Appl. Math. Mech. -Engl. Ed.   2019, Vol. 40 Issue (5): 601-620     PDF       
http://dx.doi.org/10.1007/s10483-019-2476-6
Shanghai University
0

Article Information

WANG Yanqing, LIU Yunfei, ZU J. W.
Nonlinear free vibration of piezoelectric cylindrical nanoshells
Applied Mathematics and Mechanics (English Edition), 2019, 40(5): 601-620.
http://dx.doi.org/10.1007/s10483-019-2476-6

Article History

Received Sep. 12, 2018
Revised Oct. 30, 2018
Nonlinear free vibration of piezoelectric cylindrical nanoshells
Yanqing WANG1,2, Yunfei LIU1, J. W. ZU3     
1. Department of Mechanics, Northeastern University, Shenyang 110819, China;
2. Key Laboratory of Ministry of Education on Safe Mining of Deep Metal Mines, Northeastern University, Shenyang 110819, China;
3. Schaefer School of Engineering and Science, Stevens Institute of Technology, New Jersey 07030, U. S. A
Abstract: The nonlinear vibration characteristics of the piezoelectric circular cylindrical nanoshells resting on an elastic foundation are analyzed. The small scale effect and thermo-electro-mechanical loading are taken into account. Based on the nonlocal elasticity theory and Donnell's nonlinear shell theory, the nonlinear governing equations and the corresponding boundary conditions are derived by employing Hamilton's principle. Then, the Galerkin method is used to transform the governing equations into a set of ordinary differential equations, and subsequently, the multiple-scale method is used to obtain an approximate analytical solution. Finally, an extensive parametric study is conducted to examine the effects of the nonlocal parameter, the external electric potential, the temperature rise, and the Winkler-Pasternak foundation parameters on the nonlinear vibration characteristics of circular cylindrical piezoelectric nanoshells.
Key words: piezoelectric cylindrical nanoshell    nonlinear vibration    Donnell's nonlinear shell theory    nonlocal elasticity theory    multiple-scale method    size effect    
1 Introduction

Due to the excellent electrical and mechanical properties, piezoelectric materials have been utilized in many engineering fields, e.g., aerospace engineering, medical imaging, and mechanical engineering. The coupled electromechanical properties of piezoelectric materials make them well suited for sensors, actuators, resonant ultrasonic inspection devices, micro piezoelectric power generators, etc.[1-6]. By combining piezoelectric material technology with nanotechnology, the piezoelectric nanostructure based sensing technology can be developed. Since there are more than 6 000 different types of sensors in an aircraft, the interconnected cabling and maintenance are costly. It is expected that the combination of piezoelectric nanostructure based sensing technology and wireless transmission can provide promising low-cost energy storage solution in aircrafts.

Pan et al.[7] first reported the ZnO piezoelectric nanostructures on Science. Subsequently, various piezoelectric nanomaterials and nanostructures are concerned and reported[8-12]. Piezoelectric nanostructures possess significant physical/chemical properties, e.g., thermal properties, electrical properties, and mechanical properties. Therefore, they are considered as ideal candidates for constructing nanodevices[13-15].

Piezoelectric nanostructures have the dimension ranging from a few nanometers to several hundred nanometers, where the effects of the scale parameters need to be considered. The size effect of piezoelectric nanostructures has been observed in experiments. Using a piezoresponse force microscopy, Zhao et al.[16] observed that the effective piezoelectric coefficient of ZnO nanowires was much larger than that of the bulk ZnO. Chen et al.[17] reported size-dependent Young's modulus in ZnO nanowires. They showed that Young's modulus for the nanowires with diameters smaller than 120 nm increased dramatically when the diameter decreased. These studies give important evidences for the size-dependent material properties of piezoelectric nanostructures. Therefore, the size effect must be considered in both theoretical and experimental studies on piezoelectric nanostructures.

