Acta Polytechnica DOI:10.14311/AP.2020.60.0502 Acta Polytechnica 60(6):502–511, 2020 © Czech Technical University in Prague, 2020 available online at https://ojs.cvut.cz/ojs/index.php/ap NEWMARK ALGORITHM FOR DYNAMIC ANALYSIS WITH MAXWELL CHAIN MODEL Jaroslav Schmidt∗, Tomáš Janda, Alena Zemanová, Jan Zeman, Michal Šejnoha Czech Technical University in Prague, Faculty of Civil Engineering, Department of Mechanics, Thákurova 7, 166 29 Praha, Czech Republic ∗ corresponding author: Jaroslav.Schmidt@cvut.cz Abstract. This paper investigates a time-stepping procedure of the Newmark type for dynamic analyses of viscoelastic structures characterized by a generalized Maxwell model. We depart from a scheme developed for a three-parameter model by Hatada et al. [1], which we extend to a generic Maxwell chain and demonstrate that the resulting algorithm can be derived from a suitably discretized Hamilton variational principle. This variational structure manifests itself in an excellent stability and a low artificial damping of the integrator, as we confirm with a mass-spring-dashpot example. After a straightforward generalization to distributed systems, the integrator may find use in, e.g., fracture simulations of laminated glass units, once combined with variationally-based fracture models. Keywords: Newmark method, Maxwell chain model, variational integrators. 1. Introduction The motivation of this work comes from the field of dy- namics of laminated glass structures. These sandwich structures consist of multiple glass layers connected with transparent polymer interlayers. Combining stiff, brittle glass with compliant viscoelastic polymers im- proves structural safety, but the through-thickness heterogeneity renders mechanics of laminated glass structures intricate, e.g. [2]. In particular, time- and temperature-dependent interlayer properties must be accounted for even in quasi-static analyses, e.g., [3–6] and references therein. Earlier studies [7–10] have shown that the response of commonly used interlayer materials can be captured well with the Maxwell chain model combined with the time-temperature superposition principle. Because the viscoelastic model concurrently predicts material damping, vibrations of laminated glass structures can be described more accurately than in conventional structural analyses that mostly employ the Rayleigh damping, e.g. [11, Section 12.5]. This added value has been addressed in detail for free vibrations of lami- nated glass units, e.g. [12–14]; an extension towards the response under general dynamic loads requires the development of dedicated time-stepping schemes that are in the focus of the current work. Related work. Dynamics of viscoelastic solids de- scribed by the Maxwell chain model leads to the sys- tem of initial value problems coupling the equation of motion with the local evolution of constitutive vari- ables, see Section 2.1 for illustration. Because numer- ically integrating the full system would be costly, we will follow an alternative approach in which only the equations of motion are solved approximately, whereas the evolutionary constitutive equations are resolved in the closed form, leading to an inexpensive update for- mulas for internal variables entering the equations of motion. This approach has been pioneered for quasi- static problems by Zienkiewicz et al. [15]; see also [16, Section 5.2] for a comprehensive review. To the best of our knowledge, Hatada et al. [1] were the only ones who used this strategy in dynamics, although no refer- ence to the original work [15] was made. In particular, they developed a Newmark-type [17] algorithm for the three-parameter Maxwell model and used it to predict the response of planar frames to earthquake loading. Novelty. Our work further develops the contribu- tion [1] in three aspects. First, in Section 2.2, we present a compact derivation of the Newmark scheme for a generic Maxwell chain, closely following the orig- inal exposition [15]. Second, in Section 3, we show that the algorithm can be interpreted as a variational integrator [18], in the sense that it can be derived from the Hamilton variational principle combined with a suitable time discretization. The variational structure provides the integrator with a good numerical stability and low numerical dissipation, as demonstrated in Sec- tion 4.1 with selected examples. Moreover, the scheme can easily be combined with variational approaches to fracture, e.g., [19–21], which is of independent inter- est when simulating the behaviour of laminated