Research article

Reconstruction of shear force in Atomic Force Microscopy from measured displacement of the cone-shaped cantilever tip

  • Received: 13 June 2023 Revised: 01 November 2023 Accepted: 19 January 2024 Published: 25 January 2024
  • We present a new comprehensive mathematical model of the cone-shaped cantilever tip-sample interaction in Atomic Force Microscopy (AFM). The importance of such AFMs with cone-shaped cantilevers can be appreciated when its ability to provide high-resolution information at the nanoscale is recalled. It is an indispensable tool in a wide range of scientific and industrial fields. The interaction of the cone-shaped cantilever tip with the surface of the specimen (sample) is modeled by the damped Euler-Bernoulli beam equation ρA(x)utt +μ(x)ut+(r(x)uxx+κ(x)uxxt)xx=0, (x,t)(0,)×(0,T), subject to the following initial, u(x,0)=0, ut(x,0)=0 and boundary, u(0,t)=0, ux(0,t)=0, (r(x)uxx(x,t)+κ(x)uxxt)x==M(t), ((r(x)uxx+κ(x)uxxt)x)x==g(t) conditions, where M(t):=2hcosθg(t)/π is the moment generated by the transverse shear force g(t). Based on this model, we propose an inversion algorithm for the reconstruction of an unknown shear force in the AFM cantilever. The measured displacement ν(t):=u(,t) is used as additional data for the reconstruction of the shear force g(t). The least square functional J(F)=12u(,)ν2L2(0,T) is introduced and an explicit gradient formula for the Fréchet derivative of the cost functional is derived via the weak solution of the adjoint problem. Additionally, the geometric parameters of the cone-shaped tip are explicitly contained in this formula. This enables us to construct a gradient based numerical algorithm for the reconstructions of the shear force from noise free as well as from random noisy measured output ν(t). Computational experiments show that the proposed algorithm is very fast and robust. This creates the basis for developing a numerical "gadget" for computational experiments with generic AFMs.

    Citation: Alemdar Hasanov, Alexandre Kawano, Onur Baysal. Reconstruction of shear force in Atomic Force Microscopy from measured displacement of the cone-shaped cantilever tip[J]. Mathematics in Engineering, 2024, 6(1): 137-154. doi: 10.3934/mine.2024006

    Related Papers:

    [1] Christopher Chong, Andre Foehr, Efstathios G. Charalampidis, Panayotis G. Kevrekidis, Chiara Daraio . Breathers and other time-periodic solutions in an array of cantilevers decorated with magnets. Mathematics in Engineering, 2019, 1(3): 489-507. doi: 10.3934/mine.2019.3.489
    [2] Jason Howell, Katelynn Huneycutt, Justin T. Webster, Spencer Wilder . A thorough look at the (in)stability of piston-theoretic beams. Mathematics in Engineering, 2019, 1(3): 614-647. doi: 10.3934/mine.2019.3.614
    [3] Jayme Vaz Jr., Edmundo Capelas de Oliveira . On the fractional Kelvin-Voigt oscillator. Mathematics in Engineering, 2022, 4(1): 1-23. doi: 10.3934/mine.2022006
    [4] Maurizio Garrione . Beams with an intermediate pier: Spectral properties, asymmetry and stability. Mathematics in Engineering, 2021, 3(2): 1-21. doi: 10.3934/mine.2021016
    [5] Matteo Fogato . Asymptotic finite-dimensional approximations for a class of extensible elastic systems. Mathematics in Engineering, 2022, 4(4): 1-36. doi: 10.3934/mine.2022025
    [6] Emilia Blåsten, Fedi Zouari, Moez Louati, Mohamed S. Ghidaoui . Blockage detection in networks: The area reconstruction method. Mathematics in Engineering, 2019, 1(4): 849-880. doi: 10.3934/mine.2019.4.849
    [7] Raimondo Penta, Ariel Ramírez-Torres, José Merodio, Reinaldo Rodríguez-Ramos . Effective governing equations for heterogenous porous media subject to inhomogeneous body forces. Mathematics in Engineering, 2021, 3(4): 1-17. doi: 10.3934/mine.2021033
    [8] Jinghong Li, Hongyu Liu, Wing-Yan Tsui, Xianchao Wang . An inverse scattering approach for geometric body generation: a machine learning perspective. Mathematics in Engineering, 2019, 1(4): 800-823. doi: 10.3934/mine.2019.4.800
    [9] Simone Dovetta, Angela Pistoia . Solutions to a cubic Schrödinger system with mixed attractive and repulsive forces in a critical regime. Mathematics in Engineering, 2022, 4(4): 1-21. doi: 10.3934/mine.2022027
    [10] Juan-Carlos Felipe-Navarro, Tomás Sanz-Perela . Semilinear integro-differential equations, Ⅱ: one-dimensional and saddle-shaped solutions to the Allen-Cahn equation. Mathematics in Engineering, 2021, 3(5): 1-36. doi: 10.3934/mine.2021037
  • We present a new comprehensive mathematical model of the cone-shaped cantilever tip-sample interaction in Atomic Force Microscopy (AFM). The importance of such AFMs with cone-shaped cantilevers can be appreciated when its ability to provide high-resolution information at the nanoscale is recalled. It is an indispensable tool in a wide range of scientific and industrial fields. The interaction of the cone-shaped cantilever tip with the surface of the specimen (sample) is modeled by the damped Euler-Bernoulli beam equation ρA(x)utt +μ(x)ut+(r(x)uxx+κ(x)uxxt)xx=0, (x,t)(0,)×(0,T), subject to the following initial, u(x,0)=0, ut(x,0)=0 and boundary, u(0,t)=0, ux(0,t)=0, (r(x)uxx(x,t)+κ(x)uxxt)x==M(t), ((r(x)uxx+κ(x)uxxt)x)x==g(t) conditions, where M(t):=2hcosθg(t)/π is the moment generated by the transverse shear force g(t). Based on this model, we propose an inversion algorithm for the reconstruction of an unknown shear force in the AFM cantilever. The measured displacement ν(t):=u(,t) is used as additional data for the reconstruction of the shear force g(t). The least square functional J(F)=12u(,)ν2L2(0,T) is introduced and an explicit gradient formula for the Fréchet derivative of the cost functional is derived via the weak solution of the adjoint problem. Additionally, the geometric parameters of the cone-shaped tip are explicitly contained in this formula. This enables us to construct a gradient based numerical algorithm for the reconstructions of the shear force from noise free as well as from random noisy measured output ν(t). Computational experiments show that the proposed algorithm is very fast and robust. This creates the basis for developing a numerical "gadget" for computational experiments with generic AFMs.