The nonlocal elasticity theory incorporating the size effect was proposed by Eringen[18-20]. Ke and Wang[21] first extended this nonclassical continuum theory to study the size-dependent mechanical behavior of piezoelectric nanostructures. Wang and Wang[22] studied the surface and nonlocal effects on the electromechanical coupling behavior of a piezoelectric nanowire. Ansari et al.[23] investigated the free vibration of piezoelectric nanobeams in the vicinity of postbuckling domain via the nonlocal elasticity theory. Ke et al.[24] studied the linear thermo-electro-mechanical vibration of piezoelectric cylindrical nanoshells under various boundary conditions. Based on the Kirchhoff plate theory, Jandaghian and Rahmani[25] investigated the linear vibration of functionally graded piezoelectric nanoplates. Razavi et al.[26] studied the electromechanical vibration of a functionally graded piezoelectric cylindrical nanoshell. Using Love's thin shell model, Kheibari and Beni[27] studied the linear vibration of single-walled piezoelectric nanotubes. It is noted that all the above-mentioned studies are concentrated on the linear vibrations of piezoelectric nanostructures. The nonlinear vibration investigation on piezoelectric nanostructures, however, is very limited in the literature. Ke et al.[28] studied the nonlinear vibration of piezoelectric nanobeams based on the nonlocal elasticity theory. Asemi et al.[29] analyzed the nonlinear vibration of piezoelectric nanoelectro-mechanical resonators. Ansari and Gholami[30] investigated the nonlinear free vibration of magneto-electro-thermo-elastic nanoplates. Liu et al.[31] studied the nonlinear free vibration of rectangular piezoelectric nanoplates resting on the Winkler foundation. All these studies focus on piezoelectric nanoscale beams and plates.

For solving the nonlinear vibration problem of structures, different analytical and numerical methods have been used and proposed. Wang[32] and Wang and Zu[33-34] used the harmonic balance method to investigate the nonlinear vibration of functionally graded, piezoelectric, and porous plates. Ding et al.[35] proposed a simple and efficient method to determine the convergence of the modal truncation of the vibration analysis of continuum on elastic foundations. This facilitates the vibration analysis of complex elastic systems[36]. Wang et al.[37] used the multiple-scale method to study the nonlinear vibration of moving plates immersed in fluids.

In this paper, the nonlinear vibration of piezoelectric nanoshells resting on an elastic foundation, subjected to thermo-electro-mechanical loads, is first proposed. The Donnell nonlinear shell theory and the nonlocal elasticity theory are used to model the system. The governing equations and boundary conditions are developed by using the Hamilton principle. The Galerkin method together with the multiple-scale method is used to approximately analyze the nonlinear vibration of the piezoelectric nanoshells.

2 Preliminaries 2.1 Nonlocal elasticity theory for piezoelectric materials

In Eringen's nonlocal elastic theory[20], the stress field at a reference point in an elastic continuum depends not only on the strain at that point but also on the strains at all other points. The nonlocal elasticity theory can well explain some phenomena related to the scales of atoms and molecules, e.g., high-frequency vibration and wave scattering. The most general form of the constitutive relation for nonlocal elasticity involves an integral over the whole body. The basic equations for uniform and non-local piezoelectric solids can be written as follows[18, 24]:

(1)
(2)
(3)

where i, j, k, l =1, 2, 3. σij, εij, Di, Ei, and ui represent the components of the stress, the strain, the electric displacement, the electric field, and the displacement, respectively. cijkl, ekij, sik, βij, pi, and ρ are the components of the elasticity tensor, the piezoelectric tensor, the dielectric tensor, the thermal modulus tensor, the pyroelectric vector, and the mass density, respectively. ΔT denotes the temperature change. is the electric potential. α0(|x'-x|, e0 a/l) is the nonlocal kernel function. e0a/l is the scale parameter. e0a is the scale coefficient revealing the size effect on the response of nanostructures. e0 is a material constant determined experimentally or approximated by matching the dispersion curves of the plane waves with those of the atomic lattice dynamics. a and l represent the internal and external characteristic lengths of the nanostructures, respectively. x' represents the coordinate of any material point in the area except x. |x'-x| represents the Euclidean distance.

Equivalent differential forms can be used to represent the overall constitutive relation as follows[19]:

(4)
(5)

where is the Laplace operator.

2.2 Nonlocal piezoelectric cylindrical nanoshell model

Consider a piezoelectric cylindrical nanoshell composed of PZT-4 and resting on an elastic foundation. Figure 1 shows the geometry of the nanoshell with the length L, the middle-surface radius R, and the thickness h. The elastic medium is stimulated by employing the Winkler-Pasternak foundation model. Additionally, the nanoshell is subjected to an electric potential (x, θ, z, t) and a uniform temperature change ΔT. u(x, θ, t), v(x, θ, t), and w(x, θ, t) are the displacements of the points in the middle plane of the piezoelectric nanoshell along the x-, θ-, and z-axes, respectively.

Fig. 1 Schematic diagram of a piezoelectric cylindrical nanoshell resting on an elastic foundation

According to the Kirchhoff-Love hypothesis, the displacement fields are[38]

(6)
(7)
(8)

where t is the time, and u1 (x, θ, z, t), u2 (x, θ, z, t), and u3 (x, θ, z, t) are the displacements of an arbitrary point of the nanoshell along the x-, θ-, and z-axes, respectively.