glass under impact loads. Third, in Section 4.2, we outline how to extend the algorithm to a continuum formu- lation and complement the theoretical considerations with an illustrative 3D finite element simulation. Notation. We employ the conventional notation through the text, in which scalar quantities are de- noted by a plain font, whereas bold-face letters in- 502 https://doi.org/10.14311/AP.2020.60.0502 https://ojs.cvut.cz/ojs/index.php/ap vol. 60 no. 6/2020 Newmark algorithm for dynamic analysis of viscoelastic materials k∞ ηPη3η2η1 k1 k2 k3 kP m F(t) r(t) re,1(t) rv,1(t) Figure 1. Scheme of the single-degree-of-freedom viscoelastic dynamic problem. dicate vectors or higher-order tensors. Additional nomenclature is introduced when needed. 2. Newmark method In this section, we analyse a single degree of freedom (SDOF) model of a mass supported with a Maxwell chain, consisting of the parallel connection of an elastic spring and multiple spring-dashpot cells, see Figure 1 for illustration and, e.g., [16, Section A] for further details. In particular, in Section 2.1, we review the equations of motion, which we subsequently discretize with the average acceleration version of the Newmark method [17] in Section 2.2. 2.1. Governing equations As follows from the scheme in Figure 1, the prob- lem under consideration is specified with the time- dependent load F(t), the particle mass m and the Maxwell chain model parameters: stiffness of the elas- tic spring k∞, spring stiffness kp and damper viscosity ηp of the p-th Maxwell cell; P stands for the number of Maxwell cells. Equilibrium of the forces acting on the mass requires mr̈(t) + k∞r(t) + P∑ p=1 fp(t) = F(t), (1) where r denotes the displacement of the mass, r̈ its acceleration, and fp the restoring force of the p-th cell. For the p-th Maxwell cell, the displacement r splits into an elastic part of the spring re,p and a viscous part of the damper rv,p: r(t) = re,p(t) + rv,p(t); (2) recall Figure 1. The restoring force of the p-th Maxwell cell satisfies fp(t) = kpre,p(t) = ηpṙv,p(t), (3) because of the serial arrangement of the spring and the damper in the cell. Differentiating (2) with respect to time and using (3), we obtain ḟp(t) kp + fp(t) ηp = ṙ(t). (4) In summary, the motion of the SDOF model is de- scribed with the coupled system of (P + 1) ordinary differential equations (ODEs) (1) and (4), comple- mented with the initial conditions r(0) = r0, ṙ(0) = v0, fp(0) = fp,0, (5) where r0 and v0 stand for the initial mass displacement and velocity, respectively, fp,0 is the initial force in the p-th Maxwell cell, and p = 1, . . . ,P . 2.2. Discretization The time interval of interest 〈0,T〉 is divided into (N + 1) time instants 0 = t0 < t1 < t2 < · · · < tN−1 < tN = tmax; for notational simplicity, we as- sume equidistant partitioning of the constant time step ∆t = ti+1 − ti. Considering the Newmark integration scheme [17] with an average acceleration (r̈i + r̈i+1)/2 on the time interval 〈ti, ti+1〉, the velocity and displacement within the interval vary as ṙ(ti + τ) = ṙi + 12 (r̈i + r̈i+1) τ, (6a) r(ti + τ) = ri + ṙiτ + 14 (r̈ + r̈i+1) τ 2, (6b) where τ ∈ 〈0, ∆t〉 is the local time variable within the interval and •i abbreviates •(ti) to render the notation compact. Substituting (6a) into (4) reveals that the evolution of the restoring force fp satisfies ḟp(ti + τ) kp + fp(ti + τ) ηp = ṙi + 12 (r̈i + r̈i+1)τ. (7) This Cauchy problem with the initial condition fp(ti) = fp,i has the solution fp(ti + τ) =( fp,i −ηpṙi + η2p 2kp (r̈i + r̈i+1) ) exp ( − kp ηp τ ) +ηpṙi + ηpτkp −η2p 2kp (r̈i + r̈i+1). (8) Thus, at the end of the time interval with τ = ∆t, we have fp,i+1 = Apfp,i + kpθ̂pṙi + Bp (r̈i + r̈i+1) , (9) where θ̂p = ηp kp ( 1 − exp ( − kp ηp ∆t )) (10) denotes the effective relaxation time of the p-th cell and the auxiliary factors are given by Ap = ( 1 − kp ηp θ̂p ) , Bp = 12ηp ( ∆t− θ̂p ) . (11) 503 J. Schmidt, T. Janda, A. Zemanová et al. Acta Polytechnica Finally, substituting equations (6b) and (9) into (1) expressed at ti+1 = ti + ∆t and rearranging the terms yields ( m + 14k∞∆t 2 + P∑ p=1 Bp ) r̈i+1 = Fi+1 − P∑ p=1 Apfp,i −k∞ri − ( k∞∆t + P∑ p=1 kpθ̂p ) ṙi − ( 1 4k∞∆t 2 + P∑ p=1 Bp ) r̈i. (12) After solving Eq. (12) for the acceleration r̈i+1, we update velocity ṙi+1, displacement ri+1, and restoring forces fp,i according to equations (6a), (6b), and (9), respectively, and proceed to the next time interval. Note that the initial acceleration r̈0, needed in the first step of the algorithm, is set to r̈0 = 1 m ( F(0) −k∞r0 − P∑ p=1 fp,0 ) , (13) according to the equilibrium (1) and initial (5) condi- tions. 