen verbindungsmitteln. Ph.D. thesis, Technical University of Karlsruhe (1956) Möhler, K.: Über das tragverhalten von biegeträgern und druckstäben mit zusammengestezten querschnitten und nachgiebigen verbindungsmitteln. Ph.D. thesis, Technical University of Karlsruhe (1956)
10.
Zurück zum Zitat Goodman, J.R.: Layered wood systems with interlayer slip. Ph.D. thesis, University of California, Berkeley, Berkeley, CA, USA (1967) Goodman, J.R.: Layered wood systems with interlayer slip. Ph.D. thesis, University of California, Berkeley, Berkeley, CA, USA (1967)
11.
Zurück zum Zitat Girhammar, U.A., Gopu, V.K.: Composite beam–columns with interlayer slip-exact analysis. J. Struct. Eng. 119(4), 1265–1282 (1993)CrossRef Girhammar, U.A., Gopu, V.K.: Composite beam–columns with interlayer slip-exact analysis. J. Struct. Eng. 119(4), 1265–1282 (1993)CrossRef
12.
Zurück zum Zitat Salari, M.R., Spacone, E., Shing, P.B., Frangopol, D.M.: Nonlinear analysis of composite beams with deformable shear connectors. J. Struct. Eng. 124(10), 1148–1158 (1998)CrossRef Salari, M.R., Spacone, E., Shing, P.B., Frangopol, D.M.: Nonlinear analysis of composite beams with deformable shear connectors. J. Struct. Eng. 124(10), 1148–1158 (1998)CrossRef
13.
Zurück zum Zitat Frangi, A., Fontana, M.: Elasto-plastic model for timber–concrete composite beams with ductile connection. Struct. Eng. Int. 13(1), 47–57 (2003)CrossRef Frangi, A., Fontana, M.: Elasto-plastic model for timber–concrete composite beams with ductile connection. Struct. Eng. Int. 13(1), 47–57 (2003)CrossRef
14.
Zurück zum Zitat Zhang, C., Gauvreau, P.: Timber–concrete composite systems with ductile connections. J. Struct. Eng. 141(7), 04014179 (2015)CrossRef Zhang, C., Gauvreau, P.: Timber–concrete composite systems with ductile connections. J. Struct. Eng. 141(7), 04014179 (2015)CrossRef
15.
Zurück zum Zitat Auclair, S.C., Sorelli, L., Salenikovich, A.: Simplified nonlinear model for timber–concrete composite beams. Int. J. Mech. Sci. 117, 30–42 (2016)CrossRef Auclair, S.C., Sorelli, L., Salenikovich, A.: Simplified nonlinear model for timber–concrete composite beams. Int. J. Mech. Sci. 117, 30–42 (2016)CrossRef
16.
Zurück zum Zitat Girhammar, U.A.: A simplified analysis method for composite beams with interlayer slip. Int. J. Mech. Sci. 51(7), 515–530 (2009)CrossRef Girhammar, U.A.: A simplified analysis method for composite beams with interlayer slip. Int. J. Mech. Sci. 51(7), 515–530 (2009)CrossRef
17.
Zurück zum Zitat EN 1995-1-1: Eurocode 5—Design of Timber Structures. European Committee for Standardization, Bruxelles, Belgium (2004) EN 1995-1-1: Eurocode 5—Design of Timber Structures. European Committee for Standardization, Bruxelles, Belgium (2004)
18.
Zurück zum Zitat McCutcheon, W.J.: Method for predicting the stiffness of wood-joist floor systems with partial composite action, FPL 289, Forest Products Laboratory, Forest Service, U.S. Department of Agriculture, Madison, Wisconsin, USA (1977) McCutcheon, W.J.: Method for predicting the stiffness of wood-joist floor systems with partial composite action, FPL 289, Forest Products Laboratory, Forest Service, U.S. Department of Agriculture, Madison, Wisconsin, USA (1977)
19.