    In this paper, we study the following inverse boundary value problem of reconstructing the unknown transverse shear force g(t) in

    {ρA(x)utt+μ(x)ut+(r(x)uxx+κ(x)uxxt)xx=0,(x,t)ΩT,u(x,0)=ut(x,0)=0,x(0,),u(0,t)=ux(0,t)=0,(r(x)uxx+κ(x)uxxt)x==M(t),((r(x)uxx+κ(x)uxxt)x)x==g(t),t[0,T], (1.1)

    from the measured displacement

    ν(t):=u(,t),t[0,T]. (1.2)

    Here and below, ΩT:=(0,)×(0,T), with the final time instant T>0, ρA(x):=ρ(x)As(x), where ρ(x)>0 and As(x)>0 are the mass density and the cross-sectional area of the nonhomogeneous cantilever, r(x):=E(x)I(x)>0 is the flexural rigidity (or bending stiffness) of the cantilever while E(x)>0 is the elasticity modulus and I(x)>0 is the moment of inertia. The coefficient κ(x):=cd(x)I(x) represents energy dissipated by friction internal to the beam, while cd>0 is the strain-rate damping coefficient [3]. The external and internal damping mechanisms are given by the terms μ(x)ut and (κ(x)uxxt)xx, respectively. The coefficients μ(x)0 and κ(x)>0 are called the viscous (internal) damping and the strain-rate or Kelvin-Voigt damping coefficients, respectively.

    Problem (1.1) is a mathematical model of the tip-sample interaction of an Atomic Force Microscopy (AFM) cone-shaped cantilever, within the scope of damped Euler-Bernoulli equation. The processing of the sample using an AFM cone-shaped cantilever is schematically depicted in Figure 1. This cantilever, with length >0 is clamped at the left end x=0. The tip-sample interaction is modeled by a vertical reaction force, which is the transverse shear force with the negative sign, that is g(t), at the right end x=, and the moment M(t):=2hcotθg(t)/π, generated by this force, where h,θ>0 are the tip length and half-conic angle, respectively. As a consequence, reconstructing the shear force g(t) in the inverse problem (1.1)-(1.2) actually means determining the moment M(t) as well, given the above relationship between shear force and moment.

    Figure 1.  Schematic diagram of AFM cone-shaped cantilever tip-sample interaction.

    It is well known that micro-cantilever plays a key role in nanomachining process using an AFM which was originally developed to provide surface topography information [4]. Nowadays, AFM can provide high resolution images in different settings including ambient, aqueous and vacuum environments (see [4,17,18,21] and references therein). In standard AFMs, the micro-cantilever is mounted horizontally and the devices are operated in a contact or intermittent-contact mode. The cantilever tip-sample interaction creates a transverse shear force on the tip of the cantilever [13]. Estimation of the unknown shear force signal allows better interpretation and understanding of scan results. Since this force can only be measured indirectly, via a laser based sensor system, starting from 2000 [1], various models and inversion algorithms were developed for reconstruction of the transverse shear force in atomic and dynamic force microscopy, using the measured displacement of the cantilever tip as an available data [2,5,22]. The basis of all of these models is the Euler-Bernoulli beam theory, and the majority of these models are unsuitable for non-homogeneous cantilevers since they require the constant coefficients Euler-Bernoulli equation EI(Yxx+αYt)xx+ρAytt+μyt=0 to ensure that the analytical methods used are suitable. Later, these models, along with the mathematical framework and algorithms for numerical solutions to corresponding reconstruction problems, were developed in [8,10,11,15], for the variable coefficient Euler-Bernoulli equation ρA(x)utt+μ(x)ut+(r(x)uxx+κ(x)uxxt)xx=0.

    In all of the tip-sample interaction models developed in the above cited studies, the tip geometry of the cantilever end was not taken into account; that is, it was assumed that the cantilever is a straight beam. However, the mechanical properties of silicon-based probes are governed mainly by the geometrical specifications of cantilevers. Furthermore, the mechanical features of the cantilever, such as the spring constant and resonance frequency, as well as the tip geometry, can all be changed depending on the application and sample characteristics to get the best sensing [21]. Traditionally, mechanical properties of silicon-based probes are mostly governed by geometrical specifications of cantilevers. In fact, the elastic modulus has a small tuning range, and tip shapes are often restricted to cones or pyramids with set aspect ratios because there are only a few options for etching materials [17,18]. On the other hand, geometric parameters of the cone-shaped tip impact the dynamics of AFM cantilevers. Namely, theoretical and experimental results presented in [19] show that these parameters have a significant influence on the behavior of the cantilever. The results obtained in [19] demonstrate also that the first four frequencies decrease as the slope factor increase.