3. Variational integrators Having derived the Newmark viscoelastic algorithm by conventional means, we now demonstrate its varia- tional structure, by adapting the general arguments on variational integrators by Kane et al. [18] to the current setting. We will proceed in four steps. In Section 3.1, we show that the governing equations from Section 2.1 fol- low from the Euler-Lagrange (E-L) equations of a suit- ably defined energy functional. Its discretization then provides the governing equations of the corresponding variational integrator introduced in Section 3.2. In Section 3.3, we demonstrate the equivalence of the integrator to the Newmark algorithm from Section 2.2. In the last step, Section 3.4, we comment on the en- ergy conservation properties of the time integration scheme. 3.1. Variational framework We postulate that the trajectory q : (0,T) → Q of a constrained dissipative mechanical system in the state space Q is given by the generalised Euler-Lagrange equations1, e.g., [22, Section 1.3] ∂R(q̇(t)) ∂q̇ = ∂L(t,q(t), q̇(t),λ(t)) ∂q − ∂ ∂t ∂L(t,q(t), q̇(t),λ(t)) ∂q̇ , (14a) 0 = α(q(t)). (14b) 1The term generalised corresponds to the appearance of the non-conservative forces in Eq. (14a). Here, R stands for the dissipation potential, L for the Lagrangian of the problem, and λ : (0,T) → Λ denotes the Lagrange multipliers associated with the kinematic constraint function α. Besides, these equations correspond to the stationarity conditions of the action functional S(q̂,λ̂) = ∫ T 0 L ( t, q̂(t), ̂̇q(t),λ̂(t)) dt (15) perturbed by the dissipative forces ∂q̇R. Note that the hat symbol in (15) now distinguishes the test quantities from the true trajectories defined with (14). For the problem from Figure 1, the state variable q̂(t) = [ r̂(t), {r̂(t)e,p}Pp=1, {r̂(t)v,p}Pp=1 ] T (16) collects the total displacement and the displacements of both components of each Maxwell cell; the state space Q = R2P+1. The Lagrangian has the standard form L(t, q̂, ̂̇q,λ̂) = K(̂̇q) −E(q̂) + fext(t)Tq̂ + λ̂Tα(q̂) (17) involving the kinetic energy K, potential energy of deformation E, and external forces fext given by K(̂̇q) = 12m(̂̇r)2, (18a) E(q̂) = 12k∞r̂ 2 + 12 P∑ p=1 kp(r̂e,p)2, (18b) fext(t) = [ F(t), 01×2P ] T. (18c) The kinematical constraints take the form αp(q̂) = r̂e,p + r̂v,p − r̂, p = 1, . . . ,P ; (19) the space of the Lagrange multipliers Λ then becomes RP . The last component of the general framework (14) is provided by the dissipation potential R(̂̇q) = 12 P∑ p=1 ηp(̂̇rv,p)2 (20) involving solely the viscous displacements of all cells. In this setting, the E-L equation (14a) represents the system of (1 + 2P) optimality conditions. The first one, corresponding to the total displacement r, takes the form mr̈(t) + k∞r(t) + P∑ p=1 λp(t) = F(t), (21) while the remaining 2P conditions read as λp(t) = kpre,p(t), λp(t) = ηpṙv,p(t), (22) with p = 1, . . . ,P . It is thus evident that the multipli- ers λp play a role of the viscous force fp and, because the optimality (14b) and compatibility (4) conditions coincide, the current setting is equivalent to the one of Section 2.1. 504 vol. 60 no. 6/2020 Newmark algorithm for dynamic analysis of viscoelastic materials 3.2. Discretization Recall that the incremental algorithm of Section 2.2 relies on the discretization of the total displacements, from which the evolution of cell-related variables re,p, rv,p, and fp follows in the closed form. To mimic this structure, only the total displacements r will be determined from the discrete (non-dissipative) E- L equations, whereas the remaining quantities are determined from the non-discretized optimality condi- tions (22) and (14b). To this goal, we consider the same discretization of the time interval 〈0,T〉 as in Section 2.2 and introduce the discretized action functional S(r̂,λ̂) ≈Sd ( {r̂i}Ni=0,{λ̂i} N i=0 ) = ∆t N−1∑ i=0 Ld ( r̂i, r̂i+1,λ̂i,λ̂i+1 ) , (23) with the discrete Lagrangian given by [18, Eq. (2)] Ld ( r̂i, r̂i+1,λ̂i,λ̂i+1 ) = 12m ( r̂i+1 − r̂i ∆t )2 (24) − 12k∞ ( r̂i+1 + r̂i 2 )2 + 14 (r̂i + r̂i+1)(Fi + Fi+1) − 14 P∑ p=1 (r̂i + r̂i+1)(λ̂p,i + λ̂p,i+1). The stationarity conditions at time ti, ∂Sd/∂ri = 0, with i = 1, . . . ,N − 1,2 read as 0 = ∂Ld(ri−1,ri,λi−1,λi) ∂ri + ∂Ld(ri,ri+1,λi,λi+1) ∂ri (25) which delivers the governing equations of the varia- tional integrator in the form3 m ri+1 − 2ri + ri−1 ∆t2 + 14k∞(ri+1 + 2ri + ri−1) + P∑ p=1 1 4 (λp,i+1 + 2λp,i + λp,i−1) = 14 (Fi−1 + 2Fi + Fi+1). (26) 3.3. Equivalence to Newmark We will proceed with two additional steps to show that the optimality conditions (26) correspond to the Newmark integration scheme from Section 2. First, 2Notice that the optimality conditions exclude the initial and final times t0 and tN . This mirrors the standard treatment of the time-continuous case, in which the stationarity conditions of the action functional are derived under the assumption that the state variables remain fixed at both times, e.g. [22]. 