Zurück zum Zitat McCutcheon, W.J.: Stiffness of framing members with partial composite action. J. Struct. Eng. 112(7), 1623–1637 (1986)CrossRef McCutcheon, W.J.: Stiffness of framing members with partial composite action. J. Struct. Eng. 112(7), 1623–1637 (1986)CrossRef
20.
Zurück zum Zitat van der Linden, M.L.R.: Timber–concrete composite floor systems. HERON 44(3), 215–239 (1999) van der Linden, M.L.R.: Timber–concrete composite floor systems. HERON 44(3), 215–239 (1999)
21.
Zurück zum Zitat Fernandez-Cabo, J.L., Fernandez-Lavandera, J., Diez-Barra, R., Avila-Jalvo, J.M.: Timber composite beams with a discrete connection system. Proc. Inst. Civ. Eng. Struct. Build. 166(2), 57–72 (2013)CrossRef Fernandez-Cabo, J.L., Fernandez-Lavandera, J., Diez-Barra, R., Avila-Jalvo, J.M.: Timber composite beams with a discrete connection system. Proc. Inst. Civ. Eng. Struct. Build. 166(2), 57–72 (2013)CrossRef
22.
Zurück zum Zitat Allen, D.E., Onysko, D.M., Murray, T.M.: Minimizing Floor Vibration. ATC Design Guide 1. Applied Technology Council, Redwood City, California (1999) Allen, D.E., Onysko, D.M., Murray, T.M.: Minimizing Floor Vibration. ATC Design Guide 1. Applied Technology Council, Redwood City, California (1999)
23.
Zurück zum Zitat Heimeshoff, B.: Näherungsverfahren zur berechnung von einfeldträgern mit kragarm und von zweifeldträgern, die aus nachgiebig miteinander verbundenen querschnittsteilen bestehen, im ingenieurholzbau. Holz als Roh-und Werkstoff 49(7), 277–285 (1991)CrossRef Heimeshoff, B.: Näherungsverfahren zur berechnung von einfeldträgern mit kragarm und von zweifeldträgern, die aus nachgiebig miteinander verbundenen querschnittsteilen bestehen, im ingenieurholzbau. Holz als Roh-und Werkstoff 49(7), 277–285 (1991)CrossRef
24.
Zurück zum Zitat Xu, R., Chen, D.: Variational principles of partial-interaction composite beams. J. Eng. Mech. 138(5), 542–551 (2012) Xu, R., Chen, D.: Variational principles of partial-interaction composite beams. J. Eng. Mech. 138(5), 542–551 (2012)
25.
Zurück zum Zitat Atashipour, S.R., Landel, P., Al-Emrani, M.: Accuracy evaluation of gamma-method for deflection prediction of partial composite beams. In: 2018 World Conference on Timber Engineering (WCTE 2018), Seoul, Republic of Korea (2018) Atashipour, S.R., Landel, P., Al-Emrani, M.: Accuracy evaluation of gamma-method for deflection prediction of partial composite beams. In: 2018 World Conference on Timber Engineering (WCTE 2018), Seoul, Republic of Korea (2018)
26.
Zurück zum Zitat Koiter, W.T.: Discussion: “Timoshenko beam theory is not always more accurate than elementary beam theory’’. J. Appl. Mech. 44(2), 357–358 (1977)CrossRef Koiter, W.T.: Discussion: “Timoshenko beam theory is not always more accurate than elementary beam theory’’. J. Appl. Mech. 44(2), 357–358 (1977)CrossRef
27.
Zurück zum Zitat Chui, Y., Smith, I.: Influence of rotatory inertia, shear deformation and support condition on natural frequencies of wooden beams. Wood Sci. Technol. 24(3), 233–245 (1990)CrossRef Chui, Y., Smith, I.: Influence of rotatory inertia, shear deformation and support condition on natural frequencies of wooden beams. Wood Sci. Technol. 24(3), 233–245 (1990)CrossRef
28.
