the full nonlinear crack detection problem in uniform

37
The full nonlinear crack detection problem in uniform vibrating rods Lourdes Rubio Department of Mechanical Engineering, University Carlos III of Madrid, Avda. de la Universidad 30, 28911 Legan´ es, Madrid, Spain. E-mail: [email protected] Jos´ e Fern´ andez-S´ aez Department of Continuum Mechanics and Structural Analysis, University Carlos III of Madrid, Avda. de la Universidad 30, 28911 Legan´ es, Madrid, Spain. E-mail: [email protected] Antonino Morassi Corresponding author. Universit`a degli Studi di Udine, Dipartimento di Ingegneria Civile e Architettura, via Cotonificio 114, 33100 Udine, Italy. Tel.: +39 0432 558739; fax: +39 0432 558700. E-mail: [email protected] Abstract The basic problem in structural diagnostics via dynamic methods consists in determining the position and severity of a single open crack in a beam from the knowledge of a pair of resonant or antiresonant frequencies. A well-established theory is available for longitudinally vibrating uniform beams when the severity of the crack is small. In this paper we fill the gap present in the literature by showing that the results of the linearized theory for slight damage can be extended to a crack with any level of severity. Keywords: Damage identification, cracks, natural frequencies, antiresonant frequencies, inverse problems, rods 1. Introduction Dynamic testing is commonly used as a diagnostic tool to detect damage that has occurred in a mechanical system during service. The final goal is to predict Preprint submitted to Journal of Sound and Vibration August 22, 2014

Upload: hamid-m-sedighi

Post on 20-Nov-2015

14 views

Category:

Documents


0 download

DESCRIPTION

The full nonlinear crack detection problem in uniformvibrating rods

TRANSCRIPT

  • The full nonlinear crack detection problem in uniformvibrating rods

    Lourdes Rubio

    Department of Mechanical Engineering, University Carlos III of Madrid, Avda. de la

    Universidad 30, 28911 Leganes, Madrid, Spain. E-mail: [email protected]

    Jose Fernandez-Saez

    Department of Continuum Mechanics and Structural Analysis, University Carlos III of

    Madrid, Avda. de la Universidad 30, 28911 Leganes, Madrid, Spain. E-mail:

    [email protected]

    Antonino Morassi

    Corresponding author. Universita degli Studi di Udine, Dipartimento di Ingegneria Civile eArchitettura, via Cotonificio 114, 33100 Udine, Italy. Tel.: +39 0432 558739; fax: +39

    0432 558700. E-mail: [email protected]

    Abstract

    The basic problem in structural diagnostics via dynamic methods consists in

    determining the position and severity of a single open crack in a beam from the

    knowledge of a pair of resonant or antiresonant frequencies. A well-established

    theory is available for longitudinally vibrating uniform beams when the severity

    of the crack is small. In this paper we fill the gap present in the literature

    by showing that the results of the linearized theory for slight damage can be

    extended to a crack with any level of severity.

    Keywords: Damage identification, cracks, natural frequencies, antiresonant

    frequencies, inverse problems, rods

    1. Introduction

    Dynamic testing is commonly used as a diagnostic tool to detect damage that

    has occurred in a mechanical system during service. The final goal is to predict

    Preprint submitted to Journal of Sound and Vibration August 22, 2014

  • location and level of severeness of the degradation from the measurement of the

    changes induced by the damage on the vibrational behavior of the system.5

    Within the large class of diagnostic problems arising in structural mechanics,

    the crack detection problem in vibrating beams by frequency data has received

    a lot of attention in the scientific community in last two-three decades [1]. The

    reasons for this interest are various. Firstly, the beam model describes the be-

    havior of structural members that play an important role in many civil and10

    mechanical engineering applications. Secondly, the problem of identifying a

    crack in a beam is the basic diagnostic problem and, therefore, it represents

    an important benchmark to test the effectiveness of damage identification tech-

    niques. In addition, concerning the type of input data, in most applications

    researchers have used natural frequencies or antiresonant frequencies as effec-15

    tive damage indicator. Frequencies can be measured more easily than can mode

    shapes, and are usually less affected by experimental and modelling errors.

    Among the various models that have been proposed in literature to describe

    (open) cracks in beams, localized flexibility models enable one for simple and

    effective representation of the behavior of damaged elements [2]. The results20

    of extensive series of vibration tests carried out on steel beams with a single

    and multiple cracks confirm that lower natural frequencies are predicted by

    localized flexibility models with accuracy comparable to that of the classical

    Euler-Bernoulli model for a beam without defects, see, for example, [3].

    In this paper we shall mainly concerned with the inverse problem of identi-25

    fying a single open crack in a longitudinally vibrating beam by frequency data.

    The crack is modelled by inserting a translational linearly elastic spring at the

    damage cross-section. On assuming that the undamaged configuration is com-

    pletely known, the inverse problem consists in determining the location s of

    the crack, and its magnitude or severity. This latter parameter is expressed in30

    terms of the stiffness K of the spring simulating the crack, and the undamaged

    configuration is obtained by taking the limit as the stiffness K tends to infinity

    or, equivalently, as the flexibility 1K tends to zero.

    When the crack is small, namely when the cracked rod is a perturbation of

    2

  • the undamaged rod, a well-established theory for the inverse diagnostic problem

    is available, see Gladwell ([4], Chapter 15). The cornerstone property concerns

    with the possibility to express the change (2n) of the (square of) nth natural

    frequency n produced by a small single crack as a product of the flexibility1K

    and the square of the axial force N(U)n (s) evaluated on the undamaged config-

    uration, for the relevant mode shape, at the cracked cross section [5], [6], [7]:

    (2n) = (N

    (U)n (s))2

    K. (1)

    Then, in the case of small crack, the ratios of the change in different natural

    frequencies depend on the damage location s only, not on the severity K. Hearn35

    and Testa [1], Liang et al. [8], Rubio [9], among others, have used this property

    for damage localization in beam-like structures. See also Lakshmanan et al.

    [10] for identification of localized damage in rods from the iso-eigenvalue change

    contours constructed between pairs of different frequencies.

