Shanghai University
Article Information
- ZENG Jin, MA Hui, YU Kun, XU Zhitao, WEN Bangchun
- Coupled flapwise-chordwise-axial-torsional dynamic responses of rotating pre-twisted and inclined cantilever beams subject to the base excitation
- Applied Mathematics and Mechanics (English Edition), 2019, 40(8): 1053-1082.
- http://dx.doi.org/10.1007/s10483-019-2506-6
Article History
- Received Nov. 9, 2018
- Revised Mar. 2, 2019
2. Key Laboratory of Vibration and Control of Aero-Propulsion System Ministry of Education, Northeastern University, Shenyang 110819, China
Nomenclature
A, cross-sectional area;
A0, amplitude of the base excitation;
OXYZ, global coordinate system;
OXrYrZr, rotating coordinate system;
obxbybzb, local coordinate system;
oexeyeze, element coordinate system;
Ai, coordinate-transformation matrix;
b, beam width at the arbitrary section;
b0, beam width at the root section;
c, constant;
C, Rayleigh damping;
E, Young's modulus;
Fb, base excitation force in the x-direction;
Fc, centrifugal force;
Fe, element nodal force vector in oexeyeze;
Fre, element nodal force vector in OXrYrZr;
F, global nodal force vector;
fe, excitation frequency;
G, shear modulus;
fi, concerned end frequency;
Ge, element Coriolis matrix in oexeyeze;
Gre, element Coriolis matrix in OXrYrZr;
G, global Coriolis matrix;
h, beam thickness at the arbitrary section;
h0, beam thickness at the root section;
Ix, area moment of the cross-section inertia about ox3;
Iy, area moment of the cross-section inertia about oy3;
Iz, area moment of the cross-section inertia about oz3;
J, torsional moment of the inertia of a pre-twisted rectangular beam;
Js, torsional moment of the inertia of a straight rectangular beam;
Ja, additional torsional moment of the inertia caused by γ(L);
Kee, element structural stiffness matrix in oexeyeze;
Kce, element centrifugal stiffening matrix in oexeyeze when
Kse, spin softening matrix in oexeyeze when
Ke, re, element structural stiffness matrix in OXrYrZr;
Kc, re, element centrifugal stiffening matrix in OXrYrZr when
Ks, re, element spin softening matrix in OXrYrZr when
Ke, global structural stiffness matrix;
Kc, global centrifugal stiffening matrix;
Ks, global spin softening matrix;
Kacce, stiffness matrix related to
Kacc, re, stiffness matrix related to
Kacc, global stiffness matrix related to
L, beam length;
lk, kth element length;
Me, element mass matrix in oexeyeze;
Mre, element mass matrix in OXrYrZr;
M, global mass matrix;
N, total number of the beam elements;
n, rotation speed;
Rd, disk radius;
Rk, x coordinate of oe in OXrYrZr;
rP, coordinate vector of the arbitrary point Pin OXYZ;
Tk, kinetic energy of the kth rotating Timoshenko beam element;
T, transformation matrix from OXrYrZr to oexeyeze;
t1, starting moment of the variation;
t2, ending moment of the variation;
Uk, potential energy of the kth rotating Timoshenko beam element;
u, v, w, linear displacements of the centroid of an arbitrary beam cross-section along the positive xe-, ye-, and ze-axes;
u', v', w', derivatives of u, v, and w versus the coordinate x;
Xur, base excitation along the Xr-direction;
xk, x coordinate of oe in obxbybzb;
(XP, YP, ZP), coordinate components of Point P along the X-, Y-, and Z-directions;
(x, y, z), coordinate components of Point Palong the x3-, y3-, and z3-directions;
YI, first-order flapwise frequency;
YII, second-order flapwise frequency;
ZI, first-order chordwise frequency;
ZII, second-order chordwise frequency.
Greek symbols
α, rotating angle;
α1, β1, Rayleigh damping coefficients;
β0, initial setting angle;
βk, setting angle of the kth Timoshenko beam element;
γk, pre-twisted angle of the kth Timoshenko beam element;
γ(L), pre-twisted angle at L;
δ, variational symbol;
δe, element nodal displacement vector in oexeyeze;
δre, element nodal displacement vector in OXrYrZr;
ζ, correction factor;
θ, angular displacement with respect to the xe-axis;
θ', derivative of θ versus the coordinate x;
θI, first-order torsional frequency;
ξ1, ξ2, modal damping ratios;
ρ, density;
τb, breadth taper;
τh, thickness taper;
υ, Poisson's ratio;
φ, angular displacement with respect to the ze-axis;
φ', derivative of φ versus the coordinate x;
χ, symbol representing global matrices;
χke, symbol representing element matrices;
ϕ, angular displacement with respect to the ye-axis;
ϕ', derivative of ϕ versus the coordinate x.