Based on Donnell's nonlinear shell theory, the strain components εxx, εθθ, and γxθ at an arbitrary point of the nanoshell are[39]

(9)

where εx, 0, εθ, 0, and γxθ, 0 represent the middle-surface strains, kx, kθ, and kxθ are the curvature and torsion of the middle surface, and z is the distance of the arbitrary point of the nanoshell from the middle surface.

The corresponding expressions in the above equation are

(10)
(11)

Following Wang[40], the distribution of the electric potential along the thickness of the piezoelectric nanoshell is assumed as follows:

(12)

where β=π/h, Φ(x, θ, t) is the spatial and time variation of the electric potential in the x- and θ-directions, and V0 represents the initial external electric potential applied to the piezoelectric nanoshell.

Using Eq. (12), the electric field components Ei (i=x, θ, z) can be written as follows[41]:

(13)
(14)

For the piezoelectric cylindrical nanoshell, the nonlocal constitutive relationship of Eqs. (4) and (5) can be given by[42-43]

(16)
(17)
(18)
(19)
(20)
(21)

Thereinto, , and are defined as follows:

(22)

The strain energy Πs of the piezoelectric cylindrical nanoshell can be written as follows:

(23)

in which the resultant forces and moments can be, respectively, calculated as follows:

(24)
(25)

The kinetic energy Πk is expressed as follows[44]:

(26)

where , and the rotatory inertia term is neglected due to its slight effect.

Moreover, the work ΠF1 done by the external forces can be given by

(27)

where (NTx, NTθ) and (NEx, NEθ) are the thermal and electrical forces induced by the uniform temperature rise ΔT and the uniform external electric potential V0, respectively, and

(28)

The work ΠF2 done by the nonlinear Winkler-Pasternak foundation can be calculated as follows:

(29)

where kw is the Winkler foundation parameter, and kp is the Pasternak foundation parameter.

Using Hamilton's principle

(30)

and substituting Eqs. (23), (26), (27), and (29) into Eq. (30), we can obtain the following differential equations of motion:

(31)
(32)
(33)
(34)

where

(35)

The corresponding boundary conditions are

(36)
(37)
(38)
(39)
(40)
(41)

where nx and nθ denote the directional cosines of the outward unit normal to the boundaries of the midplane.

From Eqs. (16)-(21), we obtain

(42)
(43)
(44)
(45)
(46)
(47)
(48)
(49)
(50)

where

Substituting Eqs. (42)-(50) into Eqs. (31)-(34) yields

(51)
(52)
(53)
(54)

where

(55)
(56)

It is assumed that the electric potential is equal to zero at both ends of the nanoshell. Then, the simply supported boundary condition can be written as follows:

(57)
3 Solution procedure

In this study, we consider that the piezoelectric cylindrical nanoshell is simply-supported at both ends. Therefore, the displacement functions can be written as follows:

(58)
(59)
(60)
(61)

where m is the number of the axial half waves, and n is the number of the circumferential waves. umn(t), vmn(t), wmn(t), and Φmn(t) are displacement amplitude components.

By substituting Eqs. (58)-(61) into Eqs. (51)-(54) and then utilizing the Galerkin method, we can obtain the following ordinary differential equations:

(62)
(63)
(64)
(65)

where the coefficients Gi, j are given in Appendix A.

In Eqs. (62) and (63), the inertia terms and can be neglected due to their insignificant effects compared with the transverse inertia term. Thus, by solving Eqs. (62), (63), and (65) with respect to umn, vmn, and Φmn and then inserting the results in Eq. (64), we have

(66)

where

(67)
(68)
(69)

From Eq. (66), the linear circular frequency can be obtained as follows:

(70)

The initial conditions are set as follows:

(71)

where wmax denotes the maximum of wmn(t).

For the convenience of calculation, let us introduce the following dimensionless variables:

(72)

Applying Eq. (72) into Eqs. (66) and (71) yields

(73)
(74)

where

Then, we use the multiple-scale method[45-46] to solve Eq. (73). First, let us introduce the scaled time as follows:

(75)

where ε is a small dimensionless parameter, and is used as a bookkeeping device.

The time derivatives can be written in terms of Tn as follows:

(76)

where

(77)

The response is the function of different scaled time, and thus can be written as follows:

(78)

Substituting Eqs. (76) and (78) into Eq. (73) and equating the coefficients of like powers of ε, we obtain

(79)
(80)
(81)

The solution of Eq. (79) can be assumed as follows:

(82)

Substituting Eq. (82) into Eq. (80) gives

(83)

To eliminate the secular terms in Eq. (83), should be satisfied. Thus, we have

(84)

Introducing Eqs. (82) and (84) into Eq. (81) yields

(85)

The solvability condition of the above equation requires

(86)

In order to solve the above equation, we seek the solution in the form of

(87)

Utilizing Eq. (87) in Eq. (86) and separating the real and imaginary parts, we obtain

(88)

The integral of Eq. (88) gives

(89)

Thus, Eq. (87) can be rewritten as follows:

(90)

Substituting Eq. (90) into Eq. (78) yields

(91)

where ωNL is the normalized nonlinear frequency of the piezoelectric cylindrical nanoshell.