    Concerning the rigorous, i.e., mathematically proved, identification of a small40

    crack in an axially vibrating rod, worth of mention is the result obtained by

    Narkis [11]. Narkis proved that a single crack in a uniform free-free rod can

    be uniquely localized (up to a symmetric position) by using the first two nat-

    ural frequencies. Working on a linearized version of the frequency equation,

    Narkis obtained a closed-form solution for the crack location s. Using relation45

    (1), Morassi [12] extended Narkiss result to rods with single small crack under

    different set of end conditions and for different pairs of natural frequencies, pro-

    viding closed-form expressions also for the damage severity K. Later on, Dilena

    and Morassi [13] proved that the measurement of the first natural frequency and

    the first antiresonance of the driving point frequency response function evaluated50

    at one end of a free-free uniform rod allows to uniquely determine the position

    of the crack and its severity, by means of closed-form expressions. Extensions

    to cracked rods with dissipation [14], cracked pipes [15] and multi-cracked rods

    and beams [16], [17] are also available.

    All the above mentioned results hold in the case of small damage. Therefore,55

    3

  • an important question is left open: Can the above results be extended to the

    case of not necessarily small damage? More specifically, the following question

    can be posed:

    (Q) Do the Narkiss result [11] and the Dilena-Morassis result [13] continue

    to hold even for a large crack?60

    The a priori hypothesis of small crack is often considered to be a no very

    restrictive limitation, since in several practical situations it is of interest to be

    able to identify the damage as soon as it arises in a structure. However, there

    are several motivation in support of the opportunity to extend the theory to not

    necessarily small cracks. Firstly, it is not easy to determine rigorously when a65

    crack can be considered small. The smallness of a crack is typically established

    on the basis of the crack-induced changes on the lower natural frequencies.

    However, this criterion is difficult to apply, since the vibration modes have

    different sensitivity to damage according to the position of the crack along the

    beam axis. The introduction of an average frequency shift does not simplify70

    the analysis, since it should be clarified how many data must be included in the

    calculation and how the threshold value corresponding to small damage should

    be selected. Secondly, the linearized theories by Narkis and Morassi show some

    limitation when the damage is located near a point of vanishing sensitivity for

    a vibration mode. In [12] it was shown that this aspect also prejudices the75

    reliability in assessment of the damage severity (see Table 3 in [12]). Analogous

    indeterminacy was recently encountered in identifying multiple small cracks in

    a longitudinally vibrating beam by frequency measurements [17]. Near a zero-

    sensitivity point, the first order effect of the damage on the frequencies vanishes.

    Therefore, it is expected that the indeterminacy can be removed by considering80

    the full nonlinear inverse problem instead of its linearized version. Finally, it is

    of course desirable to have a unifying general theory of the diagnostic problem

    capable to include damages ranging from small to large severity.

    The waiver to the linearized theory implies strong consequences and, in

    particular, the inability to obtain an explicit relationship like (1) expressing85

    4

  • the change of a natural frequency in terms of the damage parameters and the

    undamaged configuration. In the full nonlinear crack detection problem, the

    relationship between damage parameters (s,K) and natural frequencies is im-

    plicitly contained in the frequency equation of the cracked rod. Diagnostic

    methods based on the analysis of the frequency equation have been already in-90

    vestigated in literature. The approach generally consists in considering a pair

    of frequency data (typically, the first two natural frequencies) and then solving

    numerically the nonlinear system formed by the frequency equation written for

    the two selected frequencies in terms of the damage parameters s and K. We

    refer, among others, to the pioneering research developed by Adams et al. [5],95

    the studies by Springer et al. [18] and Lin and Chang [19] on longitudinally

    vibrating rods, and the contributions by Nandwana and Maiti [20], Cerri and

    Vestroni [21], Vestroni and Capecchi [22] on cracked beams in bending vibra-

    tion. The above results seem to suggest that the answer to question (Q) is

    positive. However, at the best of our knowledge, a rigorous proof of this general100

    property is not available, as the conclusions of the cited papers are drawn either

    on the basis of numerical analysis of specific cases or on the study of particular

    experimental situations.

    In this paper we definitively prove that the answer to question (Q) is positive,

    that is the results by Narkis [11] and Dilena-Morassi [13] continue to hold even105

    for not necessarily small cracks. The corresponding analysis is developed in

    Section 3 and in Section 4, respectively. Proofs are based on a careful analysis

    of the solutions of the system of two frequency equations written for the pair

    of frequency input data and on the use of classical results of the variational

    theory of eigenvalues for longitudinally vibrating rods. A series of applications110

    to measurements on cracked steel rods are presented in Section 5.

    2. Formulation of the diagnostic problem and some frequency bound

    Let us consider a straight thin rod under longitudinal vibration and with

    free-free end conditions (F-F). We assume that the rod is uniform and has a

    5

  • single crack at the cross-section of abscissa xd, with 0 < xd < L, where L is the115

    length of the rod. The crack is assumed to remain open during vibration and is

    modelled as a massless longitudinal linearly elastic spring with stiffness K, see,

    for example, [2] and [23]. The value of K can be determined in terms of the

    geometry of the cracked cross-section and of the material properties of the rod.

    The free undamped longitudinal vibrations of the rod with radian frequency120

    and spatial amplitude w = w(x) are governed by the following eigenvalue

    problem written in dimensionless form

    w + 2w = 0, in (0, s) (s, 1),w(s) = w(s+),

    w(s+) w(s) = Fw(s),w(0) = w(1) = 0,

    (2)

    (3)

    (4)

    (5)

    where s = xdL , s (0, 1), and 2 = L2

    E 2; E is the (constant) Youngs modulus

    of the material; is the (uniform) volume mass density of the material; F = EALK ,1250 < F < +, is the normalized flexibility associated to the crack, where A is thearea of the transversal cross-section of the rod. The eigenvalue problem (2)(5)

    has an infinite sequence of eigenvalues {n}n=0, with 0 = 0 < 1 < 2 < ... andlimnn = +. (Note that here and in the sequel we consider as eigenvaluethe positive square root of 2.) In correspondence of each eigenvalue n, there130

    exists a no trivial solution wn = wn(x) of Eqs. (2)(5), e.g., the eigenfunction

    associated to the eigenvalue n.