1 IntroductionRotating beam-like structures are widely spread over many engineering applications such as compressor blades, helicopter blades, wind blades, and propellers[1-2]. In general, almost all of the structural vibration problems such as fatigue failure, flutter, and stability are closely related to the natural frequencies of the structures[3-5]. Therefore, it is essential to have a thorough understanding of their dynamic characteristics for engineering design purposes[6].
The equations of motion of rotating beams have been studied extensively[1-10]. Unlike non-rotating beams, rotating beams have additional Coriolis, centrifugal stiffening, and spin softening effects, which will result in the couplings of structural modes among different vibration directions[9-10]. In Ref.[11], the coupled bending-bending equations of motion of a pre-twisted and tapered Euler beam with a fluctuant angular velocity were determined with the Hamilton principle, while the coupled bending-bending equations of motion of a rotating composite thin-walled beam subject to thermal stress field, humidity stress field, and aerodynamic forces were elaborately derived in Ref.[12]. The coupled flapwise-chordwise-axial equations of motion of the rotating uniform Timoshenko beam with a variable speed were established in Refs.[13] and[14]. In Refs.[5], [15], and [16], the effects of spin softening and Coriolis on the dynamics of a rotating cantilever beam were excluded when its coupled flapwise-chordwise-torsional equations of motion were derived. Such effects were also considered in the free vibration analyses of the rotating Timoshenko beam featuring flapwise-chordwise-torsional couplings in Ref.[17] and flapwise-axial-torsional couplings in Ref.[18]. In Refs.[2], [6], and [19], although the coupled flapwise-chordwise-axial-torsional equations of motion of a rotating Timoshenko beam were established, only the theoretical derivation was carried out. Some necessary verification works were deficient in Ref.[19], while the effects of centrifugal stiffening on the torsional behaviors of the beam were ignored in Refs.[2] and [6].
The equations of motion mentioned before[1-19] are partial differential equations, and their closed-form solutions are not available in most cases[2, 20]. To solve these partial differential equations, lots of numerical algorithms emerge, e.g., the dynamic stiffness method[21-23], the transfer matrix method[24], the differential transformation method[25], the Galerkin method[13, 26], the Rayleigh-Ritz method[18], the power series method[27], the Chebyshev-Ritz method[28], and the finite element (FE) method[29-30]. Generally speaking, the FE method is more suitable for handling problems such as complicated geometrical configurations, loads, materials, and complicated boundary conditions than the other methods[31-32].
The literature survey presented here indicates that little emphasis has been placed on the coupled flapwise-chordwise-axial-torsional dynamic modeling of rotating beams with the consideration of all rotating effects, i.e., the effects of Coriolis force, centrifugal stiffening, spin softening, and angular acceleration. Moreover, the modal characteristics of rotating beams are the concern of most scholars, and little attention has been paid to the coupled flapwise-chordwise-axial-torsional vibration responses under the base excitation. Aiming to these deficiencies, a rotating pre-twisted and inclined cantilever beam model (RPICBM) is established with the Hamilton principle and the FE method (see Section 2). Based on the proposed model shown in Section 2, the effects of the setting angle and the pre-twisted angle on the modal characteristics and the base excitation-induced vibration responses are analyzed theoretically (see Section 3). Finally, some conclusions are made in Section 4.
2 FE modeling of the RPICBM subject to the base excitationAn engineering example of a bladed disk structure subject to the base excitation is shown in Fig. 1(a), which consists of two first-stage fan blades, a rigid disk, and elastic supports[33]. In Fig. 1(a), the intersection angle between the unreleased blade and the released blade is 180°. Once the test rig is rotating and reaches a certain rotation speed, one of the two first-stage fan blades will be released from the rigid disk under the effect of the centrifugal force, which leads to the unbalanced bladed disk system. Then, the unbalanced bladed disk system will be whirling. Meanwhile, the motion of the unreleased blade can be viewed as the resultant motion of the rigid disk whiling and blade rotating[34], i.e., the dynamics of the unreleased rotating blade can be degenerated into analyzing the dynamic characteristics of a rotating cantilever blade subject to the base excitation[35].
![]() |
Fig. 1 FE discretization of the RPICBM: (a) first-stage fan blades of a certain aero-engine[33], (b) three coordinates of the second-kind Euler angle, and (c) discrete beam model (color online) |
|
Based on the aforementioned engineering example, a similar mathematical model is simplified and adopted in this paper, i.e., an FE model of the RPICBM subject to the base excitation is pre-supposed and then established in this paper via combining the Hamilton principle with the FE method (see Fig. 1(c)). Figure 1(c) shows that the RPICBM is discretized into a series of inclined straight rotating Timoshenko beam segments, which are so-called Timoshenko beam elements. In this section, emphasis will be placed on the derivation of the partial differential equations and the element matrices of rotating Timoshenko beams. It is worth noting that the rotating effects such as Coriolis, centrifugal stiffening, spin softening, and angular acceleration are all included in the flapwise-, chordwise-, axial-, and torsional-directions. Finally, the proposed RPICBM is verified via comparisons of the natural frequencies and the base excitation-induced vibration responses among literature and FE models by using Beam188 and Shell181 elements in ANSYS.