3Notice that we assume the Lagrange multipliers λp,i to be given arbitrary quantities, similarly to the forcing terms Fi. Once we establish the equivalence to the Newmark algorithm, their values follow from the update formula (9) from Section 2. we demonstrate that the displacements {ri}Ni=0 pro- vide definitions of velocities {ṙi}Ni=0 and accelerations {r̈i}Ni=0 consistent with the kinematic assumptions in Eq. (6). Second, we show that the discrete-in-time quantities satisfy the equations of motion (1). Kinematics. Following [18, Section 2.2], we start by introducing auxiliary accelerations mr̈i+1/2 = − k∞ 2 (ri + ri+1) + 12 (Fi + Fi+1) − P∑ p=1 1 2 (λp,i + λp,i+1), (27) for i = 0, 1, . . . ,N − 1. Summing mr̈i−1/2 with mr̈i+1/2 and comparing the result with (26) provides ri+1 − 2ri + ri−1 ∆t2 = 12 (r̈i+1/2 + r̈i−1/2), (28) with i = 1, . . . ,N − 1. The discrete linear momenta follow standardly from pi = ∂Ld(ri−1,ri,λi−1,λi) ∂ri ∆t, (29) and, using r̈i−1/2 from Eq. (27), they can be evaluated as pi = mṙi = m ri −ri−1 ∆t + m r̈i−1/2 2 ∆t. (30) Expressing pi+1 according to the previous relation and employing (28) provides pi+1 = pi + mr̈i+1/2∆t, (31) from which we obtain ṙi+1 = ṙi + ∆tr̈i+1/2. (32) Likewise, expressing ri from (28) and employing the velocity ṙi from (30) provides ri+1 = ri + ṙi∆t + 12 r̈i+1/2∆t 2. (33) Hence, expressions (33) and (32) become identical to the ones of the Newmark method (6) once setting r̈i+1/2 = 12 (r̈i + r̈i+1). (34) Equilibrium. Employing the nodal accelerations r̈i−1/2 and r̈i+1/2 from (34) in the identity (28) reveals that ri+1 − 2ri + ri−1 ∆t2 = 14 (r̈i−1 + 2r̈i + r̈i+1). (35) Furthermore, by expressing the difference m(r̈i+1/2 − r̈i−1/2) using (27), we find that 1 2m(r̈i+1 − r̈i−1) + 1 2k∞(ri+1 −ri−1) + P∑ p=1 1 2 (λp,i+1 −λp,i−1) = 1 2 (Fi+1 −Fi−1). (36) 505 J. Schmidt, T. Janda, A. Zemanová et al. Acta Polytechnica Now, after inserting the identity (35) into the discrete Euler-Lagrange equations (26) and subtracting (36) from the result, we infer that m(r̈i−1 + r̈i) + k∞(ri−1 + ri) + P∑ p=1 (λp,i−1 + λp,i) =Fi−1 + Fi, (37) which can be reduced to the final form mr̈i + k∞ri + P∑ p=1 λp,i = Fi. (38) Indeed, the equivalence between (38) and (37) for i = 1 holds because of the choice of the initial accel- eration (13), and for i = 2, . . . ,N − 1 it follows by induction. 3.4. Energy balance The variational framework (14) additionally reveals that the trajectory q satisfies the energy balance con- dition, e.g., [23, Section 5.1] Eint(t) + D(t) = Eint(0) + W(t) for 0 ≤ t ≤ T, (39) with the internal energy Eint, dissipated energy D, and the work done by external forces W given by Eint(t) = 12mṙ 2(t) + 12k∞r 2(t) + P∑ p=1 1 2kpr 2 e,p(t), (40a) D(t) = P∑ p=1 ∫ t 0 ηpṙ 2 v,p(τ)dτ, (40b) W(t) = ∫ t 0 F(τ)ṙ(τ)dτ. (40c) To later quantify the artificial dissipation induced by time discretization, we also consider the time-discrete quantities Eint(ti) = 12mṙ 2 i + 1 2k∞r 2 i + P∑ p=1 λ2p,i 2kp , (41a) Dd(ti) = i−1∑ k=0 P∑ p=1 1 2ηp ( λ2p,k + λ 2 p,k+1 ) ∆t, (41b) Wd(ti) = i−1∑ k=0 1 2 ( Fkṙk + Fk+1ṙk+1 ) ∆t; (41c) the last two expressions correspond to the approxima- tions of integrals in (40) with the trapezoidal rule and employing the identities (22). 4. Examples In this section, we demonstrate the performance of the developed Newmark algorithm on two examples. kp [kNm−1] θp [s] kp [kNm−1] θp [s] 6933.9 10−9 445.1 102 3898.6 10−8 300.1 103 2289.2 10−7 401.60 104 1672.7 10−6 348.1 105 761.60 10−5 111.6 106 2401.0 10−4 127.2 107 65.200 10−3 137.8 108 248.00 10−2 50.5 109 575.60 10−1 322.9 1010 56.30 100 100.0 1011 188.6 101 199.9 1012 Table 1. Parameters of Maxwell chain model [10], with θp = ηp/kp and k∞ = 682.18 kNm−1. Note that in Section 4.2, the stiffnesses k• correspond to shear moduli G• [MPa]. The first one in Section 4.1 addresses the accuracy and numerical energy dissipation of the integrator for the single-degree-of-freedom system from Figure 1. The follow-up example in Section 4.2 outlines an extension of the scheme towards continuum models. Data of the generalized Maxwell chain used in both examples appear in Table 1; they represent real PolyVinyl Butyral (PVB) material with sufficiently short and long relaxation times for testing the algo- rithm robustness. For more information on experimen- tal procedures to determine these parameters, see [10]. All results presented in this section are reproducible with Python-based scripts available at [24]. 