    The eigenvalues {n}n=0 are the roots of the frequency equation associatedto the cracked rod

    P(; s,F) (sinF sin(s) sin(1 s)) = 0, (6)

    where P = P(; s,F) is the corresponding characteristic polynomial. Therefore,the eigenvalues n = n(s,F) are regular functions of the damage parameterss and F .135

    The eigenvalue problem for the undamaged rod can be found by taking

    K or, equivalently, F 0+ in Eqs. (2)(5). The corresponding frequency

    6

  • equation is

    P(U)() sin = 0 (7)

    and the normalized eigenpairs {(U)n , w(U)n (x)}n=0 have the explicit expressions

    (U)n = n, w(U)n (x) =

    2 cos(nx), n = 0, 1, 2, ... (8)

    By Eq. (6) and Eq. (7), it is easy to see that the 0th eigenvalue is insensitive

    to damage, e.g., 0 = (U)0 = 0, since the corresponding vibration mode is a

    longitudinal rigid motion.

    As it is well-known, closed-form expressions for the eigenvalues of the dam-

    aged rod are in general not available. Nevertheless, useful upper and lower

    bounds on n, n 1, can be provided by the Variational Theory of eigenvalues[24]. Let us add the constraint

    l(w) w(s+) w(s) = 0 (9)

    to the no trivial solutions of the eigenvalue problem (2)(5) for the cracked rod.

    The above constraint can be interpreted as a real-valued continuous functional

    acting on the functional spaceH1(0, s)H1(s, 1) of the admissible configurationsof the cracked rod. Here, H(a, b), < a < b < , is the Hilbert space ofthe functions f : (a, b) R such that b

    af2 < and b

    a(f )2 < , where f is

    the weak derivative of f in the sense of distributions. Clearly, the eigenvalues

    of the problem (2)(5) under the constraint (9) coincide with the eigenvalues

    of the undamaged rod. The addition of the constraint (9) restricts the space

    of the admissible configurations reachable by the cracked rod, and then, by a

    Monotonicity theorem, the eigenvalues cannot decrease, namely

    n (U)n , for every n 1. (10)

    Moreover, by the Minimax Eigenvalue Principle [24], the nth eigenvalue of the

    system with the constraint (9) cannot be bigger than the (n + 1)th eigenvalue

    of the unconstrained system, namely

    (U)n n+1, for every n 1. (11)

    7

  • Then, by Eq. (10) and Eq. (11) we deduce the following upper and lower bound

    on the eigenvalues of the cracked rod

    (U)n1 n (U)n , for every n 2. (12)

    In this paper we are mainly concerned with the determination of the position

    s and the severity F of the crack by the knowledge of the two eigenvalues 1 and1402. To solve the inverse problem, we form a two-equations system by writing

    the frequency equation (6) for = 1 and = 2. More precisely, with the aim

    of reproducing the inverse problem arising in practical applications, we assume

    that specific values of the two eigenvalues have been given, say 1 = 1, 2 = 2,

    and we investigate on the solutions (s,F) of the nonlinear system of equations145

    P(1; s,F) = 0,P(2; s,F) = 0.

    (13)

    (14)

    Actually, since both 1 and 2 are strictly positive, Eqs. (13)(14) can be

    replaced by

    sin1 = F1 sin(1s) sin(1(1 s)),sin2 = F2 sin(2s) sin(2(1 s)).

    (15)

    (16)

    We note that, by the symmetry of the undamaged rod, the two damage con-

    figurations (s,F) and (1 s,F) are indistinguishable with our choice of data.Then, without loss of generality, we can assume

    0 < s 12. (17)

    In Section 4 we shall show how to select appropriately the spectral data in order150

    to remove the a priori assumption (17).

    We conclude this section with the following simple result.

    Proposition 2.1. Under the above notation, we have:

    i) If 1 = (U)1 , then there is no crack.

    8

  • ii) If 2 = (U)2 , then either there is no crack or there is a crack located at155

    s = 12 . In this latter case, the severity of the crack is indeterminate.

    Proof. Case i). Let 1 = (U)1 . Then, by Eq. (8) and Eq. (15) we have

    F sin(s) sin((1 s)) = 0. But, by Eq. (17), we have sin(s) sin((1 s)) > 0and, then, F must vanish.

    Case ii). Let 2 = (U)2 . Then, by Eq. (8) and Eq. (16) we have160

    F sin(2s) sin(2(1 s)) = 0, which implies, by Eq. (17), either s = 12 (and Fremains indeterminate) or F = 0 (i.e., there is no damage).

    By taking Proposition 2.1 into account and recalling that eigenvalues are

    simple, in the sequel we shall assume strict inequalities in Eq. (12), namely our

    data 1, 2 satisfy

    0 < 1 < , (18)

    < 2 < 2. (19)

    An additional inequality on 2 will be introduced later on.

    3. Crack identification by the first two natural frequencies

    Main result of this section is the following theorem.165

    Theorem 3.1. Let s (0, 12) and 0 < F < . Let the value 1, 2 of the firstand second eigenvalues be given and satisfying Eq. (18) and Eq. (19). There

    exists a unique pair of damage parameters (s,F) solution to Eqs. (15) and (16).

    The rest of the section is devoted to the proof of Theorem 3.1. To simplify

    the notation, we omit the overline in denoting the eigenvalue data and we simply170

    indicate by 1, 2 the given (e.g., experimental) value of the first and second

    eigenvalue.

    The proof consists of several steps. In the first step, the dependence on Fin Eqs. (15) and (16) is removed, and we write a single equation on the crack

    9

  • position s only:

    2 sin1 sin(2s) sin(2(1 s)) = 1 sin2 sin(1s) sin(1(1 s)). (20)

    This fact was already observed in the literature, even for cracked beams in

    bending vibration, see, for example, Cerri and Vestroni (2000). The relationship

    analogous to Eq. (20) for the linearized problem is the well-known property -175

    mentioned in the Introduction - that, in the case of a small crack, the ratio of

    the change in two different natural frequencies depends on the damage location

    only, not on the severity of the damage.