Zurück zum Zitat Zhang, S., Zhou, J., Niederwestberg, J., Chui, Y.H.: Effect of end support restraints on vibration performance of cross laminated timber floors: an analytical approach. Eng. Struct. 189, 186–194 (2019)CrossRef Zhang, S., Zhou, J., Niederwestberg, J., Chui, Y.H.: Effect of end support restraints on vibration performance of cross laminated timber floors: an analytical approach. Eng. Struct. 189, 186–194 (2019)CrossRef
29.
Zurück zum Zitat Timoshenko, S., Woinowsky-Krieger, S.: Theory of Plates and Shells, 2nd edn. McGraw-Hill, Inc., New York (1959)MATH Timoshenko, S., Woinowsky-Krieger, S.: Theory of Plates and Shells, 2nd edn. McGraw-Hill, Inc., New York (1959)MATH
30.
Zurück zum Zitat Smith, I.: Series type solutions for built-up timber beams with semi-rigid connections. Proc. Inst. Civ. Eng. 69(3), 707–719 (1980) Smith, I.: Series type solutions for built-up timber beams with semi-rigid connections. Proc. Inst. Civ. Eng. 69(3), 707–719 (1980)
31.
Zurück zum Zitat Amana, E.J., Booth, L.G.: Theoretical and experimental studies on nailed and glued plywood stressed-skin components, Part I. Theoretical study. J. Inst. Wood Sci. 4(1), 43–69 (1967) Amana, E.J., Booth, L.G.: Theoretical and experimental studies on nailed and glued plywood stressed-skin components, Part I. Theoretical study. J. Inst. Wood Sci. 4(1), 43–69 (1967)
32.
Zurück zum Zitat Zhang, S., Xu, L.: Exact static analysis of eccentrically stiffened plates with partial composite action. Compos. Struct. 198, 117–125 (2018)CrossRef Zhang, S., Xu, L.: Exact static analysis of eccentrically stiffened plates with partial composite action. Compos. Struct. 198, 117–125 (2018)CrossRef
33.
Zurück zum Zitat Gorman, D.J.: Vibration Analysis of Plates by the Superposition Method. Series on Stability, Vibration and Control of Systems, Series A, vol. 3. World Scientific, Singapore (1999)MATH Gorman, D.J.: Vibration Analysis of Plates by the Superposition Method. Series on Stability, Vibration and Control of Systems, Series A, vol. 3. World Scientific, Singapore (1999)MATH
34.
Zurück zum Zitat Bhaskar, K., Kaushik, B.: Simple and exact series solutions for flexure of orthotropic rectangular plates with any combination of clamped and simply supported edges. Compos. Struct. 63(1), 63–68 (2004)CrossRef Bhaskar, K., Kaushik, B.: Simple and exact series solutions for flexure of orthotropic rectangular plates with any combination of clamped and simply supported edges. Compos. Struct. 63(1), 63–68 (2004)CrossRef
35.
Zurück zum Zitat Greif, R., Mittendorf, S.: Structural vibrations and Fourier series. J. Sound Vib. 48(1), 113–122 (1976)CrossRef Greif, R., Mittendorf, S.: Structural vibrations and Fourier series. J. Sound Vib. 48(1), 113–122 (1976)CrossRef
36.
Zurück zum Zitat Wang, J.T.-S., Lin, C.-C.: Dynamic analysis of generally supported beams using Fourier series. J. Sound Vib. 196(3), 285–293 (1996)CrossRef Wang, J.T.-S., Lin, C.-C.: Dynamic analysis of generally supported beams using Fourier series. J. Sound Vib. 196(3), 285–293 (1996)CrossRef
37.
Zurück zum Zitat Li, R., Zhong, Y., Tian, B., Liu, Y.: On the finite integral transform method for exact bending solutions of fully clamped orthotropic rectangular thin plates. Appl. Math. Lett. 22(12), 1821–1827 (2009)MathSciNetMATHCrossRef Li, R., Zhong, Y., Tian, B., Liu, Y.: On the finite integral transform method for exact bending solutions of fully clamped orthotropic rectangular thin plates. Appl. Math. Lett. 22(12), 1821–1827 (2009)MathSciNetMATHCrossRef
38.
