Cylinder Cone Metal

Embed Size (px)

Citation preview

  • 8/20/2019 Cylinder Cone Metal

    1/19

    Free vibrational characteristics of isotropic coupledcylindrical–conical shells

    Mauro Caresta , Nicole J. Kessissoglou

    School of Mechanical and Manufacturing Engineering, The University of New South Wales, Sydney NSW 2052, Australia

    a r t i c l e i n f o

     Article history:

    Received 11 March 2009

    Received in revised form

    29 September 2009

    Accepted 4 October 2009

    Handling Editor: L.G. ThamAvailable online 28 October 2009

    a b s t r a c t

    This paper presents the free vibrational characteristics of isotropic coupled conical–

    cylindrical shells. The equations of motion for the cylindrical and conical shells are

    solved using two different methods. A wave solution is used to describe the

    displacements of the cylindrical shell, while the displacements of the conical sections

    are solved using a power series solution. Both Donnell–Mushtari and Flügge equations

    of motion are used and the limitations associated with each thin shell theory are

    discussed. Natural frequencies are presented for different boundary conditions. The

    effect of the boundary conditions and the influence of the semi-vertex cone angle are

    described. The results from the theoretical model presented here are compared with

    those obtained by previous researchers and from a finite element model.

    &  2009 Elsevier Ltd. All rights reserved.

    1. Introduction

    Cylindrical shells are widely reported in literature. Many researchers such as Donnell–Mushtari, Timoshenko, Reissner,

    Flügge, to name a few, have developed thin shell theory arising from different simplifying assumptions based on Love’s

    postulates. A range of general solutions for the cylindrical shell displacements and results for different boundary conditions

    have been summarised by Leissa   [1]. Conical shells have not been as widely reported in literature as in the case of 

    cylindrical shells. This is due to the increased mathematical complexity associated with the effect of the variation of the

    radius along the length of the cone on the elastic waves. The approximate location of the natural frequencies for conical

    shells has been found using the Rayleigh–Ritz method by several authors [1–5]. A transfer matrix approach was used by Irie

    et al.   [6]  to solve the free vibration of conical shells. Tong  [7]  presented a procedure for the free vibration analysis of 

    isotropic and orthotropic conical shells in the form of a power series. Guo   [8]  studied the propagation and radiation

    properties of elastic waves in conical shells.

    Very little work can be found on the vibrations of coupled cylindrical–conical shells, of which common applications are

    submarine hulls, aircraft, missiles and autonomous underwater vehicles (AUVs). Early analytical and experimental work to

    determine the natural frequencies and mode shapes of coupled conical–cylindrical shells used the finite element method

    [9]. The classic bending theory was used by Kalnins  [10]  and Rose et al.   [11]   to examine rotationally symmetric shells.

    Hu and Raney [12] examined the effects of discontinuities at the joint connecting the cone and cylinder. A transfer matrix

    approach was used by Irie et al.  [13] to solve the free vibration of coupled cylindrical–conical shells. Efraim and Eisenberger

    [14] applied a power series solution to calculate the natural frequencies of segmented axisymmetric shells. Patel et al. [15]

    presented results for laminated composite joined conical–cylindrical shells using a finite element method.

    Contents lists available at ScienceDirect

    journal homepage:   www.elsevier.com/locate/jsvi

     Journal of Sound and Vibration

    ARTICLE IN PRESS

    0022-460X/$- see front matter  &  2009 Elsevier Ltd. All rights reserved.doi:10.1016/j.jsv.2009.10.003

    Corresponding author.

    E-mail address:  [email protected] (M. Caresta).

     Journal of Sound and Vibration 329 (2010) 733–751

    http://-/?-http://www.elsevier.com/locate/jsvihttp://localhost/var/www/apps/conversion/tmp/scratch_2/dx.doi.org/10.1016/j.jsv.2009.10.003mailto:[email protected]:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_2/dx.doi.org/10.1016/j.jsv.2009.10.003http://www.elsevier.com/locate/jsvihttp://-/?-

  • 8/20/2019 Cylinder Cone Metal

    2/19

    In this work, the authors present a different approach to describe the free vibrational characteristics of coupled isotropic

    cylindrical–conical shells. The equations of motion for the cylindrical and conical shells are solved using two different

    methods. The cylindrical shell equations are solved using a wave solution while the conical shell equations are solved using

    a power series solution. These two methods are merged together for the first time to describe the dynamic response of the

    coupled shells. Two thin shell theories are used corresponding to Donnell–Mushtari and the higher order equations of 

    Flügge. In the latter case it is shown that an approximation is required in order to apply the power series solution for the

    conical shell. Results in terms of natural frequencies and mode shapes are compared with data available in literature. The

    effect of the junction between the coupled shells and the boundary conditions is also investigated.

    2. Equations of motion for thin shells

    The equations of motion to describe the vibrations of cylindrical or conical shells can be derived according to a

    particular thin shell theory using the standard derivation [1]. For completeness of the present study, the derivation of thin

    shell theory is briefly reviewed in Appendix A.

     2.1. Equations of motion for a cylindrical shell

    Using a cylindrical coordinate system ( x, y), u, v  and  w  are the orthogonal components of the shell displacement in the

    axial, circumferential and radial directions, respectively, as shown in Fig. 1. According to Flügge theory, the equations of 

    motion for a thin cylindrical shell are

    @2u

    @ x2 þ ð1 uÞ

    2a2  ð1 þ b2Þ @

    2u

    @y2 þ ð1 þ uÞ

    2a

    @2v

    @ x @yþ u

    a

    @w

    @ x  b2a @

    3w

    @ x3 þ b2 ð1 uÞ

    2a

    @3w

    @ x @y2  1

    c 2L

    @2u

    @t 2 ¼  0 (1)

    ð1 þ uÞ2a

    @2u

    @ x @yþ ð1 uÞ

    2

    @2v

    @ x2 þ   1

    a2@2v

    @y2 þ   1

    a2@w

    @y þ b2   3ð1 uÞ

    2

    @2v

    @ x2  ð3 uÞ

    2

    @3w

    @ x2 @y

    !   1

    c 2L

    @2v

    @t 2 ¼  0 (2)

    b2

    a2 @4w

    @ x4 þ 2   @

    4w

    @ x2 @y2 þ   1

    a2@4w

    @y4  a @

    3u

    @ x3 þ 1 u

    2a

    @3u

    @ x @y2  ð3 uÞ

    2

    @3v

    @ x2 @yþ   2

    a2@2w

    @y2

    !