    The reasons listed above reveal the necessity of taking geometric parameters into account in the mathematical model for studying the dynamics of the AFM cone-shaped cantilever. In this case, the cantilever tip-sample interaction creates not only a transverse shear force but also a bending moment on the tip of the cantilever. The mathematical model introduced mentioned in (1.1) results from this. In this context, we note that a simple mathematical model for the shear force reconstruction problem for the AFM cone-shaped cantilever tip-sample interaction, which takes into consideration the geometry of the tip, has first been proposed in [5]. Namely, this model considers the cutting system as the inverse problem of reconstructing the cutting force Fy(t) in

    {yxxxx+ρAEIytt=0,x(0,L),(x,t)ΩT,y(x,0)=yt(x,0)=0,x(0,),y(0,t)=yx(0,t)=0,yxx(,t)=Fx(t)hEI,yxxx(,t)=Fy(t)EI,t[0,T], (1.3)

    from the available measured displacement

    Y(,t):=y(,t),t[0,T]. (1.4)

    Here, Fx=(2hcotθFy)/π for a cone-shaped cantilever with the half-conic angle θ, and h>0 is the cantilever tip length. In accordance with the accepted terminology of the theory of inverse problems, (1.3)-(1.4) is an inverse problem defined by two Neumann inputs. It is important to emphasize that in the AFM cone-shaped cantilever tip-sample interaction model (1.3) is based on the simplified and constant coefficient Euler-Bernoulli beam equation, without the viscous external (μ(x)ut) and the internal or Kelvin-Voigt ((κ(x)uxxt)xx) damping terms. Thus, in these models, not all physical properties of the cantilever are taken into account. However, the influence of these above mentioned properties on the dynamic behavior of the AFM cantilever is enormous, and needs to be studied carefully [3,19].

    Thus, model (1.1) is a comprehensive mathematical model of AFM cone-shaped cantilever tip-sample interaction that involves variable physical quantities, damping parameters, and, most importantly, geometric parameters of the cantilever tip. Furthermore, in the studies cited above, there is an inverse problem defined with a single Neumann input, while (1.1)-(1.2) is an inverse problem with two Neumann inputs. Similar to [10,15], in this study also the weak solution theory for the direct problem (1.1), and the least square approach combined with the adjoint method for the inverse problem (1.1)-(1.2) is used. However, different from [10,15], in this study both a priori estimates and the gradient formula involves geometric parameters. All of this enables the simulation of the dynamic behavior of the tip-sample interaction of an AFM cone-shaped cantilever and, as we shall see below, the development of numerical algorithms that allow us to investigate of the effects of geometric parameters on this behavior.

    In this paper we propose a new mathematical model of the tip-sample interaction of an AFM cone-shaped cantilever. In the following meanings, this model is a generalization of known mathematical models:

    (a) The model involves all physical and geometric parameters, namely the area A and the moment of inertia I of the beam's cross section, the length , the material's Young modulus E and mass density ρ. It includes also the both damping terms μ and cd, respectively the viscous and strain-rate damping coefficients, as explained when (1.1) was presented.

    (b) The model is based on the variable coefficient Euler-Bernoulli equation.

    (c) The time interval [0,T] required to generate measured output can be arbitrarity small. Because of this, obtaining measurement data {ν(t):t[0,T]}, that is, the displacement at the free tip of the beam during the interval of time [0,T], introduced in (1.2) is made easier, and the function g(t) can be found even with data collected over a relatively short period of time.

    (d) The model allows us to derive a gradient formula that contains the geometric parameters of the cone-shaped cantilever tip explicitly.

    (e) The measured output, which is generated synthetically by an algorithm, contains random noise.

    (f) The inputs in the model may not be smooth enough as the basic conditions (1.2) below show.

    Within the proposed model, we formulate the inverse problem of reconstructing the unknown shear force from measured displacement of the cone-shaped cantilever tip. We provide a detailed mathematical and numerical analysis of the problem. Based on this analysis, we derive an explicit gradient formula for the least square functional. This allows us to construct an effective and fast reconstruction algorithm, as the presented results of computational experiments show.

    The paper is organized as follows. In Section 2 the reconstruction problem (1.1)-(1.2) is formulated as the inverse problem, introducing first the Neumann-to-Dirichlet operator and then the least square functional. Necessary estimates for the weak solution of problem (1.1) are derived in Section 3. Section 4 describes compactness and Lipschitz continuity of the Neumann-to-Dirichlet operator which leads to existence of a quasi-solution of the inverse problem. Fréchet differentiability of the least square functional is proved in Section 5. Based on this result, the gradient formula is derived. Numerical algorithm and its implementations to the direct and inverse problems are presented in Section 6, with the results of computational experiments. Some concluding remarks are given in the final Section 7.

    Consider the inverse boundary value problem of reconstructing the unknown transverse shear force g(t) and the moment M(t) in (1.1) from knowledge of the measured displacement ν(t) introduced in (1.2).

    We assume that the following basic conditions are satisfied:

    {ρA,μ,r,κL(0,),gH1(0,T),g(0)=0,0<ρ0ρA(x)ρ1,0μ0μ(x)μ1,0<r0r(x)r1,0<κ0κ(x)κ1,x(0,). (2.1)

    Introduce the set of admissible shear forces

    G:={gH1(0,T):g(0)=0,gH1(0,T)Cg}, (2.2)

    where Cg>0 is a constant independent on g(t). Denote by u(x,t;g) the solution of the forward problem (1.1) for a given gG, while u(,t;g) in defined as an output. Introduce the Neumann-to-Dirichlet operator:

    {(Ψg)(t):=u(,t;g),t[0,T],Ψ:GH1(0,T)L2(0,T), (2.3)

    defined on the set of admissible shear forces. In view of this operator, we can reformulate the inverse problem as the linear operator equation:

    u(,t;g)=ν(t),t[0,T]. (2.4)

    Since the measured output ν(t) obtained as a result of measurement, it contains random noise. Hence the exact equality between the output u(,t;g) and the measured outputs ν(t) can never be achieved. As a consequence, there can never be an exact solution to the inverse problem (1.1)-(1.2).