    By comparing the sign of the factors appearing in Eq. (20), we deduce an

    additional bound on 2. By Eqs. (17)(19) we have180

    1 (0, ) sin1 (0, 1],1s (0, 2 ) sin(1s) (0, 1),1(1 s) (0, ) sin(1(1 s)) (0, 1],2 (, 2) sin2 [1, 0),2s (0, ) sin(2s) (0, 1],2(1 s) (2 , 2) sin(2(1 s)) (0, 1) if 2(1 s) (2 , ),2(1 s) (2 , 2) sin(2(1 s)) [1, 0) if 2(1 s) (, 2).

    (21)

    (22)

    (23)

    (24)

    (25)

    (26)

    (27)

    By inserting Eqs. (21)(27) in Eq. (20) we deduce that sin(2(1 s)) < 0, thatis

    1 s < 2 0,h(0) = (22 21) sin1 sin2 < 0,h( 12 ) = 0.

    (32)

    (33)

    (34)

    (35)

    Then, since h = h(s) is a regular function, there exists sd (0, 12 ) such thath(sd) = 0, and the existence part of Theorem 3.1 is proved.185

    To conclude the proof we need to show that such a point sd is unique. To

    prove uniqueness, it is enough to prove that the following Claim is true:

    Claim: h vanishes in exactly one point inside(0,

    1

    2

    ). (36)

    In fact, if Eq. (36) is satisfied at the point sh, sh (0, 12 ), then h(sh) < 0. (Ifh(sh) 0, then h must vanish in more than one point in (0, 12 ), a contradiction.)Therefore, it is enough to apply the Intermediate Value Theorem for continuous

    functions to conclude that, since {h(sh) < 0, h( 12 ) > 0, h(s) > 0 in (sh, 12 )},there exists exactly one zero of h inside (sh,

    12 ). Observing that h(s) < 0 for190

    s (0, sh), we have the thesis.Therefore, it remains to prove Claim (36). The zeros of h(s) = 0 are the

    roots of the equation

    f(s) = g(s), s (0,

    1

    2

    ), (37)

    where

    f(s) = 22 sin1 sin(2(1 2s)), (38)

    g(s) = 21 sin2 sin(1(1 2s)). (39)

    11

  • We analyze separately the functions f and g. Concerning the function g, by

    Eqs. (29), (24) and since 1(1 2s) (0, ), we have

    g(0) < 0, g

    (1

    2

    )= 0, g(s) < 0 in

    (0,

    1

    2

    ). (40)

    The first derivative of g is

    g(s) = 231 sin2 cos(1(1 2s)) (41)

    and then g may have at most one zero in(0, 12

    ), say sg, given by

    sg =1

    2

    41. (42)

    We distinguish two cases:

    Case i) (small damage) If 2 < 1, then there exists a unique sg (0, 12 )such that g(sg) = 0, and g(s) < 0 in (0, sg), g(s) > 0 in (sg, 1).

    Case ii) (large damage) If 1 0 in (0, 12 ) and g = g(s) is195

    monotonically increasing in (0, 12 ).

    The two cases i) and ii) are sketched in Figure 1. Note that g(12

    )= 231 sin2 >

    0.

    Let us consider the function f . Since 0 < 2(1 2s) < 2 for s (0, 12

    ), the

    function f = f(s) vanishes exactly at the point sf (0, 12

    )given by

    sf =1

    2

    22. (43)

    Note that, since < 2, the point sf always belong to(0, 12

    ). Moreover, by

    direct computation we have

    f(0) < 0, f(0) g(0) = (22 21) sin1 sin2 < 0, f(1

    2

    )= 0. (44)

    Now we are in position to analyze the set of solutions to Eq. (37). Depending

    on the behavior of the function g we distinguish two cases.200

    The simplest is Case ii), corresponding to 0 < 1 sg 2 < 21,

    (45)

    The first derivative of f is

    f (s) = 232 sin1 cos(2(1 2s)). (46)

    A direct calculation shows that f vanishes at s2 = 12 42 (which alwaysbelongs to

    (0, 12

    )) and that may vanish at s1 =

    12 342 if 2 > 32 . Obviously,210

    s2 sf = sf s1 = 42 . Note also that f (12

    )= 232 sin1 < 0. The

    qualitative behavior of f is shown in Figure 3 and it is compared to that of the

    function g. One can conclude that in all the cases shown in Figure 3 there exists

    only one solution to the equation f(s) = g(s) in(0, 12

    ).

    The proof of the Claim (36) is complete.215

    Finally, once the position sd of the crack has been found, the flexibility Fcan be determined by Eq. (15), and the proof of Theorem 3.1 is complete.

    4. Crack identification by resonant and antiresonant frequency data

    Theorem 3.1 states that the knowledge of the first two natural frequencies

    of the problem (2)(5) allows to uniquely identify a single crack in a free-free220

    uniform rod, up to a symmetric position. In this section we shall prove that the

    nonuniqueness of the solution due to the symmetry of the undamaged configura-

    tion can be removed by using in connection resonant and antiresonant frequency

    data. Precisely, in addition to the first natural frequency, we use the first an-

    tiresonant frequency of the driving point frequency response function (FRF) of225

    the rod H(, 0, 0), where is the frequency variable. Antiresonances are the

    13

  • zeros of the FRF H(, 0, 0) and coincide with the natural frequencies of the

    rod when the longitudinal displacement at the cross-section x = 0 is hindered,

    namely the antiresonances are the natural frequencies of the cantilever rod sup-

    ported at x = 0. Under the same notation of Section 3, the corresponding230

    eigenvalue problem in dimensionless variables is

    w + 2w = 0, in (0, s) (s, 1),w(s) = w(s+),

    w(s+) w(s) = Fw(s),w(0) = 0, w(1) = 0,

    (47)

    (48)

    (49)

    (50)

    where s (0, 1) and 0 < F < . We denote by {n}n=1 the eigenvalues ofthe problem (47)(50), with 0 < 1 < 2 < ... and limn n = . (As inSection 2, hereinafter we consider as eigenvalues the positive roots of 2.) The235

    eigenvalue problem for the undamaged rod is obtained by taking F 0+ inEqs. (47)(50).

    The eigenvalues {n}n=1 are the roots of the frequency equation

    Q(; s,F) (cosF cos(s) sin(1 s)) = 0. (51)

    The equation analogous to Eq. (51) for the undamaged cantilever is

    Q(U)() cos = 0 (52)

    and the normalized eigenpairs {(U)n , v(U)n (x)}n=1 are

    (U)n =

    (1

    2+ (n 1)

    ), v(U)n (x) =

    2 sin

    (1

    2+ (n 1)

    )x, n = 1, 2, ...

    (53)

    With a slight change of notation with respect to Section 2 and 3, in this section it

    turns out to be convenient to enumerate the eigenvalues of the free-free cracked

    and uncracked rod from n = 1, namely

    0 = 1 < 2 < ..., {n}n=1, (54)

    0 = (U)1 <

    (U)2 < ..., {(U)n }n=1. (55)

    14

  • As in Eq. (10) we have

    n (U)n , for every n 1. (56)

    In order to compare the eigenvalues of the cantilever cracked rod with those of

    the free-free cracked rod, we consider the effect of adding the constraint

    m(w) w(0) = 0 (57)

    on the admissible set of no trivial solutions of the eigenvalue problem (2)(5) of

    the free-free cracked rod. Clearly, the eigenvalues of (2)(5) under the constraint

    (57) coincide with the eigenvalues of the cantilever cracked rod {n}n=1 and,by the Minimax Eigenvalue Principle [24], we have

    n n n+1, for every n 1. (58)

    In particular, focussing the attention on 2 and 1, by Eqs. (18), (56) (for

    n = 1) and Eq. (58) (for n = 2), we have

    0 < 2 < , (59)

    0 < 1 0,

    h(1) = 0,

    h(0) = (22 + 21) sin2 cos1 < 0,h(1) = (22 21) sin2 cos1 > 0.

    (74)

    (75)

    (76)

    (77)

    The function h = h(s) is regular and then, by Eqs. (74)(77) and the Inter-

    mediate Value Theorem, there exists a point sd (0, 1) such that h(sd) = 0(existence of a crack position). It remains to prove that such a point is unique.

    To prove that h = h(s) has exactly one zero in (0, 1), it is enough to prove that

    h = h(s) vanishes only at one point inside (0, 1). The zeros of h are the roots

    of the equation

    f(s) = g(s), (78)

    where

    f(s) = 21 sin2 cos(1(1 2s)), (79)

    g(s) = 22 cos1 sin(2(1 2s)). (80)

    We first analyze the function g = g(s). By definition we have g(1 s) = g(s),and g = g(s) is an odd function with respect to s = 12 , with g(0) = g(1) =22 cos1 sin2 < 0. Moreover, recalling Eq. (59), it easy to see that the only265zero of g belonging to (0, 1) is at s = 12 . The graph of g is sketched in Figure 4.

    Since, by Eq. (60), 1(1 2s) (2 , 2 ), we have cos(1(1 2s)) > 0,

    and the function f = f(s) is strictly positive in (0, 1). Moreover, by definition,

    f(1 s) = f(s) and f is an even function with respect to s = 12 , with f(0) =f(1) = 21 sin2 cos1 > 0, f

    (12

    )= 21 sin2 > f(0). We also have

    g(1) f(1) = (22 21) sin2 cos1 > 0, (81)

    g(0) f(0) = (22 + 21) sin2 cos1 < 0. (82)

    17

  • The graph of f is sketched in Figure 5, together with the possible behavior of the

    function g. Clearly, there exists only one point s in (0, 1) such that g(s) = f(s).

    Therefore, the function h = h(s) vanishes in (0, 1) only at s = s and this

    implies the uniqueness of the crack position. The severity coefficient F can be270determined by Eq. (63), and the proof of Theorem 4.1 is complete.

    5. Applications

    In Section 3 we have shown that the knowledge of the first two positive

    natural frequencies allows to uniquely identify a single crack in a free-free uni-

    form rod, up to a symmetric position. In Section 4 we have proved that the275

    nonuniqueness of the solution due to the symmetry of the undamaged config-

    uration can be removed by using the first positive natural frequency and the

    first antiresonant frequency of the driving point frequency response function

    H(, 0, 0) of the rod. The identification procedure has been checked on the ba-

    sis of an exhaustive numerical investigation. The results show that the method280

    allows for the exact determination of the actual solution of the diagnostic prob-

    lem in absence of errors on the data, thus confirming numerically the proofs of

    Theorem 3.1 and Theorem 4.1. This is true, in particular, for the identification

    of a crack located near a point of the beam axis of vanishing sensitivity for a

    vibration mode. Taking into account of the data used in this paper, namely285

    the first two positive resonant frequencies and the first antiresonant frequency

    of the free-free bar, the only case that falls in this class corresponds to a crack

    positioned in the vicinity of the middle-point of the bar, so that the second

    positive resonance frequency of the free-free bar is insensitive to damage.

    In this section we check the ability of the identification methodology to deal290

    with real cases in which experimental/modelling errors in the resonant/antiresonant

    data are present. In particular, two different applications shall be considered in

    the sequel. They correspond to the extreme cases in which either the analytical

    model used for the interpretation of the measurements is extremely accurate

    (Example 1) or rather large modelling errors are present in the data (Example295

    18

  • 2).

    The mechanical model of Example 1 is a free-free steel rod of length L =

    2.925 m and square solid cross-section 22 22 mm, which has been damagedat the cross-section located at xd = 1.000 m from one end, see Figure 1(a) in

    [12]. Three different damage configurations, D1, D2 and D3, were obtained300

    by introducing a notch of increasing depth. Details of the experimental setup

    and of the experimental modal analysis results can be found in [12]. Table 1

    compares the experimental natural frequencies and their corresponding analyt-

    ical estimates for the undamaged and damaged rod. For the definition of the

    analytical model for the damaged rod, the theoretical value of the stiffness K,305

    for each damage configuration, was obtained by assuming that the position xd

    of the damage is known and by taking the measured value for the fundamen-

    tal (positive) frequency of the damaged rod. Thus, the actual values of K for

    the different damage configurations are the following: K = 3.09119 1010N/m(case D1), K = 7.84984109N/m (case D2), K = 4.37183108N/m (case D3).310The analytical model turns out to be extremely accurate for all the configura-

    tions under investigation and the discrepancy between measured and analytical

    values of the first two positive natural frequencies is lower than 0.04 per cent.