    þ ua

    @u

    @ x þ  1

    a2@v

    @yþ wð1 þ b2Þ

    þ   1

    c 2L

    @2w

    @t 2 ¼ 0 (3)

    where a  is the radius of the middle surface of the shell,  b ¼ h= ffiffiffiffiffiffi12p    a is the thickness parameter,  h  is the shell thickness andf ¼ @w=@ x is the slope. c L ¼ ½E =rð1 u2Þ1=2 is the longitudinal wave speed. For Donnell–Mushtari theory, the equations of motion simplify to

    @2u

    @ x2 þ ð1 uÞ

    2a2@2u

    @y2 þ ð1 þ uÞ

    2a

    @2v

    @ x @yþ u

    a

    @w

    @ x    1

    c 2L

    @2u

    @t 2 ¼  0 (4)

    ð1 þ uÞ2a

    @2u

    @ x @yþ ð1 uÞ

    2

    @2v

    @ x2 þ   1

    a2@2v

    @y2 þ   1

    a2@w

    @y   1

    c 2L

    @2v

    @t 2 ¼  0 (5)

    b2

    a2 @4w

    @ x4 þ 2   @

    4w

    @ x2 @y2 þ   1

    a2@4w

    @y4

    !þ u

    a

    @u

    @ x þ  1

    a2@v

    @yþ  w

    a2 þ 1

    c 2L

    @2w

    @t 2 ¼ 0 (6)

    ARTICLE IN PRESS

    u

    θ 

    w

    φ 

     x 

     v

    Fig. 1.  Coordinate system for a thin walled cylindrical shell.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751734

  • 8/20/2019 Cylinder Cone Metal

    3/19

     2.2. Equations of motion for a conical shell

    For a conical shell, the coordinate system ( xc , yc ) is defined in Fig. 2. The equations of motion are given in terms of  uc  and

    vc  that are the orthogonal components of the displacement in the  xc  and yc  directions, respectively. wc  is the displacementnormal to the shell surface.  a   is the semi-vertex angle of the cone.   s   is the coordinate used in the standard derivationpresented in Appendix A. R  is the radius of the cone at location  xc . R0  is the mean radius of the shell and corresponds to the

    origin of the coordinate system.  Lc  is the length of the cone along its generator, and R1, R2  are, respectively, the radii at the

    smaller and larger ends of the cone. According to Fl ügge theory, the equations of motion are

    ðL11 þ   ~L11Þuc  þ L12vc  þ ðL13 þ   ~L13Þwc     1c 2cL

    @2uc @t 2

     ¼ 0 (7)

    L21uc  þ ðL22 þ   ~L22Þvc  þ ðL23 þ   ~L23Þwc     1c 2cL

    @2vc @t 2

     ¼ 0 (8)

    ðL31

     þ  ~L31

    Þuc 

     þ ðL32

     þ  ~L32

    Þvc 

     þ ðL33

     þ  ~L33

    Þwc 

     

      1

    c 2

    cL

    @2wc 

    @t 2

      ¼ 0 (9)

    c cL ¼ ½E c =rc ð1 u2c Þ1=2 is the longitudinal wave speed. E c , rc  and  uc  are, respectively, Young’s modulus, density and Poisson’sratio. Using Donnell–Mushtari theory, the differential operators   ~L ij  are zero. The Flügge equations of motion are more

    complicated due to the numerous higher order terms. The differential operators  L ij  and   ~L ij  are given in Appendix B.

    3. Solutions to the equations of motion

    The equations of motion for the cylindrical and conical shells are solved using two different methods, and are then

    merged together to provide the complete response of the coupled conical–cylindrical structure. The cylindrical shell

    equations are solved using a wave solution whilst the conical shell equations are solved using a power series method.

    Furthermore, expressions for the conical shell displacements are obtained for both the Donnell–Mushtari and Flügge

    theories.

     3.1. General solutions for the cylindrical shell

    General solutions to the equations of motion for a cylindrical shell can be assumed as  [1]

    uð x; y; t Þ ¼ U e jkn x cosðnyÞe jot  (10)

    vð x;y; t Þ ¼ V e jkn x sinðnyÞe jot  (11)

    wð x; y; t Þ ¼ W e jkn x cosðnyÞe jot  (12)where kn  is the axial wavenumber and n  is the circumferential mode number. Substituting the general solutions given by

    Eqs. (10)–(12) into the Flügge equations of motion given by Eqs. (1)–(3) results in three linear equations in terms of  U , V  and

    W . These linear equations can be arranged in matrix form as  AU ¼ 0, where  U ¼ ½U V W T contains the unknown waveamplitudes. The elements of the matrix A are given in Appendix C. For a non-trivial solution, the determinant of the matrix A  must be zero. The expanded determinant results in an eighth order characteristic equation in   kn. For each value of 

    ARTICLE IN PRESS

     Lc

     R2

    uc, x c

    wc

     R1 R0 R

    vc

    s

    c

    Fig. 2.  Coordinate system for a thin walled conical shell.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   735

  • 8/20/2019 Cylinder Cone Metal

    4/19

    kn;i ði ¼ 1  :  8Þ, the axial and circumferential amplitude ratios can be obtained as   C n;i ¼ U n;i=W n;i   and   Gn;i ¼  V n;i=W n;i,respectively. For harmonic motion, the complete solutions are given by

    uð x; y; t Þ ¼X1n¼0

    X8i¼1

    C n;iW n;ie jkn;i x cosðnyÞe jot  (13)

    v

    ð x; y; t 

    Þ ¼X1

    n¼0X8

    i¼1Gn;iW n;ie

     jkn;i x sin

    ðny

    Þe jot  (14)

    wð x; y; t Þ ¼X1n¼0

    X8i¼1

    W n;ie jkn;i x cosðnyÞe jot  (15)

     3.2. General solutions for the conical shell

    The equations of motion for the conical shell are solved using the power series approach presented by Tong  [7] for

    shallow shell theory. This approach is applied here to the equations of motion for both the Donnell–Mushtari and the

    Flügge thin shell theories. General solutions to the equations of motion for a conical shell given by Eqs. (7)–(9) can be

    expressed as

    uc 

    ð xc ;yc ; t 

    Þ ¼ uc 

    ð xc 

    Þcos

    ðnyc 

    Þe jot  (16)

    vc ð xc ; yc ; t Þ ¼ vc ð xc Þ sinðnyc Þ e jot  (17)

    wc ð xc ;yc ; t Þ ¼ wc ð xc Þcosðnyc Þe jot  (18)where the xc -dependent component of the displacement can be expressed in terms of a power series by

    uc ð xc Þ ¼X1m¼0

    am xmc    (19)

    vc ð xc Þ ¼X1m¼0

    bm xmc    (20)

    wc ð xc Þ ¼ X

    1

    m¼0 c m xm

    c    (21)

    Solutions for the conical shell displacements for the two thin shell theories are presented in what follows.

     3.2.1. Donnell–Mushtari equations

    Using the low order Donnell–Mushtari theory, the equations of motion given by Eqs. (7) and (8) (and where the

    differential operators   ~L ij  are zero) are multiplied by  R2 while Eq. (9) is multiplied by  R4 [7], resulting in

    R2L11uc  þ R2L12vc  þ R2L13wc   R2

    c 2cL

    @2uc @t 2

     ¼ 0 (22)

    R2L21uc  þ R2L22vc  þ R2L23wc   R2

    c 2cL

    @2vc @t 2

     ¼ 0 (23)

    R4L31uc  þ R4L32vc  þ R4L33wc   R4

    c 2cL

    @2wc @t 2

      ¼ 0 (24)

    Substituting Eqs. (16)–(21) into Eqs. (22)–(24) results in the following recurrence relations for  m ¼ 0; 1; 2; . . .:

    amþ2 ¼X4i¼1

     Aa;iam3þi þX2i¼1

    Ba;ibm1þi þX2i¼1

    C a;ic m1þi   (25)

    bmþ2 ¼X2i¼1

     Ab;iam1þi þX4i¼1

    Bb;ibm3þi þ C b;1c m   (26)

    c mþ4 ¼

    X4

    1

     Ac ;iam3þi þ

    X3

    1

    Bc ;ibm3þi þ

    X8

    1

    C c ;ic m5þi   (27)

    The coefficients in Eqs. (25)–(27) are given in Appendix D.