Substituting the initial conditions into Eq. (74), we obtain

(92)

The results are presented as the nonlinear frequency ratio calculated as follows:

(93)

where

4 Results and discussion

For the purpose of examining the validity of the present analysis, the linear natural frequencies of a macroscopic isotropic thin cylindrical shell are calculated and compared with the literature in Table 1. The material and geometric parameters of the shell are as follows:

Table 1 Comparison of the linear natural frequency ωL (Hz) of a macroscopic isotropic thin cylindrical shell

The results show good agreement.

The second comparison is related to a nanoshell incorporating the size effect and piezoelectric effect. Table 2 lists the comparison of the linear natural frequencies of a PZT-4 piezoelectric nanoshell with

Table 2 Comparison of the linear natural frequency ωL (GHz) of a simply supported homogeneous piezoelectric nanoshell (μ=e0a/L)

The material properties of PZT-4 are shown in Table 3. It is found that the present results agree well with those given by Ke et al.[24], bespeaking the validity of the present study.

Table 3 Material properties of PZT-4[24]

In order to validate the present nonlinear analysis, we consider anisotropic homogeneous cylindrical shell to make a comparison of nonlinear results in Table 4. The following parameters are used:

Table 4 Comparison of the nonlinear frequency ratio of an anisotropic homogeneous cylindrical shell with the initial condition of wmax/h=1

The comparison study shows that the results obtained from the present method are in good agreement with the existing ones. The small differences between the present results and those of Raju and Rao[50] and Rafiee et al.[51] might be due to the different shell theories and different solving processes.

In what follows, we will analyze the nonlinear vibration of the piezoelectric cylindrical nanoshell shown in Fig. 1. The material properties of the nanoshell are displayed in Table 3. If not specified, the following parameters of the nanoshell are considered:

Figure 2 shows the variations of the linear natural frequency against the circumferential wave number of the piezoelectric nanoshell. It can be found that the natural frequency of the nanoshell initially decreases and then increases when the circumferential wave number increases. The fundamental natural frequency occurs at n=5. Therefore, the mode (m=1, n=5) corresponds to the lowest natural frequency of the piezoelectric nanoshell. This mode is chosen as a representative mode in the following discussion.

Fig. 2 Variations of the linear natural frequency ωL against the circumferential wave number n, where kp=0, and kw=0

Figure 3 illustrates the nonlinear frequency ratio versus the dimensionless vibration amplitude of the piezoelectric nanoshell for different nonlocal parameters. Obviously, the nonlinear frequency ratio increases with the increase in the dimensionless vibration amplitude. The nanoshell exhibits a hardening behavior. At a given vibration amplitude, a larger nonlocal parameter leads to smaller linear and nonlinear frequencies but a higher nonlinear frequency ratio. The reason is that the nonlocal effect tends to decrease the stiffness of the nanoshell and hence decrease the values of the linear and nonlinear frequencies. Moreover, the nonlinear frequency drops faster than the linear frequency, which results in the increase in the nonlinear frequency ratio.

Fig. 3 Nonlinear frequency ratio versus the dimensionless vibration amplitude wmax/h for different nonlocal parameters, where kp=0, and kw=0 (color online)

Figure 4 presents the effects of the external electric potential on the nonlinear frequency ratio of the piezoelectric nanoshell. As we can see, the positive/negative potential decreases/increases the linear and nonlinear frequencies of the piezoelectric nanoshell. This is due to the fact that the axial compressive and tensile forces are generated in the nanoshell by the applied positive and negative potentials, respectively. At a given vibration amplitude, it is found that a change in the external electric potential from -0.000 2 V to 0.000 2 V leads to the increase in the nonlinear frequency ratio.

Fig. 4 Effects of the external electric potential V0 on the nonlinear frequency ratio of the piezoelectric nanoshell (color online)

Figure 5 shows the frequency ratio versus the external electric potential for different dimensionless vibration amplitudes of the piezoelectric nanoshell, where the frequency ratio is defined as the non-classical nonlinear frequency to its classical counterpart. Since the classical nonlinear frequency is calculated without the consideration of the nonlocal effect, i.e., e0a=0, the frequency ratio increases with the increase in the external electric potential. This characteristic indicates that, at a larger external electric potential, the size effect on the vibration characteristics of the piezoelectric nanoshell is more notable.