4.1. Discrete problem We consider the following two types of loading: F(t) = F for t ≥ 0, (42a) F(t) = F sin t for t ≥ 0, (42b) corresponding to step and harmonic loads, respectively. In both cases, we set the amplitude F = 1 MN and the mass m = 106 kg to scale the displacement amplitude to ≈ 1 m. Initial displacement, velocity, and forces in Maxwell cells were set to zero; recall (5). As for the Newmark algorithm, we set the time steps ∆t to 1.0, 0.5, and 0.2 s. Accuracy of the Newmark algorithm is checked by comparing its trajectories with the reference ones, obtained with the adaptive solver lsoda [25] — avail- able through odeint function of Scipy library [26] — applied to the full initial value problem (1), (4), and (5). The results appear in Figure 3 and demonstrate that the Newmark algorithm is stable even for coarse time steps. The errors behave consistently with findings for Newmark-family methods applied to linearly damped systems, e.g. [27, Section B.II.5]. In particular, the numerical dispersion (understood as the error in peri- ods) and dissipation (error in amplitudes) decays as 506 vol. 60 no. 6/2020 Newmark algorithm for dynamic analysis of viscoelastic materials 10-2 10-1 100 Time step ∆t [s] 10-6 10-5 10-4 10-3 10-2 10-1 100 Re la tiv e er ro r [ -] Numerical errors Quadratic order Figure 2. Convergence diagram for the step loading. Crosses represent the relative numerical errors for dif- ferent time steps, the solid line a quadratic function. O((ω∆t)2), where ω stands for the eigenfrequency of the undamped system. For ∆t = 0.2 s, the trajectories predicted by the Newmark scheme closely match the reference ones. To demonstrate the second-order accuracy qualita- tively, in Figure 2, we plot the decay of a relative error with a decreasing time step ∆t. The relative errors were calculated according to ||r∆t − rref∆t||2/||r ref ∆t||2, where r∆t stores displacements of all time instants for time step ∆t, rref∆t the reference solution of odeint at the same times, and ‖•‖ denotes the Euclidean norm. A comparison with the quadratic function (solid line) confirms the second-order accuracy. Note that the deviation visible for short time steps results from a prescribed tolerance of the odeint solver. Numerical dissipation. As follows from the en- ergy equality (39), the additional dissipation induced by the integrator can be estimated as ∆d(ti) = |Eint(0) + Wd(ti) −Eint(ti) −Dd(ti)| , (43) with the individual terms provided by Eq. (41). The evolution of these quantities for the step and harmonic loading appears in Figure 4, considering the time interval 〈0, 300〉 s. For both loads, we observe that the work done by external forces eventually distributes between the internal and dissipated energies; the ratio Dd/Wd stabilizes at 0.4 for the step load and for the harmonic loading, the ratio reaches about 0.9. The artificial dissipation is only significant for the coarsest step of ∆t = 1.0 s; for ∆t = 0.1 s it reaches the value of about 1 ‰ and further deteriorates with a decreasing time step. This confirms excellent energy conservation properties of the scheme, especially when taking into account that the error introduced by the trapezoidal rule in (41) is of order O(∆t2). 4.2. Generalization Additional constitutive assumptions must be adopted to extend the SDOF models into a continuum for- mulation. Here, we assume that the Maxwell model applies when modelling the response under shear. The spring stiffnesses k• thus become shear moduli G•, and the Poisson ratio ν is a time-independent mate- rial constant. This assumption considerably simpli- fies the multi-dimensional constitutive law, e.g., [16, Section 2.4], and provides the same results as the conventional volumetric-deviatoric split for our target applications [5]. Under these assumptions, the weak form of the equations of motion take the form, e.g. [11, Part III]: δFext(t,δu) = ∫ Ω δu ·ρü(t) dΩ + ∫ Ω δε : ( G∞Dν : ε(t) + P∑ p=1 σp(t) ) dΩ, (44) where δFext stands for the virtual work done by exter- nal loads on a virtual displacement δu, ü denotes the acceleration, and the small strain tensor ε is obtained as the symmetric part of the displacement gradient, ε = ∇su; virtual strain δε is defined in the same way. The material is characterized by its density ρ, long-term shear modulus of the Maxwell chain G∞, and the dimensionless tensor Dν corresponding to the stiffness tensor of an isotropic material of the unit shear modulus and the Poisson ratio of ν. The stresses σp carried by individual cells follow as the solution of initial value problems σ̇p(t) Gp + σp(t) ηp = Dνε̇(t), p = 1, 2, . . . ,P. (45) The evolution of the state variables u and σp is speci- fied with the initial conditions on the displacements, velocities, and cell stresses: u(0) = u0, u̇(0) = v0, σp(0) = σp,0. (46) The comparison of the initial value problems spec- ified with Eqs. (1), (4), and (5) and Eqs. (44), (45), and (46) reveals that the derivation of the Newmark- type scheme follows exactly the steps as in Section 2.2. As a result, the following variational problem needs to be solved at a time ti+1:∫ Ω δu ·ρüi+1 dΩ + ∫ Ω δε : ( 1 4G∞∆t 2 + P∑ p=1 Bp ) Dν : ε̈i+1 dΩ = δFext(ti+1,δu) − ∫ Ω δε : ( P∑ p=1 Apσp,i ) dΩ − ∫ Ω δε : G∞Dν : εi dΩ 507 J. Schmidt, T. Janda, A. Zemanová et al. Acta Polytechnica 0 5 10 15 20 25 30 time t [s] 0.0 0.1 0.2 0.3 0.4 0.5 0.6 di sp la ce m en t r (t ) [ m ] Reference Newmark 0 5 10 15 20 25 30 time t [s] 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 di sp la ce m en t r (t ) [ m ] Reference Newmark 0 5 10 15 20 25 30 time t [s] 0.0 0.1 0.2 0.3 0.4 0.5 0.6 di sp la ce m en t r (t ) [ m ] Reference Newmark 0 5 10 15 20 25 30 time t [s] 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 di sp la ce m en t r (t ) [ m ] Reference Newmark 0 5 10 15 20 25 30 time t [s] 0.0 0.1 0.2 0.3 0.4 0.5 0.6 di sp la ce m en t r (t ) [ m ] Reference Newmark 0 5 10 15 20 25 30 time t [s] 0.6 0.4 0.2 0.0 0.2 0.4 0.6 di sp la ce m en t r (t ) [ m ] Reference Newmark Figure 3. Accuracy of the viscous Newmark method for step (left) and harmonic (right) loadings defined by Eq. (42). Top, center, and bottom graphs show trajectories for time steps ∆t = 1.0 s, ∆t = 0.5 s, and ∆t = 0.2 s respectively. − ∫ Ω δε : ( G∞∆t + P∑ p=1 Gpθ̂p ) Dν : ε̇i dΩ − ∫ Ω δε : ( 1 4G∞∆t 2 + P∑ p=1 Bp ) Dν : ε̈i dΩ, (47) with the parameters θ̂p, Ap, and Bp provided by Eqs. (10) and (11); recall that δFext denotes the vir- tual work done by external forces. Once the the accel- erations üi+1 are obtained from the weak form (47), the displacements ui+1, velocities u̇i+1, and the cell stresses σp,i+1 are updated according to Eqs. (6) and (9), respectively. The outlined formulation (47) was further dis- cretized with the finite element method and imple- mented in FEniCS project [28, 29] version 2018.1. As an indicative example, we consider a unit cube, see Figure 5, fixed on the bottom surface and subjected to a step load (42a) with the tensile traction of intensity 1.0 Nm−2 perpendicular to the top surface. The ma- terial response is characterized by the Maxwell chain parameters from Table 1 and the value of the Pois- son ratio ν = 0.49. In the numerical resolution, we 508 vol. 60 no. 6/2020 Newmark algorithm for dynamic analysis of viscoelastic materials 0 50 100 150 200 250 300 time t [s] 0.0 0.2 0.4 0.6 0.8 1.0 Re la tiv e en er gy [- ] Eint/Wd Dd/Wd 0 50 100 150 200 250 300 time t [s] 10-9 10-8 10-7 10-6 10-5 10-4 10-3 10-2 10-1 100 Re la tiv e di ss ip at io n ∆ d / W d [- ] ∆t = 1.00 ∆t = 0.10 ∆t = 0.01 0 50 100 150 200 250 300 time t [s] 0.0 0.2 0.4 0.6 0.8 1.0 Re la tiv e en er gy [- ] Eint/Wd Dd/Wd 0 50 100 150 200 250 300 time t [s] 10-5 10-4 10-3 10-2 10-1 100 Re la tiv e di ss ip at io n ∆ d / W d [- ] ∆t = 1.00 ∆t = 0.10 ∆t = 0.01 Figure 4. Normalized energies corresponding to SDOF response to step (top) and harmonic loads (bottom). Left: the evolution of internal energy Eint and dissipation Dd, normalized by the work done by external forces Wd for time step ∆t = 0.5 s. Right: the evolution of numerical dissipation ∆d, normalized by the work done by external forces Wd. discretize the sample into identical 1,000 hexahedron elements and consider the time step of 0.01 s. The snapshots of the vibrations reveal a similar behaviour to the SDOF example, recall Figure 3, namely the attenuation of the propagating waves by viscous damping. An interested reader is invited to the dataset [24] for full details on the simulation. 5. Conclusions In this contribution, we have developed a Newmark integration scheme for viscoelastic solids character- ized by the generalized Maxwell model. Besides the direct derivation, we have shown the scheme can be derived from the Hamilton variational principle com- bined with a suitable structure-preserving time dis- cretization. This variational structure is then reflected in low numerical energy dissipation of the resulting scheme, which has been confirmed with selected nu- merical examples. As the next step, we will combine the continuum framework outlined in Section 4.2 with Newmark- type solvers for variational fracture models, e.g. [30, 31], to extend the currently available approaches to simulating the response of laminated glass structures under impact. Acknowledgements This publication was supported by the Czech Science Foundation, the grant No. 19-15326S. We wish to thank an anonymous reviewer for her/his constructive criticism that helped us clarify the original manuscript at several places. References [1] T. Hatada, T. Kobori, M. Ishida, N. Niwa. Dynamic analysis of structures with Maxwell model. Earthquake Engineering & Structural Dynamics 29(2):159–176, 2000. doi:10.1002/(SICI)1096- 9845(200002)29:2<159::AID-EQE895>3.0.CO;2-1. [2] M. Haldimann, A. Luible, M. Overend. Structural Use of Glass, vol. 10 of Structural Engineering Documents. IABSE, Zürich, Switzerland, 2008. [3] A. van Duser, A. Jagota, S. J. Bennison. Analysis of glass/Polyvinyl Butyral laminates subjected to uniform pressure. Journal of Engineering Mechanics 125(4):435–442, 1999. doi:10.1061/(ASCE)0733-9399(1999)125:4(435). 509 http://dx.doi.org/10.1002/(SICI)1096-9845(200002)29:2<159::AID-EQE895>3.0.CO;2-1 http://dx.doi.org/10.1002/(SICI)1096-9845(200002)29:2<159::AID-EQE895>3.0.CO;2-1 http://dx.doi.org/10.1061/(ASCE)0733-9399(1999)125:4(435) J. Schmidt, T. Janda, A. Zemanová et al. Acta Polytechnica time t [s] di sp la ce m en t r (t )[ m ] Figure 5. Snapshots of deformations of a viscoelastic cube subjected to step load. [4] L. Galuppi, G. Royer-Carfagni. The design of laminated glass under time-dependent loading. International Journal of Mechanical Sciences 68:67–75, 2013. doi:10.1016/j.ijmecsci.2012.12.019. [5] A. Zemanová, J. Zeman, M. Šejnoha. Comparison of viscoelastic finite element models for laminated glass beams. International Journal of Mechanical Sciences 131-132:380–395, 2017. doi:10.1016/j.ijmecsci.2017.05.035. [6] A. Zemanová, J. Zeman, T. Janda, M. Šejnoha. Layer-wise numerical model for laminated glass plates with viscoelastic interlayer. Structural Engineering and Mechanics 65(4):369–380, 2018. doi:10.12989/sem.2018.65.4.369. [7] L. Andreozzi, S. B. Bati, M. Fagone, et al. Dynamic torsion tests to characterize the thermo-viscoelastic properties of polymeric interlayers for laminated glass. Construction and Building Materials 65:1–13, 2014. doi:10.1016/j.conbuildmat.2014.04.003. [8] Y. Shitanoki, S. Bennison, Y. Koike. A practical, nondestructive method to determine the shear relaxation modulus behavior of polymeric interlayers for laminated glass. Polymer Testing 37:59–67, 2014. doi:10.1016/j.polymertesting.2014.04.011. [9] I. Mohagheghian, Y. Wang, L. Jiang, et al. Quasi-static bending and low velocity impact performance of monolithic and laminated glass windows employing chemically strengthened glass. European Journal of Mechanics – A/Solids 63:165–186, 2017. doi:10.1016/j.euromechsol.2017.01.006. [10] T. Hána, T. Janda, J. Schmidt, et al. Experimental and numerical study of viscoelastic properties of polymeric interlayers used for laminated glass: Determination of material parameters. Materials 12(14):2241, 2019. doi:10.3390/ma12142241. [11] R. W. Clough, J. Penzien. Dynamics of Structures. Computers & Structures, Inc, Berkeley, 3rd edn., 2003. [12] Y. Koutsawa, et al. Static and free vibration analysis of laminated glass beam on viscoelastic supports. International Journal of Solids and Structures 44(25- 26):8735–8750, 2007. doi:10.1016/j.ijsolstr.2007.07.009. [13] M. L. Aenlle, F. Pelayo. Frequency Response of Laminated Glass Elements: Analytical Modeling and Effective Thickness. Applied Mechanics Reviews 65(2):020802–020802–13, 2013. doi:10.1115/1.4023929. [14] A. Zemanová, J. Zeman, T. Janda, et al. On modal analysis of laminated glass: Usability of simplified methods and Enhanced Effective Thickness. Composites Part B: Engineering 151:92–105, 2018. doi:10.1016/j.compositesb.2018.05.032. [15] O. C. Zienkiewicz, M. Watson, I. P. King. A numerical method of visco-elastic stress analysis. International Journal of Mechanical Sciences 10(10):807– 827, 1968. doi:10.1016/0020-7403(68)90022-2. [16] Z. P. Bažant, M. Jirásek. Creep and Hygrothermal Effects in Concrete Structures, vol. 225 of Solid Mechanics and Its Applications. Springer, Dordrecht, 2018. doi:10.1007/978-94-024-1138-6. [17] N. M. Newmark. A Method of Computation for Structural Dynamics. Journal of the Engineering Mechanics Division 85(3):67–94, 1959. [18] C. Kane, J. E. Marsden, M. Ortiz, M. West. Variational integrators and the Newmark algorithm for conservative and dissipative mechanical systems. International Journal for Numerical Methods in Engineering 49(10):1295–1325, 2000. doi:10.1002/1097-0207(20001210)49:10<1295::AID- NME993>3.0.CO;2-W. [19] B. Bourdin, G. A. Francfort, J.-J. Marigo. The variational approach to fracture. Journal of Elasticity 91(1-3):5–148, 2008. doi:10.1007/s10659-007-9107-3. [20] M. Buliga. Hamiltonian inclusions with convex dissipation with a view towards applications. Annals of the Academy of Romanian Scientists Series on Mathematics and its Applications 1(2):228–251, 2009. 510 http://dx.doi.org/10.1016/j.ijmecsci.2012.12.019 http://dx.doi.org/10.1016/j.ijmecsci.2017.05.035 http://dx.doi.org/10.12989/sem.2018.65.4.369 http://dx.doi.org/10.1016/j.conbuildmat.2014.04.003 http://dx.doi.org/10.1016/j.polymertesting.2014.04.011 http://dx.doi.org/10.1016/j.euromechsol.2017.01.006 http://dx.doi.org/10.3390/ma12142241 http://dx.doi.org/10.1016/j.ijsolstr.2007.07.009 http://dx.doi.org/10.1115/1.4023929 http://dx.doi.org/10.1016/j.compositesb.2018.05.032 http://dx.doi.org/10.1016/0020-7403(68)90022-2 http://dx.doi.org/10.1007/978-94-024-1138-6 http://dx.doi.org/10.1002/1097-0207(20001210)49:10<1295::AID-NME993>3.0.CO;2-W http://dx.doi.org/10.1002/1097-0207(20001210)49:10<1295::AID-NME993>3.0.CO;2-W http://dx.doi.org/10.1007/s10659-007-9107-3 vol. 60 no. 6/2020 Newmark algorithm for dynamic analysis of viscoelastic materials [21] B. Bourdin, C. J. Larsen, C. L. Richardson. A time-discrete model for dynamic fracture based on crack regularization. International Journal of Fracture 168(2):133–143, 2011. doi:10.1007/s10704-010-9562-x. [22] A. Bedford. Hamilton’s principle in continuum mechanics. Pitman Publishing, Boston, 1985. [23] A. Mielke, T. Roubíček. Rate-independent systems: theory and application, vol. 655 of Applied Mathematical Sciences. Springer, New York, 2015. doi:10.1007/978-1-4939-2706-7. [24] J. Schmidt, T. Janda, A. Zemanová, et al. Source codes for preprint Newmark algorithm for dynamic analysis with Maxwell chain model, 2019. doi:10.5281/zenodo.3531802. [25] L. Petzold. Automatic selection of methods for solving stiff and nonstiff systems of ordinary differential equations. SIAM Journal on Scientific and Statistical Computing 4(1):136–148, 1983. doi:10.1137/0904010. [26] E. Jones, T. Oliphant, P. Peterson, et al. SciPy: Open source scientific tools for Python, 2001–. http://www.scipy.org, [accessed January 12, 2021]. [27] T. J. R. Hughes. Analysis of transient algorithms with particular reference to stability behavior. In T. Belytschko, T. J. R. Hughes (eds.), Computational Methods for Transient Analysis, chap. 2, pp. 67–155. North-Holland, Amsterdam, 1983. [28] A. Logg, K.-A. Mardal, G. Wells (eds.). Automated Solution of Differential Equations by the Finite Element Method: The FEniCS Book. Lecture Notes in Computational Science and Engineering. Springer-Verlag, Berlin Heidelberg, 2012. doi:10.1007/978-3-642-23099-8. [29] M. S. Alnæs, J. Blechta, J. Hake, et al. The FEniCS Project Version 1.5. Archive of Numerical Software 3(100), 2015. doi:10.11588/ans.2015.100.20553. [30] T. Li, J.-J. Marigo, D. Guilbaud, S. Potapov. Numerical investigation of dynamic brittle fracture via gradient damage models. Advanced Modeling and Simulation in Engineering Sciences 3(1):26, 2016. doi:10.1186/s40323-016-0080-x. [31] T. Roubíček. Models of dynamic damage and phase- field fracture, and their various time discretisations. In M. Hintermüller, J. F. Rodrigues (eds.), Topics in Applied Analysis and Optimisation, pp. 363–396. Springer International Publishing, Cham, 2019. 511 http://dx.doi.org/10.1007/s10704-010-9562-x http://dx.doi.org/10.1007/978-1-4939-2706-7 http://dx.doi.org/10.5281/zenodo.3531802 http://dx.doi.org/10.1137/0904010 http://www.scipy.org http://dx.doi.org/10.1007/978-3-642-23099-8 http://dx.doi.org/10.11588/ans.2015.100.20553 http://dx.doi.org/10.1186/s40323-016-0080-x Acta Polytechnica 60(6):502–511, 2020 1 Introduction 2 Newmark method 2.1 Governing equations 2.2 Discretization 3 Variational integrators 3.1 Variational framework 3.2 Discretization 3.3 Equivalence to Newmark 3.4 Energy balance 4 Examples 4.1 Discrete problem 4.2 Generalization 5 Conclusions Acknowledgements References