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751736

  • 8/20/2019 Cylinder Cone Metal

    5/19

     3.2.2. Flügge equations

    Using the Flügge equations, the application of the power series solution is more complicated compared with the

    Donnell–Mushtari theory due to the higher order terms. The coefficients of the operators   Lij   and   ~L ij fi; j ¼ 1  :  3g   includeterms of the form 1=Rk, k=1:4. Hence, the equations of motion given by Eqs. (17)–(19) are multiplied by R4 in order to apply

    the power series solution. Furthermore, the terms with h2 in the membrane force N s given by Eq. (A.6), given in Appendix A,

    are neglected, as in the Donnell–Mushtari theory. The approximation of  N s  results in a new   ~L13  term, given by

    ~L13 ¼ h2c 12

    sin2

    a cos3 aR4

      h2c  sin2

    a cosa12R3

      þ   @@ xc 

    1 uc 2R

    h2c 12

    @3

    @ xc  @y2c 

    3 u2

    h2c 12

    sina cosaR4

    @2

    @y2c 

    (28)

    Substituting Eqs. (16)–(21) into Eqs. (7)–(9) (multiplied by   R4) results in the following recurrence relations for

    m ¼ 0; 1; 2; . . .:

    amþ2 ¼X6i¼1

    ~ Aa;iam5þi þX4i¼1

    ~Ba;ibm3þi þX4i¼1

    ~C a;ic m3þi   (29)

    bmþ2 ¼X4i¼1

    ~ Ab;iam3þi þX6i¼1

    ~Bb;ibm5þi þX5i¼1

    ~C b;ic m3þi   (30)

    c mþ4 ¼ X6i¼1

    ~ Ac ;iam3þi þX5i¼1

    ~Bc ;ibm3þi þX8i¼1

    ~C c ;ic m5þi   (31)

    The recurrence coefficients in Eqs. (29)–(31) are given in Appendix E. It is important to note that if the h2 term in Eq. (A.6)

    is not neglected, the power series method cannot be applied since the recurrence relation given by Eq. (29) would

    become

    amþ2 ¼X6i¼1

    ~ Aa;iam5þi þX4i¼1

    ~Ba;ibm3þi þX6i¼1

    ~C a;ic m3þi   (32)

    The new terms c mþ2  and  c mþ3  are not compatible with the term  amþ2  on the left side of the equation. It can be concludedthat the use of the power series solution with the Flügge equations is only possible if the approximated membrane force N s,

    as in the Donnell–Mushtari theory, is used.

     3.2.3. Conical shell displacements

    Tong   [7]   showed that the   xc -dependent part of the displacements can be expressed in terms of eight unknown

    coefficients a0; a1; b0; b1; c 0; c 1; c 2; c 3;  which can be determined from the boundary conditions at both ends on the conical

    shell. In terms of the unknown coefficients, Eqs. (19)–(21) can be written as follows:

    uc ð xc Þ ¼ u  x ;   vc ð xc Þ ¼ v  x ;   wc ð xc Þ ¼ w  x    (33)

    where

    u ¼ ½u1ð xc Þ   u8ð xc Þ   (34)

     v ¼ ½v1ð xc Þ   v8ð xc Þ   (35)

     w ¼ ½w1ð xc Þ   w8ð xc Þ   (36)

     x  ¼ ½a0   a1   b0   b1   c 0   c 1   c 2   c 3T (37) x  is the vector of the eight unknown coefficients. In Eqs. (34)–(36),  uið xc Þ,  við xc Þ and  w ið xc Þ are the base functions of  uc ð xc Þ,vc ð xc Þ and wc ð xc Þ, respectively. The convergence property of the series solutions  uc ð xc Þ, vc ð xc Þ, wc ð xc Þ given by Eqs. (19)–(21)has been previously discussed by Tong [7]  and are maintained for the thin-shell theories presented here.

    4. Boundary and continuity conditions

    The two different methods corresponding to the wave solution and power series method both require the application of 

    four boundary conditions at each end of the shell to determine the unknown coefficients. Thus, the cylindrical and conical

    shells can be coupled together by applying the required continuity and equilibrium conditions at the interface.

    The remaining boundary conditions are applied at the ends of the coupled cylindrical–conical shell. The forces, momentsand displacements at the junction and at the boundaries of the coupled shells are given in accordance with the sign

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   737

  • 8/20/2019 Cylinder Cone Metal

    6/19

    convention shown in Fig. 3. At the cylinder–cone junction, continuity of displacements, slope, forces and bending moment

    are given by

    u ¼ U c    (38)

    w ¼ W c    (39)

    v ¼ V c    (40)

    @w

    @ x ¼  @wc 

    @ xc (41)

    ~N  x;c   N  x ¼ 0 (42)

    N  xy;c  þM  xy;c 

    R2

      N  xy þ

    M  xya

    ¼ 0 (43)

    M  x;c   M  x ¼ 0 (44)

    ~V  x;c   V  x ¼ 0 (45)To take into account the change of curvature between the cylinder and the cone, the following notation was introduced:

    U c  ¼  uc  cosa wc  sina   (46)

    W c  ¼  uc  sinaþ wc  cosa   (47)

    ~N  x;c  ¼  N  x;c  cosa V  x;c  sina   (48)

    ~V  x;c  ¼  V  x;c  cosaþ N  x;c  sina   (49)The membrane forces N  x,  N y  and  N  xy, bending moments  M  x,  M y  and M  xy, transverse shearing Q  x  and the Kelvin–Kirchhoff 

    shear force V  x can be derived for both shells. Different boundary conditions can be applied to the extremities of the coupled

    shells. In this work, three different boundary conditions have been considered, corresponding to free, clamped and shear

    diaphragm. The various boundary conditions, for example for the cylindrical shell, are

    Free end  :  N  x ¼   N  xy þM  xy

    a

    ¼ M  x ¼  V  x ¼ 0 (50)

    Clamped end  :  u ¼ w ¼ @w@ x

     ¼  v ¼ 0 (51)

    Shear-diaphragm ðSDÞ end  :  N  x ¼  v ¼ M  x ¼ w ¼ 0 (52)

    ARTICLE IN PRESS

    vc

    v

    wc

    w

    u

    uc

     N  x  ,c

     N  x  ,c

    Q x,c

     N  x,c

     N  x,c

    Q x,c

     M  x,c

     M  x,c

     M  x 

    Q x 

    Q x 

     N  x 

     N  x 

     N  x,

     N  x,

     M  x 

    Fig. 3.  Positive directions for the forces, moments and displacements of the cone and cylinder.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751738

  • 8/20/2019 Cylinder Cone Metal

    7/19

    ARTICLE IN PRESS

     L

     R1

    a

    Fig. 4.  Coupled cylindrical–conical shell.

     Table 1

    Frequency parameters for the free–clamped cylindrical–conical shell.