2.1 Establishment of the partial differential equationsThe kth Timoshenko beam element is taken as an example to establish the corresponding coupled flapwise-chordwise-axial-torsional equations of motion (see Fig. 1(c)). In Fig. 1(c), OXYZ, OXrYrZr, obxbybzb, oexeyezeand are the global coordinate system, the rotating coordinate system, the local coordinate system, and the element coordinate system, respectively, and α is the rotation angle. In Fig. 1(b), three coordinates of the second-kind Euler angles θ, ϕ, and φ are adopted to describe the angular displacements of an arbitrary rigid beam cross-section in oexeyeze, β0 is the initial setting angle, and γk is the pre-twisted angle of the kth Timoshenko beam element. Here,
![]() |
is defined as the setting angle of the kth Timoshenko beam element. Let
![]() |
where Rk is the x-coordinate of oe in OXrYrZr, Rd is the disk radius, xk is the x-coordinate of oe in obxbybzb, and Xur is the base excitation along the Xr-direction. Then, the coordinates of an arbitrary point P on the beam cross-section in OXYZ can be expressed as follows:
![]() |
(1) |
where (x, y, z) is the coordinates of Point P in ox3y3z3 (see Fig. 1(b)). u, v, and w are the linear displacements of the centroid of an arbitrary beam cross-section along the positive xe-, ye-, and ze-axes, respectively. A1, A2, A3, A4, and A5 are the coordinate-transformation matrices from ox3y3z3 to OXYZ, which can be expressed as follows:
![]() |
(2) |
In Eq.(2), the first-order approximation is adopted based on the small deformation assumption, i.e.,
![]() |
According to Eqs.(1) and (2), the kinetic energy Tk of the kth rotating Timoshenko beam element can be calculated as follows:
![]() |
(3) |
where ρ, lk, and A are the density, the kth element length, and the cross-sectional area, respectively. An over-dot in Eq.(3) represents the differentiation with respect to time.
According to Refs.[4], [17], [36], and [37], the potential energy Uk of the kth rotating Timoshenko beam element is shown as follows:
![]() |
(4) |
where E and G are the Young's modulus and the shear modulus, respectively. Ix, Iy, and Iz are the area moments of the inertia of the cross-section about ox3, oy3, and oz3, respectively. J is the torsional moment of inertia, and its calculation is shown in Subsection 2.4.
![]() |
is the centrifugal force.
![]() |
is the base excitation force in the x-direction.
![]() |
is a constant. In addition, a prime in Eq.(4) represents the differentiation with respect to the coordinate x. In Eq.(4), it should be noted that the underlined term is the effect of the centrifugal stiffening on the torsional behavior, while the double underlined term is the contribution of the pre-twisted angle to the potential energy[37].
Taking the functionals u, v, w, θ, ϕ, and φ as independent variables, the Hamilton principle is applied to the derived energy equation
![]() |
and six partial differential equations are then determined. The xe-axial displacement u is
![]() |
(5) |
The ye-bending direction displacement v is
![]() |
(6) |
The ze-bending direction displacement w is
![]() |
(7) |
The angular displacement θ with respect to the xe-axis is
![]() |
(8) |
The angular displacement ϕ with respect to the ye-axis is
![]() |
(9) |
The angular displacement φ with respect to the ze-axis is
![]() |
(10) |
It should be noted that the terms including
In order to solve Eqs.(5)–(10), the displacement field functions in the u-, v-, w-, θ-, ϕ-, and φ-directions are assumed, and the corresponding expressions are shown in Appendix A. Substituting Eqs.(A1)–(A4) into Eqs.(5)–(10), the following second-order ordinary differential equations are determined:
![]() |
(11) |
where Me and kee represent the mass and structural stiffness matrices, respectively. Ge, Kce, and Kse are the Coriolis, centrifugal stiffening, and spin softening matrices when the angular velocity
All the element matrices, i.e., Me, Ge, kee, Kce, Kse, Kacce,
![]() |
(12) |
where T is the transformation matrix from OXrYrZr to oexeyeze.
Based on Eqs.(11) and (12), the element matrices in OXrYrZr such as Mre, Gre, Ke, r e, Kc, re, Ks, re, Kacc, re, and Fre can be calculated as follows:
![]() |
(13) |
The global matrices
A detailed derivation process for the calculation of the torsional moment of inertia of a straight rectangular beam Js with the consideration of the effect of the thickness-width ratio (h/b, see Fig.C1 in Appendix C) is shown in Appendix C. Here, the expressions for calculating Js are directly shown as follows:
![]() |
(14) |
Furthermore, an additional torsional moment of inertia Ja caused by a pre-twisted angle γ(L) is determined as follows[40]:
![]() |
(15) |
where γ(L) is the pre-twisted angle at the distance L (see Fig. 1).