Zurück zum Zitat Zhang, S., Xu, L.: Bending of rectangular orthotropic thin plates with rotationally restrained edges: a finite integral transform solution. Appl. Math. Model. 46, 48–62 (2017)MathSciNetMATHCrossRef Zhang, S., Xu, L.: Bending of rectangular orthotropic thin plates with rotationally restrained edges: a finite integral transform solution. Appl. Math. Model. 46, 48–62 (2017)MathSciNetMATHCrossRef
39.
Zurück zum Zitat Wu, Y.-F., Xu, R., Chen, W.: Free vibrations of the partial-interaction composite members with axial force. J. Sound Vib. 299(4), 1074–1093 (2007)CrossRef Wu, Y.-F., Xu, R., Chen, W.: Free vibrations of the partial-interaction composite members with axial force. J. Sound Vib. 299(4), 1074–1093 (2007)CrossRef
40.
Zurück zum Zitat Gjelsvik, A.: Analog-beam method for determining shear-lag effects. J. Eng. Mech. 117(7), 1575–1594 (1991) Gjelsvik, A.: Analog-beam method for determining shear-lag effects. J. Eng. Mech. 117(7), 1575–1594 (1991)
41.
Zurück zum Zitat Betti, R., Gjelsvik, A.: Elastic composite beams. Comput. Struct. 59(3), 437–451 (1996)MATHCrossRef Betti, R., Gjelsvik, A.: Elastic composite beams. Comput. Struct. 59(3), 437–451 (1996)MATHCrossRef
42.
Zurück zum Zitat Ellakany, A.M., Tablia, H.A.: A numerical model for static and free vibration analysis of elastic composite beams with end shear restraint. Meccanica 45(4), 463–474 (2010)MathSciNetMATHCrossRef Ellakany, A.M., Tablia, H.A.: A numerical model for static and free vibration analysis of elastic composite beams with end shear restraint. Meccanica 45(4), 463–474 (2010)MathSciNetMATHCrossRef
43.
Zurück zum Zitat Tolstov, G.P.: Fourier Series. Dover Publications, New York (1962) Tolstov, G.P.: Fourier Series. Dover Publications, New York (1962)
44.
Zurück zum Zitat Bromwich, T.J.I.: An Introduction to the Theory of Infinite Series, 2nd edn. Macmillan and Company, New York (1965) Bromwich, T.J.I.: An Introduction to the Theory of Infinite Series, 2nd edn. Macmillan and Company, New York (1965)
45.
Zurück zum Zitat Zhang, S.: Vibration serviceability of cold-formed steel floor systems. Ph.D. thesis, University of Waterloo, Waterloo, ON, Canada (2017) Zhang, S.: Vibration serviceability of cold-formed steel floor systems. Ph.D. thesis, University of Waterloo, Waterloo, ON, Canada (2017)
46.
Zurück zum Zitat Bary, N.K.: A Treatise on Trigonometric Series, vol. 1. Pergamon Press Ltd., Oxford (1964)MATH Bary, N.K.: A Treatise on Trigonometric Series, vol. 1. Pergamon Press Ltd., Oxford (1964)MATH
47.
Zurück zum Zitat Xu, L.: Second-order analysis for semirigid steel frame design. Can. J. Civ. Eng. 28(1), 59–76 (2001)CrossRef Xu, L.: Second-order analysis for semirigid steel frame design. Can. J. Civ. Eng. 28(1), 59–76 (2001)CrossRef
Metadaten
Titel
Exact series solutions of composite beams with rotationally restrained boundary conditions: static analysis
verfasst von
Yuan Zhang
Lei Zhang
Sigong Zhang
Publikationsdatum
08.10.2022
Verlag
Springer Berlin Heidelberg
Erschienen in
Archive of Applied Mechanics / Ausgabe 12/2022
Print ISSN: 0939-1533
Elektronische ISSN: 1432-0681
DOI
https://doi.org/10.1007/s00419-022-02277-0

Weitere Artikel der Ausgabe 12/2022

Archive of Applied Mechanics 12/2022 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.