    Mode order   Frequency parameter Oc 

    n   Z   Irie et al. [13]   Efraim et al. [14]   Pre se nt ( Donnell –M ushta ri ) Pre sent ( Flügge)

    0 1 0.5047 0.503779 0.503752 0.505354

    T    – 0.609852 0.609855 0.609816

    2 0.9312 0.930942 0.930916 0.930904

    3 0.9566 0.956379 0.956315 0.956292

    4 0.9718 0.971634 0.971596 0.971538

    5 1.0122 1.012090 1.011884 1.011873

    1 1 0.2930 0.292875 0.292908 0.2933572 0.6368 0.635834 0.635819 0.636844

    3 0.8116 0.811454 0.811446 0.811434

    4 0.9316 0.931565 0.931481 0.931458

    5 0.9528 0.952178 0.952189 0.952120

    6 0.9922 0.992175 0.991959 0.991936

    2 1 0.1010 0.099968 0.102034 0.100087

    2 0.5032 0.502701 0.502899 0.502819

    3 0.6916 0.691305 0.691479 0.691353

    4 0.8592 0.859114 0.859017 0.858971

    5 0.9164 0.915870 0.916072 0.915877

    6 0.9608 0.960702 0.960475 0.960429

    3 1 0.09076 0.087603 0.093771 0.087330

    2 0.3921 0.391569 0.392199 0.391450

    3 0.5148 0.514478 0.515184 0.514424

    4 0.7537 0.753402 0.753595 0.7532955 0.7970 0.796590 0.796983 0.796557

    6 0.9197 0.919635 0.919391 0.919369

    4 1 0.1477 0.144619 0.150574 0.144478

    2 0.3312 0.330354 0.331698 0.330177

    3 0.3965 0.395649 0.397604 0.395495

    4 0.6473 0.646678 0.647700 0.646548

    5 0.6932 0.692805 0.693197 0.692690

    6 0.8720 0.871812 0.871555 0.871532

    5 1 0.2021 0.199546 0.203896 0.199540

    2 0.2966 0.296020 0.296330 0.295939

    3 0.3730 0.370901 0.376227 0.370707

    4 0.5805 0.579750 0.581667 0.579581

    5 0.6138 0.613363 0.614222 0.613231

    6 0.8187 0.817951 0.819801 0.818014

    ‘T ’ denotes the purely torsional frequency.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   739

  • 8/20/2019 Cylinder Cone Metal

    8/19

    The four boundary conditions at each end of the coupled cylindrical–conical shell together with the eight continuity

    equations at the junction can be arranged in matrix form  BX  ¼ 0, where  X  is the vector of the 16 unknown coefficientsgiven by

    X  ¼ ½a0   a1   b0   b1   c 0   c 1   c 2   c 3   W n;1     W n;8T (53)

    The vanishing of the determinant of matrix  B  gives the undamped natural frequencies of the joined shells.

    ARTICLE IN PRESS

    n = 0; Ωc = 0.6672

    n =1; Ωc = 0.4779

    n =2;Ωc = 0.3466

    n =3;Ωc = 0.2587

    n = 4; Ωc = 0.211

    n = 5; Ωc = 0.2097

    Fig. 5.  Lowest order mode shapes corresponding to  n=0:5 SD–SD case.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751740

  • 8/20/2019 Cylinder Cone Metal

    9/19

    5. Results

    5.1. Natural frequencies

    To confirm the validity of the presented method, results available in Refs. [13,14] are reproduced here. A coupled

    cylindrical–conical shell shown in Fig. 4 with free boundary conditions at the cone end and a clamped boundary for the

    ARTICLE IN PRESS

    n = 0; Ωc = 0.8371

    n = 1; Ωc = 0.5267

    n = 3; Ωc = 0.2875

    n = 4; Ωc = 0.2362

    n = 5; Ωc = 0.2252

    n = 2; Ωc = 0.3771

    Fig. 6.  Lowest order mode shapes corresponding to  n=0:5 clamped–clamped case.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   741

  • 8/20/2019 Cylinder Cone Metal

    10/19

    cylinder is examined, with the following data:   L=a ¼ 1,   h=a ¼ 0:01,   R1=a ¼ 0:4226,  a ¼ 303. The shells are of the samematerial with Young’s modulus E  ¼ 2:11 1011 N m2, Poisson’s ratio  u ¼ 0:3 and density r ¼ 7800kg m3.

    The dimensionless frequency parameter  Oc  ¼ oa=c L   for the lowest six values of the circumferential mode numberðn ¼ 0; . . . ; 5Þ are given in Table 1. The values of the frequency parameters agree well with those presented previously by Irieet al.   [13]  and Efraim et al.   [14]. A very small difference is observed between the two shell theories except at lower

    frequencies, where the Donnell–Mushtari theory is not as accurate as the Flügge theory [1]. When  n=0, the equation of 

    motion for the circumferential displacement is uncoupled from the equations of motion for the axial and radial

    displacements for both the conical and cylindrical shells, yielding a purely torsional mode   [1]. The frequency value of the mode with order ½n   Z ¼ ½0   T  corresponds to the first purely torsional mode. This frequency is omitted in the work of Irie et al. [13] since they did not consider the torsional solution. The purely torsional frequency is reported in Efraim et al.

    ARTICLE IN PRESS

    n = 0; Ωc = 0.6618

    n = 1; Ωc = 0.7200

    n = 2; Ωc = 0.01001

    n = 3; Ωc = 0.02566

    n = 4; Ωc = 0.04649

    n = 5; Ωc = 0.07271

    Fig. 7.  Lowest order mode shapes corresponding to  n=0:5 free–free case.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751742

    http://-/?-http://-/?-http://-/?-

  • 8/20/2019 Cylinder Cone Metal

    11/19

    [14], but their corresponding mode shape appears to be in contrast with a purely torsional solution. To further validate the

    analytical method presented in this work, a computational finite element model (FEM) was developed using Patran/

    Nastran. Quadratic eight node (CQUAD8) elements were used for the 2D thin shell elements; the cylinder and the cone

    were meshed with 20 elements in the axial direction and 30 in the circumferential direction. The Lanczos extraction

    method was adopted in the analysis   [16]. The lowest order mode shapes corresponding to   n=0:5 for SD–SD,

    clamped–clamped and free–free boundary conditions have been normalised and are, respectively, shown in Figs. 5–7.

    The analytical results represented by the continuous line are practically indistinguishable to those obtained from the FE

    model (represented by dots). Screenshots of the mode shapes from the FE model are also shown. The mode shapes for theSD–SD and clamped–clamped cases are similar with a large deformation at the cylinder/cone junction for   n=0. As the

    circumferential mode number increases, a larger deformation of the cone with respect to the cylinder can be observed. For

    the free–free case and n=0:1, the mode shapes show similar characteristics to the other boundary conditions while for nZ2,

    a larger deformation for the cylindrical shell is observed compared to the displacement of the conical section.

    5.2. Effect of the boundary conditions

    The effect of boundary conditions on the free vibrational characteristics of a coupled conical–cylindrical shell is

    examined. In Fig. 8, the lowest frequency parameter is plotted versus the circumferential mode number  n. The frequencies

    calculated using both the Donnell–Mushtari and Flügge equations are compared with the results given by the FE model. A

    logarithmic scale was used to emphasize the small differences in the results. It can be seen that the results given by the

    Flügge equations of motion match almost perfectly with the FE results. The results given by the Donnell–Mushtari

    equations are affected by several issues for coupled cylindrical–conical shells. Firstly, it can be observed that they performless well at low frequencies compared with the results given by both the Fl ügge equations of motion and the finite element

    model. Furthermore, they are in error for the free–free case. For this boundary condition, the equations for  n=1 give two

    incorrect frequencies of very low values that should be the zeroes associated with rigid body rotation. This is due to the

    inconsistency of Donnell–Mushtari theory with free body motion, as reported by Kadi  [17] and Kraus [18].

    For clamped–clamped and SD–SD boundary conditions, the lowest frequency parameter decreases with   n. For

    free–clamped boundary conditions, the frequencies initially decrease and then increase after  n=3. For the free–free case, a

    very low frequency occurs at  n=2.