Fig. 5 Frequency ratio versus the external electric potential V0 for different dimensionless vibration amplitudes (color online)

The effects of the temperature change on the nonlinear frequency ratio of the piezoelectric nanoshell are depicted in Fig. 6. It can be found that when the temperature change increases, the linear and nonlinear frequencies decrease accordingly. The reason is that a larger temperature change results in a reduction in the nanoshell stiffness, and hence leads to lower linear and nonlinear frequencies of the nanoshell. Additionally, at a given vibration amplitude, the nonlinear frequency ratio increases with the increase in the temperature change. Comparing the effects of the external electric potential and temperature change on the nonlinear vibration behaviors (see Figs. 4 and 6), we find that the external electric potential has a greater effect than the temperature change from the numerical point of view. This phenomenon can be explained because the coefficients of the thermal force NT and the electrical force NE have the relationship , as deduced above.

Fig. 6 Effects of the temperature change ΔT on the nonlinear frequency ratio of the piezoelectric nanoshell (color online)

The frequency ratio versus the temperature change for different dimensionless vibration amplitudes of the piezoelectric nanoshell is displayed in Fig. 7. It is observed that the increasing temperature change leads to the increase in the frequency ratio. This characteristic is quite obvious when the dimensionless vibration amplitude is large, showing that the nonlocal effect is quite significant when the dimensionless vibration amplitude is large.

Fig. 7 Frequency ratio versus temperature change ΔT for different dimensionless vibration amplitudes wmax/h of the piezoelectric nanoshell (color online)

The nonlinear frequency ratio versus the dimensionless vibration amplitude of the piezoelectric nanoshell for different Winkler foundation parameters is plotted in Fig. 8. It can be found that when the Winkler foundation parameter increases, the linear and nonlinear frequencies increase, whereas the nonlinear frequency ratio decreases.

Fig. 8 Nonlinear frequency ratio versus the dimensionless vibration amplitude wmax/h for different Winkler foundation parameters (kp=0) (color online)

Figure 9 illustrates the effects of the Winkler foundation parameter on the frequency ratio of the piezoelectric nanoshell. One can find that with the increase in the Winkler foundation parameter, the frequency ratio decreases. This indicates that the nonlocal effect is weaker when the foundation parameter becomes larger. In addition, at a given Winkler foundation parameter, the nonlocal effect is more pronounced when the dimensionless vibration amplitude is larger.

Fig. 9 Effects of the Winkler foundation parameter on the frequency ratio of the piezoelectric nanoshell (kp=0) (color online)

Figure 10 presents the nonlinear frequency ratio versus the dimensionless vibration amplitude of the piezoelectric nanoshell for different Pasternak foundation parameters. It is found that when the Pasternak foundation parameter increases, the linear and nonlinear frequencies increase while the nonlinear frequency ratio decreases. Figure 11 shows the effects of the Pasternak foundation parameter on the frequency ratio of the piezoelectric nanoshell. It can be found that the frequency ratio decreases with the increase in the Pasternak foundation parameter, which means that the Pasternak foundation can weaken the nonlocal effect.

Fig. 10 Nonlinear frequency ratio versus the dimensionless vibration amplitude wmax /h for different Pasternak foundation parameters (kw=0) (color online)
Fig. 11 Effects of the Pasternak foundation parameter kp on the frequency ratio of the piezoelectric nanoshell (kw=0) (color online)

The effects of the length-to-radius ratio on the nonlinear frequency ratio are presented in Fig. 12. At a given vibration amplitude, the nonlinear frequency ratio increases when the length-to-radius ratio increases. Figure 13 shows the frequency ratio versus the length-to-radius ratio of the piezoelectric nanoshell for different dimensionless vibration amplitudes. It is found that when the length-to-radius ratio increases, the frequency ratio increases initially, and then tends to keep constant.

Fig. 12 Effects of the length-to-radius ratio on the nonlinear frequency ratio of the piezoelectric nanoshell (color online)
Fig. 13 Frequency ratio versus the length-to-radius ratio for different dimensionless vibration amplitudes (color online)
5 Conclusions

The nonlinear free vibration of piezoelectric cylindrical nanoshells resting on a Winkler-Pasternak foundation, subjected to thermal and electrical loads, is investigated. The theoretical model is established based on Donnell's nonlinear shell theory and Eringen's nonlocal elasticity theory. The nonlinear governing equations and boundary conditions are derived by using Hamilton's principle. With the Galerkin method, the partial differential equations are discretized into a set of nonlinear ordinary differential equations. Afterwards, the method of multiple scales is used to approximately analyze the nonlinear free vibration of the piezoelectric nanoshells. The results are summerized as follows:

(ⅰ) Larger vibration amplitudes lead to higher nonlinear frequency ratios, and higher nonlocal parameters lead to lower linear and nonlinear frequencies but higher nonlinear frequency ratios.