    We introduce the least square functional

    J(g):=12u(,;g)ν2L2(0,T),gG,νL2(0,T) (2.5)

    and look for the quasi-solution of the inverse problem (1.1)-(1.2): Find gG such that

    J(g)=inf˜gGJ(˜g). (2.6)

    In the case when M(t)=0, the existence and uniqueness of the weak solution uL2(0,T;V2(0,)), with utL2(0,T;L2(0,)) and uttL2(0,T;H2(0,)) of the initial boundary value problem (1.1) is proved in [15], where

    V2(0,):={vH2(0,):v(0)=v()=0}.

    For the direct problem (1.1) the same results can be proved in the same way. We derive here some a priori estimates for the weak solution which are necessary in the analysis of the inverse problem (1.1)-(1.2).

    Note that in none of the previous studies, including articles [10,15], a priori estimates were made on constants containing geometric parameters. However, such estimates are needed to see the role of these parameters in the behavior of the solution and in the continuity of the least square functional.

    Theorem 1. Assume that the inputs in (1.1) satisfy the basic conditions (2.1). Then the following estimates holds:

    uxx2L(0,T;L2(0,))C21g2L2(0,T),uxx2L2(0,T;L2(0,))C22g2L2(0,T),ut2L2(0,T;L2(0,))r02ρ0C22g2L2(0,T),uxxt2L2(0,T;L2(0,))r04κ0C22g2L2(0,T), (3.1)

    where

    C21=C20(1+C2θ)C2e,C22=C20(1+C2θ)(1+C2e),C2e=exp(T),C20=4ˆ(1+T)r20,C2θ=(2hcotθπ)2,ˆ=+3/3, (3.2)

    and r0,ρ0,κ0>0 are the constants introduced in (2.1)

    Proof. Multiply both sides of Eq (1.1) by 2ut(x,t), integrate it over Ωt:=(0,)×(0,t), t(0,T], and employ the identities

    2t00(r(x)uxx)xxuτdxdτ=2t00[(r(x)uxx)xuτr(x)uxxuxτ]xdxdτ+t00(r(x)u2xx)τdxdτ,t(0,T],2t00(κ(x)uxxτ)xxuτdxdτ=2t00[(κ(x)uxxτ)xuτκ(x)uxxτuxτ]xdxdτ+2t00κ(x)(uxxτ)2dxdτ,t(0,T].

    Applying the integration by parts formula multiple times, using the initial and boundary conditions in (1.1) we obtain the following energy identity:

    0[ρA(x)u2t+r(x)u2xx]dx+2t00μ(x)u2τdxdτ+2t00κ(x)u2xxτdxdτ=2t0M(τ)uxτ(,τ)dτ+2t0g(τ)uτ(,τ)dτ,

    for all t(0,T]. Now, applying the integration by parts to the right hand side integrals, taking into account the condition M(0)=0 and then the ε-inequality 2|ab|εa2+b2/ε, we get:

    2t0M(τ)uxτ(,τ)dτ+2t0g(τ)uτ(,τ)dτε[u2x(,t)+u2(,t)+t0u2x(,τ)dτ+t0u2(,τ)dτ]+1ε[M2(t)+g2(t)+t0(M(τ))2dτ+t0(g(τ))2dτ],t(0,T].

    Use also the auxiliary inequalities ([12, Chapter 11, formula (11.1.30)]):

    u2(,t)330u2xx(x,t)dx,u2x(,t)0u2xx(x,t)dx,uV2(0,),M2(t)TM2L2(0,T),g2(t)Tg2L2(0,T),gH1(0,T),g(0)=0, (3.3)

    for all t[0,T]. Taking these inequalities with the identity

    M2L2(0,T)=C2θg2L2(0,T),

    into account in the energy identity above, we get the following inequality:

    ρ00u2tdx+(r0ˆε)0u2xxdx+2t00μ(x)u2τdxdτ+2t00κ(x)u2xxτdxdτεˆt00u2xxdxdτ+1+Tε(1+C2θ)T0(g(t))2dt,t[0,T],

    where ˆ,Cθ>0 are the constants introduced in (3.2). Choosing here the arbitrary parameter ε>0 from the condition r0ˆε>0 as ε=r0/(2ˆ) we finally obtain the main integral inequality:

    ρ00u2tdx+r020u2xxdx+2t00μ(x)u2τdxdτ+2t00κ(x)u2xxτdxdτr02t00u2xxdxdτ+r02C20(1+C2θ)T0(g(t))2dt,t[0,T], (3.4)

    where C0>0 is the constant introduced in (3.2).

    The first consequence of (3.4) is the inequality

    0u2xxdxt00u2xxdxdτ+C20(1+C2θ)T0(g(t))2dt,t[0,T].

    With the Grönwall-Bellmann inequality ([12, Appendix B, Lemma B.0.1]) this implies:

    0u2xxdxC20(1+C2θ)g2L2(0,T)exp(t),t[0,T]. (3.5)

    Both of the first two estimates in (3.1) are easily derived from this inequality.

    The second consequence of (3.4) is the inequality

    ρ00u2tdxr02t00u2xxdxdτ+r02C20(1+C2θ)T0(g(t))2dt,t[0,T].

    With (3.5) this leads to the third estimate in (3.1).

    The fourth estimate in (3.1) is proved in the same way.

    Remark 1. The results of Theorem 1 are valid, with slightly different from the constants introduced in (3.2), also for the case where the consistency condition g(0)=0 in (2.1) is not met.

    Corollary 1. Assume that conditions of Theorem 1 hold. Then for the H1-norm of the output u(,t;g) the following trace estimate holds:

    u(,;g)2H1(0,T)C23g2L2(0,T),C23=33(C21+r04κ0C22). (3.6)

    Proof. Proof follows from the trace inequalities

    u(,;g)2L2(0,T)33C21g2L2(0,T),ut(,;g)2L2(0,T)33r04κ0C22g2L2(0,T),

    which are the consequence of the first inequality in (3.3) and estimates in (3.1).