    Frequency shifts induced by the damage are significantly larger than the mod-

    elling errors, and are about 0.08, 0.32, 5 per cent of the undamaged value for315

    damage configurations D1, D2 and D3, respectively.

    Table 2 shows the estimated values of the position xNLd of the cracked sec-

    tion and the spring stiffness KNL. The results of the linearized version of the

    identification procedure proposed by Morassi [12] are also included in Table 2

    for comparison.320

    It is possible to observe that the estimated values obtained using the full

    nonlinear solution agree well with the actual solution of the problem: the max-

    imum difference is about 3% for the estimated position and less than 4% for

    the spring stiffness. The discrepancies between identified and actual damage

    parameters connected with this solution are exclusively due to the experimen-325

    tal errors associated to frequency measurements and to modelling errors. De-

    19

  • viations are typically bigger for less severe levels of damage. The maximum

    discrepancy occurs for damage configuration D1. It is worthy to note that in

    this case the linearized estimation procedure gives better results than the full

    nonlinear method, whereas, as it was expected, the linearized solution deviates330

    increasingly from the actual solution as the damage severity increases.

    Example 2 is based on the experimental results presented in [13]. The me-

    chanical model is a double T free-free steel rod of series HE100B. The length

    is L = 2.747 m and the linear mass density is 20.4 kg/m. The damage was

    obtained by saw cutting the rod at the cross-section at xd = 0.550 m far from335

    the left end, see Figure 5(a) in [13]. Two different damage configurations, called

    D4 and D5 in the following, were obtained by introducing a notch of increas-

    ing depth. Details of the experimental setup and discussion on the results of

    experimental modal analysis can be found in [13].

    Table 3 compares the experimental values of the first two (positive) natural340

    frequencies and the first antiresonance of H(, 0, 0) with their corresponding an-

    alytical estimates, both for the undamaged and damaged rod. The actual value

    of the spring stiffness K of the cracked rod was defined by assuming the position

    xd of the damage as known and providing that, for each damage configuration,

    the measured and the analytical fundamental (positive) natural frequency co-345

    incide. Thus, the actual values of K for the different damage configurations

    are the following: K = 3.507 1010N/m (case D4), K = 1.736 109N/m(case D5). Frequency shifts induced by damage in the first two resonances are

    around 0.20.6 and 410 per cent for D4 and D5 configurations, respectively.

    Antiresonance decreasing is about 7 per cent for configuration D5. As mod-350

    elling errors are concerned, the analytical model turns out to be accurate for

    natural frequencies, with maximum differences between experimental and the-

    oretical values equal to 0.25 per cent. Modelling error for the first antiresonant

    frequency is also around 1 per cent, with the exception of the configuration

    D4 in which a discrepancy around 6 per cent was noticed. The source of this355

    disagreement was not explained in [13], and it is expected that the important

    error on first antiresonance will produce wrong estimates of the damage param-

    20

  • eters for damage configuration D4. It should be also noticed that, although the

    analytical model can be considered accurate in predicting natural frequencies,

    percentage crack-induced changes in natural frequencies are comparable with360

    the accuracy of the rod model for damage level D4.

    Table 4 collects the estimated values of the position of the cracked section,

    xNLd , and the spring stiffness, KNL, when the first two positives resonant fre-

    quencies are used in the identification process. As expected, for configuration

    D4, the inaccuracy of the data prejudices the reliability of the reconstruction.365

    The results obtained by the linearized technique proposed in [12] are also in-

    cluded for comparison in Table 4. As the small damage case (D4) is concerned,

    the estimation obtained from the linearized procedure is better, whereas the

    full nonlinear procedure gives an estimation closer to the actual solution as the

    damage severity increases.370

    Finally, Table 5 shows the results when the first positive resonant and the

    first antiresonant frequency are used in the identification process. It can be

    noted that the employment of the antiresonance, which is affected by large

    modelling error in configuration D4, prevents obtaining accurate estimation of

    the damage parameters. Conversely, the solution predicted by the theory is a375

    satisfactory estimate of the actual solution of the damage problem for configu-

    ration D5. In this case, in fact, modelling errors are small with respect to the

    shifts induced by the crack and, as it was expected, the full nonlinear solution

    gives a more accurate estimate of the damage parameters than the linearized

    procedure.380

    6. Conclusions

    In this paper we have considered the problem of identifying a single open

    crack in a uniform longitudinally vibrating rod by minimal frequency data. The

    crack is modelled by a massless translational linearly elastic spring placed at

    the damaged cross section. Two parameters, the position of the crack and its385

    severity, are the unknowns of the inverse problem, and two spectral information

    21

  • are considered as input data.

    We have shown that the crack can be uniquely identified - up to a symmet-

    rical position - by the knowledge of the first two positive natural frequencies of

    the rod under free-free end conditions. Moreover, it was also shown that the390

    symmetric position can be removed by replacing the second positive natural

    frequency with the first antiresonant frequency of the driving point frequency

    response function evaluated at one end of the rod.

    The above results were known to hold in the case of small crack. The analy-

    sis developed in this paper allows to extend the existing theory to a crack with395

    any level of severity. The methodology used for the proof is completely different

    from that adopted in the case of small damage. It is based on a careful analysis

    of the solutions of the nonlinear system formed by the frequency equation writ-

    ten for the pair of spectral input data, coupled with suitable upper and lower

    bounds derived within the variational theory of eigenvalues. Numerical results400

    are in agreement with the theory when exact analytical data are employed in

    identification. The theory has been also checked on a series of dynamic tests on

    cracked steel beams. Experiments show that if frequency data used in identifi-

    cation are affected by relatively small errors with respect to the shifts induced

    by the crack, then the proposed full nonlinear solution gives a more accurate405

    estimate of the position and severity of the crack than the linearized procedure

    based on the assumption of small damage.