In terms of Eqs.(14) and (15), J can be determined via Js+Ja. In Fig. 2, the torsional stiffness GJ and the first-order torsional frequency θI obtained from this paper agree well with those obtained from Ref.[40], and this verifies the derived J.
![]() |
Fig. 2 Effects of γ(L) on the torsional stiffness and frequency (color online) |
|
Based on Subsections 2.1–2.4, the FE model of the RPICBM is established in Fig. 3. In addition, some measures should be taken to validate the developed model. In terms of the effects of the profile tapper and the pre-twisted angle on the dynamic characteristics of a beam-like structure, various beam models have been proposed[11, 19, 21, 23, 30, 41]. Here, a pre-twisted cantilever beam model in Ref.[41] is taken as an example to illustrate the availability of the developed model. The corresponding material and geometrical parameters are defined as follows[41]:
![]() |
Fig. 3 FE model of the RPICBM (color online) |
|
![]() |
where v is Poisson's ratio, L is the beam length, b0 is the beam root width, h0 is the beam root thickness, and τh is the thickness taper. In order to verify the proposed model, comparisons of the natural frequencies and the vibration responses are made among the proposed model, Ref.[41], and FE models by using Beam188 and Shell181 elements (see Figs. 4 and 5).
![]() |
Fig. 4 Effects of the breadth taper τb on the natural frequencies, where γ(L) = 45°, β0 = 0°, and n = 0 r·min−1 (color online) |
|
![]() |
Fig. 5 Vibration responses, where γ(L) = 45°, β0 = 0°, τb = 0.3, and n = 20 000πr·min-1 (color online) |
|
In Fig. 4, YI, ZI, YII, and ZII obtained from the proposed model, the FE models by using shell181 and Beam188 elements, and Ref.[41] agree well with each other under different τb, which verifies the proposed model.
Based on the different models, the X-, Y-, Z-, and θ-vibration responses of the tip node (see Fig. 3) under the base excitation
![]() |
It should be noted that the vibration responses of 17 excitation periods are calculated, and each period is divided into 60 time steps. In addition, the Rayleigh damping C is adopted, and the corresponding expressions are listed as follows[42]:
![]() |
(16) |
where
Generally speaking, the X-, Y-, and Z-vibration responses agree well with each other (see Figs. 5(a)–5(c)), while the θ-vibration response obtained from the proposed model is closer to that obtained from the Shell181 elements than the Beam188 elements (see Fig. 5(d)), which indicates that there may exist some errors during the process of simulation for the torsional behaviors of the RPICBM when the Beam188 elements are used.
3 Dynamic characteristics of the RPICBMIn this section, the pre-stressed modal[43], where the rotation speed-induced Coriolis, centrifugal stiffening, and spin softening effects are considered, and the transient dynamic analyses are performed to study the effects of β0 and γ(L) on the natural frequencies and the vibration responses. The corresponding simulation parameters are listed in Table 1.
![]() |
With the consideration of the rotation speed-induced Coriolis, centrifugal stiffening, and spin softening effects, the effects of β0 and γ(L) on YI, ZI, and θI are analyzed, respectively (see Figs. 6 and 7).
![]() |
Fig. 6 Effects of β0 on the first-order dynamic frequencies of Case 1 (color online) |
|
![]() |
Fig. 7 Effects of γ(L) on the first-order dynamic frequencies of Case 2 (color online) |
|
In Fig. 6, it can be seen that when n≠0r·min-1,
YI and θI always increase while ZI decreases when β0 increases. The larger β0 is, the larger the increment of YI and θI is (see Figs. 6(a) and 6(c)) and the larger the decrement of ZI is (see Fig. 6(b)). This is because the spin softening effect decreases in the Y- and θ-directions (see
Some typical features in Fig. 7 are concluded as follows:
(ⅰ) When γ(L) increases, YI and θI increase, while ZI decreases (see Fig. 7). This is because the element structural stiffness increases in the Y- and θ-directions but decreases in the Z-direction when γ(L) increases.
(ⅱ) The larger γ(L) is, the larger the increment of YI and θI is (see Figs. 7(a) and 7(c)) and the smaller the increment of ZI is (see Fig. 7(b)). This is because the increasing γ(L) reduces the spin softening effects in the Y- and θ-directions (see
Considering the effect of the X-harmonic base excitation (see Fig. 3), we perform a transient dynamic analysis to obtain the amplitude frequency responses in the X-, Y-, Z-, and θ-directions as well as the constant components in the X- and θ-directions. The corresponding simulation parameter settings can refer to Cases 3 and 4 in Table 1.