    The compactness property is one of the main properties of the input-output operators corresponding to problems, since the ill-posedness of an inverse problem is the result of this property. For the simplified version, with one Neumann input (M(t)=0) and with κ(t)=0, the compactness of the Neumann-to-Dirichlet operator (2.3) is proven in [10] for the regular weak solution. For the model (1.1) we are considering, the regularity condition is not necessary, as we shall see below. That is, this property is also preserved in the case of the weak solution, which shows the role of the Kelvin–Voigt damping coefficient κ(x)>0.

    Lemma 1. Under the basic conditions (2.1), the Neumann-to-Dirichlet operator Ψ:GH1(0,T)L2(0,T) introduced in (2.3) is a linear compact operator.

    Proof. Let {gm}G, m=1,2,..., be a sequence of inputs, bounded in the norm of H1(0,T), according to the definition (2.2) of set of admissible shear forces. Denote by {u(m)(x,t)}, where u(m)(x,t):=u(x,t;gm), the corresponding sequence of weak solutions of the direct problem (1.1). By the estimate (3.6), the sequence of outputs {u(m)(x,t)} is bounded in H1(0,T). Then by the Rellich-Kondrachov compactness theorem (see for instance [6]), ψ is compact operator.

    Lemma 2. Assume that the basic conditions (2.1) hold. Then the Neumann-to-Dirichlet operator is Lipschitz continuous, that is

    Φg1Φg2L2(0,T)L0g1g2L2(0,T),for allg1,g2G, (4.1)

    with here L0=3/3C1>0 is the Lipschitz constant and C1>0 is the constant introduced in (3.2).

    Proof. Let uk(x,t):=u(x,t;gk), k=1,2, be two weak solutions of the direct problem (1.1) corresponding to the inputs g1,g2G. Then the function δu(x,t)=u1(x,t)u2(x,t) solves the problem

    {ρA(x)δutt+μ(x)δut+(r(x)δuxx+κ(x)δuxxt)xx=0,(x,t)ΩT,δu(x,0)=δut(x,0)=0,x(0,),δu(0,t)=δux(0,t)=0,(r(x)δuxx+κ(x)δuxxt)x==δM(t),((r(x)δuxx+κ(x)δuxxt)x)x==δg(t),t[0,T], (4.2)

    subject to the inputs δM(t)=C2θδg(t) and δg(t)=g1(t)g2(t). By the definition (2.3) of the input-output operator we have:

    Φg1Φg22L2(0,T)=δu(,)2L2(0,T).

    In view of the first inequality in (3.3) and the second estimate in (3.1) applied to the weak solution δu(x,t) of problem (4.2) we deduce that

    δu(,)2L2(0,T)33C21δg2L2(0,T). (4.3)

    This leads to (4.1).

    The Lipschitz continuity of the Neumann-to-Dirichlet operator leads to the Lipschitz continuity of the least square functional introduced in (2.5), and this, in turn, leads to the existence of the quasi-solution of the inverse problem (1.1)-(1.2), by Theorem 6.5.2 [12].

    Theorem 2. Assume that the inputs in (1.1) satisfy the basic conditions (2.1). Suppose that the measured output ν(t) belongs to L2(0,T). Then there exists a quasi-solution of the inverse problem (1.1)-(1.2) in the set of admissible shear forces G.

    For g,g+δgG we find the increment δJ(g):=J(g+δg)J(g) of the least square functional introduced in (2.5) is

    δJ(g)=0[u(,t;g)ν(t)]δu(,t)dt+120(δu(,t))2dt, (5.1)

    where δu(x,t) is the solution of the sensitivity problem (4.2).

    Multiplying both sides of Eq (4.2) by arbitrary function ϕ(x,t), integrating it over (0,T) and applying the integration by parts formula multiple times, we obtain:

    T00[ρA(x)ϕttμ(x)ϕt+(r(x)ϕxxκ(x)ϕxxt)xx]δudxdt+0[ρA(x)δutϕρA(x)δuϕt+μ(x)δuϕ+κ(x)δuxxϕxx]t=Tt=0dx+T0[(r(x)δuxx)xϕr(x)δuxxϕx+r(x)δuxϕxxδu(r(x)ϕxx)x]x=x=0dt+T0[(κ(x)δuxxt)xϕκ(x)δuxxtϕxκ(x)ϕxxtδux+(κ(x)ϕxxt)xδu]x=x=0dt=0. (5.2)

    We require now ϕ(x,t) solves the well-posed backward problem

    {ρA(x)ϕttμ(x)ϕt+(r(x)ϕxxκ(x)ϕxxt)xx=0,(x,t)ΩT,ϕ(x,T)=0,ϕt(x,T)=0,x(0,),ϕ(0,t)=ϕx(0,t)=0,(r(x)ϕxxκ(x)ϕxxt)x==0,((r(x)ϕxxκ(x)ϕxxt)x)x==p(t),t[0,T]. (5.3)

    The control function p(t) here is the arbitrary Neumann input and is specified below.

    In view of the initial, final and boundary conditions in (1.1) and (4.2) we deduce from (5.3) the following integral relationship:

    T0p(t)δu(,t)dt=T0[ϕ(,t)+C2θϕx(,t)]δg(t)dt, (5.4)

    where Cθ>0 is the constant introduced in (3.2).

    Taking into account the increment formula (5.1) we choose the control function p(t) as follows:

    p(t)=u(,t;g)ν(t),t[0,T]. (5.5)

    The backward problem with this input, i.e., the problem

    {ρA(x)ϕttμ(x)ϕt+(r(x)ϕxxκ(x)ϕxxt)xx=0,(x,t)ΩT,ϕ(x,T)=0,ϕt(x,T)=0,x(0,),ϕ(0,t)=ϕx(0,t)=0,(r(x)ϕxxκ(x)ϕxxt)x==0,((r(x)ϕxxκ(x)ϕxxt)x)x==u(,t;g)ν(t),t[0,T], (5.6)

    is called the adjoint problem corresponding to the inverse problem (1.1)-(1.2).