    The method proposed in this paper can be used, in principle, also to deal

    with the corresponding diagnostic problem for a uniform beam in bending vi-

    bration with a single open crack. In that case, the crack is represented by the410

    insertion of a massless rotational spring at the damaged cross-section. A pre-

    liminary analysis of the problem shows that the study of the solutions of the

    system formed by the frequency equation written for, say, the first two natu-

    ral frequencies is significantly more difficult with respect to the axial case. No

    general results seem to be available yet on this challenging diagnostic problem.415

    As a final remark, we point out that another aspect worth of investigation,

    both from the theoretical and practical point of view, stands on the possibility

    22

  • of extending our analysis to axially vibrating rods with variable profile. Such an

    extension seems to be not trivial at all, since, for the variable coefficient case, no

    closed-form expression for the eigenpairs is available and the frequency equation420

    is written in implicit form. It is likely that new ideas and mathematical tools

    must be developed to deal with this class of problems.

    Acknowledgement

    The work of A. Morassi is supported by the University Carlos III of Madrid-

    Banco de Santander Chairs of Excellence Programme for the 2013-2014 Aca-425

    demic Year.

    A. Morassi wishes to thank the colleagues of the University Carlos III of

    Madrid, especially Professors L. Rubio and J. Fernandez-Saez, for the warm

    hospitality at the Department of Engineering Mechanics.

    References430

    [1] G. Hearn, R.B. Testa, Modal analysis for damage detection in structures.

    Journal of Structural Engineering ASCE 117(10) (1991) 30423063.

    [2] L.B. Freund, G. Herrmann, Dynamic fracture of a beam or plate in plane

    bending. Journal of Applied Mechanics 76-APM-15 (1976) 112116.

    [3] S. Caddemi, A. Morassi, Multi-cracked Euler-Bernoulli beams: mathemati-435

    cal modelling and exact solutions. International Journal of Solids and Struc-

    tures 50(6) (2013) 944956.

    [4] G.M.L. Gladwell, Inverse problems in vibration, Kluwer Academic Publish-

    ers, Dordrecht, The Netherlands, Second edition, 2004.

    23

  • [5] R.D. Adams, P. Cawley, C.J. Pye, B.J. Stone, A vibration technique for440

    non-destructively assessing the integrity of structures. Journal of Mechanical

    Engineering Science 20(2) (1978) 93100.

    [6] P. Gudmundson, The dynamic behaviour of slender structures with cross-

    sectional cracks. Journal of the Mechanics and Physics of Solids 31 (1983)

    329345.445

    [7] A. Morassi, Crack-induced changes in eigenparameters of beam structures.

    Journal of Engineering Mechanics ASCE 119(9) (1993) 17981803.

    [8] R.Y. Liang, J. Hu, F. Choy, Quantitative NDE technique for assessing dam-

    ages in beam structures. Journal of Engineering Mechanics ASCE 118(7)

    (1992) 1468-1487.450

    [9] L. Rubio L, An efficient method for crack identification in simply supported

    Euler-Bernoulli beams. Journal of Vibration and Acoustics 131 (2009) Paper

    051001.

    [10] N. Lakshmanan, B.K. Raghuprasad, N. Gopalakrishnan, K. Sathishkumar,

    S.G.N. Murthy, Detection of contiguous and distributed damage through455

    contours of equal frequency change. Journal of Sound and Vibration 329

    (2010) 13101331.

    [11] Y. Narkis, Identification of crack location in vibrating simply supported

    beams. Journal of Sound and Vibration 172(4) (1994) 549558.

    [12] A. Morassi, Identification of a crack in a rod based on changes in a pair of460

    natural frequencies. Journal of Sound and Vibration 242(4) (2001) 577596.

    [13] M. Dilena, A. Morassi, The use of antiresonances for crack detection in

    beams. Journal of Sound and Vibration 276(1-2) (2004) 195214.

    [14] M. Dilena, A. Morassi, Detecting cracks in a longitudinally vibrating beam

    with dissipative boundary conditions. Journal of Sound and Vibration 267465

    (2003) 87103.

    24

  • [15] M. Dilena, M. Fedele DellOste, A. Morassi, Detecting cracks in pipes filled

    with fluid from changes in natural frequencies. Mechanical Systems and Sig-

    nal Processing 25(8) (2011) 31863197.

    [16] A. Morassi, M. Rollo, Identification of two cracks in a simply supported470

    beam from minimal frequency measurements. Journal of Vibration and Con-

    trol 7(5) (2001) 729740.

    [17] L. Rubio, J. Fernandez-Saez, A. Morassi, Identification of two cracks in a

    rod by minimal natural frequency and antiresonant frequency data. Submit-

    ted, 2013.475

    [18] W.T. Springer, K.L. Lawrence, T.J. Lawley, Damage assessment based on

    the structural frequency-response function. Experimental Mechanics 28(1)

    (1998) 34-37.

    [19] H.-P. Lin, S.-C. Chang, Identification of crack location in a rod from mea-

    surements of natural frequencies, Preprint, 2004.480

    [20] B.P. Nandwana, S.K. Maiti, Detection of the location and size of a crack

    in stepped cantilever beams based on measurements of natural frequencies.

    Journal of Sound and Vibration 192(3) (1997) 324335.

    [21] M.N. Cerri, F. Vestroni, Detection of damage in beams subjected to diffused

    cracking. Journal of Sound and Vibration 234(2) (2000) 259276.485

    [22] F. Vestroni, D. Capecchi, Damage detection in beam structures based

    on frequency measurements. Journal of Engineering Mechanics ASCE 126

    (2000) 761-768.

    [23] E. Cabib, L. Freddi, A. Morassi, D. Percivale, Thin notched beams. Journal

    of Elasticity 64(2/3) (2001) 157178.490

    [24] R. Courant, D. Hilbert, Methods of Mathematical Physics (volume I), First

    English Edition, Interscience Publishers, Inc., New York, 1966.

    25

  • Table Captions

    Table 1. Example 1: First two positive natural frequencies fn =n2 , n = 1, 2,

    of the undamaged free-free rod and their values associated to the damaged cases495

    Di, i = 1, 2, 3. Frequency values in Hz. Note: Modelling errors are reported

    in brackets, = 100 (fModel fExp.)/fExp..