3.2.1 Case 3: the effects of β0 on the vibration responses under the base excitationThe amplitude frequency responses in the X-, Y-, Z-, and θ-directions are shown in Fig. 8. Some main conclusions are made as follows:
![]() |
Fig. 8 Amplitude frequency responses of Case 3 (color online) |
|
(ⅰ) When β0 = 0° and γ(L)= 0°, the first-order resonance is only embodied in the X- and Y-directions, respectively, (see Figs. 8(a)–8(d)), while the amplitude frequency responses in Figs. 8(c) and 8(d) are zero. This is because there exists the coupling between the X- and Y-directions via the Coriolis effect (see
(ⅱ) When β0 = ± 20° and γ(L)=0°, there are couplings among the X-, Y-, Z-, and θ-directions, which leads to the appearance of the resonance peak. The corresponding coupling terms are mentioned in Eq.(1).
(ⅲ) When β0 = ± 45° and γ(L)=0°, only vibration responses in the X-, Y-, Z-, and θ-directions occur, and there is no resonance peak. This is mainly due to that the excitation frequency n/60 is lower than the first-order flapwise dynamic frequency of the model under β0 = 45° or -45° and γ(L)=0° (see Fig. 9).
![]() |
Fig. 9 First-order flapwise dynamic frequency curves of Case 3 (color online) |
|
Furthermore, it is worth noting that the constant components in the X- and θ-directions under different β0 and n are also extracted (see Fig. 10). Due to the non-existence of constant loads in the Y- and Z-directions (see Eqs.(6), (7), (9), and (10)), the constant components in the Y- and Z-directions are zero and omitted here.
![]() |
Fig. 10 Constant components under different β0 and n of Case 3 (color online) |
|
In Fig. 10(a), the X-constant components are not affected by β0, and increase with the increase in n. This is because the element centrifugal force obtained from the proposed model (see
In Fig. 10(b), the θ-constant component is the largest when β0 = ± 45° and the smallest when β0 = 0°. Moreover, a positive/negative β0 leads to a positive/negative θ-constant component (see Figs. 10(b) and 11). This can be easily illustrated via the constant element torque
![]() |
Fig. 11 Deformation nephogram in the θ-direction under n = 40 000π r·min-1 of Case 3 (color online) |
|
Different from the effects of β0 on the vibration responses under the base excitation (see Subsection 3.2.1), γ(L) (γ(L) ≠ 0°) directly causes couplings among the X-, Y-, Z-, and θ-directions, and further leads to the occurrence of the first-order resonance in all the directions (see Figs. 12 and 13). The corresponding coupling terms can refer to Subsection 3.2.1.
![]() |
Fig. 12 Amplitude frequency responses of Case 4 (color online) |
|
![]() |
Fig. 13 Amplitude frequency responses of Case 4 (color online) |
|
In Fig. 14(a), the X-constant component obtained from the proposed model is not affected by γ(L). This is because the element centrifugal force (see
![]() |
Fig. 14 Constant components under different β0 and n of Case 4 (color online) |
|
In Fig. 14(b), when the summation of two different γ(L)
![]() |
Fig. 15 Deformation nephogram in the θ-direction under n = 40 000π r·min-1 of Case 4 (color online) |
|
(Ⅰ) The element torque in the θ-direction is
![]() |
(see Eq.(8),
(Ⅱ) A negative/positive γ(L) can determine a positive/negative element torque, thus leading to a positive or negative θ-constant component (see Figs. 5(d), 14(b), and 15).
4 ConclusionsThe coupled flapwise-chordwise-axial-torsional dynamic model of the RPICBM is established with the Hamilton principle and the FE method. The proposed model is verified via comparisons with those in the literature. In the FE models, Beam188 and Shell181 elements are used in the ANSYS software. Based on the proposed model, the effects of the setting angle and the pre-twisted angle on the natural frequencies and the base excitation-induced vibration responses of the RPICBM are discussed. Some conclusions are summarized as follows:
(ⅰ) When the setting angle or pre-twisted angle increases, the first-order flapwise and torsional frequencies increase while the first-order chordwise frequency decreases. Moreover, the first-order flapwise, chordwise, and torsional frequencies are affected by the pre-twisted angle rather than the setting angle when the rotation speed is 0 r·min-1.
(ⅱ) Under the effect of the axial base excitation, there exists the flapwise-axial coupling or even the flapwise-chordwise-axial-torsional coupling of the RPICBM. In addition, a positive or negative setting angle can result in a positive or negative constant component, while a positive or negative pre-twisted angle can result in a negative or positive constant component. When the pre-twisted angle is 0°, and the setting angle is ± 45°, the constant component in the torsional direction reaches the maximum. When both the setting angle and the summation of two pre-twisted angles are 0°, the constant components in the torsional direction under two pre-twisted angles are the same in the magnitude but opposite in the axial directions.
(ⅲ) Within the scope of linear elastic theory, the FE model using Shell181 elements in the ANSYS software and the proposed model may be more suitable for simulating the dynamic characteristics of the RPICBM than the FE model using Beam188 elements in the ANSYS software, especially for the simulation of torsional behaviors.
Appendix AThe derivation of the shape functions in the ye-bending, ze-bending, xe-axial, and xe-torsional directions are as follows.
(ⅰ) The shape function in the xe-direction is
![]() |
(A1) |
(ⅱ) The torsional shape function in the θ-direction is
![]() |
(A2) |
(ⅲ) The bending shape function in the xeoeye-plane is
![]() |
(A3) |
(ⅳ) The bending shape function in the xeoeze-plane is
![]() |
(A4) |
In Eqs.(A1)-(A4), the expressions of the shape functions Nmi(ξ) (m=u, θ; i=1, 2), Nmk(ξ)(m=v, φ, w, ϕ; k=1, 2, 3, 4) can refer to Ref.[38]. However, there exist minor modifications on the ϕ-displacement shape function, and the modified ϕ -displacement shape function is shown as follows:
![]() |
(A5) |
![]() |
(A6) |
![]() |
(A7) |
![]() |
(A8) |
where
![]() |
The matrices and vectors related to the rotating Timoshenko beam element are as follows.
The mass matrix Me is
![]() |
where
![]() |
The Coriolis matrix Ge is
![]() |
where
![]() |
The structural stiffness matrix Kec is
![]() |
where
![]() |
The centrifugal stiffening matrix Kce is
![]() |
where
![]() |
The spin softening matrix Kse is
![]() |
where
![]() |
The external force vector Fe is
![]() |
where
![]() |
For the torsional moment of the inertia of a rectangular beam, the torsional stress function τ(y, z) of an arbitrary rectangular cross-sections is[39]
![]() |
(C1) |
where h is the length of the short edge of the beam cross-section (see Fig.C1), and τ1(y, z) is the additional stress correction term considering the effect of h/b.
![]() |
Fig. C1 Rectangular section size |
|
Furthermore, τ(y, z) should satisfy the Poisson's equation as follows:
![]() |
(C2) |
Substituting Eq.(C1) into Eq.(C2), the following expression can be determined:
![]() |
(C3) |
where τ(y, z) should satisfy the following equations at the boundaries of the rectangular cross-section:
![]() |
(C4) |
Substituting Eq.(C1) into Eq.(C4), the following expressions can be determined:
![]() |
(C5) |
Letting τ1(y, z)=f(y) f(z) and substituting it into Eq.(C3), the following formula can be determined:
![]() |
(C6) |
where λ represents an arbitrary constant.
According to Eq.(C6), the following two expressions can be determined:
![]() |
(C7) |
The general solutions of Eq.(C7) can be obtained as follows:
![]() |
(C8) |
Referring to the theory of membrane analogy, τ1(y, z) should be an even function with respect to the coordinates y and z. Therefore, τ1(y, z) should be expressed as follows:
![]() |
(C9) |
Substituting Eq.(C9) into the second expression in Eq.(C5), the following equation can be determined:
![]() |
(C10) |
The general solutions of Eq.(C10) can be expressed as follows:
![]() |
(C11) |
Substituting Eq.(C11) into Eq.(C9), τ1(y, z) can be expanded in the following series:
![]() |
(C12) |
Substituting Eq.(C12) into the first expression in Eq.(C5), we can obtain
![]() |
(C13) |
Multiplying both sides of Eq.(C13) by cos(λly) and then integrating it with respect to y over [-b/2, b/2], we have
![]() |
(C14) |
Substituting Eq.(C14) into Eq.(C12), we have
![]() |
(C15) |
Combining Eq.(C1) with Eq.(C15), we can obtain the torsional moment of inertia Js as follows:
![]() |
(C16) |
where
![]() |
is a correction factor for Js. In this paper,
![]() |
when l=0.