    Substituting (5.5) into (5.4) we obtain the input-output relationship:

    T0[u(,t;g)ν(t)]δu(,t)dt=T0[ϕ(,t)+C2θϕx(,t)]δg(t)dt, (5.7)

    which contains the output u(,t;g) and the measured output ν(t). Comparing (5.1) and (5.7) we deduce that

    δJ(g)=T0[ϕ(,t)+C2θϕx(,t)]δg(t)dt+120(δu(,t))2dt,gG. (5.8)

    Theorem 3. Assume that the inputs in (1.1) satisfy the basic conditions (2.1). Suppose, in addition, the measured output ν(t) belongs to H1(0,T). Then the least square functional introduced in (2.5) is Fréchet differentiable. Furthermore, for the Fréchet gradient of this functional the following gradient formula holds:

    J(g)(t)=ϕ(,t)+C2θϕx(,t),t(0,T),gG. (5.9)

    Proof. Applying the first inequality in (3.3) with the second estimate in (3.1) to the weak solution δu(x,t) of problem (4.2) we conclude that the second right hand side integral in (5.8) of the order O(gL2(0,T)). This means that the least square functional is Fréchet differentiable.

    The constant Cθ=(2hcotθ)/π in formula (5.9) shows that the tip length h>0 and half-conic angle θ>0, which are the fundamental geometric properties of the cone-shaped cantilever, are now involved in the gradient formula, in contrast to all previous gradient formulas. Furthermore, this formula (5.9) expressed in terms of the weak solution ϕ(x,t) of the adjoint problem (5.6) forms the basis of the algorithm for numerical solving the inverse problem (1.1)-(1.2).

    In this section, a detailed description of an efficient numerical method is presented to solve the inverse problem (1.1)-(1.2). This process has several steps and each of them should be considered carefully due to the sensitivity of the identification process. First, measured data ν(x):=u(,t) is generated by solving the direct problem. It is critical to keep the error as low as possible in this step. This requires a successful algorithm for the solution of the direct problem (1.1). Due to the effectiveness of the method of lines approach used in our several published previous studies [8,9,10,11] on an optimized mesh, an improved version of this method is employed here.

    Basically, the MOL is based on the principle of independent discretization of space and time variables. More specifically, a semi-analytical structure is obtained by expressing the variational formulation in finite dimensional space denoted by Vh. The method here is a finite element approximation with cubic Hermite basis functions which ensures continuity of both deflection and slope throughout the beam. These shape functions is defined on uniformly discretizing spatial domain 0=x0<x1<<xM= (where h=/M). Formally, the solution Uh(t):=uh(,t)u(,t) satisfies the following semi-discrete version of the variational formulation of (1.1).

    For all t(0,T], find Uh(t)VhV2(0,) such that vhVh

    {(ρA()Uh(t),vh)+(μ()Uh(t),vh)+ar(Uh(t),vh)+aκ(Uh(t),vh)=0,Uh(0)=0,Uh(0)=0. (6.1)

    Here the symmetric bilinear functional aψ:H2(0,)×H2(0,)R is defined, for ψL2(0,), by

    aψ(u,v):=(ψuxx,vxx).

    The next discretization step is performed for temporal derivatives. At this level the second order system of ODE in (6.1) can be approximately solved by using any temporal finite difference method. It is crucial that the approach to be used here have to be practical, fast and stable. These requirements can be met through the following second order backward finite difference approximations of Uh and Uh with uniform temporal discretization 0=t0<t1<<tN=T (where τ=T/N).

    Uh(tj)ττUjh:=2Ujh5Uj1h+4Uj2hUj3hτ2,U(tj)τUjh:=3Ujh4Uj1h+Uj2h2τ,

    The full-discrete algebraic systems of equations are obtained by substituting these difference expressions with Uh(t) and Uh(t) in (6.1). Solutions of the resulted equations are provided desired approximations Ujhu(x,tj) for j=0:N. Note that for j=1,2, the necessary a priori approximations can be obtained by combining the ghost point technique within the central difference scheme.

    Finally, several numerical tests in [14] are compared to determine the effective values of the pair (M,N) and optimized with the ratio h/τ142.

    The explicit gradient formula in (5.9) is very important in determining the minimizer of the least square functional (2.3) for any unconstrained optimization techniques. Here we use CGA, one of the most suitable and stable one. It is known that this method is based on the conjugate directions and these directions are determined by the solution of the adjoint problem (5.6). This requires the MOL technique at each iteration step. Although CGA is a self-stabilized method, the quality of the reconstruction process also depends on the success of solving both adjoint and direct problem. The details of the CGA is as follows.

    ● From g(i)(t), calculate the decent direction

    p(i)(t)=J(g(i))2L2(0,T)J(g(i1))2L2(0,T)p(i1)(t)J(g(i))(t).

    ● Define the next iteration g(i+1)(t)=g(i)(t)+α(i)p(i)(t). Here αi solution of the minimization problem

    J(g(i)(t)+α(i)p(i)(t))=minα>0J(g(i)(t)+αp(i)(t));

    and has the following explicit form,

    α(i)=||J(g(i))||2L2(0,T)||u(,,p(i))||2L2(0,T).

    ● If the following stopping condition based on Mozorov's discrepancy principle holds,

    u(,;g(i))νγL2(0,T)εγ<u(,;g(i1))νγL2(0,T)

    for known parameter ε>0, stop the iteration; otherwise, repeat the process by taking g(i)(t):=g(i+1)(t).

    For the first iteration, an arbitrary choice of g(0)(t) can be made, but if there is no prior knowledge it is better to choose g(0)(t)=0 and p(i)(t):=J(g(0))(t). As a note, the first iteration has no significant effect to the success of the algorithm.