    Table 2. Example 1: Damage parameter estimates xNLd , KNL obtained by the

    first two natural frequencies fn =n2 , n = 1, 2. Actual solution: xd = 1.000

    m; K = 3.09119 1010N/m (case D1), K = 7.84984 109N/m (case D2),500K = 4.37183 108N/m (case D3). xLd and KL are the corresponding estimatesobtained by the linearized procedure presented in [12].

    Table 3. Example 2: First two positive natural frequencies fn =n2 , n =

    2, 3, (note the ordering relation in Eq. (54)) and first antiresonant frequency

    f1A =12 of the undamaged free-free rod, and their values associated to the505

    damaged cases Di, i = 4, 5. Frequency values in Hz. Note: Modelling errors

    are reported in brackets, = 100 (fModel fExp.)/fExp..

    Table 4. Example 2: Damage parameter estimates xNLd , KNL obtained by

    the first two positive natural frequencies fn =n2 , n = 2, 3 (note the ordering

    relation in Eq. (54)). Actual solution: xd = 0.550 m; K = 3.507 1010N/m510(case D4), K = 1.736 109N/m (case D5). xLd and KL are the correspondingestimates obtained by the linearized procedure presented in [12].

    Table 5. Example 2: Damage parameter estimates xNLd , KNL obtained by the

    first positive natural frequency f2 =22 (note the ordering relation in Eq. (54))

    and the first antiresonant frequency f1A =12 . Actual solution: xd = 0.550 m;515

    K = 3.507 1010N/m (case D4), K = 1.736 109N/m (case D5). xLd and KLare the corresponding estimates obtained by the linearized procedure presented

    in [13].

    26

  • Figure Captions

    Figure 1. The function g = g(s).520

    Figure 2. Case ii): unique solution of f(s) = g(s).

    Figure 3. Case i): unique solution of f(s) = g(s). Left: 32 < 2(< 2); right:

    (

  • Table 1: Example 1: First two positive natural frequencies fn =n2

    , n = 1, 2, of the undam-

    aged free-free rod and their values associated to the damaged cases Di, i = 1, 2, 3. Frequency

    values in Hz. Note: Modelling errors are reported in brackets, = 100 (fModel fExp.)/fExp..

    Data Undamaged D1 D2 D3

    Exper. Model Exper. Model Exper. Model Exper. Model

    f1 882.25 882.25 881.5 881.5 879.3 879.3 831.0 831.0

    (0.00) (0.00) (0.00) (0.00)

    f2 1764.6 1764.5 1763.3 1763.1 1759.0 1759.2 1679.5 1680.1

    (-0.006) (-0.011) (0.011) (0.036)

    28

  • Table 2: Example 1: Damage parameter estimates xNLd , KNL obtained by the first two natural

    frequencies fn =n2

    , n = 1, 2. Actual solution: xd = 1.000 m; K = 3.09119 1010N/m (caseD1), K = 7.84984 109N/m (case D2), K = 4.37183 108N/m (case D3). xLd and KL arethe corresponding estimates obtained by the linearized procedure presented in [12].

    Position [m] Spring stiffness [N/m]

    Case xNLd xLd K

    NL KL

    D1 1.031 1.012 3.20720 1010 3.13548 1010

    D2 0.992 0.989 7.78180 109 7.73130 109

    D3 0.998 1.021 4.36188 108 4.77239 108

    29

  • Table 3: Example 2: First two positive natural frequencies fn =n2

    , n = 2, 3, (note the

    ordering relation in Eq. (54)) and first antiresonant frequency f1A =12

    of the undamaged

    free-free rod, and their values associated to the damaged cases Di, i = 4, 5. Frequency values

    in Hz. Note: Modelling errors are reported in brackets, = 100(fModelfExp.)/fExp..Data Undamaged D4 D5

    Exper. Model Exper. Model Exper. Model

    f2 941.1 941.1 939.3 939.3 901.8 901.8

    (0.00) (0.00) (0.00)

    f3 1879.1 1882.2 1868.3 1872.5 1693.3 1697.6

    (0.16) (0.23) (0.25)

    f1A 468.6 470.6 439.5 468.2 432.9 427.7

    (0.41) (6.52) (-1.19)

    30

  • Table 4: Example 2: Damage parameter estimates xNLd , KNL obtained by the first two

    positive natural frequencies fn =n2

    , n = 2, 3 (note the ordering relation in Eq. (54)).

    Actual solution: xd = 0.550 m; K = 3.507 1010N/m (case D4), K = 1.736 109N/m(case D5). xLd and K

    L are the corresponding estimates obtained by the linearized procedure

    presented in [12].

    Position [m] Spring stiffness [N/m]

    Case xNLd xLd K

    NL KL

    D4 0.253 0.461 8.387 109 2.626 1010

    D5 0.539 0.623 1.679 109 2.075 109

    31

  • Table 5: Example 2: Damage parameter estimates xNLd , KNL obtained by the first positive

    natural frequency f2 =22

    (note the ordering relation in Eq. (54)) and the first antiresonant

    frequency f1A =12

    . Actual solution: xd = 0.550 m; K = 3.507 1010N/m (case D4),K = 1.736 109N/m (case D5). xLd and KL are the corresponding estimates obtained by thelinearized procedure presented in [13].

    Position [m] Spring stiffness [N/m]

    Case xNLd xLd K

    NL KL

    D4 0.143 0.157 2.760 109 3.301 109

    D5 0.598 0.672 1.968 109 2.348 109

    32

  • Figure 1: The function g = g(s).

    33

  • Figure 2: Case ii): unique solution of f(s) = g(s).

    34

  • Figure 3: Case i): unique solution of f(s) = g(s). Left: 32 < 2(< 2); right: (
  • Figure 4: The function g = g(s) of Eq. (80).

    36

  • Figure 5: The function f = f(s) of Eq. (79) and the intersection between f = f(s) and

    g = g(s).

    37

    IntroductionFormulation of the diagnostic problem and some frequency boundCrack identification by the first two natural frequenciesCrack identification by resonant and antiresonant frequency dataApplicationsConclusions