(ⅱ) A change in the external electric potential from a negative value to a positive one leads to the increase in the nonlinear frequency ratio, and the nonlinear frequency ratio increases with the increase in the temperature change. Moreover, when the external electric potential and the temperature change increase, the nonlocal effects on the vibration characteristics of the piezoelectric nanoshells increase.

(ⅲ) The nonlinear frequency ratio decreases when the length-to-radius ratio decreases. The nonlocal effect is quite significant at small length-to-radius ratios, but does not change notably when the length-to-radius ratio reaches a certain value.

(ⅳ) The Winkler-Pasternak foundation results in the increase in the linear and nonlinear frequencies but the decrease in the nonlinear frequency ratio of the piezoelectric nanoshells. Moreover, the Winkler-Pasternak foundation can weaken the nonlocal effect on the piezoelectric nanoshells.

Appendix A
References
[1]
GAO, Z., ZHU, X., FANG, Y., and ZHANG, H. Active monitoring and vibration control of smart structure aircraft based on FBG sensors and PZT actuators. Aerospace Science and Technology, 63, 101-109 (2017) doi:10.1016/j.ast.2016.12.027
[2]
SCHINDEL, D. W., HUTCHINS, D. A., and GRANDIA, W. A. Capacitive and piezoelectric air-coupled transducers for resonant ultrasonic inspection. Ultrasonics, 34, 621-627 (1996) doi:10.1016/0041-624X(96)00063-7
[3]
LU, F., LEE, H. P., and LIM, S. P. Modeling and analysis of micro piezoelectric power generators for micro-electromechanical-systems applications. Smart Materials and Structures, 13, 57-63 (2003)
[4]
YANG, M. and QIAO, P. Modeling and experimental detection of damage in various materials using the pulse-echo method and piezoelectric sensors/actuators. Smart Materials and Structures, 14, 1083-1100 (2005) doi:10.1088/0964-1726/14/6/001
[5]
GUPTA, V., SHARMA, M., and THAKUR, N. Optimization criteria for optimal placement of piezoelectric sensors and actuators on a smart structure:a technical review. Journal of Intelligent Material Systems and Structures, 21, 1227-1243 (2010) doi:10.1177/1045389X10381659
[6]
AKSEL, E. and JONES, J. L. Advances in lead-free piezoelectric materials for sensors and actuators. Sensors, 10, 1935-1954 (2010) doi:10.3390/s100301935
[7]
PAN, Z. W., DAI, Z. R., and WANG, Z. L. Nanobelts of semiconducting oxides. Science, 291, 1947-1949 (2001) doi:10.1126/science.1058120
[8]
WU, C., KAHN, M., and MOY, W. Piezoelectric ceramics with functional gradients:a new application in material design. Journal of the American Ceramic Society, 79, 809-812 (1996)
[9]
PARK, K. I., XU, S., LIU, Y., HWANG, G. T., KANG, S. J. L., WANG, Z. L., and LEE, K. J. Piezoelectric BaTiO3 thin film nanogenerator on plastic substrates. Nano Letters, 10, 4939-4943 (2010) doi:10.1021/nl102959k
[10]
XU, S. and WANG, Z. L. One-dimensional ZnO nanostructures:solution growth and functional properties. Nano Research, 4, 1013-1098 (2011) doi:10.1007/s12274-011-0160-7
[11]
ZHU, C. S., FANG, X. Q., LIU, J. X., and LI, H. Y. Surface energy effect on nonlinear free vibration behavior of orthotropic piezoelectric cylindrical nano-shells. European Journal of MechanicsA/Solids, 66, 423-432 (2017) doi:10.1016/j.euromechsol.2017.08.001
[12]
FANG, X. Q., ZHU, C. S., LIU, J. X., and LIU, X. L. Surface energy effect on free vibration of nano-sized piezoelectric double-shell structures. Physica B:Condensed Matter, 529, 41-56 (2018) doi:10.1016/j.physb.2017.10.038
[13]
GAO, P. X., SONG, J., LIU, J., and WANG, Z. L. Nanowire piezoelectric nanogenerators on plastic substrates as flexible power sources for nanodevices. Advanced Materials, 19, 67-72 (2007) doi:10.1002/(ISSN)1521-4095
[14]
SEOL, M. L., IM, H., MOON, D. I., WOO, J. H., KIM, D., CHOI, S. J., and CHOI, Y. K. Design strategy for a piezoelectric nanogenerator with a well-ordered nanoshell array. ACS Nano, 7, 10773-10779 (2013) doi:10.1021/nn403940v