    Here a standard method is used for the derivation of synthetic noise with a given noise level γ>0. In deed, the formula νγ(tj):=ν(tj)+γνL2(0,T)Rj for j=1,,N generates measured noisy data. The vector R has M random numbers array normally distributed with mean 0 and standard deviation σ=1.

    In CGA steps, both Fréchet derivative J(g) and L2(0,T) norms are computed by Simpson's numerical integration while in MOL algorithm, a three-point Gauss quadrature rule is employed for all computation on each element.

    In the reconstruction process, we work on two different test problems. One of them is based on engineering applications (realistic parameters), while the other one is preferred to test the applicability of the method.

    It is a general approach to use error analysis when comparing the quality of the methods. In the literature, two quantities are frequently used. These are Convergence and Accuracy Errors as follows.

    e(i;g;γ)=ν(;g(i))νγL2(0,T)ConvergenceError

    and

    E(i;g;γ)=gg(i)L2(0,T)AccuracyError.

    As can be seen from their definitions, the Accuracy Error determines the success of the reconstruction. On the other hand, especially in the case of noisy data, the stop criterion is very crucial to prevent divergence of the approximation and it is completely related to Convergence Error. Therefore, these quantities should be evaluated together to analyze the process.

    For the first test problem, the parameters are selected in accordance with real engineering applications and are based on those proposed in [7,16,20]. We take a beam of length of 200nm and observe it for a time interval of 103s. After a simple change of variables, to re-scale the problem so that the length of the beam and the time observation length interval become =1 and T=1 respectively, the numerical values adopted for this study become as follows:

    ρA(x)=1.864×107kg/nm,μ(x)=8.16×106kg  s1/nm,r(x)=2.265×103kg  nm3/s2,κ(x)=3.5875×105kg  nm3/s,

    and domain parameters are =1 and T=1, both non-dimensional. As for the tip length, it usually ranges from 5 nm to 50 nm [13]. After the re-scaling for doing our numerical simulations, we take as a reasonable value h=0.2 (non-dimensional).

    We tested the performance of the algorithm for the unknown shear force g(t)=t sin(7πt/2) with θπ=(cot(π/36))/(5π). The graph on the left in Figure 2 shows noisy free as well as random noisy output data with the noise levels γ=3% and 6%. Then unknown target g(t) is identified by using each of these data. Results can be seen on the right in Figure 2.

    Figure 2.  Synthetic noise free and noisy output data (left), reconstruction of smooth shear force g(t)=t sin(7πt/2) (right).

    Figure 3 reveals the general characteristics of an iteration. Especially the rapid deterioration in the Accuracy Error indicates that the sensitivity of the stopping which is directly determined by the Convergence Error. In case this balance is not determined appropriately, the success of the construction process can be adversely affected.

    Figure 3.  Convergence error (left) and accuracy error (right) for g(t)=t sin(7πt/2).

    The second computational experiment aims to test the accuracy of CGA regardless of the realizability of the parameters. For this goal, reconstruction of the following discontinues target source g(t) is studied under high noise levels.

    g(t)=12H(1/2t)+sin(3πt)H(t1/2)

    and

    θπ=2cot(π/4)π.

    Here H(x) is the Heaviside step function. Moreover, all problem parameters are imposed as non-constant case as follows with unit domain parameters =1 and T=1.

    ρA(x)=exp(x),μ(x)=sin(πx),
    r(x)=2+x2,κ(x)=1+exp(x).

    Synthetic noise free and noisy data are plotted in Figure 4 (left) with noise levels γ=5% and 10%. Then CGA is applied for identification of the temporal function g(t) and results are illustrated in Figure 4 (right). Here, due to the effect of high noise levels and discontinuity on g(t), non-physical distortions are naturally observed in the reconstruction.

    Figure 4.  Synthetic noise free and noisy output data (left), reconstruction of non-smooth shear force g(t)=12H(1/2t)+sin(3πt)H(t1/2) (right).

    Convergence and Accuracy Errors are plotted in Figure 5 on the left and on the right, respectively. Similar behavior of these error quantities examined in the first problem is also observed in this second experiment.

    Figure 5.  Convergence error (left) and accuracy error (right) for g(t)=12H(1/2t)+sin(3πt)H(t1/2).

    The results of the two experiments presented here show that CGA is effective and successful for the solution of the inverse problem under consideration, provided that certain sensitivities are taken into account. Nevertheless, the algorithm may need to be improved for further applications. Especially in realistic cases, it is required to choose a small final time such as T=103 for stable calculations using the Finite Element Method. Since the method suggested here is just a preliminary numerical study of the inverse problem related to Atomic Force Microscopy, we only aimed to present the general principles.

    In this study, a novel mathematical model of tip-sample processing with AFM cone-shaped cantilever is proposed. Compared to the models known in the literature, this model is a fairly advanced model, and takes into account not only both viscous and internal damping parameters, but also the geometric parameters of the cone-shaped tip. A detailed mathematical analysis of the model has been carried out. A new gradient formula for the Fréchet derivative of the least square functional that specifically takes the geometric parameters into account is derived. This allows us to construct the fast Conjugate Gradient Algorithm for the numerical reconstruction of the shear force. Numerical experiments carried out with real physical and geometric parameters show the high accuracy of the algorithm.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The research of the first and second authors have been supported by FAPESP, through the Visiting Researcher Program, proc. 2021/08936-1, in Escola Politécnica, University of São Paulo, Brazil, during the period November 02 - December 18, 2022. We thank the referees for their careful reading of the manuscript and their insightful comments, which helped to improve the revised manuscript significantly.

    All authors declare no conflicts of interest in this paper.