    5.3. Effect of the semi-vertex cone angle

    Figs. 9–11 present the effect of the semi-vertex angle  a  of the conical shell on the frequency parameter Oc , for different

    boundary conditions of the coupled shell. The following data for the coupled shells were used:  L=a ¼ 1,  h=a ¼ 0:01, Lc  ¼ 1,a 2 ½0; 903. For extreme values of the semi-vertex angle corresponding to  a=01 and 901, the conical shell degenerates to acylindrical shell and a circular plate, respectively. For the  n=0 mode, the behaviour of the coupled shell is similar for all

    ARTICLE IN PRESS

    0 1 2 3 4 5

    100

    Circumferential mode number

       L  o  w  e  s   t   f  r  e  q  u  e  n  c  y

      p  a  r  a  m  e   t  e  r     Ω

          c

    FEM

    D.Mushtari

    Flügge

    Free−Free

    Free−Clamped

    SD−SD

    Clamped−Clamped

    Wrong results from the D.−Mushtari

    equations (Free−Free)

    10−1

    10−2

    10−3

    Fig. 8.  Lowest frequency parameter Oc  for different boundary conditions calculated using the Donnell–Mushtari and Flügge equations and compared withthe results from an FE model.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   743

  • 8/20/2019 Cylinder Cone Metal

    12/19

    boundary conditions considered, resulting in a relatively constant value in the frequency parameter for increasing values of 

    a and then a mainly linear decrease in Oc . The corresponding motion is primarily axial. As the conical shell changes from acylindrical shell (at a=0) to a plate-like structure (at a=901), a decrease in axial stiffness occurs resulting in a decrease in thefrequency parameter. A similar behaviour to the n=0 mode is observed for the  n=1 bending mode for a free–free coupled

    shell. For nZ2, the frequency parameter is almost constant for all values of  a for the free–free shell. For a coupled shell witha free boundary at the conical shell end and clamped at the cylindrical shell end, for  n=1 a slight increase of the frequency

    parameter with   a   is observed, showing a small mass effect. A stiffening effect then dominates after   a=651. In thefree–clamped shell, higher order circumferential modes result in a wavelike behaviour due to greater shape complexity.

    ARTICLE IN PRESS

    0 10 20 30 40 50 60 70 80 90

    0

    0.2

    0.4

    0.6

    0.8

    1

       F  r  e  q  u  e  n  c  y  p  a  r  a  m  e   t  e  r     Ω

          c

    n = 0 

    n = 1

    n = 2n = 3

    n = 4

    n = 5

    Semi−vertex angle α

    Fig. 10.  Lowest frequency parameter  Oc   for free–clamped boundary conditions.

    0 10 20 30 40 50 60 70 80 90

    0

    0.2

    0.4

    0.6

    0.8

    1

       F  r  e  q  u  e  n

      c  y  p  a  r  a  m  e   t  e  r     Ω

          c

    n = 0

    n = 1

    n = 2

    n = 3

    n = 4

    n = 5

    Semi−vertex angle α

    Fig. 9.  Lowest frequency parameter  Oc   for free–free boundary conditions.

    0 10 20 30 40 50 60 70 80 900

    0.2

    0.4

    0.6

    0.8

    1

       F  r  e  q  u  e  n  c  y  p  a  r  a  m  e   t  e  r     Ω

          c

    n = 0n = 1n = 2n = 3n = 4n = 5

    Semi−vertex angle α

    Fig. 11.  Lowest frequency parameter  Oc  for clamped–clamped boundary conditions.

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751744

  • 8/20/2019 Cylinder Cone Metal

    13/19

    6. Conclusions

    A different approach to obtain the free vibrational characteristics of coupled cylindrical–coupled shells has been

    introduced. Two different methods corresponding to a wave solution and power series method were used to obtain the

    shell displacements. The shells were then coupled by means of continuity conditions at the cone/cylinder junction. Results

    in terms of natural frequencies were compared for two different thin shell theories, corresponding to Donnell–Mushtari

    and Flügge equations, as well as with data presented previously in Refs. [13,14]. It was shown that in order to use the Flügge

    equations of motion with the power series solution, an approximation of the shear force is required. The effect of fourclassical boundary conditions at the ends of the coupled shells on the natural frequencies was investigated.

    In general, little difference was observed between the results given by the two shell theories. The Flügge equations were

    shown to be in very close agreement with results from a finite element model, but the Donnell–Mushtari equations were

    less accurate at low frequencies. Furthermore, for free–free boundary conditions of the coupled shells, the Donnell–Mushtari

    equations generate errors in the values of the lowest frequency parameter for circumferential mode number  n=1.

    The method described in this work can also be applied to the coupled shells of different materials and thickness.

     Appendix A. Equations of motion for thin isotropic shells

    According to Flügge theory, the equations of motion for a thin shell are given by

    @ðBN sÞ

    @s

      þ

    @ð AN ysÞ

    @y

      þ

    @ A

    @y

    N sy

     

    @B

    @s

     N y

     þ

     AB

    Rs

    Q s

      ABrh

    @2u

    @t 2

     ¼ 0 (A.1)

    @ð AN yÞ@y

      þ @ðBN syÞ@s

      þ @B@s

     N ys @ A

    @yN s þ AB

    RyQ y  ABrh

    @2v

    @t 2 ¼  0 (A.2)

     ABRs

    N s  AB

    RyN y þ

    @ðBQ sÞ@s

      þ @ð AQ yÞ@y

       ABrh @2w

    @t 2 ¼ 0 (A.3)

    where u,  v  and  w, respectively, denote the orthogonal component of the displacement.  r  is the density and  h  is the shellthickness. The equations are given in terms of two independent coordinates  s  and y. The parameters A, B, Rs  and  Ry  depend

    on the type of shell. The forces and moments in Eqs. (A.1)–(A.3) are given by  [1]

    Q s ¼   1 AB

    @ðBM sÞ@s

      þ @ð AM ysÞ@y

      þ @ A@y

    M sy @B

    @s M y   (A.4)

    Q y ¼   1 AB @ð AM yÞ@y   þ @ðBM syÞ@s   þ @B@s M ys  @ A@yM s   (A.5)

    N s ¼  Eh

    1 u2   es þ uey  h2

    12

    1

    Rs   1

    Ry

      ks 

     esRs

      (A.6)

    N y ¼  Eh

    1 u2   ey þ ues  h2

    12

    1

    Ry   1

    Rs

      ky 

     eyRy

      (A.7)

    N sy ¼  Eh

    2ð1 þ uÞ   esy  h2

    12

    1

    Rs   1

    Ry

      t

    2 esy

    Rs

      (A.8)

    N ys ¼  Eh

    2

    ð1

    þu

    Þ  esy 

     h2

    12

    1

    Ry  1

    Rs   t

    2 esy

    Ry   (A.9)

    M s ¼  Eh3

    12ð1 u2Þ   ks þ uky   1

    Rs   1

    Ry

    es

      (A.10)

    M y ¼  Eh3

    12ð1 u2Þ   ky þ uks   1

    Ry  1

    Rs

    ey

      (A.11)

    M sy ¼  Eh3

    24ð1 þ uÞ   tesyRs

      (A.12)

    M ys ¼  Eh3

    24ð1 þ uÞ   tesyRy

      (A.13)

    V s ¼ Q s þ 1B

    @M  xy@y

      (A.14)

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   745

  • 8/20/2019 Cylinder Cone Metal

    14/19

    E  is Young’s modulus and  u  is Poisson’s ratio of the material. Eq. (A.14) is the Kelvin–Kirchhoff shearing force. The normal

    strain es, ey, and shear strain esy  of the middle surface and the rotations of the normal to the middle surface denoted by  Wsand Wy  are given by

    es ¼ 1 A

    @u

    @s þ   v

     AB

    @ A

    @y þ w

    Rs(A.15)

    ey ¼ 1

    B

    @v

    @y þ  u

     AB

    @B

    @s þ w

    Ry(A.16)

    esy ¼ A

    B

    @ðu= AÞ@y

      þ B A

    @ðv=BÞ@s

      (A.17)

    Ws ¼  u

    Rs 1

     A

    @w

    @s  (A.18)

    Wy ¼  v

    Ry 1

    B

    @w

    @y  (A.19)

    The mid-surface changes in curvature  ks,  ky  and twist t  are given by

    ks

     ¼

     1

     A

    @Ws

    @s þ

     Wy

     AB

    @ A

    @y

      (A.20)

    ky ¼ 1

    B

    @Wy@y

     þ  Ws AB

    @B

    @s  (A.21)

    t ¼ AB

    @ðWs= AÞ@y

      þ B A

    @ðWy=BÞ@s

      þ  1Rs

    1

    B

    @u

    @y   v

     AB

    @B

    @s

    þ   1

    Ry

    1

     A

    @v

    @s    u

     AB

    @ A

    @y

      (A.22)

    According to the Donnell–Mushtari theory, the terms including   Q s   and   Q y   in Eqs. (A.1) and (A.2) are neglected

    and Eqs. (A.6)–(A.13), respectively, simplify to   N s ¼ ½Eh=ð1 u2Þðes þ ueyÞ,   N y ¼ ½Eh=ð1 u2Þðey þ uesÞ,   N sy ¼ N ys ¼½Eh=2ð1 þ uÞesy,   M s ¼ ½Eh3=12ð1 u2Þðks þ ukyÞ,  M y ¼ ½Eh3=12ð1 u2Þðky þ uksÞ,   M sy ¼ M ys ¼ ½Eh3=24ð1 þ uÞt. Furthermore,the mid-surface changes in curvature  ks,  ky  and twist t  simplify to the following expressions:

    ks ¼ 1 A

    @

    @s

    1

     A

    @w

    @s   1

     AB2@ A

    @y

    @w

    @y  (A.23)

    ky ¼ 1

    B

    @

    @y

    1

    B

    @w

    @y

      1

    BA2@B

    @s

    @w

    @s  (A.24)

    t ¼ B A

    @

    @s

    1

    B2@w

    @y

     A

    B

    @

    @y

    1

     A2@w

    @s

      (A.25)

    For a cylindrical shell, the equations of motion for   u,   v  and   w   can be derived from Eqs. (A.1)–(A.3) using the following

    parameters; A ¼ a, B ¼ a, Rs ¼ 1, Ry ¼ a  and s ¼ x=a, where a  is the mean radius of the shell and u, v, w and x  are defined asin  Fig. 1. For a conical shell, the equations of motion for   uc ,   vc   and   wc   can be derived using   A ¼ 1,   B ¼ s sina,   Rs ¼ 1,Ry ¼ s tana, as well as using the change of coordinate given by  s ¼ R=sina  and  R ¼ R0 þ xc  sina. R ,  R0,  uc ,  vc ,  wc  and  xc  aredefined in Fig. 2.

     Appendix B. Differential operators for the conical shell

    For the conical shell, omitting the subindex c , the differential operators are given by

    L11 ¼ sin2 a

    R2  þ sina

    R

    @

    @ xþ   @

    2

    @ x2 þ 1 u

    2R2@2

    @y2

      (B.1)

    ~L11 ¼ 1 u

    2R2h2

    12

    cos2 a

    R2@2

    @y2  h

    2

    12

    cos2 a sin2 a

    R4  (B.2)

    L12 ¼ 1 þu

    2R

    @2

    @ x @y 3 u

    2

    sina

    R2@

    @y  (B.3)

    L13 ¼ sina cosaR

      þ u cosaR

    @@ x

      (B.4)

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751746

  • 8/20/2019 Cylinder Cone Metal

    15/19

    ~L13 ¼ h2

    12

    sin2 a cos3 a

    R4  h

    2 sin2 a cosa

    12R3   h

    2

    12

    1

    R

    @3

    @ x3 þ   @

    @ x

    1 u2R

    h2

    12

    @3

    @ x @y2  ð3 uÞ

    2

    h2

    12

    sina cosa

    R4@2

    @y2

      (B.5)

    L21 ¼ 3 u

    2

    sina

    R2@

    @yþ 1 þ u

    2R

    @2

    @ x @y  (B.6)

    L22 ¼

    1 u

    2

    sin2 a

    R2

      þ1 u

    2

    sina

    R

    @

    @ x þ  1

    R2

    @2v

    @y2

     þ1 u

    2

    @2

    @ x2

      (B.7)

    ~L22 ¼ h2

    12

    sin2 a cos2 a

    R43

    2ð1 uÞ  h

    2

    12

    sina cos2 a

    R3@

    @ xþ h

    2

    12

    cos2 a

    R23

    2ð1 uÞ  @

    2

    @ x2  (B.8)

    L23 ¼ cosa

    R2@

    @y  (B.9)

    ~L23 ¼ h2

    12

    sina cosa

    R4@

    @yþ h

    2

    12

    cosa

    R33

    2ð1 uÞ   @

    2

    @ x @y h

    2

    12

    cosa

    R23 u

    2

    @3

    @ x2 @y  (B.10)

    L31 ¼ sina cosa

    R2  u cosa

    R

    @

    @ x  (B.11)

    ~L31 ¼ h2

    122sin

    3

    a cosaR4

      þ h2

    12sin

    2

    a cosaR3

    @@ x

     h2

    12sina cos

    3

    aR4

      þ h2

    12cosaR

    @3

    @ x3  h

    2

    12sina cosa

    R41 þ u

    2@

    2

    @y2

     h2

    12

    cosa

    R31 u

    2

    @3

    @ x @y2

      (B.12)

    L32 ¼ cosa

    R2@v

    @y  (B.13)

    ~L32 ¼ h2

    12

    sina cosa

    R33 þ u

    2

    @2

    @ x @yþ h

    2

    12

    sin2 a cosa

    R43 þ u

    2

    @

    @yþ h

    2

    12

    cosa

    R23 u

    2

    @3

    @ x2 @y  (B.14)

    L33 ¼ cos2 a

    R2   h

    2

    12r 4 (B.15)

    ~L33 ¼ h2

    12

    cos2 a

    R4  2 þ cos2 aþ 2   @

    2

    @y2

    !  (B.16)

    r 4 ¼ r 2r 2;r 2 ¼   @2

    @ x2 þ sina

    R

    @

    @ xþ   1

    R2@2

    @y2

      (B.17)

     Appendix C. Elements of matrix A

    The elements of matrix  A   for the cylindrical shell are given by

     A11 ¼ O2 ðknaÞ2 ð1 uÞ

    2  n2ð1 þ b2Þ   (C.1)

     A12 ¼ jnknað1 þ uÞ=2 (C.2)

     A13 ¼ jukna þ jb2½ðknaÞ3 n2ðknaÞð1 uÞ=2   (C.3)

     A21 ¼  A12   (C.4)

     A22 ¼ O2 ðknaÞ2ð1 uÞ=2ð1 þ 3b2Þ n2 (C.5)

     A23 ¼ n b2nðknaÞ2ð3 uÞ=2 (C.6)

     A31 ¼  A13   (C.7)

     A32 ¼  A23   (C.8)

     A33 ¼ 1 O2 þ b2f½ðknaÞ2 þ n22 þ ð1 2n2Þg   (C.9)

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   747

  • 8/20/2019 Cylinder Cone Metal

    16/19

    O ¼ oa=c L is the dimensionless frequency parameter. Using Donnell–Mushtari theory, Eqs. (C.1), (C.3), (C.5), (C.6) and (C.9),respectively, reduce to   A11 ¼ O2 ðknaÞ2ð1 uÞn2=2,   A13 ¼ jukna,   A22 ¼ O2 ðknaÞ2ð1 uÞ=2 n2 ,   A23 ¼ n   and A33 ¼ 1 O2 þ b2½ðknaÞ2 þ n22.

     Appendix D. Recurrence terms for the Donnell–Mushtari equations

    The recurrence terms for the Donnell–Mushtari equations are given by

     Aa;1 ¼ rho2 sin2 a=Da   (D.1)

     Aa;2 ¼ 2rho2R0 sina=Da   (D.2)

     Aa;3 ¼ ½Gðm2 1Þsin2 aþ rho2R20  Ehn2=2ð1 þ uÞ=Da   (D.3)

     Aa;4 ¼ GR0 sinaðm þ 1Þð2m þ 1Þ=Da   (D.4)

    Ba;1 ¼ Gn sinaðum þ m 3 þ uÞ=2Da   (D.5)

    Ba;2 ¼ GR0nðm þ 1Þðu þ 1Þ=2Da   (D.6)

    C a;1 ¼ G sina cosaðum 1Þ   (D.7)

    C a;2 ¼ GR0 sina cosaðm þ 1Þ   (D.8)

     Ab;1 ¼ Gn sinað3 þ um þ m uÞ=2Db   (D.9)

     Ab;2 ¼  GR0nðm þ 1Þðu þ 1Þ=2Db   (D.10)

    Bb;1 ¼ rho2 sin2 a=Db   (D.11)

    Bb;2 ¼ 2rho2R0 sina=Db   (D.12)

    Bb;3 ¼ ½Gðum2 þ m2 1 þ uÞsin2 a Gn2 þ rho2R20=2Db   (D.13)

    Bb;4

     ¼ GR0 sina

    ðm

    þ1

    Þð2m

    þ1

    Þ=2Db   (D.14)

    C b;1 ¼  Ehn cosa=ð1 þ uÞDb   (D.15)

     Ac ;1 ¼ G cosa sin3 að1 þ um 2uÞ=Dc    (D.16)

     Ac ;2 ¼ GR0 cosa sin2 að2 þ 3um 3uÞ=Dc    (D.17)

     Ac ;3 ¼ GR20 cosa sinað1 þ 3umÞ=Dc    (D.18)

     Ac ;4 ¼ uGR30 cosaðm þ 1Þ=Dc    (D.19)

    Bc ;1 ¼ Gn cosa sin2 a=Dc    (D.20)

    Bc ;2 ¼ 2GR0n cosa sina=Dc    (D.21)Bc ;3 ¼ GR20n cosa=Dc    (D.22)

    C c ;1 ¼ rho2 sin4 a=Dc    (D.23)

    C c ;2 ¼ 4rho2R0 sin3 a=Dc    (D.24)

    C c ;3 ¼ sin2 aðG cos2 aþ 6rho2R20Þ=Dc    (D.25)

    C c ;4 ¼ 2R0 sinaðG cos2 aþ 2rho2R20Þ=Dc    (D.26)

    C c ;5 ¼ ½Dð4m2 þ 4m3 m4Þsin4 a GR20 cos2 aþ Dð2n2m2 þ 4n2 4n2mÞsin2 a Dn4 þ rho2R40=Dc    (D.27)

    C c ;6 ¼ DR0 sinað2sin2 am2 þ 2sin2 am þ sin2 aþ 2n2Þðm þ 1Þð1 þ 2mÞ=Dc    (D.28)

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751748

  • 8/20/2019 Cylinder Cone Metal

    17/19

    C c ;7 ¼ DR20ð6sin2 am2 þ sin2 aþ 2n2Þðm þ 2Þðm þ 1Þ=Dc    (D.29)

    C c ;8 ¼ 2DR30 sinað2m þ 1Þðm þ 3Þðm þ 2Þðm þ 1Þ=Dc    (D.30)where

    Da ¼  GR20ðm þ 2Þðm þ 1Þ   (D.31)

    Db ¼ GR2

    0ðm þ 2Þðm þ 1Þðu 1Þ=2 (D.32)Dc  ¼ DR40ðm þ 4Þðm þ 3Þðm þ 2Þðm þ 1Þ   (D.33)

    G ¼ Eh=ð1 u2Þ; D ¼ Eh3=12ð1 u2Þ   (D.34)

     Appendix E. Recurrence terms for the Flügge equations

    The recurrence terms for the Flügge equations are given bye Aa;1 ¼ rho2s4=eDa   (E.1)e

     Aa;2 ¼ 4rho2R0s3=eDa   (E.2)

    e Aa;3 ¼ fG½ðm 2Þ2 1s4 þ ðGn2u=2 þ 6rho2R20  Gn2=2Þs2g=eDa   (E.3)e Aa;4 ¼ sR0ð9Gs2m þ 3Gs2 þ Gn2u þ 4Gs2m2 Gn2 þ 4rho2R20Þ=eDa   (E.4)

    e Aa;5 ¼ ½Gð6R20m2 3mR20  R20  c 2h2=2Þs2 þ Dn2uc 2=2 þ rho2R40 þ Gn2uR20=2 Dn2c 2=2 Gn2R20=2=eDa   (E.5)e Aa;6 ¼ GR30sðm þ 1Þð1 þ 4mÞ=eDa   (E.6)eBa;1 ¼ Gs3nðum u þ m 5Þ=2eDa   (E.7)

    eBa;2 ¼  GR0s2nð3um u þ 3m 9Þ=2

    eDa   (E.8)

    eBa;3 ¼  GR20snð3um þ 3m 3 þ uÞ=2eDa   (E.9)eBa;4 ¼ GR30nðm þ 1Þðu þ 1Þ=2eDa   (E.10)eC a;1 ¼ Gcs3ð1 þ um 2uÞ=eDa   (E.11)

    eC a;2 ¼  GR0cs2ð2 þ 3um 3uÞ=eDa   (E.12)eC a;3 ¼ Gcsð2h2n2 þ 72umR20 þ h2n2 2c 2h2 þ h2n2um h2n2m 24R20  h2n2u 2h2s2mÞ=24eDa   (E.13)

    eC a;4 ¼ GR0c ðm þ 1Þð24R20u 2h2s2 þ h2n2u h2n2Þ=24eDa   (E.14)e Ab;1 ¼ Gs3nðum 3u þ m þ 1Þ=2eDb   (E.15)e Ab;2 ¼ GR0s2nð3um 5u þ 3m þ 3Þ=2eDb   (E.16)

    e Ab;3 ¼ snð36c 2D þ 36GumR20  Guh2c 2 þ 3Gh2c 2 þ 36GmR20 þ 36GR20  12GuR20 þ 12c 2DuÞ=24eDb   (E.17)e Ab;4 ¼ GR30ðm þ 1Þðu þ 1Þ=2eDb   (E.18)

    eBb;1 ¼ rho2s4=eDb   (E.19)eBb;2 ¼ 4rho2R0s3=eDb   (E.20)

    eBb;3 ¼ Gð1 þ u uðm 2Þ2 þ ðm 2Þ2Þs4=2 þ ð6rho2R20  Gn2Þs2=

    eDb   (E.21)

    eBb;4 ¼ R0sð8rho2R20  3s2G þ 4Gn2 þ 4Gus2m2 9Gus2m þ 3Gus2 4s2Gm2 þ 9s2GmÞ=2eDb   (E.22)

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   749

  • 8/20/2019 Cylinder Cone Metal

    18/19

    eBb;5 ¼ ½ð3GuR20m2 c 2Du=2 þ Dc 2 þ c 2D=2 GR20=2 þ GuR20=2 Duc 2 Gh2mc 2=8 þ umR203G=2 3=2GmR20þ 1=8Gh2umc 2 þ 3GR20mq þ 1=24c 2Gh2m2 1=24Gh2uc 2m2 3=2c 2Dm þ c 2Dm2 þ 3=2c 2Dum c 2Dum2Þs2 Gn2R20 þrho2R40=eDb   (E.23)

    eBb;6 ¼ R0sðm þ 1Þð24GR20m þ 24c 2Dm þ c 2Gh2m þ 6GR20  c 2Gh2 6c 2DÞðu 1Þ=12

    eDb   (E.24)

    eC b;1 ¼ Gcs2n=eDb   (E.25)eC b;2 ¼ 2GR0csn=eDb   (E.26)

    eC b;3 ¼ cnð24GR20 þ 24Dm2s2 þ 4Gh2ums2 þ 2n2Gh2 þ 3Gh2s2 24n2D Gh2um2s2 6Gh2ms2 þ 24c 2D 2c 2Gh2 3Gh2us2 þ Gh2m2s2Þ=24eDb   (E.27)

    eC b;4 ¼ R0csnðm þ 1Þð48Dm 2Gh2m þ 2Gh2um 24D 3Gh2u þ 5Gh2Þ=24eDb   (E.28)eC b;5 ¼ DR20cnðm þ 2Þðm þ 1Þð1 þ uÞ=2eDb   (E.29)

    e Ac ;1 ¼ Gcs3ð1 þ um 2uÞ=eDc    (E.30)e Ac ;2 ¼ GR0cs2ð2 þ 3um 3uÞ=eDc    (E.31)e Ac ;3 ¼ ½1=12cDð24 36m 12m3 þ 36m2Þs3

    1=12c ð12GR20  6Dn2u þ c 2Gh2 þ 6Dn2um þ 36GuR20m Dn2 6Dn2mÞs=eDc    (E.32)e Ac ;4 ¼ 1=2R0c ðm þ 1ÞðDn2 þ 6Dms2 2Ds2 6Dm2s2 þ Dn2u þ 2GuR20Þ=eDc    (E.33)

    e Ac ;5 ¼ 3DR20csðm þ 2Þðm þ 1Þm=eDc    (E.34)

    e Ac ;6 ¼  DR30c ðm þ 3Þðm þ 2Þðm þ 1Þ=

    eDc    (E.35)

    eBc ;1 ¼ Gcs2n=eDc    (E.36)eBc ;2 ¼ 2GR0csn=eDc    (E.37)

    eBc ;3 ¼ cnðDum2s2 þ 6Dms2 þ 2GR20  3Dm2s2 3Ds2 Ds2uÞ=2eDc    (E.38)eBc ;4 ¼ Dcnðm þ 1ÞsR0ð3 þ 2um þ u 6mÞ=2eDc    (E.39)

    eBc ;5 ¼ DR20cnðm 1Þmð3 þ uÞ=2eDc    (E.40)

    eC c ;1 ¼ rho2s4=

    eDc    (E.41)

    eC c ;2 ¼ 4rho2R0s3=eDc    (E.42)eC c ;3 ¼ s2ðc 2G 6rho2R20Þ=eDc    (E.43)eC c ;4 ¼ 2R0sðc 2G 2rho2R20Þ=eDc    (E.44)

    eC c ;5 ¼ ½Dð4m3 m4 4m2Þs4 þ Dðmc 2 c 2m 2c 2 4n2m þ 2n2m2 þ 4n2Þs2 þ rho2R40  c 2GR20 þ Dc 2n2þ Dn2c 2 c 4D Dn4=eDc    (E.45)

    eC c ;6 ¼ DR0sðm þ 1Þð12s2 48s2m3 þ 72s2m2 24n2 12c 2 þ 48n2m þ 12c 2Þ=12eDc    (E.46)eC c ;7 ¼ DR20ð6s2m2 þ s2 þ 2n2Þðm þ 2Þðm þ 1Þ=

    eDc    (E.47)

    eC c ;8 ¼ 2DR30sð2m þ 1Þðm þ 3Þðm þ 2Þðm þ 1Þ=eDc    (E.48)

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751750

  • 8/20/2019 Cylinder Cone Metal

    19/19

    where eDa ¼ GR40ðm þ 2Þðm þ 1Þ   (E.49)eDb ¼  R20ð12GR20 þ 24Dc 2 þ Gh2c 2Þðm þ 2Þðm þ 1Þðu 1Þ=24 (E.50)

    eDc  ¼ DR40ðm þ 4Þðm þ 3Þðm þ 2Þðm þ 1Þ   (E.51)

    c  ¼ cos a;   s ¼ sina   (E.52)

    References

    [1] A.W. Leissa, in:  Vibration of Shells, American Institute of Physics, New York, 1993.[2] H. Saunders, E.J. Paslay, P.R. Wisniewski, Vibrations of conical shells,  J. Acoust. Soc. Am.  32 (1960) 765–772.[3] H. Garnet, J. Kempner, Axisymmetric free vibration of conical shells, J. Appl. Mech.  31 (1964) 458–466.[4] R.A. Newton, Free vibrations of rocket nozzles,  Am. Inst. Aeronaut. Astronaut. J.  4 (1966) 1303–1305.[5] C.C. Siu, C.W. Bert, Free vibrational analysis of sandwich conical shells with free edges, J. Acoust. Soc. Am.  47 (1970) 943–945.[6] T. Irie, G. Yamada, Y. Kaneko, Free vibration of a conical shell with variable thickness, J. Sound Vib.  82 (1982) 83–94.[7] L. Tong, Free vibration of orthotropic conical shells, Int. J. Eng. Sci.  31 (1993) 719–733.[8] Y.P. Guo, Normal mode propagation on conical shells,  J. Acoust. Soc. Am.  96 (1994) 256–264.[9] M. Lashkari, V.I. Weingarten, Vibrations of segmented shells,  Exp. Mech. 13 (1973) 120–125.

    [10] A. Kalnins, Free vibration of rotationally symmetric shells,  J. Acoust. Soc. Am.  36 (1964) 1355–1365.[11] J.L. Rose, R.W. Mortimer, A. Blum, Elastic-wave propagation in a joined cylindrical–conical–cylindrical shell, Exp. Mech.  13 (1973) 150–156.

    [12] W.C.L. Hu, J.P. Raney, Experimental and analytical study of vibrations of joined shells, AIAA J.  5 (1967) 976–980.[13] T. Irie, G. Yamada, Y. Muramoto, Free vibration of joined conical–cylindrical shells, J. Sound Vib.  95 (1984) 31–39.[14] E. Efraim, M. Eisenberger, Exact vibration frequencies of segmented axisymmetric shells, Thin-Walled Struct.  44 (2006) 281–289.[15] B.P. Patel, M. Ganapathi, S. Kamat, Free vibration characteristics of laminated composite joined conical–cylindrical shells, J. Sound Vib.  237 (2000)

    920–930.[16] MSC/NASTRAN Basic Dynamic Analysis User’s Guide, 1997.[17] A.S. Kadi, A Study and Comparison of the Equations of Thin Shell Theories, Ph.D. Thesis, Ohio State University, Columbus, 1970.[18] K. Kraus, in:  Thin Elastic Shells, Wiley, New York, 1967.

    ARTICLE IN PRESS

    M. Caresta, N.J. Kessissoglou / Journal of Sound and Vibration 329 (2010) 733–751   751