[15]
WANG, X. and SHI, J. Piezoelectric nanogenerators for self-powered nanodevices. Piezoelectric Nanomaterials for Biomedical Applications, Springer, New York, 135-172 (2012)
[16]
ZHAO, M. H., WANG, Z. L., and MAO, S. X. Piezoelectric characterization of individual zinc oxide nanobelt probed by piezoresponse force microscope. Nano Letters, 4, 587-590 (2004) doi:10.1021/nl035198a
[17]
CHEN, C. Q., SHI, Y., ZHANG, Y. S., ZHU, J., and YAN, Y. J. Size dependence of Young's modulus in ZnO nanowires. Physical Review Letters, 96, 75505 (2006) doi:10.1103/PhysRevLett.96.075505
[18]
ERINGEN, A. C. Nonlocal Continuum Field Theories, Springer, New York (2002)
[19]
ERINGEN, A. C. On differential equations of nonlocal elasticity and solutions of screw dislocation and surface waves. Journal of Applied Physics, 54, 4703-4710 (1983) doi:10.1063/1.332803
[20]
ERINGEN, A. C. Nonlocal polar elastic continua. International Journal of Engineering Science, 10, 1-16 (1972) doi:10.1016/0020-7225(72)90070-5
[21]
KE, L. L. and WANG, Y. S. Thermoelectric-mechanical vibration of piezoelectric nanobeams based on the nonlocal theory. Smart Materials and Structures, 21, 025018 (2012) doi:10.1088/0964-1726/21/2/025018
[22]
WANG, K. F. and WANG, B. L. The electromechanical coupling behavior of piezoelectric nanowires:surface and small-scale effects. Europhysics Letters, 97, 66005 (2012) doi:10.1209/0295-5075/97/66005
[23]
ANSARI, R., OSKOUIE, M. F., GHOLAMI, R., and SADEGHI, F. Thermo-electro-mechanical vibration of postbuckled piezoelectric Timoshenko nanobeams based on the nonlocal elasticity theory. Composites Part B:Engineering, 89, 316-327 (2016) doi:10.1016/j.compositesb.2015.12.029
[24]
KE, L. L., WANG, Y. S., and REDDY, J. N. Thermo-electro-mechanical vibration of sizedependent piezoelectric cylindrical nanoshells under various boundary conditions. Composite Structures, 116, 626-636 (2014) doi:10.1016/j.compstruct.2014.05.048
[25]
JANDAGHIAN, A. A. and RAHMANI, O. Size-dependent free vibration analysis of functionally graded piezoelectric plate subjected to thermo-electro-mechanical loading. Journal of Intelligent Material Systems and Structures, 28, 3039-3053 (2017) doi:10.1177/1045389X17704920
[26]
RAZAVI, H., BABADI, A. F., and BENI, Y. T. Free vibration analysis of functionally graded piezoelectric cylindrical nanoshell based on consistent couple stress theory. Composite Structures, 160, 1299-1309 (2017) doi:10.1016/j.compstruct.2016.10.056
[27]
KHEIBARI, F. and BENI, Y. T. Size dependent electro-mechanical vibration of single-walled piezoelectric nanotubes using thin shell model. Materials and Design, 114, 572-583 (2017) doi:10.1016/j.matdes.2016.10.041
[28]
KE, L. L., WANG, Y. S., and WANG, Z. D. Nonlinear vibration of the piezoelectric nanobeams based on the nonlocal theory. Composite Structures, 94, 2038-2047 (2012) doi:10.1016/j.compstruct.2012.01.023
[29]
ASEMI, S. R., FARAJPOUR, A., and MOHAMMADI, M. Nonlinear vibration analysis of piezoelectric nanoelectromechanical resonators based on nonlocal elasticity theory. Composite Structures, 116, 703-712 (2014) doi:10.1016/j.compstruct.2014.05.015
[30]
ANSARI, R. and GHOLAMI, R. Size-dependent nonlinear vibrations of first-order shear deformable magneto-electro-thermo elastic nanoplates based on the nonlocal elasticity theory. International Journal of Applied Mechanics, 8, 1650053 (2016) doi:10.1142/S1758825116500538
[31]
LIU, C., KE, L. L., WANG, Y. S., and YANG, J. Nonlinear vibration of nonlocal piezoelectric nanoplates. International Journal of Structural Stability and Dynamics, 15, 1540013 (2015) doi:10.1142/S0219455415400131
[32]
WANG, Y. Q. Electro-mechanical vibration analysis of functionally graded piezoelectric porous plates in the translation state. Acta Astronautica, 143, 263-271 (2018) doi:10.1016/j.actaastro.2017.12.004
[33]
WANG, Y. Q. and ZU, J. W. Nonlinear steady-state responses of longitudinally traveling functionally graded material plates in contact with liquid. Composite Structures, 164, 130-144 (2017) doi:10.1016/j.compstruct.2016.12.053
[34]
WANG, Y. Q. and ZU, J. W. Nonlinear dynamics of a translational FGM plate with strong mode interaction. International Journal of Structural Stability and Dynamics, 18, 1850031 (2018) doi:10.1142/S0219455418500311
[35]
DING, H., CHEN, L. Q., and YANG, S. P. Convergence of Galerkin truncation for dynamic response of finite beams on nonlinear foundations under a moving load. Journal of Sound and Vibration, 331, 2426-2442 (2012) doi:10.1016/j.jsv.2011.12.036
[36]
DING, H., YANG, Y., CHEN, L. Q., and YANG, S. P. Vibration of vehicle-pavement coupled system based on a Timoshenko beam on a nonlinear foundation. Journal of Sound and Vibration, 333, 6623-6636 (2014) doi:10.1016/j.jsv.2014.07.016
[37]
WANG, Y. Q., HUANG, X. B., and LI, J. Hydroelastic dynamic analysis of axially moving plates in continuous hot-dip galvanizing process. International Journal of Mechanical Sciences, 110, 201-216 (2016) doi:10.1016/j.ijmecsci.2016.03.010
[38]
SOEDEL, W. Vibrations of Shells and Plates, CRC Presss, Boca Raton (2004)
[39]
AMABILI, M. Nonlinear Vibrations and Stability of Shells and Plates, Cambridge University Press, Cambridge (2008)
[40]
WANG, Q. On buckling of column structures with a pair of piezoelectric layers. Engineering Structures, 24, 199-205 (2002) doi:10.1016/S0141-0296(01)00088-8
[41]
LI, H. Y., LIN, Q. R., LIU, Z. X., and WANG, C. Free vibration of piezoelastic laminated cylindrical shells under hydrostatic pressure. International Journal of Solids and Structures, 38, 7571-7585 (2001) doi:10.1016/S0020-7683(01)00008-7
[42]
ZHANG, D. P., LEI, Y. J., and SHEN, Z. B. Thermo-electro-mechanical vibration analysis of piezoelectric nanoplates resting on viscoelastic foundation with various boundary conditions. International Journal of Mechanical Sciences, 131, 1001-1015 (2017)
[43]
ZHAO, M. H., QIAN, C. F., LEE, S. W. R., TONG, P., SUEMASU, H., and ZHANG, T. Y. Electro-elastic analysis of piezoelectric laminated plates. Advanced Composite Materials, 16, 63-81 (2007) doi:10.1163/156855107779755273
[44]
WANG, Y. Q., YE, C., and ZU, J. W. Identifying temperature effect on vibrations of functionally graded cylindrical shells with porosities. Applied Mathematics and Mechanics (English Edition), 39(11), 1587-1604 (2018) doi:10.1007/s10483-018-2388-6
[45]
NAYFEH, A. H. and MOOK, D. T. Nonlinear Oscillations, John Wiley & Sons, New York (2008)
[46]
WANG, Y. Q. and ZU, J. W. Instability of viscoelastic plates with longitudinally variable speed and immersed in ideal liquid. International Journal of Applied Mechanics, 9, 1750005 (2017) doi:10.1142/S1758825117500053
[47]
GONÇALVES, P. B. and RAMOS, N. R. S. S. Numerical method for vibration analysis of cylindrical shells. Journal of Engineering Mechanics, 123, 544-550 (1997) doi:10.1061/(ASCE)0733-9399(1997)123:6(544)
[48]
DYM, C. L. Some new results for the vibrations of circular cylinders. Journal of Sound and Vibration, 29, 189-205 (1973) doi:10.1016/S0022-460X(73)80134-8
[49]
SHEN, H. S. and XIANG, Y. Nonlinear vibration of nanotube-reinforced composite cylindrical shells in thermal environments. Computer Methods in Applied Mechanics and Engineering, 213, 196-205 (2012)
[50]
RAJU, K. K. and RAO, G. V. Large amplitude asymmetric vibrations of some thin shells of revolution. Journal of Sound and Vibration, 44, 327-333 (1976) doi:10.1016/0022-460X(76)90505-8
[51]
RAFIEE, M., MOHAMMADI, M., ARAGH, B. S., and YAGHOOBI, H. Nonlinear free and forced thermo-electro-aero-elastic vibration and dynamic response of piezoelectric functionally graded laminated composite shells, part Ⅱ:numerical results. Composite Structures, 103, 188-196 (2013) doi:10.1016/j.compstruct.2012.12.050