[1] |
YARDIMOGLU, B. and YILDIRIM, T. Finite element model for vibration analysis of pre-twisted Timoshenko beam. Journal of Sound and Vibration, 273, 741-754 (2004) doi:10.1016/j.jsv.2003.05.003 |
[2] |
YANG, X. D., WANG, S. W., ZHANG, W., QIN, Z. H., and YANG, T. Z. Dynamic analysis of a rotating tapered cantilever Timoshenko beam based on the power series method. Applied Mathematics and Mechanics (English Edition), 38(10), 1425-1438 (2017) doi:10.1007/s10483-017-2249-6 |
[3] |
RAO, J. S. and CARNEGIE, W. Solution of the equations of motion of coupled-bending bending torsion vibrations of turbine blades by the method of Ritz-Galerkin. International Journal of Mechanical Sciences, 12, 875-882 (1970) doi:10.1016/0020-7403(70)90024-X |
[4] |
J. C. and BROOKS, G. W. Differential equations of motion for combined flapwise bending, chordwise bending, and torsion of twisted nonuniform rotor blades. NACA Report, TN 3905, NASA Langley Research Center, Hampton (1958)
|
[5] |
ŞKAR, G. and SABUNCU, M. Dynamic stability of a rotating asymmetric cross-section blade subjected to an axial periodic force. International Journal of Mechanical Sciences, 45, 1467-1482 (2003) doi:10.1016/j.ijmecsci.2003.10.006 |
[6] |
YANG, X. D., WANG, S. W., ZHANG, W., YANG, T. Z., and LIM, C. W. Model formulation and modal analysis of a rotating elastic uniform Timoshenko beam with setting angle. European Journal of Mechanics-A/Solids, 72, 209-222 (2018) doi:10.1016/j.euromechsol.2018.05.014 |
[7] |
RAFIEE, M., NITZSCHE, F., and LABROSSE, M. Dynamics, vibration and control of rotating composite beams and blades:a critical review. Thin-Walled Structures, 119, 795-819 (2017) doi:10.1016/j.tws.2017.06.018 |
[8] |
QIN, Y., LI, X., YANG, E. C., and LI, Y. H. Flapwise free vibration characteristics of a rotating composite thin-walled beam under aerodynamic force and hygrothermal environment. Composite Structures, 153, 490-503 (2016) doi:10.1016/j.compstruct.2016.06.057 |
[9] |
TIAN, J. J., SU, J. P., ZHOU, K., and HUA, H. X. A modified variational method for nonlinear vibration analysis of rotating beams including Coriolis effects. Journal of Sound and Vibration, 426, 258-277 (2018) doi:10.1016/j.jsv.2018.04.027 |
[10] |
LIN, S. C. and HSIAO, K. M. Vibration analysis of a rotating Timoshenko beam. Journal of Sound and Vibration, 240, 303-322 (2001) doi:10.1006/jsvi.2000.3234 |
[11] |
YOUNG, T. H. Dynamic response of a pretwisted, tapered beam with non-constant rotating speed. Journal of Sound and Vibration, 150, 435-446 (1991) doi:10.1016/0022-460X(91)90896-R |
[12] |
QIN, Y., WANG, L., and LI, Y. H. Coupled vibration characteristics of a rotating composite thinwalled beam subjected to aerodynamic force in hygrothermal environment. International Journal of Mechanical Sciences, 140, 260-270 (2018) doi:10.1016/j.ijmecsci.2018.03.002 |
[13] |
MA, H., XIE, F. T., NAI, H. Q., and WEN, B. C. Vibration characteristics analysis of rotating shrouded blades with impacts. Journal of Sound and Vibration, 378, 92-108 (2016) doi:10.1016/j.jsv.2016.05.038 |
[14] |
XIE, F. T., MA, H., CUI, C., and WEN, B. C. Vibration response comparison of twisted shrouded blades using different impact models. Journal of Sound and Vibration, 397, 171-191 (2017) doi:10.1016/j.jsv.2017.02.056 |
[15] |
ŞKAR, G. and SABUNCU, M. Buckling and dynamic stability of a rotating pretwisted asymmetric cross-section blade subjected to an axial periodic force. Finite Elements in Analysis and Design, 40, 1399-1415 (2004) doi:10.1016/j.finel.2003.09.005 |
[16] |
SABUNCU, M. and EVRAN, K. Dynamic stability of a rotating pre-twisted asymmetric crosssection Timoshenko beam subjected to an axial periodic force. International Journal of Mechanical Sciences, 48, 579-590 (2006) doi:10.1016/j.ijmecsci.2006.01.010 |
[17] |
OZGUMUS, O. O. and KAYA, M. O. Energy expressions and free vibration analysis of a rotating Timoshenko beam featuring bending-bending-torsion coupling. Archive of Applied Mechanics, 83, 97-108 (2013) doi:10.1007/s00419-012-0634-4 |
[18] |
SINHA, S. K. Combined Torsional-bending-axial dynamics of a twisted rotating cantilever Timoshenko beam with contact-impact loads at the free end. Journal of Applied Mechanics, 74, 505-522 (2006) |
[19] |
GEORGIADES, F., LATALSKI, J., and WARMINSKI, J. Equations of motion of rotating composite beam with a nonconstant rotation speed and an arbitrary preset angle. Meccanica, 49, 1833-1858 (2014) doi:10.1007/s11012-014-9926-9 |
[20] |
BANERJEE, S. and RAO, J. S. Coupled bending-torsion vibrations of rotating blades. Gas Turbine and Fluids Engineering Conference, American Society of Mechanical Engineers, New Orleans (1976)
|
[21] |
BANERJEE, J. R. Development of an exact dynamic stiffness matrix for free vibration analysis of a twisted Timoshenko beam. Journal of Sound and Vibration, 270, 379-401 (2004) doi:10.1016/S0022-460X(03)00633-3 |
[22] |
BANERJEE, J. R., SU, H., and JACKSON, D. R. Free vibration of rotating tapered beams using the dynamic stiffness method. Journal of Sound and Vibration, 298, 1034-1054 (2006) doi:10.1016/j.jsv.2006.06.040 |
[23] |
BANERJEE, J. R. and JACKSON, D. R. Free vibration of a rotating tapered Rayleigh beam:a dynamic stiffness method of solution. Computers and Structures, 124, 11-20 (2013) doi:10.1016/j.compstruc.2012.11.010 |
[24] |
STAFFORD, R. O. and GIURGIUTIU, V. Semi-analytic methods for rotating Timoshenko beams. International Journal of Mechanical Sciences, 17, 719-727 (1975) doi:10.1016/0020-7403(75)90075-2 |
[25] |
KAYA, M. O. Free vibration analysis of a rotating Timoshenko beam by differential transform method. Aircraft Engineering and Aerospace Technology, 78, 194-203 (2006) doi:10.1108/17488840610663657 |
[26] |
QIN, Y. and LI, Y. H. Influences of hygrothermal environment and installation mode on vibration characteristics of a rotating laminated composite beam. Mechanical Systems and Signal Processing, 91, 23-40 (2017) doi:10.1016/j.ymssp.2016.12.041 |
[27] |
DU, H., LIM, M. K., and LIEW, K. M. A power series solution for vibration of a rotating Timoshenko beam. Journal of Sound and Vibration, 175, 505-523 (1994) doi:10.1006/jsvi.1994.1342 |
[28] |
FANG, J. S. and ZHOU, D. Free vibration analysis of rotating axially functionally graded tapered Timoshenko beams. International Journal of Structural Stability and Dynamics, 16, 1-9 (2015) |
[29] |
RAO, S. S. and GUPTA, R. S. Finite element vibration analysis of rotating Timoshenko beams. Journal of Sound and Vibration, 242, 103-124 (2001) doi:10.1006/jsvi.2000.3362 |
[30] |
BAZOUNE, A. Effect of tapering on natural frequencies of rotating beams. Shock and Vibration, 14, 169-179 (2007) doi:10.1155/2007/865109 |
[31] |
REDDY, J. N. Energy Principles and Variational Methods in Applied Mechanics. John Wiley and Sons, New York, 475-483 (2002) |
[32] |
SUN, Q., MA, H., ZHU, Y. P., HAN, Q. K., and WEN, B. C. Comparison of rubbing induced vibration responses using varying-thickness-twisted shell and solid-element blade models. Mechanical Systems and Signal Processing, 108, 1-20 (2018) doi:10.1016/j.ymssp.2018.02.002 |
[33] |
HE, Q., XUAN, H. J., LIU, L. L., HONG, W. R., and WU, R. R. Perforation of aero-engine fan casing by a single rotating blade. Aerospace Science and Technology, 25, 234-241 (2013) doi:10.1016/j.ast.2012.01.010 |
[34] |
HAMMOND, C. E. An application of Floquet theory to prediction of mechanical instability. Journal of the American Helicopter Society, 19, 14-23 (1974) doi:10.4050/JAHS.19.14 |
[35] |
TAN, T. H., LEE, H. P., and LENG, G. S. B. Dynamic stability of a radially rotating beam subjected to base excitation. Computer Methods in Applied Mechanics and Engineering, 146, 265-279 (1997) doi:10.1016/S0045-7825(96)01238-8 |
[36] |
OZGUMUS, O. O. and KAYA, M. O. Energy expressions and free vibration analysis of a rotating double tapered Timoshenko beam featuring bending-torsion coupling. International Journal of Engineering Science, 45, 562-586 (2007) doi:10.1016/j.ijengsci.2007.04.005 |
[37] |
HODGES, D. H. Torsion of pretwisted beams due to axial loading. Journal of Applied Mechanics, 47, 393-397 (1980) doi:10.1115/1.3153675 |
[38] |
BAZOUNE, A., KHULIEF, Y. A., and STEPHEN, N. G. Shape functions of three-dimensional Timoshenko beam element. Journal of Sound and Vibration, 259, 473-480 (2003) doi:10.1006/jsvi.2002.5122 |
[39] |
WANG, M. Z., WANG, W., and WU, J. K. Course of Elastic Mechanics (in Chinese) (in Chinese), Beijing University Press,, Beijing, 146-150 (2002)
|
[40] |
WILLIAM, C. Vibrations of pre-twisted cantilever blading. Proceedings of the Institution of Mechanical Engineers, 173, 343-374 (1959) doi:10.1243/PIME_PROC_1959_173_038_02 |
[41] |
RAO, J. S. Flexural vibration of pretwisted tapered cantilever blades. Journal of Engineering for Industry, 94, 343-346 (1972) doi:10.1115/1.3428148 |
[42] |
ZENG, J., CHEN, K. K., MA, H., DUAN, T. T., and WEN, B. C. Vibration response analysis of a cracked rotating compressor blade during run-up process. Mechanical Systems and Signal Processing, 118, 568-583 (2019) doi:10.1016/j.ymssp.2018.09.008 |
[43] |
ANSYS Inc. ANSYS Mechanical APDL and Mechanical Applications Theory Reference, ANSYS Release 13.0, ANSYS Inc., Lincoln (2010)
|