    [1] M. Antognozzi, Investigation of the shear force contrast mechanism in transverse dynamic force microscopy, Ph.D. Thesis, University of Bristol, UK, 2000.
    [2] M. Antognozzi, D. R. Binger, A. D. L. Humphris, P. J. James, M. J. Miles, Modeling of cylindrically tapered cantilevers for transverse dynamic force microscopy (TDFM), Ultramicroscopy, 86 (2001), 223–232. https://doi.org/10.1016/S0304-3991(00)00087-5 doi: 10.1016/S0304-3991(00)00087-5
    [3] H. T. Banks, D. J. Inman, On damping mechanisms in beams, J. Appl. Mech., 58 (1991), 716–723. https://doi.org/10.1115/1.2897253 doi: 10.1115/1.2897253
    [4] G. Binnig, C. F. Quate, C. Gerber, Atomic force microscopy, Phys. Rev. Lett., 56 (1986), 930. https://doi.org/10.1103/PhysRevLett.56.930 doi: 10.1103/PhysRevLett.56.930
    [5] W. J. Chang, T. H. Fang, C. I. Weng, Inverse determination of the cutting force on nanoscale processing using atomic force microscopy, Nanotechnology, 15 (2004), 427–430. https://doi.org/10.1088/0957-4484/15/5/004 doi: 10.1088/0957-4484/15/5/004
    [6] L. C. Evans, Partial differential equations, 2 Eds., Providence: American Mathematical Society, 2010.
    [7] B. Geist, J. R. McLaughlin, The effect of structural damping on nodes for the Euler-Bernoulli beam: a specific case study, Appl. Math. Lett., 7 (1994), 51–55. https://doi.org/10.1016/0893-9659(94)90112-0 doi: 10.1016/0893-9659(94)90112-0
    [8] A. Hasanov, O. Baysal, Identification of an unknown spatial load distribution in a vibrating cantilever beam from final overdetermination, J. Inverse Ill-Posed Probl., 23 (2015), 85–102. https://doi.org/10.1515/jiip-2014-0010 doi: 10.1515/jiip-2014-0010
    [9] A. Hasanov, O. Baysal, Identification of unknown temporal and spatial load distributions in a vibrating Euler-Bernoulli beam from Dirichlet boundary measured data, Automatica, 71 (2016), 106–117. https://doi.org/10.1016/j.automatica.2016.04.034 doi: 10.1016/j.automatica.2016.04.034
    [10] A. Hasanov, O. Baysal, C. Sebu, Identification of an unknown shear force in the Euler-Bernoulli cantilever beam from measured boundary deflection, Inverse Probl., 35 (2019), 115008. https://doi.org/10.1088/1361-6420/ab2a34 doi: 10.1088/1361-6420/ab2a34
    [11] A. Hasanov, O. Baysal, H. Itou, Identification of an unknown shear force in a cantilever Euler-Bernoulli beam from measured boundary bending moment, J. Inverse Ill-posed Probl., 27 (2019), 859–876. https://doi.org/10.1515/jiip-2019-0020 doi: 10.1515/jiip-2019-0020
    [12] A. H. Hasanoglu, A. G. Romanov, Introduction to inverse problems for differential equations, 2 Eds., New York: Springer, 2021.
    [13] G. Haugstad, Atomic force microscopy: understanding basic modes and advanced applications, Wiley, 2012.
    [14] O. Baysal, A. Hasanov, K. Sakthivel, Determination of unknown shear force in transverse dynamic force microscopy from measured final data, J. Inverse Ill-posed Probl., 2023. https://doi.org/10.1515/jiip-2023-0021 doi: 10.1515/jiip-2023-0021
    [15] S. Kumarasamy, A. Hasanov, A. Dileep, Inverse problems of identifying the unknown transverse shear force in the Euler-Bernoulli beam with Kelvin-Voigt damping, J. Inverse Ill-posed Probl., 2023. https://doi.org/10.1515/jiip-2022-0053 doi: 10.1515/jiip-2022-0053
    [16] H. L. Lee, W. J. Chang, Effects of damping on the vibration frequency of atomic force microscope cantilevers using the Timoshenko beam model, Jpn. J. Appl. Phys., 48 (2009), 065005. https://doi.org/10.1143/JJAP.48.065005 doi: 10.1143/JJAP.48.065005
    [17] J. S. Lee, J. Song, S. O. Kim, S. Kim, W. Lee, J. A. Jackman, et al., Multifunctional hydrogel nano-probes for atomic force microscopy, Nat. Commun., 7 (2016), 11566. https://doi.org/10.1038/ncomms11566 doi: 10.1038/ncomms11566
    [18] M. P. Scherer, G. Frank, A. W. Gummer, Experimental determination of the mechanical impedance of atomic force microscopy cantilevers in fluids up to 70 kHz, J. Appl. Phys., 88 (2000), 2912–2920. https://doi.org/10.1063/1.1287522 doi: 10.1063/1.1287522
    [19] K. Shen, D. C. Hurley, J. A. Turner, Dynamic behaviour of dagger-shaped cantilevers for atomic force microscopy, Nanotechnology, 15 (2004), 1582–1589. https://doi.org/10.1088/0957-4484/15/11/036 doi: 10.1088/0957-4484/15/11/036
    [20] J. A. Turner, J. S. Wiehn, Sensitivity of flexural and torsional vibration modes of atomic force microscope cantilevers to surface stiffness variations, Nanotechnology, 12 (2001), 322–330. https://doi.org/10.1088/0957-4484/12/3/321 doi: 10.1088/0957-4484/12/3/321
    [21] N. R. Wilson, J. V. Macpherson, Carbon nanotube tips for atomic force microscopy, Nature Nanotechnology, 4 (2009), 483–491. https://doi.org/10.1038/nnano.2009.154 doi: 10.1038/nnano.2009.154
    [22] K. Zhang, T. Nguyen, C. Edwards, M. Antognozzi, M. Miles, G. Herrmann, Real-time force reconstruction in a transverse dynamic force microscope, IEEE Trans. Ind. Electron., 69 (2022), 11403–11413. https://doi.org/10.1109/TIE.2021.3120487 doi: 10.1109/TIE.2021.3120487
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1112) PDF downloads(146) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog