1 Introduction

This paper addresses the construction of mass matrices for dynamic models of structures treated by the finite element method (FEM). The goal is to introduce a general approach through which customized mass matrices can be constructed for specific structural elements. This is called the method of templates. The qualifier “customized” is defined more precisely later.

Two standard procedures for constructing FEM mass matrices are well known. They are outlined in Appendix 2 to make this paper reasonably self-contained. They lead to consistent and diagonally lumped forms, respectively. Conventional forms of those models are denoted by \(\mathbf {M}_C^{}\) and \(\mathbf {M}_L^{}\), respectively, in the sequel, with additional subscripts or superscripts as necessary or convenient. Abbreviations consistent mass matrix (CMM) and diagonally lumped mass matrices (DLMM), respectively, are also used. Collectively those two models take care of many engineering applications in structural dynamics. Occasionally, however, they fall short. The gap can be filled with a more general approach that relies on templates. These are algebraic forms that carry free parameters. The set of parameters is called the template signature. When given numerical values, the signature uniquely characterizes a mass matrix instance.

This paper presents basic concepts and techniques that underlie the template approach. This methodology is applied to several one-dimensional (1D) structural elements as expository examples.

The template approach has the virtue of generating a set of mass matrices that satisfy certain a priori constraints; for example symmetry, nonnegativity, invariance and linear momentum conservation. A mass matrix that satisfies those will be called admissible. In particular, the diagonally lumped and consistent mass matrices should be obtained as instances. Thus those standard models are not excluded. Availability of free parameters, however, allows the mass matrix to be customized to special requirements.

Several customization scenarios are listed in Table 1, along with their acronyms. The last one: reduction of directional anisotropy in wave propagation, is not applicable to 1D elements and therefore not treated in this paper.

Table 1 Template customization scenarios

The versatility of application will be evident from the examples. It will be also seen that optimizing templates for one scenario generally does not help with others, and in fact may make things worse. Thus, ability to adapt the mass matrix to particular needs as well as problem regions is an important virtue. Note that mesh and freedom configuration need not be modified in any way; only template signatures are adjusted.

An attractive feature of templates for FEM programming is that each “custom mass matrix” need not be coded and tested individually. It is sufficient to implement the template as a single element-level module, with free parameters as arguments (alternatively, useful instances may be identified by predefined mnemonic character strings, and converted to numerical signatures internally). The signature is adjusted according to goals and needs. In particular the same module should be able to produce the conventional DLMM and CMM models as instances. This can provide valuable crosschecking with well established programs while doing benchmarks.

In problems characterized by rapid transients, such as contact-impact and fragmentation, templates allow a flexible customization: reduced high-frequency pollution in elements in or near shock regions while maintaining low-frequency continuum fit away from such regions. In those scenarios, signatures may evolve in time.

2 Is Customization Worth the Trouble?

The ability to customize a mass matrix is not free of development costs. The presence of free parameters makes template derivations considerably more complicated than those based on the two standard procedures noticed in Sect. 1 and outlined in Appendix 2. Reason: everything must be carried along symbolically: geometry, material and fabrication properties, in addition to the free parameters. Consequence: hand computations rapidly become unfeasible, even for fairly simple 1D elements. Help from a computer algebra system (CAS) is needed to get timely results. A key issue is: when is this additional work justified? Two specific cases may be mentioned.

One is high fidelity systems. Dynamic analysis covers a wide range of applications. There is a subclass that calls for a level of simulation accuracy beyond that customary in engineering analysis. Examples are deployment of precision space structures, resonance analysis of machinery or equipment, adaptive active control systems, medical imaging, phononics (long wave guidance at molecular level), vehicle signature detection, radiation loss in layered circuits, and molecular- and crystal-level simulations in micro- and nano-mechanics.

In static structural analysis an error of 20 or 30 % in peak stresses is not cause for alarm—such discrepancies are usually covered adequately by safety factors. But a similar error in frequency analysis or impedance response of a high fidelity system can be disastrous. Achieving acceptable precision with a fine mesh, however, can be expensive. Model adaptivity comes to the rescue in statics. This approach is less effective in dynamics, however, on account of the time dimension and the fact that irregular meshes are prone to develop numerical pollution. Customized elements may provide a practical solution: achieving adequate accuracy with a coarse regular mesh.

Another possibility is that the stiffness matrix comes from a method that avoids displacement shape functions (DSF). For example, assumed-stress or strain elements. Or, it could simply be an array of numbers provided by a black-box program, with no documentation explaining its source. If this happens the concept of CMM, in which velocity shape functions (VSF) are taken to coincide with DSF, loses the comfortable variational meaning outlined in  Sect. 1. An expedient way out is to choose an element with similar geometry and freedom configuration derived with DSF and take those as VSF. But which element to pick? If time allows, constructing and customizing a template avoids uncritically rolling the dice.

3 Mass Parametrization Techniques

There are several ways to parametrize mass matrices. Techniques found effective in practice are summarized below. Most of them are illustrated in the worked out examples of ensuing sections.

It is often advantageous to have several template expressions for the same element configuration. For example, to study the subset of DLMM it may be convenient to streamline the general form to one that produces only such matrices. Likewise for singular mass matrices. In that case we speak of template variants. These may overlap totally or partially: the DLMM variant is plainly a subset of the general mass template. The key difference between a template subset and a variant is that the latter redefines free parameters from scratch.

For the reader’s convenience, acronyms often used in this paper are listed in Table 2. A set of definitions and abbreviations pertaining to templates are collected in Table 3.

Table 2 Acronyms used in paper
Table 3 Template related nomenclature

Notational conventions for mathematical expressions that appear in this paper are summarized in Table 4. Specific conventions used for free template parameters are given in Table 5. As a general rule, template parameters are always dimensionless and denoted by lower case Greek letters.

3.1 Matrix-Weighted Parametrization

A matrix-weighted (MW) mass template for element \(e\) is a linear combination of \((k+1)\) component mass matrices, \(k\ge 1\) of which are weighted by parameters \(\mu _i\), (\(i=1,\ldots k\)):

$$\begin{aligned} \mathbf {M}^e\,\,{\mathop {=}\limits ^{\mathrm{def}}}\,\,\mathbf {M}_0^e+\mu _1 \mathbf {M}_1^e +\cdots \mu _k\mathbf {M}_k^e. \end{aligned}$$
(1)

Here \(\mathbf {M}_0^e\) is the baseline mass matrix. This should be an admissible mass matrix on its own if \(\mu _1=\ldots \mu _k=0\). The simplest instance of (1) is a linear combination of the CMM and a DLMM:

$$\begin{aligned} \mathbf {M}^e \,\,{\mathop {=}\limits ^{\mathrm{def}}}\,\,(1-\mu )\mathbf {M}_C^e+\mu \mathbf {M}_L^e. \end{aligned}$$
(2)

This can be reformatted as (1) by writing \( \mathbf {M}^e = \mathbf {M}_C^e + \mu (\mathbf {M}_L^e-\mathbf {M}_C^e) = \mathbf {M}_0^e + \mu \,\mathbf {M}_1^e\). Here \(k=1\), the baseline is \(\mathbf {M}_0^e\equiv \mathbf {M}_C^e,\, \mu \equiv \mu _1\) and \(\mathbf {M}_1^e\) is the “mass deviator” \(\mathbf {M}_L^e-\mathbf {M}_C^e\). The specialization (2) is often abbreviated to “linear combination of consistent and diagonally lumped masses,” with acronym LCDM; cf. Table 2. The rationale behind (2) is that the CMM typically overestimates natural frequencies while a DLMM usually underestimates them. Thus a linear combination has a good chance of improving low-frequency accuracy for some \(\mu \in [0,1]\).

A MW mass template represents a tradeoff. It cuts down on the number of free parameters. Such a reduction is essential for 2D and 3D elements. It makes it easier to satisfy conservation and nonnegativity conditions through appropriate choice of the \(\mathbf {M}_i^e\). On the minus side it generally spans only a subspace of admissible mass matrices.

Table 4 General notational conventions for mathematical expressions
Table 5 Notational conventions for template parameters

3.2 Spectral Parametrization

A spectrally parametrized (SP) mass template has the form

$$\begin{aligned} \mathbf {M}^e \,\,{\mathop {=}\limits ^{\mathrm{def}}}\,\,\mathbf {H}^T\,\mathbf {D}_{\mu }^{}\,\mathbf {H}, \qquad \mathbf {D}_\mu ^{}=\mathbf{diag}{\left[ \begin{array}{lllll}c_0\mu _0 &{} c_1\mu _1 &{} \ldots &{} c_k\mu _k\\ \end{array}\right] }. \end{aligned}$$
(3)

in which \(\mathbf {H}\) is typically a full matrix. Parameters \(\mu _0 \ldots \mu _k\) appear as entries of the diagonal matrix \(\mathbf {D}_\mu \). Scaling coefficients \(c_i\) may be introduced for convenience so the \(\mu _i\) are dimensionless. Often the values of \(\mu _0\) and/or \(\mu _1\) are preset from conservation conditions.

Configuration (3) occurs naturally when \(\mathbf {M}^e\) is constructed first in generalized coordinates, followed by congruential transformation to physical coordinates via \(\mathbf {H}\). If the generalized mass is derived using mass-orthogonal functions (for example, Legendre polynomials in 1D elements), the unparametrized generalized mass \(\mathbf {D}=\mathbf{diag}\left[ \begin{array}{lllll}c_0&{}\quad c_1&{}\quad \ldots c_k\\ \end{array}\right] \) is diagonal. Parametrization is effected by scaling its entries. As noted, some entries may be left fixed to satisfy a priori constraints such as mass conservation.

Expanding (3) and collecting matrices that multiply each \(\mu _i^{}\) leads to a matrix weighted combination form (1) in which each \(\mathbf {M}_i^e\) is a rank-one matrix. The analogy with the spectral representation theorem of symmetric matrices is obvious. But in practice it is usually better to work directly with the congruent representation (3).

As remarked later in Sect. 3.6, SP is especially convenient for constructing singular mass matrices under customization scenario RHFP of Table 1.

3.3 Entry-Weighted Parametrization

An entry-weighted (EW) mass template applies free parameters directly to each entry of the mass matrix, except for a priori constraints on symmetry, invariance and conservation. As an example, for a 1D element with three translational DOF we may start from

$$\begin{aligned} \mathbf {M}^e \,\,{\mathop {=}\limits ^{\mathrm{def}}}\,\,m^e \left[ \begin{array}{lll} \mu _{11} &{}\quad \mu _{12} &{}\quad \mu _{13}\\ \mu _{12} &{}\quad \mu _{22} &{}\quad \mu _{23} \\ \mu _{13} &{}\quad \mu _{23} &{}\quad \mu _{33} \end{array}\right] \!, \end{aligned}$$
(4)

in which \(m^e\) is the total element mass, and the sum of all row sums is one. EW is often applied to entries of a “deviator matrix” that measures the change from a baseline matrix such as \(\mathbf {M}_C\). For example, see the three-node bar template (39).

Because of its generality, EW parametrization can be expected to lead to optimal customized instances. But it is restricted to simple (usually 1D) elements because the number of parameters grows quadratically in the matrix size, whereas for the foregoing two schemes either it grows linearly, or stays constant.

3.4 Multilevel Parametrization

A hierarchical combination of parametrization schemes can be used to advantage if the kinetic energy can be naturally decomposed from physical considerations. For example, the Timoshenko beam element covered in Sect. 7 uses a two-matrix-split template combined by a weighted form similar to (2) as top level (the energy split is between translational and rotational inertia). The two components are constructed by spectral and EW parametrization, respectively. Such combinations fall under the scope of multilevel (ML) parametrization.

3.5 Selective Mass Scaling

Selective mass scaling (SMS), is a method proposed recently (references given in Appendix Sect. 1.4), in which the mass matrix is modified by a scaled version of the stiffness matrix. Thus \(\mathbf {M}\) becomes

$$\begin{aligned} \mathbf {M}_K=\mathbf {M}+{\mu _K\over \omega _{ref}^2}\,\mathbf {K}. \end{aligned}$$
(5)

Here \(\mu _K\ge 0\) is a dimensionless scaling factor whereas \(\omega _{ref}^2\) is a “reference” frequency used to homogenize physical dimensions. The modification (5) may be done at the element or system level. The objective is to “filter down” high frequencies in explicit DTI for applications such as contact-impact; e.g., vehicle crash simulation. Filtering aims to reduce spurious noise as well as increasing the stable timestep. It thus follows under customization scenarios RHFP and MSTS of Table 1. The basic idea can be explained as follows. Let \(\omega _i\) and \(\mathbf {v}_i\) denote the natural frequencies and associated orthonormalized eigenvectors, respectively, whereas \(\hat{\omega }_i\) and \(\hat{\mathbf {v}}_i\) are their counterparts for the modified eigenproblem \((\mathbf {M}_K+\hat{\omega }_i^2\,\mathbf {K})\,\hat{\mathbf {v}}_i=\mathbf {0}\). By inspection the eigenvectors are preserved: \(\hat{\mathbf {v}}_i=\mathbf {v}_i\). Taking the Rayleigh quotient shows that the modified frequencies are

$$\begin{aligned} \hat{\omega }_i^2={\omega _i^2\over R_i}, \quad \hbox { in which }\quad R_i=1+\mu _K{\omega _i^2\over \omega _{ref}^2}. \end{aligned}$$
(6)

Choosing \(\mu _K>0\) cuts down each frequency by \(R_i>0\). For low frequencies the modification is negligible if \(\mu _K\) and \(\omega _{ref}^2\) are appropriately selected so that \(R_i\approx 1\). For nonphysical high frequencies (mesh modes) the reduction can be significant In fact note that if \(\omega _i^2>>\mu _K/\omega _{ref}^2\), \(\hat{\omega }_i^2\) cannot exceed the fixed bound \(\omega _{max}^2=\omega _{ref}^2/\mu _K\). The downside is that low frequency accuracy may suffer significantly, as illustrated later.

Although SMS may be presented as a variant of the MW parametrization technique of Sect. 3.1, it deserves to be considered on its own for the reasons stated in Appendix Sect. 1.4.

3.6 Singular Mass Matrices

A thread linked to SMS but independently developed is that of singular mass matrices. This has been primarily advocated for multibody dynamics, as well as dynamical systems leading to differential-difference equations of motion (EOM) that occur in active or passive control with time lags. References are provided in Appendix Sect. 1.5. The objective is roughly similar to SMS: reduce high frequency noise pollution triggered by rapid transients and/or time lags. But now this is done by raising the optical branch (OB) (or branches) so as to widen the acoustoptical gap pictured described in Sect. 5.1 and illustrated in Fig. 8. Noisy frequencies that fall in the gap decay exponentially.

There are several ways to produce such matrices. Under the template framework, the use of SP is particularly convenient, as observed in Sect. 3.2. Other approaches include reduced numerical integration, as well as injection of a convenient null space using mass matrix projection.

3.7 Constant Optical Branch Variant

Instead of rising the OB (or branches) by making \(\mathbf {M}^e\) singular, one may try to make the OB frequency independent. Templates that accomplish that feat are tagged as having a constant optical branch (COB) for short. They form subsets collectively identified as the COB variant. The group velocity pertaining to a COB vanishes, so associated waveforms with that particular frequency do not propagate. COB templates were discovered during the course of this work, and are briefly studied in Sect. 5.13 for the three-node bar element.

3.8 Mass-Stiffness Template Pairs

The concept of template was first developed for element stiffness matrices, as a natural generalization of its decomposition into basic and higher order parts. A brief historical account is provided in Appendix Sect. 1.6. Normally the stiffness template is optimized by imposing superconvergence conditions dealing with higher order patch tests while element aspect ratios are kept arbitrary. That optimal instance, if found, is kept fixed while a mass matrix template is subsequently investigated.

Maximum customization for dynamics can be expected if both stiffness and mass matrix templates can be simultanously adjusted. This is known as a mass-stiffness (MS) template. These may be of interest when improving dynamic behavior is paramount. Presently there is relatively little experience with this more ambitious approach. A note of caution: highly optimized MS templates may be abnormally sensitive to geometric or material perturbations away from a regular mesh.

3.9 Frequency Dependent Templates

One final generalization should be mentioned: allowing free parameters to be function of the frequency. If this is done for the mass matrix, we speak of a frequency dependent mass (FDM) template. If this is done for both the mass and stiffness matrices, we call the combination a FDM-stiffness (FDMS) template. Both cases are illustrated in Sects.  4.114.13 for the two-node bar element.

Although this ultimate complication is largely a curiosity, it might be occasionally useful in problems that profit from transformation to the frequency domain. For example: a linear dynamic system driven by a harmonic excitation of slowly varying frequency, if only the long term (steady-state) response is considered. Such systems may arise in parametric stability and active control.

4 The Two-Node Bar Element

The template approach is best grasped through an example that involves the simplest nontrivial structural finite element: a two-node prismatic bar of mass density \(\rho \), area \(A\) and length \(\ell \), that can only move along the longitudinal axis \(x\). See Fig. 1a. This element is often acronymed Bar2 for brevity’s sake. The well known consistent and DLMM forms are

$$\begin{aligned} \mathbf {M}_C^e = {m^e\over 6}\left[ \begin{array}{ll} 2 &{}\quad 1 \\ 1 &{}\quad 2 \\ \end{array}\right] , \qquad \mathbf {M}_L^e = {m^e\over 2} \left[ \begin{array}{ll} 1 &{}\quad 0 \\ 0 &{}\quad 1\\ \end{array}\right] . \end{aligned}$$
(7)

in which \(m^e=\rho A\ell \) is the total element mass. These are derived in Sect. 1.

Fig. 1
figure 1

The two-node prismatic bar element: a element configuration, b direct mass lumping to end nodes

4.1 Bar2 Entry Weighted Template

The most general mass matrix form for Bar2 is the EW template

$$\begin{aligned} \mathbf {M}^e&= \left[ \begin{array}{ll} M_{11}^e &{}\quad M_{12}^e\\ M_{21}^e &{}\quad M_{22}^e\\ \end{array}\right] = \rho A\,\ell \left[ \begin{array}{ll} \mu _{11} &{}\quad \mu _{12} \\ \mu _{21} &{}\quad \mu _{22}\\ \end{array}\right] \nonumber \\&= m^e\, \left[ \begin{array}{ll} \mu _{11} &{}\quad \mu _{12}\\ \mu _{21} &{}\quad \mu _{22} \end{array}\right] \!. \end{aligned}$$
(8)

The first form is merely a list of entries. Next the element mass \(m^e=\rho A\,\ell \) is factored out. The emerging parameters \(\mu _{11}\) through \(\mu _{22}\) are numbers, which illustrates a general rule: template free parameters should be dimensionless. This simplifies analysis and implementation. To cut down on parameters one looks at configuration constraints. The most obvious ones are:

Matrix symmetry \(\mathbf {M}^e=(\mathbf {M}^e)^T\). For the expression (8) this requires \(\mu _{21}=\mu _{12}\).

Physical symmetry: For a prismatic bar, \(\mathbf {M}^e\) in (8) must exhibit antidiagonal symmetry: \(\mu _{22}=\mu _{11}\).

Conservation of total translational mass: same as conservation of linear momentum or of kinetic energy. Apply the uniform velocity field \(\dot{u}=v\) to the bar. The associated nodal velocity vector is \(\dot{\mathbf {u}}^e=\mathbf {v}^e=v \left[ \begin{array}{ll} 1&{}\quad 1 \\ \end{array}\right] ^T\). The kinetic energy is \(T^e={\textstyle {1\over 2}}(\mathbf {v}^e)^T\mathbf {M}^e\mathbf {v}^e={\textstyle {1\over 2}}m^e v^2(\mu _{11}+\mu _{12}+\nu _2+\mu _{22})\). This must equal \({\textstyle {1\over 2}}m^e v^2\), whence \(\mu _{11}+\mu _{12}+\mu _{21}+\mu _{22}=2(\mu _{11}+\mu _{12})=1\).

Nonnegativity: \(\mathbf {M}^e\) should not be indefinite (this is not an absolute must, and it is actually relaxed in some elements discussed later). Whether checked by computing eigenvalues or principal minors, this constraint is nonlinear and of inequality type. Consequently it is not often applied ab initio, unless the element is quite simple, as in this case, or can be stated through simple expressions.

4.2 Bar2 One Parameter Mass Template

On applying the symmetry and conservation rules three parameters of (8) are eliminated. The remaining one, called \(\mu \), is taken for convenience to be \( \mu _{11}=\mu _{22}=(2+\mu )/6\) and \(\mu _{12}=\mu _{21}= (1-\mu )/6\). This rearrangement gives

$$\begin{aligned} \mathbf {M}^e_\mu = {\textstyle {1\over 6}}\rho A\,\ell \left[ \begin{array}{ll} 2+\mu &{}\quad 1-\mu \\ 1-\mu &{}\quad 2+\mu \end{array}\right] = (1-\mu )\mathbf {M}^e_C+\mu \mathbf {M}^e_L. \end{aligned}$$
(9)

Expression (9) shows that the general Bar2 mass template can be recast as a linear combination of the CMM and DLMM instances listed in (7). Summarizing, we end up with a one-parameter, matrix-weighted (MW) template that befits the LCD form (2). If \(\mu =0\) and \(\mu =1\), (9) reduces to \(\mathbf {M}^e_C\) and \(\mathbf {M}^e_L\), respectively. This illustrates another requirement: the CMM and DLMM forms must be instances of the mass template.

Finally we can apply the nonnegativity constraint. For the two principal minors of \(\mathbf {M}^e_\mu \) to be nonnegative, \(2+\mu \ge 0\) and \((2+\mu )^2-(1-\mu )^2=3+6\mu \ge 0\). Both are satisfied if \(\mu \ge -1/2\). Unlike the others, this constraint is of inequality type, and only limits the range of \(\mu \).

The remaining task is to select the parameter. This is done by introducing an optimality criterion that fits the problem at hand. This is where customization comes in. Even for this simple case the answer is not unique. Thus the statement “the best mass matrix for Bar2 is so-and-so” has to be qualified. Two specific optimization criteria are considered in Sects. 4.4 and 4.5.

4.3 Bar2 Alternative Parametrization

An alternative template expression that is useful in some investigations, such as those undertaken in Appendix 6, is obtained by reparametrizing via \(\chi =1+2\mu \), the inverse of which is \(\mu =(\chi -1)/2\). The resulting form is

$$\begin{aligned} \mathbf {M}^e_\chi = {\textstyle {1\over 12}}\,\rho A\,\ell \left[ \begin{array}{ll} 3+\chi &{}\quad 3-\chi \\ 3-\chi &{}\quad 3+\chi \\ \end{array}\right] . \end{aligned}$$
(10)

This is called the “\(\chi \) form” of the general Bar2 mass template. Because the determinant is \(\rho A\, \ell \chi \), \(\mathbf {M}_{\chi }^e\) is seen to be singular if \(\chi =0\), and nonnegative if \(\chi \ge 0\).

4.4 Bar2 Angular Momentum Conservation

This criterion can only be applied in multiple dimensions, since element-transverse angular rotations do not exist in 1D. Accordingly we allow the bar to move in the \(\{x,y\}\) plane by expanding its nodal DOF to \(\mathbf {u}^e= \left[ \begin{array}{llll} u_{x1}&\quad u_{y1}&\quad u_{x2}&\quad u_{y2} \end{array}\right] ^T\), whence (9) becomes a \(4\times 4\) matrix

$$\begin{aligned} \mathbf {M}^e_\mu = {\textstyle {1\over 6}}\rho A\,\ell \left[ \begin{array}{llll} 2+\mu &{}\quad 0 &{}\quad 1-\mu &{}\quad 0\\ 0 &{}\quad 2+\mu &{}\quad 0 &{}\quad 1-\mu \\ 1-\mu &{}\quad 0 &{}\quad 2+\mu &{}\quad 0\\ 0 &{}\quad 1-\mu &{}\quad 0 &{}\quad 2+\mu \\ \end{array}\right] \end{aligned}$$
(11)

Apply a uniform angular velocity \(\dot{\theta }\) about the midpoint. The associated node velocity vector at \(\theta =0\) is \(\dot{\mathbf {u}}^e={\textstyle {1\over 2}}\ell \dot{\theta }\left[ \begin{array}{lllll}0&{}-1&{}0&{}1\\ \end{array}\right] ^T\). The discrete and continuum energies are

$$\begin{aligned} T^e_\mu&= {\textstyle {1\over 2}}(\dot{\mathbf {u}}^e)^T\mathbf {M}^e_\mu \dot{\mathbf {u}}^e= {\textstyle {1\over 24}}\rho A\ell ^3(1+2\mu ), \nonumber \\ T^e&= \int _{-\ell /2}^{\ell /2}\rho A\big (\dot{\theta }\,x\big )^2\,dx = {\textstyle {1\over 24}}\rho A\ell ^3. \end{aligned}$$
(12)

Matching \(T^e_\mu =T^e\) gives \(\mu =0\). So according to this criterion the optimal mass matrix is the consistent one (CMM). Note that if \(\mu =1,\, T^e_\mu =3T^e\), whence the DLMM overestimates the element rotational (rotary) inertia by a factor of three.

4.5 Bar2 Fourier Analysis

For longitudinal motions, a more useful customization criterion is to improve accuracy in the long wavelength, low-frequency limit; this is labeled low frequency continuum fit (LFCF) in Table 1. This is carried out by a well known tool: Fourier analysis. Physical interpretation: probe the fidelity with which planes waves are propagated over a FEM-discretized regular lattice, when compared to the propagation over a continuum bar. The essentials are illustrated in Fig. 2. The top half depicts the continuum bar whereas the bottom half shows stages of the Fourier analysis of its FEM-discretized counterpart.

Fig. 2
figure 2

Propagation of a harmonic plane wave over an infinite, prismatic, elastic bar: a propagation over a continuum bar, b FEM discretization as infinite regular lattice, c propagation of plane wave over Bar2-discretized lattice, d extraction of a typical two-element patch. For visualization convenience, the wave-profile axial displacement \(u(x,t)\) is plotted normal to the bar

Symbols used for the analysis of plane wave propagation are collected in Table 6 for the reader’s convenience (the same notation is reused in later Sections). Corresponding nomenclature for the FEM-discretized two-node bar lattice is collected in Table 7. The continuum-versus-lattice notational rule is: corresponding quantities use the same symbol but the zero subscript is suppressed in the lattice. For example, the continuum wavelength \(\lambda _0\) becomes the lattice wavelength \(\lambda \).

Table 6 Nomenclature for harmonic plane wave propagation over continuum bar
Table 7 Nomenclature for harmonic plane wave propagation over Bar2 lattice

Plane wave propagation over a regular spring-mass lattice is governed by the semidiscrete linear EOM:

$$\begin{aligned} \mathbf{M}\ddot{\mathbf{u}}\mathbf{+K u=0,} \end{aligned}$$
(13)

in which \(\mathbf {M}\) and \(\mathbf {K}\) are infinite, tridiagonal Toeplitz matrices. This EOM can be solved by Fourier methods. Figure 2b displays two characteristic lengths: \(\lambda \) and \(\ell \). The element length-to-wavelength ratio is called \(\Upsilon =\ell /\lambda \). The floor function of its inverse: \(N_{e\lambda }=\left\lfloor \lambda /\ell \right\rfloor \) is the number of elements per wavelength. Those ratios characterize the fineness of the discretization, as illustrated in Fig. 2b.

Within constraints noted later the lattice can propagate real, travelling, harmonic plane waves of wavelength \(\lambda \) and grpup velocity \(c\), as depicted in Fig. 2b, c. The wavenumber is \(k=2\pi /\lambda \) and the circular frequency \(\omega =2\pi /T=2\pi c/\lambda =k\,c\). The range of wavelengths that the lattice may transport is illustrated in Fig. 3.

Fig. 3
figure 3

Selected plane waves of various wavelengths, illustrating the physical meaning of the dimensionless wavenumber (DWN) \(\kappa =k\ell =2\pi \ell /\lambda \). a, b, c, d show cases \(\kappa =0, \pi /4, \pi \) and \(2\pi \), respectively. The number of elements per wavelength is \(N_{e\lambda }= \left\lfloor \lambda /\ell \right\rfloor = \left\lfloor 2\pi /\kappa \right\rfloor \), in which \(\left\lfloor .\right\rfloor \) denotes the floor function. (This is equivalent to the spatial sampling rate of filter technology). The case \(\lambda =2\ell \) pictured in c pertains to the folding or Nyquist frequency, at which \(\kappa =\pi ,\, N_{e\lambda }=2\), and the group velocity \(c\) vanishes

To study plane wave solutions it is sufficient to extract a two-element patch, a process depicted in Fig. 2d. A harmonic plane wave of amplitude \(B\) is described by the function

$$\begin{aligned} u(x,t)&= B\,\exp \left[ j(k x -\omega \,t)\right] \nonumber \\&= B\,\exp \left[ j\big (\kappa x-\Omega \,c_0 \,t\big )\big /\ell \right] , \quad j=\sqrt{-1}. \end{aligned}$$
(14)

Here the dimensionless wavenumber \(\kappa \) and dimensionless circular frequency \(\Omega \) were introduced as \(\kappa =k\,\ell =2\pi \ell /\lambda =2\pi \chi \) and \(\Omega =\omega \,\ell /c_0\), respectively. Here \(c_0=\sqrt{E/\rho }\) is the elastic bar group velocity, which for the continuum is the same as the phase velocity (in physical acoustics \(c_0\) is the sound speed of the material). Using the well-known Bar2 static stiffness matrix and the mass template (9) gives the patch equations

$$\begin{aligned}&{\rho A\ell \over 6} \left[ \begin{array}{ccc} 2+\mu &{}\quad 1-\mu &{}\quad 0\\ 1-\mu &{}\quad 4+2\mu &{}\quad 1-\mu \\ 0&{}\quad 1-\mu &{}\quad 2+\mu \\ \end{array}\right] \left[ \begin{array}{l} \ddot{u}_{j-1}\\ \ddot{u}_j\\ \ddot{u}_{j+1} \end{array}\right] \nonumber \\&\qquad +{EA\over \ell } \left[ \begin{array}{ccc} 1&{}\quad -1&{}\quad 0 \\ -1&{}\quad 2&{}\quad -1\\ 0&{}\quad -1&{}\quad 1 \end{array}\right] \left[ \begin{array}{l} u_{j-1}\\ u_j\\ u_{j+1} \end{array}\right] =0. \end{aligned}$$
(15)

From this one takes the middle (node \(j\)) equation, which repeats in the infinite lattice:

$$\begin{aligned}&{\rho A\ell \over 6} \left[ \begin{array}{lll} 1-\mu&\quad 4+2\mu&\quad 1-\mu \end{array}\right] \left[ \begin{array}{l} \ddot{u}_{j-1}\\ \ddot{u}_j\\ \ddot{u}_{j+1} \end{array}\right] \nonumber \\&\qquad +{EA\over \ell } \left[ \begin{array}{lll} -1&{}\quad 2&{}\quad -1\\ \end{array}\right] \left[ \begin{array}{l} u_{j-1}\\ u_j\\ u_{j+1} \end{array}\right] =0. \end{aligned}$$
(16)

Evaluate the wave motion (14) at \(x=x_{j-1}=x_j-\ell ,\, x=x_j\) and \(x=x_{j+1}=x_j+\ell \,\) while keeping \(t\) continuous. Substitution into (16) gives the wave propagation condition

$$\begin{aligned}&{\rho A \,c_0^2\over 3\ell } \Big [6-(2 +\mu )\Omega ^2 -\big (6-(1 -\mu )\Omega ^2\big )\cos \kappa \Big ] \nonumber \\&\qquad \qquad \times \, \left( \cos {\Omega \,c_0 \,t\over \ell }- i\sin {\Omega \, c_0 \,t\over \ell }\right) B=0. \end{aligned}$$
(17)

If this is to be zero for any \(t\) and \(B\), the expression in brackets, called the characteristic equation, must vanish. Solving gives the dimensionless frequency versus wavenumber relation

$$\begin{aligned} \Omega ^2&= {6(1-\cos \kappa )\over 2+\mu +(1-\mu )\cos \kappa } \nonumber \\&= \kappa ^2+{1-2\mu \over 12} \,\kappa ^4+C_6\,\kappa ^6+\cdots \end{aligned}$$
(18)

in which \(C_6=(1-10\mu +10\mu ^2)/360\). Its inverse is

$$\begin{aligned} \kappa&= \arccos \left[ {6-(2 +\mu )\Omega ^2\over 6+(1 -\mu )\Omega ^2}\right] \nonumber \\&= \Omega -{ 1-2\mu \over 24}\Omega ^3+{9-20\mu +20\mu ^2\over 1920}\Omega ^5+\cdots \end{aligned}$$
(19)

Transforming (18) to physical wavenumber \(k=\kappa /\ell \) and circular frequency \(\omega =\Omega \,c_0/\ell \) gives

$$\begin{aligned} \omega ^2&= \left( {6c_0^2\over \ell ^2}\right) {1-\cos (k\ell )\over 2+\mu +(1{-}\mu )\cos (k\ell )}\nonumber \\&= c_0^2 \,k^2 \left( 1+{1{-}2\mu \over 12}k^2\,\ell ^2+C_6\,k^4\ell ^4+\cdots \right) \end{aligned}$$
(20)

4.6 Bar2 Dispersion Diagrams

An equation that links frequency and wavenumber: \(\Omega =\Omega (\kappa )\) as in (18), or \(\omega =\omega (k)\), as in (20), is a dispersion relation. A plot of the dispersion relation with \(k\) and \(\omega \) along horizontal and vertical axes, respectively, is called a dispersion diagram. When this is done in terms of dimensionless wavenumber \(\kappa \) and dimensionless frequency \(\Omega \), the plot is called a dimensionless dispersion diagram (DDD). Such diagrams exhibit a \(2\pi \) period: \(\Omega (\kappa )=\Omega (\kappa +2\pi n)\) for integer \(n\). Thus it is enough to plot \(\Omega (\pi )\) over either \([-\pi ,\pi ]\) or \([0,2\pi ]\), a range called a Brillouin zone. All DDD in this paper use the \([0,2\pi ]\) range choice.

Why is \(\Omega =0\) at \(\kappa =2\pi \)? The wavelenth \(\lambda =\ell \) pictured in Fig. 3d has the same value at all nodes for each time \(t\). This nodal sampling cannot be distinguished from the case \(\lambda =\infty \) (that is, \(\kappa =0\)) shown in Fig. 3a. They must share the same frequency, which is zero. Associated plane waves propagate with the same speed but in opposite directions. Similar arguments can be adduced to justify the dispersion curve symmetry about wavenumber \(\kappa =\pi \), as well as the \(2\pi \) periodicity.

4.7 Best \(\mu \) By Low Frequency Fitting

An oscillatory dynamical system is nondispersive if \(\omega \) is linear in \(k\), in which case \(c=\omega /k\) is constant and the wavespeed (the group velocity) is the same for all frequencies. The physical dispersion relation for the continuum bar is \(c_0=\omega _0/k_0=\sqrt{E/\rho }\). Hence all waves propagate with the same speed in this model. Group and phase velocities coalesce.

The FEM-discretized lattice group velocity is \(c=\partial \omega /\partial k=c_0(\partial \Omega /\partial \kappa )\), which differs from \(c_0\) except at \(\omega =\kappa =0\). The Bar2 discrete model is dispersive for any fixed \(\mu \), since from (20) we get

$$\begin{aligned} \gamma _c&= {c\over c_0} = {\partial \Omega \over \partial \kappa }= {1\over \kappa } \sqrt{6(1-\cos \kappa ) \over 2 +\mu +(1-\mu ) \cos \kappa } \nonumber \\&= 1+{1-2\mu \over 24}\kappa ^2+{1-20\mu +20\mu ^2\over 1920} \kappa ^4+\cdots \end{aligned}$$
(21)

Plainly the best fit to the continuum for small wavenumbers \(\kappa =k\ell {<}{<}1\) is obtained by taking \(\mu =1/2\), which makes the second term of the series (18) or (21) vanish. So for LFCF customization the best mass matrix is the average of the lumped and consistent ones:

$$\begin{aligned} \mathbf {M}^e_{BLFM}=\left. \mathbf {M}^e_\mu \right| _{\mu ={\textstyle {1\over 2}}}= {\textstyle {1\over 2}}\mathbf {M}^e_C+{\textstyle {1\over 2}}\mathbf {M}^e_L= {\rho A\ell \over 12} \left[ \begin{array}{ll} 5&{}\quad 1 \\ 1 &{}\quad 5\\ \end{array}\right] \!. \end{aligned}$$
(22)

This instance is labeled BLFM, for best low-frequency match. Figure 4a plots the dimensionless dispersion relation (18) for the CMM (\(\mu =0\)), DLMM (\(\mu =1\)) and BLFM (\(\mu ={\textstyle {1\over 2}}\)) mass matrices, along with the continuum-bar relation \(\Omega _0=\kappa _0\). The superior small-\(\kappa \) fit provided by the BLFM is evident.

Fig. 4
figure 4

Results from Fourier analysis of Bar2 infinite regular lattice for three choices of \(\mu \), plus continuum: a DDD, b dimensionless group velocity diagram (DGVD)

4.8 Folding Frequency

The maximum lattice frequency occurs at the folding wavenumber \(\kappa =k\ell =\pi \) or \(\lambda =2\ell \), which is waveform (c) in Fig. 3. The sampling rate \(N_{e\lambda }\) is then 2 elements per wavelength. This is called the folding or Nyquist frequency, and is denoted as

$$\begin{aligned} \Omega ^2_{af}={12\over 1+2\,\mu }. \end{aligned}$$
(23)

(The “a” in the subscript stands for AB; this notation is explained in Sect. 5.1). This varies from \(\Omega _{af}=\sqrt{12}=2\sqrt{3}\) for the CMM through \(\Omega _{af}=2\) for the DLMM. Frequencies higher than \(\Omega _{af}\) cannot be propagated over the lattice. As shown in Fig. 4b, the lattice wavespeed vanishes at the folding wavenumber \(\kappa =\pi \), and is negative over the range \((\pi ,2\pi ]\). Waveforms in that rage move with negative speed: \(c<0\). As discussed in Sect. 4.6, the waveform with \(\ell =\lambda \), or \(\kappa =2\pi \), cannot be distinguished from a rigid motion such as that pictured in Fig. 3a, and the lattice frequency falls to zero.

4.9 Bar2 Test: Vibrations of a Fixed-Free Bar Member

Natural frequency predictions of three Bar2 template instances are compared for predicting natural frequencies of longitudinal vibrations of the fixed-free elastic bar member pictured in Fig. 5. The member is prismatic, with constant \(E=1,\, A=1\), and \(\rho =1\). The total member length is taken as \(L=\pi /2\) for convenience. With those numerical properties the continuum eigenfrequencies are

Fig. 5
figure 5

Fixed-free homogeneous prismatic elastic bar member used in vibration test for Bar2 and Bar3 template instances. Both pictured discretizations display 4 elements. a Member modeled with Bar2 elements; results reported in Table 8 and Fig. 6, b member modeled with Bar3 elements; results reported in Fig. 12

Fig. 6
figure 6

Performance of selected Bar2 template instances in predicting the first three natural frequencies \(\omega _i,\, i=1,2,3\) of the fixed-free prismatic homogeneous bar shown in Fig. 5a. a, b, c compare against frequencies \(\omega 1=1, \omega 2=3\) and \(\omega 3=5\), respectively. This is a graphical, log-log representation of the results of Table  8. Horizontal axis shows number of elements while vertical axis displays correct digits of computed frequency. See text for details of what is shown along each axis

Table 8 Bar2 instance results for vibrations of a fixed-free bar member
$$\begin{aligned} \omega _{0i}={(2i-1)\pi \over 2L}\sqrt{E\over \rho }= 2i-1, \quad i=1,2,3,\ldots \end{aligned}$$
(24)

The member is divided into \(N_e\) identical elements, with \( N_e=1,2,\ldots 16\). Figure 5a pictures the case \(N_e=4\). Three template instances are compared: CMM (\(\mu =0\)), DLMM (\(\mu =1\)) and BLFM (\(\mu =1/2\)). Numerical results obtained for the first three frequencies are collected in Table 10. The \(O(\kappa ^4)\) convergence of BLFM is obvious. For example, 4 elements give \(\omega _2\) correct to 4 digits while both CMM and DLMM, which converge as \(O(\kappa ^2)\), give only 2. As expected, CMM overestimates the continuum frequencies while DLMM underestimates them.

The results of Table 8 are graphically reformatted in Fig. 6, as accuracy versus elements log-log plots. The horizontal axis shows number of elements \(N_e\) in \(\log _2\) scale. The vertical axis displays correct digits of computed frequency, obtained as

$$\begin{aligned} d=-\log _{10}\left| \Delta \omega _i\right| ,\quad \hbox {in which}\quad \Delta \omega _i=\omega _i-\omega _{0i}. \end{aligned}$$
(25)

Here \(\Delta \omega _i\) is the frequency error of computed values with respect to continuum frequencies \(\omega _{0i}=2\,i-1\), given by (24). The plots clearly show at a glance that, for the same \(N_e\), BLFM roughly doubles the number of correct digits provided by the other two instances. It also illustrates that CMM and DLMM give the same error magnitude (within plot accuracy) although of different signs. Thus log-log plots such as those in Fig. 6 are unable to show whether the convergence is from above or below, because of the taking of absolute values in (25). That visualization deficiency should be kept in mind should error signs become important.

4.10 Other Customization Options

The last three customization options listed in Table 1 are not relevant to this element. RHFP is unnecessary because the dispersion diagram does not have an OB. MSTS is pointless because the DLMM in (7) is unique. Finally, RDAW does not apply to 1D elements.

4.11 Bar2 Frequency Dependent Mass

As noted in Sect. 3.9, it is occasionally useful to make the mass and/or stiffness matrix frequency dependent. The goal is to exactly match the continuum dispersion relation \(\Omega =\kappa \) for all frequencies, or at least a finite range that includes \(\Omega =\kappa =0\). Such an exact fit allows for coarser discretizations. The cost paid is that matrix entries become trigonometric functions of frequency. Both the EOM and associated eigenproblems become trascendental.

Unless the frequency is specified beforehand (for example, in pure harmonic excitation) an iterative process is unavoidable. Therefore “exactness” gains might be illusory: the dog chases its own tail. Early publications that follow this approach are cited in Sect. 1. For reasons indicated there, those formulations are not necessarily instances of the general template derived in Sect. 4.12.

The simplest way to introduce frequency dependency is to allow the mass template parameter \(\mu \) in (9) to be frequency dependent, while the stiffness matrix is held fixed. To find the expression of \(\mu \), set \(\kappa \rightarrow \Omega \) in the characteristic equation extracted from (17):

$$\begin{aligned} 6-(2 +\mu ^\omega )\Omega ^2 -\big (6-(1 -\mu ^\omega )\Omega ^2\big )\cos \Omega =0, \end{aligned}$$
(26)

in which \(\mu \) has been renamed \(\mu ^\omega \). Solving for it gives

$$\begin{aligned} \mu ^\omega = 1+{6\over \Omega ^2}-{3\over 1{-}\cos \Omega }= {1\over 2}-{\Omega ^2\over 40}-{\Omega ^4\over 1008}-{\Omega ^6\over 28800}-\cdots \end{aligned}$$
(27)

Since \(\kappa =\Omega \) for the continuum,

$$\begin{aligned} \mu ^\omega = 1+{6\over \kappa ^2}-{3\over 1{-}\cos \kappa } = {1\over 2}-{\kappa ^2\over 40}-{\kappa ^4\over 1008}-{\kappa ^6\over 28800}-\cdots \end{aligned}$$
(28)

As \(\Omega \rightarrow 0\) or \(\kappa \rightarrow 0\) both (27) and (28) approach \(0/0\). The indeterminacy is removed by the Taylor expansions given above, which show that the limit is \(\mu ^\omega \rightarrow {\textstyle {1\over 2}}\), as may be expected. As \(\Omega \) or \(\kappa \) grows, \(\mu ^\omega \) decreases so the template gradually favors the CMM more. Two interesting values should be noted. If \(\kappa =3.38742306673364,\, \mu ^\omega =0\), which makes the CMM frequency exact; this occurs at the intersection of the continuum and CMM dispersion curves in Fig. 4a. If \(\kappa =\kappa _{lim}=4.05751567622863,\, \mu ^\omega =-1/2\), which makes \(\mathbf {M}^e\) singular. If \(\kappa >\kappa _{lim},\, \mathbf {M}^e\) becomes indefinite. It follows that the match (27) or (28) is practically limited to the DWN range \(0\le \kappa < 4\).

4.12 Bar2 Frequency Dependent Mass-Stiffness Pair

The most general FDMS template for Bar2 has 8 free parameters. These are chosen as deviations from the optimal frequency-independent matrices:

$$\begin{aligned} \mathbf {M}^e&= C_M\left( \left[ \begin{array}{ll} 5 &{} \quad 1\\ 1&{} \quad 5\\ \end{array}\right] + \left[ \begin{array}{cc} \mu _{11}^\omega &{} \quad -\mu _{12}^\omega \\ -\mu _{21}^\omega &{}\quad \mu _{22}^\omega \\ \end{array}\right] \right) ,\,\, \mathbf {K}^e\nonumber \\&= C_K\left( \left[ \begin{array}{cc} 1 &{} \quad -1\\ -1&{} \quad 1\\ \end{array}\right] + \left[ \begin{array}{cc} \beta _{11}^\omega &{}\quad -\beta _{12}^\omega \\ -\beta _{21}^\omega &{}\quad \beta _{22}^\omega \\ \end{array}\right] \right) , \end{aligned}$$
(29)

in which \(C_M=\rho \,A\ell /12\) and \(C_K=E\,A/\ell \). All parameters may be frequency dependent. For brevity that dependency will not be explicitly shown unless necessary. If all \(\mu _{ij}^\omega \) vanish, \(\mathbf {M}^e\) reduces to (22), which is BLFM optimal. If all \(\beta _{ij}^\omega \) vanish, \(\mathbf {K}^e\) reduces to the well known stiffness of a 2-node prismatic bar. Thus in the zero-frequency (static) limit all parameters must vanish, which provides useful checks. To cut down on parameters, we impose diagonal and antidiagonal symmetry conditions a priori: \(\mu _{21}^\omega =\mu _{12}^\omega \), \(\mu _{22}^\omega =\mu _{11}^\omega \), \(\beta _{21}^\omega =\beta _{12}^\omega \), and \(\beta _{22}^\omega =\beta _{11}^\omega \). In addition setting \(\beta _{12}^\omega =\beta _{21}^\omega =\beta _{11}^\omega \) avoids singularities in the static limit, as noted later. Thus (29) reduces to

$$\begin{aligned} \mathbf {M}^e&= C_M\left( \left[ \begin{array}{ll} 5 &{}\quad 1\\ 1&{}\quad 5\\ \end{array}\right] + \left[ \begin{array}{cc} \mu _{11}^\omega &{}\quad -\mu _{12}^\omega \\ -\mu _{12}^\omega &{}\quad \mu _{11}\\ \end{array}\right] \right) ,\quad \mathbf {K}^e\nonumber \\&= C_K\left( \left[ \begin{array}{cc} 1 &{}\quad -1\\ -1&{}\quad 1\\ \end{array}\right] +\left[ \begin{array}{cc} \beta _{11}^\omega &{}\quad -\beta _{11}^\omega \\ -\beta _{11}^\omega &{}\quad \beta _{11}^\omega \\ \end{array}\right] \right) \!. \end{aligned}$$
(30)

which has 3 free parameters: \(\mu _{11}^\omega \), \(\mu _{12}^\omega \) and \(\beta _{11}^\omega \). These matrices are nonnegative if

$$\begin{aligned} 4+\mu _{11}^\omega +\mu _{12}^\omega \ge 0, \quad 6+\mu _{11}^\omega -\mu _{12}^\omega \ge 0, \quad 1+\beta _{11}^\omega \ge 0.\nonumber \\ \end{aligned}$$
(31)

Imposing the plane wave motion (14) on a two-element patch, extracting the middle node equation and dropping extraneous factors yields the complex characteristic equation

$$\begin{aligned} \Big [12 (1{+}\beta _{11})&- (5{+}\mu _{11}^\omega )\,\Omega ^2-\big (12+12 \beta _{11}\nonumber \\&+ (1{-}\mu _{12}^\omega )\,\Omega ^2\big )\cos \kappa \Big ]\,\exp (j\kappa )=0. \end{aligned}$$
(32)

Since the complex exponential never vanishes, it may be dropped and (32) reduces to the real equation

$$\begin{aligned} 12 \,(1{+}\beta _{11}^\omega )&- (5{+}\mu _{11}^\omega )\,\Omega ^2\nonumber \\&- \big (12+12\beta _{11}^\omega + (1{-}\mu _{12}^\omega )\,\Omega ^2\big )\cos \kappa =0. \end{aligned}$$
(33)

To match the continuum, \(\Omega \) is replaced by \(\kappa \), whence

$$\begin{aligned} f_{cm}&= 12 \,(1{+}\beta _{11}^\omega )-(5{+}\mu _{11}^\omega )\,\kappa ^2- (12+12\beta _{11}^\omega \nonumber \\&+(1{-}\mu _{12}^\omega )\,\kappa ^2)\cos \kappa =0. \end{aligned}$$
(34)

This establishes a linear constraint among the 3 parameters. Consider these as functions of \(\kappa \): \(\beta _{11}^\omega =\beta _{11}^\omega (\kappa )\), \(\mu _{11}^\omega =\mu _{11}^\omega (\kappa )\), and \(\mu _{12}^\omega =\mu _{12}^\omega (\kappa )\). Expanding in Taylor series about \(\kappa =0\) yields

$$\begin{aligned} f_{cm}&= \big (6 \left. \beta _{11}^\omega \right| _0 - \left. \mu _{11}^\omega \right| _0 + \left. \mu _{12}^\omega \right| _0\big )\,\kappa ^2\nonumber \\&+\left( 6\left. {\partial \beta _{11}^\omega \over \partial \kappa }\right| _0 - \left. {\partial \mu _{11}^\omega \over \partial \kappa }\right| _0 + \left. {\partial \mu _{12}^\omega \over \partial \kappa }\right| _0\right) \,\kappa ^3 + \cdots = 0,\nonumber \\ \end{aligned}$$
(35)

This shows that in the static limit \(\Omega =\kappa =0\) the continuum equation is identically satisfied. If \(\left. \beta _{12}\right| _0\ne \left. \beta _{11}^\omega \right| _0\), however, a term in \(\kappa ^{-2}\) appears in (35); this is the reason for presetting \(\beta _{12}^\omega =\beta _{11}^\omega \).

Further developments depend on which parameter pair is kept. Table 9 lists three possibilities: (\(\beta _{11}^\omega ,\mu _{11}^\omega \)), (\(\beta _{11}^\omega ,\mu _{12}^\omega \)), and (\(\mu _{11}^\omega ,\mu _{12}^\omega \)).

Table 9 General FDMS template for Bar2

4.13 Bar2 Frequency Dependent Mass Instances

Some relatively simple FDM instances can be obtained by setting \(\beta _{11}^\omega =0\) in (30). Taking \(\mu _{12}^\omega =\mu _{11}^{\omega }\) and solving for the latter gives

$$\begin{aligned} \mu _{11}^\omega =\mu _{12}^{\omega }&= 1 + {12\over \kappa ^2} - {6\over 1-\cos \kappa }\nonumber \\&= -{\kappa ^2\over 20}-{\kappa ^4\over 540}-{\kappa ^6\over 14400}-\cdots \end{aligned}$$
(36)

The resulting \(\mathbf {M}^e\) is indefinite if \(\,\kappa >4.05752\). This is the equivalent of the FDM instance considered in Sect. 4.11. The difference between (36) and (28) lies in the choice of baseline matrix for null free parameters. On the other hand, setting \(\mu _{12}^\omega =0\) along with \(\beta _{11}^\omega =0\) yields

$$\begin{aligned} \mu _{11}^\omega&= {-12 + 5\kappa ^2 +(12+\kappa ^2)\,\cos \kappa \over \kappa ^2}\nonumber \\&= -{\kappa ^4\over 40}-{11\kappa ^6\over 14400}-\cdots , \end{aligned}$$
(37)

This correction is smaller than (36) if \(\kappa <\pi /2\). The resulting \(\mathbf {M}^e\) is indefinite if \(\,\kappa >4.46192\).

5 The Three-Node Bar Element

The three-node bar element configuration is shown in Fig. 7a. The element is prismatic with length \(\ell =L^e\), uniform cross section area \(A\) and mass density \(\rho \). Midnode 3 is at the center. The element DOFs are arranged as \(\mathbf {u}^e = \left[ \begin{array}{lllll} u_1 &{} u_2 &{} u_3\\ \end{array}\right] ^T\). The element name is often abbreviated to Bar3 in the sequel. We will consider only frequency independent templates here.

Fig. 7
figure 7

Three-node prismatic bar element: a configuration, b extraction of two-element patch from a regular lattice

Despite its simplicity, the Bar3 template is sufficiently feature-rich so it can be used to illustrate most of the customization scenarios listed in Table 1. Two reasons: it has multiple dispersion branches, and the stiffness has a free parameter. But additional terminology on dispersion diagrams has to be introduced first. Readers familiar with that topic should skip to Sect. 5.2.

5.1 Dispersion Diagram Terminology

The characteristic equation of the Bar3 element, derived in Sect. 5.5 below, gives two positive real frequencies for each plane wave wavelength. The dimensionless forms are identified by \(\Omega _a\) and \(\Omega _o\), ordered so \(\Omega _a\le \Omega _o\). The functions \(\Omega _a(\kappa )\) and \(\Omega _o(\kappa )\), in which \(\kappa \) is the dimensionless wavenumber, are called acoustic and optical branches, respectively, of the DDD. This terminology originated in crystal physics, in which both branches have physical meaning in modeling molecular level oscillations (in crystallography, acoustic waves are lower frequency waves caused by sonic-like disturbances, in which adjacent molecules move in the same direction. Optical waves are higher frequency oscillations caused by interaction with light or electromagnetics, in which adjacent molecules move in opposite directions. Textbook references are provided in Sect. 1).

In FEM discretization work, only the AB has physical meaning because for small \(\kappa \) (that is, long wavelengths) it approaches the continuum bar relation \(\Omega =\kappa \), as plainly illustrated by the \(\Omega _a^2\) series in Sect. 4.5. On the other hand, the OB is spurious. It is caused by the discretization and pertains to higher frequency lattice oscillations, also known as “mesh modes.’

Figure 8 displays nomenclature used for a two-branch dispersion diagram, such as that exhibited by the Bar3 element. As noted, the AB is the long-wavelength counterpart of the continuum model, for which \(\Omega =\kappa \); thus \(\left. \Omega _a\right| _{\kappa \rightarrow 0}=0\). On the other hand, the OB has a nonzero frequency \(\Omega _{oc}=\left. \Omega _o\right| _{\kappa \rightarrow 0}\) called the cutoff frequency (COF) that cannot vanish, although it may go to infinity under certain conditions; such as a singular mass matrix. Also of interest are the values of \(\Omega _a\) and \(\Omega _o\) at the folding frequency \(\kappa _f=\pi \); these are denoted by \(\Omega _{af}\) and \(\Omega _{of}\), respectively. The lowest and highest values of \(\Omega _o\) are called \(\Omega _o^{max}\) and \(\Omega _o^{min}\), respectively, whereas the largest \(\Omega _a\) is called \(\Omega _a^{max}\). For the plots drawn in Fig. 8 (note disclaimer on the right):

$$\begin{aligned} \Omega _a^{max}=\Omega _{af}, \quad \Omega _o^{min}=\Omega _{of}, \quad \Omega _o^{max}=\Omega _{oc}. \end{aligned}$$
(38)

Often, but not always, \(\Omega _o^{min}\) and \(\Omega _a^{max}\) occur at \(\kappa =\pi \), the folding (Nyquist) wavenumber at which group velocities vanish. In any case, if \(\Omega _o^{min}>\Omega _a^{max}\), the frequency range \(\Omega _o^{min}>\Omega >\Omega _a^{max}\) is called the acoustoptical frequency gap. Frequencies within the gap are said to pertain to portion I of the stopping band or stopband, a term derived from filter technology. Frequencies \(\Omega >\Omega _o^{max}\) pertain to portion II of the stopping band.

Fig. 8
figure 8

Nomenclature for a two-branch dispersion diagram typical of 1D structural elements such as Bar3. The stopping band is the union of I and II (symmetry and monotonicity about the folding wavenumber \(\kappa _f=\pi \) is typical of prismatic, simple 1D elements in regular lattices; else those features are typically lost. In addition, multiple optical branches will appear for characteristic equations with more than two roots for each \(\kappa \))

A frequency that falls within a stopping band cannot propagate as plane wave over the FEM lattice, since there the characteristic equation has complex roots with negative real parts. This causes exponential attenuation so any periodic disturbance with that frequency will die out.

5.2 Bar3 General Mass-Stiffness Template

We begin by introducing a general template for the MS pair. The mass template is given four parameters:

$$\begin{aligned} \mathbf {M}_\mu ^e&= {m^e \over 30} \left[ \begin{array}{ccc} 4+\mu _1 &{} \quad -1+\mu _3 &{} \quad 2+\mu _4\\ -1+\mu _3 &{} \quad 4+\mu _1 &{} \quad 2+\mu _4 \\ 2+\mu _4 &{} \quad 2+\mu _4 &{} \quad 16+\mu _2 \\ \end{array}\right] \nonumber \\&= \mathbf {M}_{CMM}^e+ {m^e\over 30} \left[ \begin{array}{ccc} \mu _1 &{} \quad \mu _3 &{} \quad \mu _4 \\ \mu _3 &{} \quad \mu _1 &{} \quad \mu _4 \\ \mu _4 &{} \quad \mu _4 &{} \quad \mu _2 \end{array}\right] \!. \end{aligned}$$
(39)

in which \(m^e=\rho \,A\,\ell \). Here \(\mathbf {M}_{CMM}^e\) is the CMM, obtained for \(\mu _1=\mu _2=\mu _3=\mu _4=0\), which is derived in Sect. 1. The template (39) incorporates matrix, geometric and fabrication symmetries ab initio. It includes all DLMM by setting \(\mu _3=1\) and \(\mu _4=-2\). Because of its practical importance, however, that “lumped mass subset” is studied in Sect. 5.8 using a two-parameter template variant.

For (39) to be nonnegative definite (NND), three inequality constraints have to be satisfied. Those are more elegantly expressed in terms of the alternative “\(\chi \)-form” derived in Sect. 5.3.

Conservation of total element mass \(\rho \,A\,\ell \) (invariance of linear momentum) imposes the following homogeneous constraint:

$$\begin{aligned} 2\mu _1+\mu _2+2\mu _3+4\mu _4=0. \end{aligned}$$
(40)

This constraint is not always preimposed as it may complicate intermediate expressions, but it is eventually applied at some point. Conservation of angular momentum in 2D or 3D requires \(\mu _1=\mu _3\), as verified by the CMM. This is ignored, however, as it hinders customization.

As regards the stiffness matrix, the following one- parameter template is used:

$$\begin{aligned} \mathbf {K}^e&= \mathbf {K}_b^e+\mathbf {K}_h^e= k^e \left[ \begin{array}{ccc} 1 &{} \quad -1 &{} \quad 0 \\ -1 &{} \quad 1 &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 0 \\ \end{array}\right] \nonumber \\&+\, {4\,k^e\beta \over 3} \left[ \begin{array}{ccc} 1 &{} \quad 1 &{} \quad -2 \\ 1 &{} \quad 1 &{} \quad -2 \\ -2 &{} \quad -2 &{} \quad 4 \\ \end{array}\right] , \end{aligned}$$
(41)

in which \(k^e=EA/\ell \). Here \(\mathbf {K}_b^e\) and \(\mathbf {K}_h^e\) denote the basic and higher-order stiffness matrices, respectively. This decomposition was introduced by Bergan and coworkers in the 1980s for the development of the Free Formulation; references are provided in Appendix Sect. 1.7. The higher order stiffness is scaled by the free parameter \(\beta \ge 0\). Setting \(\beta =1\) produces the well known stiffness of the quadratic (isoparametric) displacement model, whereas \(\beta =0\) reduces \(\mathbf {K}^e\) to the Bar2 stiffness considered in Sect. 4.2.

5.3 Bar3 Alternative Mass Template

An alternative configuration of the general Bar3 mass template is obtained by changing the four \(\mu _i\) free parameters to three: \(\chi _1,\, \chi _2\), and \(\chi _3\), through the replacement rule

$$\begin{aligned} \mu _1&= \chi _1+\chi _2-4,\quad \mu _2=14+4\chi _1-4\chi _{13},\nonumber \\&\mu _3=\chi _1-\chi _2+1,\quad \mu _4=\chi _{13}-2\chi _1-2. \end{aligned}$$
(42)

in which \(\chi _{13}=\sqrt{30}\sqrt{\chi _1{-}\chi _3}\). The mass conservation condition (40) is identically satisfied by (42), which is why the number of free parameters can be cut by one. Conversely, if the \(\mu _i\) are given and do satisfy (40), the \(\chi _i\) can be computed from

$$\begin{aligned} \chi _1&= {\textstyle {1\over 2}}\,(3+\mu _1+\mu _3), \quad \chi _2={\textstyle {1\over 2}}(5+\mu _1-\mu _3),\nonumber \\ \chi _3&= \frac{1}{480}\,\big [4\mu _1(40+\mu _2-2\mu _3)+ 40\,(8+\mu _2+4\mu _3)\nonumber \\&-4\mu _1^2-(\mu _2-2\mu _3)^2\big ]. \end{aligned}$$
(43)

If \(\mu _1=\mu _2=\mu _3=0\), this gives \(\chi _1=3/2,\, \chi _2=5/2\) and \(\chi _3=2/3\) for the CMM. On inserting (42) into (39) the so-called “\(\chi \)-form” of the Bar3 mass template emerges:

$$\begin{aligned} \mathbf {M}_{\chi }^e&= {m^e\over 30} \nonumber \\&\times \left[ \begin{array}{ccc} \chi _1+\chi _2 &{} \quad \chi _1-\chi _2 &{} \quad -2 \chi _1+\chi _{13} \\ \chi _1-\chi _2 &{} \quad \chi _1+\chi _2 &{} \quad -2 \chi _1+\chi _{13} \\ -2 \chi _1+\chi _{13} &{} \quad -2 \chi _1+\chi _{13} &{} \quad 30+4 \chi _1-4 \chi _{13} \\ \end{array}\right] \nonumber \\ \end{aligned}$$
(44)

An attractive feature of (44) is that mass matrix admissibility can be readily correlated to parameter values. Specifically, \(\mathbf {M}_{\chi }^e\) is positive definite (PD) if and only if \(\chi _1,\, \chi _2\) and \(\chi _3\) are positive. This can be proven from the following properties:

$$\begin{aligned} \lambda _1(\mathbf {M}_{\chi }^e)&= {m^e\,\chi _2\over 15}, \quad \det (\mathbf {M}_{\chi }^e)={m^e\,\chi _2\,\chi _3\over 225},\quad \det (\mathbf {M}_{\chi \,2\times 2}^e)\nonumber \\&= {m^e\,\chi _1\,\chi _2\over 225}, \end{aligned}$$
(45)

The first equality gives one eigenvalue of \(\mathbf {M}_{\chi }^e\) (the other two have more complicated expressions), whence PD mandates \(\chi _2>0\). The last equalities give the determinants of \(\mathbf {M}_{\chi }^e\) and of its \(2\times 2\) upper principal minor, respectively. Accordingly, PD requires also \(\chi _3>0\) and \(\chi _1>0\). For \(\mathbf {M}_{\chi }^e\) to be NND, simply change \(>\) to \(\ge \). Those conditions can be harked back to the \(\mu _i\) of \(\mathbf {M}_\mu ^e\) using (43), but the expressions are noticeably messier. The second equality of (45) gives another nice feature: \(\mathbf {M}_{\chi }^e\) becomes singular if and only if \(\chi _2=0\), or \(\chi _3=0\), or both.

The main advantage of \(\mathbf {M}_\mu ^e\) over \(\mathbf {M}_{\chi }^e\) is the linear dependence of entries on the \(\mu _i\). This simplifies patch analysis as well as reparametrization for several template variants studied later.

5.4 Bar3 Patch Equations

To assess wave propagation and dispersion performance of the MS template defined by (39) and (41), we carry out the Fourier analysis of the infinite bar lattice shown in Fig. 7b. Extract a typical two node patch as illustrated. The patch has five nodes: three endpoints and two midpoints, which are assigned global numbers \(j{-}2,\, j{-}1, \,\ldots \,j{+}2\). The unforced semidiscrete dynamical equations of the patch are \(\mathbf {M}_p \,\ddot{\mathbf {u}}_P+\mathbf {K}_p\,\mathbf {u}_P=\mathbf {0}\), in which

$$\begin{aligned} \mathbf {M}_p&= {m^e\over 30}\nonumber \\&\times \left[ \begin{array}{lllll} 4 + \mu _1 &{} \quad 2 + \mu _4 &{} \quad -1 + \mu _3 &{} \quad 0 &{} \quad 0 \\ 2 + \mu _4 &{} \quad 16 + \mu _2 &{} \quad 2 + \mu _4 &{} \quad 0 &{} \quad 0 \\ -1 + \mu _3 &{} \quad 2 + \mu _4 &{} \quad 2 (4 + \mu _1) &{} \quad 2 + \mu _4 &{} \quad -1 + \mu _3 \\ 0 &{} \quad 0 &{} \quad 2 + \mu _4 &{} \quad 16 + \mu _2 &{} \quad 2 + \mu _4 \\ 0 &{} \quad 0 &{} \quad -1 + \mu _3 &{} \quad 2 + \mu _4 &{} \quad 4 + \mu _1 \end{array}\right] ,\nonumber \\ \mathbf {K}_p&= {k^e\over 3}\nonumber \\&\times \left[ \begin{array}{lllll} 3 + 4\,\beta &{} \quad -8\,\beta &{} \quad -3 + 4\,\beta &{} \quad 0 &{} \quad 0 \\ -8\,\beta &{} \quad 16\,\beta &{} \quad -8\,\beta &{} \quad 0 &{} \quad 0 \\ -3 + 4\,\beta &{} \quad -8\,\beta &{} \quad 6 + 8\,\beta &{} \quad -8\,\beta &{} \quad -3 + 4\,\beta \\ 0 &{} \quad 0 &{} \quad -8\,\beta &{} \quad 16\,\beta &{} \quad -8\,\beta \\ 0 &{} \quad 0 &{} \quad -3 + 4\,\beta &{} \quad -8\,\beta &{} \quad 3 + 4\,\beta \\ \end{array}\right] ,\nonumber \\ \mathbf {u}_P&= \left[ \begin{array}{ccccc} u_{j-2}&\quad u_{j-1}&\quad u_j&\quad u_{j+1}&\quad u_{j+2} \end{array}\right] ^T. \end{aligned}$$
(46)

Note that the element mass conservation constraint (40) is not preimposed as it would complicate intermediate expressions. It is enforced later. Keep the second and third equations, namely those for nodes \(j-1\) and \(j\). This selection picks the equations for a typical corner and midpoint node. Accordingly, the patch equations are

$$\begin{aligned} \widehat{\mathbf {M}}_p \ddot{\mathbf {u}}_P+\widehat{\mathbf {K}}_p \mathbf {u}_P=\mathbf {0}. \end{aligned}$$
(47)

The \(2\times 5\) matrices \(\widehat{\mathbf {M}}_p\) and \(\widehat{\mathbf {K}}_p\) result on deleting rows 1,4,5 of \(\mathbf {M}_p\) and \(\mathbf {K}_p\), respectively.

5.5 Bar3 Fourier Analysis

We study the propagation of harmonic plane waves of wavelength \(\lambda \), wavenumber \(k=2\pi /\lambda \), and circular frequency \(\omega \) over the lattice of Fig. 7b. For convenience they are separated into corner and midpoint waves:

$$\begin{aligned} u_c(x,t)&= B_c \,\exp \,\big (j (k x-\omega t)\big ), \nonumber \\ u_m(x,t)&= B_m \,\exp \,\big (j (k x-\omega t)\big ), \quad j=\sqrt{-1}. \end{aligned}$$
(48)

Wave \(u_c(x,t)\) propagates only over corner nodes and vanishes at midpoints, whereas \(u_m(x,t)\) propagates only over midpoints and vanishes at corner nodes. Both have the same wavenumber and frequency but different amplitudes and phases. (The wave pair (48) can be combined to form a single waveform that propagates over all nodes. The combination has two components that propagate at the same speed but in opposite directions. This is useful when studying boundary conditions or transitions in finite lattices, but unecessary for a periodic infinite lattice.)

As in Sect. 4.5, we will work with the dimensionless frequency \(\Omega =\omega \,\ell /c_0\) with \(c_0=\sqrt{E/\rho }\), and the dimensionless wavenumber \(\kappa =k\ell \). Inserting (48) into (47), passing to dimensionless variables, removing scale factors, and requiring that solutions exist for any \(t\) yields the characteristic equation

$$\begin{aligned}&\left[ \begin{array}{ll} {\textstyle {1\over 2}}A_3 (160 \beta {-}(16 {+}\mu _2) \Omega ^2) &{} \quad {-} A_1 A_3 (80 \beta {+}(2 {+}\mu _4) \Omega ^2) \\ {-} A_1 (80 \beta {+}(2 {+}\mu _4) \Omega ^2) &{} \quad 30 {+}40 \beta {-}(4{+}\mu _1) \Omega ^2 {+} A_1 ( {-}30 {+}40 \beta {+}(1{-}\mu _3) \Omega ^2)\\ \end{array}\right] \left[ \begin{array}{c} B_c\\ B_m\\ \end{array}\right] =\mathbf {0}. \end{aligned}$$
(49)

in which \(A_1=\cos \kappa ,\, A_2=\cos (\kappa /2)\) and \(A_3=\cos (\kappa /2) -j\sin (\kappa /2)\). For nontrivial solutions the determinant of the characteristic matrix must vanish, which provides a quadratic equation in \(\Omega ^2\). For each wavenumber \(\kappa \), solving the equation gives two squared frequencies. Their expressions, found by Mathematica, are

$$\begin{aligned} \Omega ^2_a=5 \,{P-\sqrt{Q}\over R}, \quad \Omega ^2_o=5 \,{P+\sqrt{Q}\over R}, \end{aligned}$$
(50)

in which coefficients \(P,\, Q\), and \(R\) are given by

$$\begin{aligned} P&=c_1+3\,c_2+c_9\,\cos \kappa , \nonumber \\ R&=c_5-4\,\mu _4-\mu _4^2+c_8\,\cos \kappa , \nonumber \\ Q&=192\,\beta \,(\cos \kappa -1)\,(c_5-c_3+c_8\,\cos \kappa )\nonumber \\&\quad +(c_1+3\,c_2 +c_9\,\cos \kappa )^2,\nonumber \\ c_1&=4\,\beta \,(40+4\,\mu _1+\mu _2+4\,\mu _4), \quad c_2=16+\mu _2,\nonumber \\ c_3&=\mu _4\,(4+\mu _4), \nonumber \\ c_4&=4\,(5+\mu _3+\mu _4), \quad c_5=60+4\,\mu _2+\mu _1\,c_2, \nonumber \\ c_6&=16\,\mu _3-c_3, \nonumber \\ c_7&=4\,\beta \,(\mu _2+c_4), \quad c_8=c_6+\mu _2\,(\mu _3-1)-20, \nonumber \\ c_9&=c_7-3\,c_2. \end{aligned}$$
(51)

Subscripts \(a\) and \(o\) stand for acoustic and optical branches, respectively, a terminology explained in Sect. 5.1. If \(\mathbf {M}^e\) is PD and the conservation condition (40) holds, the branch frequencies (50) have small \(\kappa \) (low frequency, long wavelength) Taylor series of the generic form

$$\begin{aligned} \Omega ^2_a&= \kappa ^2+ {C_4\,\kappa ^4\over 4!}+{C_6\,\kappa ^6\over 6!}+ {C_8\,\kappa ^8\over 8!} +\cdots , \nonumber \\ \Omega ^2_o&= D_0+{D_2\,\kappa ^2\over 2!}+{D_4\,\kappa ^4\over 4!} +\cdots , \end{aligned}$$
(52)

Coefficients \(C_n\) and \(D_n\) were obtained through the Mathematica built-in Series function up to \(n=10\) and are displayed for some interesting instances below.

5.6 Bar3 Standard Template Instances

We start by considering two instances available in the FEM literature since the mid 1960s. The CMM instance \(\mathbf {M}^e_C\) is obtained for \(\mu _1=\mu _2=\mu _3=\mu _4=0\). Using \(\beta =1\) for \(\mathbf {K}^e\) we get \(P=208+32\,\cos \kappa \), \(Q=128\,(237+224\cos \kappa -11\,\cos (2\kappa ))\), and \(R=20\,(3-\cos \kappa )\). The squared frequencies have the small-\(\kappa \) expansions

$$\begin{aligned} \Omega ^2_{a}&= \kappa ^2 + {\kappa ^6\over 720}- {11\,\kappa ^8\over 151200} + \cdots , \nonumber \\ \Omega ^2_{o}&= 60 - 20\,\kappa ^2 + {19\,\kappa ^4\over 3} + \cdots . \end{aligned}$$
(53)

The SLMM (Simpson-lumped diagonal mass matrix) instance derived in (146) of Appendix 3.1 results if \(\mu _1=1,\, \mu _2=4,\, \mu _3=1\), and \(\mu _4=-2\). Using \(\beta =1\) in (41) gives \(P=220+20\cos \kappa \), \(Q=200\,(147+140+\cos (2\kappa ))\), and \(R=100\). The squared frequencies have the small-\(\kappa \) expansions

$$\begin{aligned} \Omega ^2_{a} = \kappa ^2 - {\kappa ^6\over 1440} - {\kappa ^8\over 48384}\, + \cdots , \quad \Omega ^2_{o} = 24 - 2\kappa ^2 + {\kappa ^4\over 12}+ \cdots . \end{aligned}$$
(54)

For small \(\kappa \), SLMM fits the continuum better than CMM. Dispersion diagrams for the foregoing instances are plotted in Fig. 9a, c. Corresponding group velocity diagrams are shown in Fig. 9b, d. As in the case of the two-node bar, the consistent mass overestimates the continuum frequency \(\Omega =\kappa \) for \(0\le \kappa \le \pi \), whereas the lumped mass underestimates it.

Fig. 9
figure 9

DDD and DGVD plots for two Bar3 template instances treated in Sect. 5.6. a, b Diagrams for CMM instance, c, d diagrams for SLMM (Simpson DLMM) instance. Acoustic and optical branches shown in red and blue, respectively. Continuum case \(\Omega =\kappa \) and \(\gamma _c=1\) shown in black. (Color figure online)

5.7 Bar3 Low-Frequency Fitting

Inspection of the AB coefficient of \(\kappa ^6\) in (53) and (54) suggests combining one third of \(\mathbf {M}^e_C\) with two thirds of \(\mathbf {M}^e_L\) to cancel it. Setting

$$\begin{aligned} \mu _1=2/3, \quad \mu _2=8/3, \quad \mu _3=-2/3, \quad \mu _4=4/3, \quad \beta =1, \end{aligned}$$
(55)

in (41) gives

$$\begin{aligned} \mathbf {M}^e_{BLCD}= {\textstyle {1\over 3}}\,\mathbf {M}^e_C+{\textstyle {2\over 3}}\,\mathbf {M}^e_{L}= {\rho A \ell \over 90} \left[ \begin{array}{ccc} 14 &{} \quad -1 &{} \quad 2 \\ -1 &{} \quad 14 &{} \quad 2 \\ 2 &{} \quad 2 &{} \quad 56 \\ \end{array}\right] \!. \end{aligned}$$
(56)

For this instance, labeled BLCD, \(P=24\,(9+\cos \kappa )\), \(Q=32\,(927 + 884\,\cos \kappa -11\,\cos 2\kappa )\), and \(R=20\,(13-\cos \kappa )/3\). It has the small \(\kappa \) expansions

$$\begin{aligned} \Omega ^2_{a} = \kappa ^2 - {\kappa ^8\over 37800} + \cdots , \quad \Omega ^2_{o} = 30 - {15\,\kappa ^2\over 4} + {11\,\kappa ^4\over 32}+ \cdots . \end{aligned}$$
(57)

Dispersion and group velocity diagrams are shown in Fig. 10a, b. Despite the \(O(\kappa ^8)\) accuracy achieved in the AB of BLCD, it is shown next that this instance is not optimal.

Fig. 10
figure 10

DDD and DGVD plots for RHFP instances derived in Sect. 5.7 and Sect. 5.8. a, b Diagrams for the BLCD instance (56), c, d diagrams for the BLFM instance (60), d, e diagrams for the BLFD instance (66). Acoustic and optical branches shown in red and blue, respectively. Continuum case \(\Omega =\kappa \) and \(\gamma _c=c/c_0=1\) shown in black. (Color figure online)

Considering next the general MS template (39) –(41), let us find the MS pair for which the AB \(\Omega _a\) best matches the continuum \(\Omega =\kappa \) for small \(\kappa \). Given the expansion of \(\Omega _a^2\) in (52), the goal is to make as many coefficients beyond \(\kappa ^2\) vanish as possible, and to minimize the magnitude of the first surviving one. The analysis was actually performed using the \(\chi \)-form (44) of the general mass template. Four free parameters are available: \(\chi _1,\, \chi _2, \,\chi _3\), and \(\beta \). Only a procedural summary and final results are given.

It is possible to make \(C_4=C_6=0\) without difficulty, which permits elimination of \(\chi _1\) and \(\chi _2\). But all solutions of \(C_8(\chi _3,\beta )=0\) are imaginary, so the term in \(\kappa ^8\) cannot be cancelled. Extremization of \(C_8\) with respect to \(\chi _3\) and \(\beta \) gives only one constraint: \(160\beta ^2-120\beta \chi _3+9\chi _3^2=0\), from which \(\chi _3=4\,(5\mp \sqrt{15}\beta )/3\). Both signs give the same \(C_8\). Taking \(\beta =1\) for convenience, the \(-\) sign in \(\chi _3\), and working back we get

$$\begin{aligned} \chi _1&= {85\over 6}-2\sqrt{3\over 5}\mp 2\root 4 \of {375},\quad \chi _2=5-{\textstyle {1\over 2}}\,\sqrt{15}, \nonumber \\ \chi _3&= {4\over 3}\left( 5-\sqrt{15}\right) , \quad \beta =1. \end{aligned}$$
(58)

The \(-\) sign for \(\chi _1\) gives better conditioned mass matrices and still the same \(C_8\), so we pick that one. The numeric values to 16 places are \(\chi _1=2.7835604012611213\), \(\chi _2=3.0635083268962915\), and \(\chi _3=1.50268887172344\). The resulting minimum of \(|C_8|\) is \(C_{8best}=64/3-28\sqrt{3/5}=-0.355373\). This is about 3 times smaller than the \(C_8=-8!/37800=-16/15\) from (56). Converting to the \(\mu _i\) parameters via (42) yields

$$\begin{aligned} \mu _1&= {91 {-} 12\,a_1 {-} 7\,a_2\over 6}, \, \mu _2={32 {-} 8\,a_2\over 3},\nonumber \\ \mu _3&= {61 {-} 12\,a_1 {-} a_2\over 6}, \,\mu _4= {-46 {+} 6\,a_1 {+} 4\,a_2\over 3}, \end{aligned}$$
(59)

in which \(a_1=3^{1/4}\,5^{3/4}=\root 4 \of {375}\) and \(a_2=\sqrt{15}\). Numerical values to 16 places are \( \mu _1=1.8470687281574132\), \(\mu _2=0.3387110767802213\), \( \mu _3= 0.7200520743648302\), and \(\mu _4=-1.3682381704561768\). The resulting mass matrix, labeled BLFM (for best low frequency match), given to 16 places, is

$$\begin{aligned} \mathbf {M}^e_{BLFM}= m^e\, \left[ \begin{array}{ccc} 0.1949022909385804 &{} \quad -0.0093315975211724 &{} \quad 0.0210587276514608 \\ -0.0093315975211724 &{} \quad 0.1949022909385804 &{} \quad 0.0210587276514608 \\ 0.0210587276514608 &{} \quad 0.0210587276514608 &{} \quad 0.5446237025593408 \\ \end{array}\right] . \end{aligned}$$
(60)

Its eigenvalues (to 6 places) are positive: \(0.547077\), \(0.204234\), and \(0.183117\) times \(m^e=\rho A\,\ell \). Hence this mass matrix is admissible and well scaled. For this instance \(P=8 \,(35-2 \sqrt{15}-(5-2 \sqrt{15}) \cos \kappa )\), \(Q=1920 \,(15+4 \sqrt{15}+(15-4 \sqrt{15}) \cos \kappa ) \cos (\kappa /2)^2\), and \(R=20 \,(53-10 \sqrt{15}+(7-2 \sqrt{15}) \cos \kappa )/3\). It has the small-\(\kappa \) expansions

$$\begin{aligned} \Omega ^2_a&= \kappa ^2 + {1127- 291\sqrt{15}\over 3024 (5-\sqrt{15})^3}\kappa ^8+\cdots \nonumber \\ \Omega ^2_{o}&= {30\over 5-\sqrt{15}} -{5(29 - 8\sqrt{15}))\over 32(5-\sqrt{15})^3}\,\kappa ^2 - \cdots \end{aligned}$$
(61)

Dispersion and group velocity diagrams are shown in Fig. 10c, d. Note that values at the folding (Nyquist) wavenumber \(\kappa =\pi \) are identical: \(\Omega _a^2(\pi )= \Omega _o^2(\pi )= 12(10-\sqrt{15})/(23-4\sqrt{15})=9.792694126734647\), which is amazingly close to the continuum value of \(\pi ^2=9.86960440108935\) (in fact, the AB for \(\kappa <\pi \) and the continuum are indistinguishable at plot resolution). There is no acoustoptical gap; instead we observe a bifurcation point.

5.8 Bar3 Lumped Mass Template Variant

Although DLMM plainly form a subset of the general template (39), their practical importance justifies the use of a more compact two-parameter form. This is done by taking

$$\begin{aligned} \mu _{1}=\mu _{L1}+1, \quad \mu _{2}=\mu _{L2}+4, \quad \mu _{3}=1, \quad \mu _{4}=-2. \end{aligned}$$
(62)

Replacing into (39) produces the lumped mass template variant

$$\begin{aligned} \mathbf {M}_{L}^e&= {\rho A \ell \over 30} \left[ \begin{array}{lll} 5+\mu _{L1} &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad 5+\mu _{L1} &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 20+\mu _{L2} \\ \end{array}\right] \nonumber \\&= \mathbf {M}_{SLMM}^e+ {\rho A \ell \over 30} \left[ \begin{array}{lll} \mu _{L1} &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad \mu _{L1} &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad \mu _{L2} \\ \end{array}\right] \!. \end{aligned}$$
(63)

The baseline mass matrix is the DLMM (146) produced by Simpson’s 3-point integration rule, and now labeled SLMM. The stiffness matrix template is still (41). Parameters \(\mu _{L2}\) and \(\beta \) can be eliminated in favor of \(\mu _{L1}\) through

$$\begin{aligned} 2\mu _{L1}+\mu _{L2}=0, \qquad \beta ={(10-\mu _{L1})^2\over 20\,(5-\mu _{L1})}. \end{aligned}$$
(64)

The first constraint expresses element mass conservation while the second one enforces \(C_4=0\) and makes the AB agree with the continuum through \(\kappa ^4\) in the expansion (52) (this agreement is considered essential as otherwise there would be no advantage in using this element instead of Bar2). As only one parameter remains, customization is straightforward. The admissible range in \(\mu _{L1}\) for PD mass is \(-5<\mu _{L1}<10\), but if \(\mu _{L1}>5\), \(\beta <0\) and \(\mathbf {K}^e\) becomes indefinite.

On applying (64), the first nonzero term in the ABTS beyond \(\kappa ^2\) is \(C_6\,\kappa ^6/6!\), in which \(C_6=(5-3\mu _{L1}+\mu _{L1}^2)/(\mu _{L1}-10)\). Trying to attain \(O(\kappa ^6)\) accuracy by setting \(C_6=0\) is futile since the \(\mu _{L1}\) roots are complex conjugate. Solving \(\partial C6/\partial \mu _{L1}=0\) gives two real solutions: \(\mu _{L1}=5(2\pm \sqrt{3})\). Of these only the one with \(-\sqrt{3}\) keeps \(\mathbf {M}^e\) admissible. Replacing into (64) gives the signature

$$\begin{aligned} \mu _{L1}=5(2-\sqrt{3}), \quad \mu _{L2}=-2 \quad \mu _{L1}&= -10(2-\sqrt{3}),\nonumber \\ \beta&= {3\over 4(\sqrt{3}-1)}. \end{aligned}$$
(65)

Numerical values to 16 places are \(\mu _{L1}=1.339745962155614\), \( \mu _{L2}=-2.679491924311228\), and \(\beta =1.024519052838329\). The \(\kappa ^6\) term in the AB is about 36 % smaller than that of SLMM: \(\approx -\kappa ^6/2246\) versus \(-\kappa ^6/1440\). The template instance, labeled BLFD, is

$$\begin{aligned} \mathbf {M}^e_{BLFD}&= {\rho A \ell \over 30} \left[ \begin{array}{lll} 5\sqrt{3}-5 &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad 5\sqrt{3}-5 &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 20-10\sqrt{3} \\ \end{array}\right] , \nonumber \\ \mathbf {K}^e_{BLFD}&= {EA\over \ell } \left[ \begin{array}{ccc} 1 &{} \quad -1 &{} \quad 0 \\ -1 &{} \quad 1 &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 0 \\ \end{array}\right] \nonumber \\&+ {4EA\over \ell \,(\sqrt{3}-1)} \left[ \begin{array}{ccc} 1 &{} \quad 1 &{} \quad -2 \\ 1 &{} \quad 1 &{} \quad -2 \\ -2 &{} \quad -2 &{} \quad 4 \\ \end{array}\right] \!. \end{aligned}$$
(66)

The Taylor series of the dispersion branches are

$$\begin{aligned} \Omega _a^2&= \kappa ^2-{10\sqrt{3}-10\over 720}\kappa ^6- \cdots , \nonumber \\ \Omega _o^2&= {12\over (\sqrt{3}-1)^2} - {\sqrt{3}\over 4\sqrt{3}-6}\kappa ^2 - \cdots \end{aligned}$$
(67)

The DDD and DGVD are shown in Fig. 10e, f. As in the case of the BLFM, pictured in (c,d) of that figure, the branches intersect at \(\kappa =\pi \), where \(\Omega ^2_{af}=\Omega ^2_{of}=(6-2\sqrt{3})/(2-\sqrt{3})\approx 9.4641\).

5.9 Bar3 Maximum Stable Time Step Customization

Since DLMM are often used in explicit DTI, it is of some interest to find whether the stable time stepsize can be maximized while still satisfying \(O(\kappa ^4)\) accuracy. This goal pertains to the MSTS customization of Table 1. Let \(\Omega _{max}\) be the maximum of \(\Omega _a\) and \(\Omega _o\) over the Brillouin zone \(\kappa \in [0,2\pi ]\). To maximize the stable time step, one minimizes \(\Omega _{max}\) over free parameters, while trying to keep both mass and stiffness admissible. This procedure can be streamlined by assuming a DDD configured as in Fig. 8, whence only frequency values at \(\kappa =0\) and \(\kappa =\pi \) need to be considered. Since \(\left. \Omega _a\right| _{\kappa =0}=0\), the search only involves \(\Omega _{oc}\), \(\Omega _{af}\) and \(\Omega _{of}\), or (for convenience) their squares. For the DLMM template variant (63) under the accuracy constraints (64) the process boils down to solving the max-min problem in one variable:

$$\begin{aligned}&\min _{\mu _{L1}} \max \,\left( \Omega _{oc}^2,\Omega _{af}^2,\Omega _{of}^2\right) \nonumber \\&\quad = \min _{\mu _{L1}} \max \,\left( {60 (10 - \mu _{L1})\over 25 - \mu _{L1}^2}, {4 (10 - \mu _{L1})\over \mu _{L1}-5 }, {60\over 5 + \mu _{L1} }\right) .\nonumber \\ \end{aligned}$$
(68)

A simple plot shows that the cutoff frequency dominates for admissible \(\mu _{L1}\in (-5,5)\), so it is sufficient to minimize with respect to \(\Omega _{oc}^2\). This again leads to the solution (65). Consequently the BLFD instance also maximizes the explicit DTI time step. The reward, however, is marginal with respect to SLMM: only about a 3.5 % gain.

To get a more significant improvement, it is necessary to keep \(\beta \) free, and accept that \(O(\kappa ^4)\) accuracy is lost. It may be verified that the largest possible stable timestep is produced by the signature

$$\begin{aligned} \mu _{L1}=5/2, \quad \mu _{L2}=-2\mu _{L1}=-5, \quad \beta =3/8. \end{aligned}$$
(69)

which apportions nodal masses as 1:1:2, while substantially modifying the stiffness matrix. Setting (69) gives an instance with a COB \(\Omega _o^2=8\). Its stable stepsize is 1.673 times that of BLFD. But its LF performance is exactly the same as that of the lumped-mass Bar2, which does not have an OB. So it is largely a curiosity.

5.10 Reducing High Frequency Pollution

The presence of the OB does not affect vibration calculations in structural dynamics. One simply ignores those eigenfrequencies as nonphysical. However, the OB may become a nuisance in direct time integration (DTI) for problems that involve discontinuities, such as pulse propagation, or contact-impact, because it may feed spurious noise. To alleviate this problem three approaches may be tried at the template level:

  1. 1.

    Singular Mass Matrix If \(\mathbf {M}^e\) is made singular with an appropriate null eigenvector, the OB is raised. In fact it becomes infinite at \(\kappa =0\). The net effect is that the acoustoptical gap is increased at low wavenumbers. This heps to filter out frequencies that fall in the gap, since they will decay exponentially. One drawback of singularity is that explicit DTI is excluded, even if \(\mathbf {M}^e\) is diagonal.

  2. 2.

    SMS A scaled stiffness matrix is added to the mass. As discussed in Sect. 3.5, eigenvectors are unchanged but higher natural frequencies are effectively reduced. The effect is similar to that of adding stiffness proportional damping, but without altering vibration modes. It may be done at the element or assembly (master) level. In the study of Sect. 5.12 it is done at the element level.

  3. 3.

    COB A constant optical branch (COB) is one independent of wavenumber. It stays at a constant frequency \(\Omega _{o}=\Omega _{oc}\) over the entire Brillouin zone, and has zero group velocity since \(\partial \Omega _{oc}/\partial \kappa =0\). To be effective in cutting noise pollution, \(\Omega _{oc}\ge \Omega _{af}\), in which \(\Omega _{af}\) is the folding acoustic frequency. If that holds, the stopping band above \(\Omega _{af}\) is effectively maximized (even if a mesh frequency hits \(\Omega _{oc}\) exactly, it will not propagate since its group velocity vanishes). A COB template is one that possesses that property (for each OB in case there are multiple ones).

The three foregoing approaches are studied below for the Bar3 element.

5.11 Bar3 Spectral Mass Variant

Making \(\mathbf {M}^e\) singular is not sufficient. It is important to have the correct null eigenvector. To achieve that it is convenient to use the SP outlined in Sect. 3.2. Select three generalized coordinates: \(g_0,\, g_1\) and \(g_2\) as amplitudes of three physically transparent eigenmotions:

\(g_0\) :

Amplitude of rigid body motion: \(\mathbf {v}_0=[\begin{array}{ccc} 1&\quad 1&\quad 1 \end{array}]^T\).

\(g_1\) :

Amplitude of acoustic bar motion: : \(\mathbf {v}_1=[\begin{array}{ccc} -1&\quad 1&\quad 0 \end{array}]^T\).

\(g_2\) :

Amplitude of optical bar motion: : \(\mathbf {v}_2= [\begin{array}{ccc} 1&\quad 1&\quad -2 \end{array}]^T\).

Those three vectors are mutually orthogonal. To make them orthonormal, divide by \(\sqrt{3},\, \sqrt{2}\) and \(\sqrt{6}\), respectively. Stacking the orthonormalized vectors as columns, the linkage between physical and generalized coordinates can be expressed as

$$\begin{aligned} \mathbf {u}^e&= \left[ \begin{array}{ccc} u_1&\quad u_2&\quad u_3 \end{array}\right] = \left[ \begin{array}{lll} 1/\sqrt{3} &{} \quad -1/\sqrt{2} &{} \quad 1/\sqrt{6} \\ 1/\sqrt{3} &{} \quad 1/\sqrt{2} &{} \quad 1/\sqrt{6} \\ 1/\sqrt{3} &{} \quad 0 &{} \quad -2/\sqrt{6} \\ \end{array}\right] \nonumber \\&= \left[ \begin{array}{c} g_0 \\ g_1 \\ g_2\\ \end{array}\right] =\mathbf {H}^T\mathbf {g}. \end{aligned}$$
(70)

The inverse relation is \(\mathbf {g}=\mathbf {H}\,\mathbf {u}^e\) because \(\mathbf {H}\) is orthogonal by construction, and thus \(\mathbf {H}^{-1}=\mathbf {H}^T\). As mass matrix in generalized coordinates we stipulate the \(3\times 3\) diagonal matrix \(\mathbf {D}_\mu \) of entries \(m^e\mu _{S0}/3\), \(m^e\mu _{S1}/45\), and \(m^e\mu _{S2}/15\), in which \(m^e=\rho \,A\,\ell \), and the scaling factors were chosen for convenience in cleaning up downstream expressions. Element mass conservation will require \(\mu _{S0}=1\), so the first entry is simply \(m^e/3\). Transforming to physical coordinates yields the spectral mass template variant

$$\begin{aligned}&\mathbf {M}_S^e={m^e\over 45} \mathbf {H}^T \left[ \begin{array}{lll} 15 &{} \quad 0&{} \quad 0\\ 0&{} \quad \mu _{S1}&{} \quad 0 \\ 0&{} \quad 0&{} \quad 3\mu _{S2} \end{array}\right] \nonumber \\&\mathbf {H}= {m^e\over 90}\, \left[ \begin{array}{lll} 10+\mu _{S1}+\mu _{S2} &{} \quad 10-\mu _{S1}+\mu _{S2} &{} \quad 10-2\,\mu _{S2} \\ 10-\mu _{S1}+\mu _{S2} &{} \quad 10+\mu _{S1}+\mu _{S2} &{} \quad 10-2\,\mu _{S2} \\ 10-2\,\mu _{S2} &{} \quad 10-2\,\mu _{S2} &{} \quad 10+4\,\mu _{S2} \end{array}\right] .\nonumber \\ \end{aligned}$$
(71)

The variant (71) is a subset of the general template (39) that results by taking

$$\begin{aligned} \mu _1&= {\textstyle {1\over 3}}(\mu _{S1}{+}\mu _{S2}{-}2),\,\, \mu _2={\textstyle {1\over 3}}(4\,\mu _{S2}{-}38),\nonumber \\ \mu _3&= {\textstyle {1\over 3}}(13{-}\mu _{S1}{+}\mu _{S2}),\,\, \mu _4={\textstyle {1\over 3}}(4{-}2\,\mu _{S2}). \end{aligned}$$
(72)

By construction, the eigenvalues of (71) are \(m^e/3\), \(m^e\mu _{S1}/45\) and \(m^e\,\mu _{S2}/15\), whence the nonnegativity condition is fulfilled if \(\mu _{S1}\) and \(\mu _{S2}\) are nonnegative. To make \(\mathbf {M}^e_S\) singular, set \(\mu _{S2}=0\), which produces

$$\begin{aligned} \mathbf {M}_S^e=\mathbf {H}^T\,\mathbf {D}_\mu \,\mathbf {H}={m^e\over 90}\, \left[ \begin{array}{lll} 10+\mu _{S1} &{} \quad 10-\mu _{S1} &{} \quad 10 \\ 10-\mu _{S1} &{} \quad 10+\mu _{S1} &{} \quad 10 \\ 10 &{} \quad 10 &{} \quad 10 \end{array}\right] . \end{aligned}$$
(73)

Solving \(C_4=0\) and \(C_6=0\) yields two solutions for \(\beta \) and \(\mu _{S1}\), of which we pick that with larger \(\mu _{S1}\) (to get a better conditioned nonsingular subspace). This gives

$$\begin{aligned}&\mu _{S1}={3(5+\sqrt{10})\over 2}=12.24341649025257, \nonumber \\&\beta ={5+\sqrt{10}\over 12}=0.6801898050140316, \end{aligned}$$
(74)

in addition to \(\mu _{S2}=0\). Inserting into (73) gives the instance labeled BSSM for Best Singular Spectral Mass. The mass matrix, with numerical values given to 6 places, is

$$\begin{aligned} \mathbf {M}^e_{BSSM}&= {m^e\over 180} \left[ \begin{array}{ccc} M_{11} &{} \quad M_{12} &{} \quad M_{13} \\ M_{12} &{} \quad M_{22} &{} \quad M_{23} \\ M_{13} &{} \quad M_{23} &{} \quad M_{33} \end{array}\right] \nonumber \\&\approx m^e \left[ \begin{array}{ccc} 0.247149 &{} \quad -0.024927 &{} \quad 0.111111 \\ -0.024927 &{} \quad 0.247149 &{} \quad 0.111111 \\ 0.111111 &{} \quad 0.111111 &{} \quad 0.111111 \end{array}\right] \!.\nonumber \\ \end{aligned}$$
(75)

in which \(M_{11}=M_{22}=35+\sqrt{10}\), \(M_{12}=5-3\sqrt{10}\), and \(M_{13}=M_{23}=M_{33}=20\). The associated stiffness matrix, with numerical values given to 6 places, is

$$\begin{aligned} \mathbf {K}^e_{BSSM}&= {k^e\over 9} \left[ \begin{array}{ccc} K_{11} &{} \quad K_{12} &{} \quad K_{13} \\ K_{12} &{} \quad K_{22} &{} \quad K_{23} \\ K_{13} &{} \quad K_{23} &{} \quad K_{33} \end{array}\right] \nonumber \\&\approx k^e \left[ \begin{array}{ccc} 1.906920 &{} \quad -0.093080 &{} \quad -1.813839 \\ -0.093080 &{} \quad 1.906920 &{} \quad -1.813839 \\ -1.813839 &{} \quad -1.813839 &{} \quad 3.627679 \end{array}\right] \!.\nonumber \\ \end{aligned}$$
(76)

in which \(k^e=E\,A/\ell \), \(K_{11}=K_{22}=14+\sqrt{10}\), \(K_{12}=-4+\sqrt{10}\), \(K_{13}=K_{23}=-10-2\sqrt{10}\) and \(K_{33}=20+4\sqrt{10}\). Dispersion and group velocity diagrams are shown in Fig. 11a, b. The Taylor series of the dispersion branches are

$$\begin{aligned} \Omega _a^2=\kappa ^2-{\kappa ^8\over 6048}- \cdots , \qquad \Omega _o^2=240\,\kappa ^{-2} + {5\kappa ^4\over 126} + \cdots \nonumber \\ \end{aligned}$$
(77)

The \(O(\kappa ^8)\) AB accuracy of this element is comparable to that of BLCD and BLFM, but its OB gets out of the way. Is this the template instance for all seasons? Only future experimentation in DTI will tell.

Fig. 11
figure 11

DDD and DGVD plots for three RHFP instances derived in Sects. 5.115.13. a, b Diagrams for the BSSM instance (56), c, d diagrams for the SMS2 instance (60), e, f diagrams for the COB0 instance: first of (90). Acoustic and optical branches shown in red and blue, respectively. Continuum case \(\Omega =\kappa \) and \(\gamma _c=c/c_0=1\) shown in black. (Color figure online)

5.12 Bar3 Selective Mass Scaling Variant

In the SMS approach outlined in Sect. 3.5, the mass matrix is modified by adding a scaled version of the stiffness matrix:

$$\begin{aligned} \mathbf {M}_K^e=\mathbf {M}^e_u + c_K\,\mathbf {K}^e. \end{aligned}$$
(78)

Here \(\mathbf {M}^e\) is an unmodified mass matrix, and \(c_K\) a scaling coefficient with appropriate physical dimensions. Both \(\mathbf {M}^e\) and \(\mathbf {K}^e\) may be template forms. Since \(\mathbf {M}^e\) and \(\mathbf {K}^e\) have different physical dimensions, it is convenient to change the raw expression (78) to

$$\begin{aligned} \mathbf {M}_K^e=\mathbf {M}^e_u + \mu _K\,s^e\,\mathbf {K}^e, \end{aligned}$$
(79)

in which \(s^e\) is a scaling coefficient with dimension of mass-over-stiffness (equivalently, \(1/s^e\) has dimensions of squared physical frequency) while \(\mu _K\) is a dimensionless free parameter. For the Bar3 element we take \(s^e=(\rho \,A\,\ell )/(EA/\ell )=\rho \,\ell ^2/E\). This can be maneuvered to the following equivalent form, which is convenient for implementation:

$$\begin{aligned} \mathbf {M}_K^e=\mathbf {M}^e_u + \mu _K\,m^e\,\widehat{\mathbf {K}}^e. \end{aligned}$$
(80)

Here \(m^e=\rho A \ell \) is (as usual) the element mass, whereas \(\widehat{\mathbf {K}}^e_u\) is a dimensionless stiffness matrix obtained by setting \(E=1,\, A=1\) and \(\ell =1\). To reduce the overall number of parameters, we pick \(\mathbf {M}_u^e\) to be the diagonally lumped template subset (63); this agrees with the common use of SMS in explicit DTI. The general stiffness template (41) with unit \(E,\, A\) and \(\ell \) is used for \(\widehat{\mathbf {K}}^e\). Hence

$$\begin{aligned} \mathbf {M}_K^e&= m^e\,\left( \,{1\over 30}\left[ \begin{array}{lll} 5+ \mu _{L1} &{} \quad 0 &{} \quad 0 \\ 0 &{} \quad 5+\mu _{L1} &{} \quad 0 \\ 0 &{} \quad 0 &{} \quad 20+\mu _{L2} \\ \end{array}\right] \right. \nonumber \\&\left. + {4\beta \,\mu _K \over 3\,\ell } \left[ \begin{array}{ccc} 1 &{} \quad 1 &{} \quad -2 \\ 1 &{} \quad 1 &{} \quad -2 \\ -2 &{} \quad -2 &{} \quad 4 \\ \end{array}\right] \,\right) \!. \end{aligned}$$
(81)

Mass conservation is enforced if \(\mu _{L2}=-2 \mu _{L1}\). Inserting this in (81) we have three free parameters: \( \mu _{L1},\, \mu _K\) and \(\beta \). This \(\mathbf {M}^e_K\) with \(\mu _{L2}=-2\mu _{L1}\) is a particular case of the general mass template if

$$\begin{aligned} \mu _1&=1+\mu _{L1}+10 \,(4 \beta +3) \mu _K,\nonumber \\ \mu _2&= 4-2 \mu _{L1}+160 \,\beta \mu _K, \nonumber \\ \mu _3&= 1+10 \,(4 \beta -3) \mu _K,\quad \mu _4=-2-80 \,\beta \mu _K\!. \end{aligned}$$
(82)

Unlike previous variants, now \(\beta \) appears in the mass template. The linkage (82) becomes linear if \(\beta \) is preset, for example to 1, and nonlinear otherwise.

Further experimentation with the SMS template variant (81) was confined to \(\mu _{L1}=\mu _{L2}=0\), which takes SLMM as original mass matrix. That leaves out two free parameters: \(\mu _K\) and \(\beta \). Suppose \(\mu _K\) is chosen. Then \(O(\kappa ^4)\) AB accuracy can be maintained by taking

$$\begin{aligned} \beta = {1\over 1- 12\mu _K}. \end{aligned}$$
(83)

If \(\mu _K>1/12,\, \beta <0\) and \(\mathbf {K}^e\) becomes indefinite. But setting \(0\le \mu _K\le 1/12\) hardly change the higher frequencies. For that one needs a much larger \(\mu _K\); say \(\mu _K=O(1)\). If so, adjusting \(\mathbf {K}^e\) as per (83) is precluded: the cure is worst than the disease. One may as well set \(\beta =1\). The high frequencies are cut down, but LF accuracy is seriously lost.

This tradeoff is vividly displayed in the vibration benchmarks reported in Sect. 5.14. Three instances labeled SMS1, SMS2 and SMS3, are tested there. Their signatures are \(\{\mu _K=1/24,\beta =2\}\), \(\{\mu _K=1/2,\beta =1\}\), and \(\{\mu _K=2,\beta =1\}\), respectively. Dispersion and group velocity diagrams for SMS2 are shown in Fig. 11c, d. The poor LF fit is obvious.

5.13 Bar3 Constant Optical Branch Variant

The investigation of the general Bar3 template (39)–(41) for COB instances was done under two preset conditions: \(C_4=0\), which enforces order \(O(\kappa ^4)\) accuracy in the AB (AB), and \(\beta =1\) in the stiffness template (41). Several one-parameter families satisfying these conditions were found. The two that produced simpler mass matrices were retained, reparametrized, and labeled COBA and COBB. Associated mass matrices are subscripted accordingly.

The COBA family is defined by

$$\begin{aligned} \mathbf {M}_{COBA}^e= {m^e\over 12}\left[ \begin{array}{ccc} 6- \nu _A &{} \quad 2- \nu _A &{} \quad -2+2\nu _A \\ 2- \nu _A &{} \quad 6- \nu _A &{} \quad -2+2\nu _A \\ -2+2\nu _A &{} \quad -2+2\nu _A &{} \quad 4-4\nu _A \end{array}\right] . \end{aligned}$$
(84)

in which \(m^e=\rho \,A\,\ell \). The determinant is \((1-\nu _A)/18\). \(\mathbf {M}_{COBA}\) is PD if \(\nu _A<1\). Parameter \(\nu _A\) is linked to those of the general template (39) by

$$\begin{aligned} \mu _1&= 11-{5\nu _A\over 2},\quad \mu _2=-2 (3+5\nu _A),\quad \mu _3=6-{5\nu _A\over 2},\nonumber \\ \mu _4&= -7+5\nu _A. \end{aligned}$$
(85)

The COBB family is defined by

$$\begin{aligned}&\mathbf {M}_{COBB}^e\nonumber \\&\quad = {m^e\over 432} \left[ \begin{array}{lll} 96-36\nu _B-\nu _B^2 &{} \quad 24-12\nu _B+\nu _B^2 &{} \quad -48+24\nu _B \\ 24-12\nu _B+\nu _B^2 &{} \quad 96-36\nu _B-\nu _B^2 &{} \quad -48+24\nu _B \\ -48+24\nu _B &{} \quad -48+24\nu _B &{} \quad 384 \end{array}\right] .\nonumber \\ \end{aligned}$$
(86)

The determinant is \((36-12\nu _B-\nu _B^2)^2/34992\). \(\mathbf {M}_{COBB}\) is PD if \(-6(\sqrt{2}+1)<\nu _B<6(\sqrt{2}-1)\). Parameter \(\nu _B\) is linked to those of the general template (39) by

$$\begin{aligned} \mu _1&= {8\over 3}-{5\nu _B\over 2}-{5\nu _B^2\over 72},\quad \mu _2={32\over 3},\nonumber \\ \mu _3&= {8\over 3}-{5\nu _B\over 6}+{5\nu _B^2\over 72},\quad \mu _4 ={-16+5\nu _B\over 3}.\nonumber \\ \end{aligned}$$
(87)

These two families are taken to collectively define the Bar3 template variant identified as COB. They coalesce only for \(\nu _A=-5/3\) and \(\nu _B=-6\), which produces an instance discussed below. An interesting result is that the AB is identical for all COB instances:

$$\begin{aligned} \Omega _a^2= {12(1-\cos \kappa )\over 5+\cos \kappa } = \kappa ^2 - {\kappa ^6\over 240} - {\kappa ^8\over 6048} -\cdots \end{aligned}$$
(88)

whereas the constant OB value is family and parameter dependent:

$$\begin{aligned} \Omega ^2_{ocA}= {16\over 1-\nu _A}, \qquad \Omega ^2_{ocB}= {432\over 36-12\nu _B-\nu _B^2}. \end{aligned}$$
(89)

It follows that the only role played by \(\nu _A\) and \(\nu _B\) is to adjust the “OB height” along the vertical DDD axis. As noted in Sect. 5.10, it should equal or exceed the folding acoustic frequency \(\Omega _{af}^2=6\), which is the same for all COB instances on account of (88). This requires \(\nu _A\ge -5/3\) and \(\nu _B\ge -6\). As \(\nu _A\rightarrow 1\) and \(\nu _B\rightarrow 12\) the OB moves to \(\infty \) and the mass matrices assume different limits. For COBA, \(\left. \mathbf {M}^e_{COBA}\right| _{\nu _A\rightarrow 1}\) is the optimal Bar2 matrix (22), which is PD. On the other hand the limit \(\left. \mathbf {M}^e_{COBB}\right| _{\nu _B\rightarrow 12}\) falls in the indefinite range. Three noteworthy instances of the mass matrices produced by these two families are

$$\begin{aligned} \mathbf {M}^e_{COB0}&= {m^e\over 36}\left[ \begin{array}{ccc} 23 &{} \quad 11 &{} \quad -16 \\ 11 &{} \quad 23 &{} \quad -16 \\ -16 &{} \quad -16 &{} \quad 32\\ \end{array}\right] ,\nonumber \\ \mathbf {M}^e_{COB1}&= {m^e\over 6}\left[ \begin{array}{ccc} 3 &{} \quad 1 &{} \quad -1 \\ 1 &{} \quad 3 &{} \quad -1 \\ -1 &{} \quad -1 &{} \quad 2\\ \end{array}\right] , \nonumber \\ \mathbf {M}^e_{COB2}&= {m^e\over 18} \left[ \begin{array}{ccc} 4 &{} \quad 1 &{} \quad -2 \\ 1 &{} \quad 4 &{} \quad -2 \\ -2 &{} \quad -2 &{} \quad 16 \\ \end{array}\right] \!. \end{aligned}$$
(90)

\(\mathbf {M}^e_{COB0}\) is the unique mass matrix for which \(\Omega ^2_{oc}=\Omega _a^2=6\); that is, the COB passes through the folding (Nyquist) frequency. It emerges by setting either \(\nu _A=-5/3\) in (84) or \(\nu _B=-6\) in (86). \(\mathbf {M}^e_{COB1}\) which gives \(\Omega _{ocA}^2=16\), is the simplest mass matrix that produces a COB. It is obtained by setting \(\nu _A=0\) in (84). Finally \(\mathbf {M}_{COB2}\), which yields \(\Omega _{ocB}^2=12\), was the first COB instance discovered, as noted in Appendix Sect. 1.7. It is obtained by setting \(\nu _B=0\) in (86).

Dispersion and group velocity diagrams for COB0 are shown in Fig. 11e, f. The DDD for COB1 and COB2 would possess an identical AB branch but the flat OB would appear higher, whereas the DGVD would be identical. Those four diagrams are omitted to save space.

5.14 Bar3 Test: Vibrations of a Fixed-Free Bar Member

The natural frequency benchmark test presented in Sect. 4.9 for three Bar2 discretizations is repeated for the ten Bar3 template instances listed in Table 10. The fixed-free bar member is pictured in Fig. 5b. It is prismatic, with constant \(E=1,\, A=1\, \rho =1\). The total member length is \(L=\pi /2\). With those numerical properties, the continuum eigenfrequencies \(\omega _{0i}\) are given by (24). The member is divided into \(N_e\) identical elements, with \(N_e=1,2,\ldots 16\).

Table 10 Bar3 instances compared in fixed-free bar vibrations tests

To reduce cluttering the instances in Table 10 are divided into two groups of five each. Results are presented in number of correct digits versus number of elements for the first three frequencies, exactly as described for the Bar2 test in Sect. 4.9. Group 1 include CMM and SLMM as well as instances constructed with optimal LFF customization in mind: BLCD, BLFM and BLFD. Results are displayed in Fig. 12a–c. Group 2 includes instances derived with RHFP in mind: BSSM, SMSx (\(x=1,2,3\)) and COB0. Results are displayed in Fig. 12d–f.

Fig. 12
figure 12

Performance of ten Bar3 template instances in predicting the first three natural frequencies \(\omega _i,\, i=1,2,3\) of the fixed-free prismatic homogeneous bar shown in Fig. 5b. See text in Sect. 4.9 for a detailed description of the log-log plots

BLFM is the clear winner in the first group, with BLCD close behind, while the others, with only \(O(\kappa ^4)\) AB accuracy, lag appreciably. In the second group, BSSM is the clear winner, with performance comparable to BLFM and BLCD of the first group. SMS1 and COB0 are way behind, while SMS2 and SMS3 are highly inaccurate (as observed in Sect. 5.12, SMS1 would hardly effect any HF reduction, so its reasonable LF accuracy is misleading).

6 The Bernoulli–Euler Plane Beam Element

This Section and the next one study templates for two-node plane beam elements constructed from the Bernoulli–Euler (BE) and Timoshenko models, respectively. To keep the material relatively compact, two restrictions are observed:

  • Only mass matrix templates are developed.

  • The only customization is LFCF

To enforce the first one, the optimal stiffness matrix for statics (“optimal” means that it satisfies the homogeneous static equilibrium equations over the element) is chosen and kept fixed. Simultaneous adjustment of the mass and stiffness templates to form MS pairs is relegated to future research. Prior experience in this regard, cited in Appendix Sect. 1.7, suggests that the improvement is marginal.

6.1 The BE Beam Mass Template

The BE beam model is a special case of the Timoshenko model treated in Sect. 7. Nevertheless it is useful to build its mass template separately, since results provide a valuable cross check with the more complicated Timoshenko beam. The well known CMM of this element is derived in Appendix Sect. 3.2, to which the reader is referred for notation; the derivation assumes a prismatic two-node element with four nodal DOF with the standard cubic shape functions. This matrix is augmented to produce the following EW template:

$$\begin{aligned}&\mathbf {M}^e_\mu \nonumber \\&\quad =m^e\!\left[ \!\begin{array}{cccc} \frac{13}{35}+\mu _{11} &{} \quad \left( \frac{11}{210}+\mu _{12}\right) \ell &{} \quad \frac{9}{70}+\mu _{13} &{} \quad -\left( \frac{13}{420}+\mu _{14}\right) \ell \\ &{} \quad \left( \frac{1}{105}+\mu _{22}\right) \ell ^2 &{} \quad \left( \frac{13}{420}+\mu _{23}\right) \ell &{} \quad -\left( \frac{1}{140}+\mu _{24}\right) \ell ^2 \\ &{} \quad &{} \quad \frac{13}{35}+\mu _{11} &{} \quad -\left( \frac{11}{210}+\mu _{12}\right) \ell \\ \hbox {symm} &{} \quad &{} \quad &{} \quad \left( \frac{1}{105}+\mu _{22}\right) \ell ^2 \\ \end{array}\!\right] \nonumber \\ \end{aligned}$$
(91)

in which \(m^e=\rho \,A\,\ell \). The parameters in (91) are \(\mu _{ij}\), in which \(ij\) identifies the mass matrix entry. The template (91) accounts for matrix symmetry and some physical symmetries. Three more conditions can be imposed right away:

$$\begin{aligned}&\mu _{14}=\mu _{23}, \quad \mu _{13}=-\mu _{11}, \nonumber \\&2\mu _{12}=\mu _{11}+2\mu _{22}+2\mu _{23}-2\mu _{24}. \end{aligned}$$
(92)

The first comes from prismatic fabrication, and the others from conservation of total translational mass and angular momentum, respectively. Four free parameters remain: \(\{\mu _{11},\mu _{22},\mu _{23},\mu _{24}\}\). For the stiffness matrix we take the well known one for a plane prismatic homogeneous BE beam element

$$\begin{aligned} \mathbf {K}^e ={EI\over \ell ^3}\left[ \begin{array}{llcc} 12 &{} 6\ell &{} \quad -12 &{} \quad 6\ell \\ &{} 4\ell ^2 &{} \quad -6\ell &{} \quad 2\ell ^2 \\ &{} &{} \quad 12 &{} \quad -6\ell \\ \hbox {symm} &{} \quad &{} &{}\quad 4\ell ^2 \\ \end{array}\right] \end{aligned}$$
(93)

in which \(I=I_{zz}\) is the second moment of inertial of the cross section with the respect to \(z\), which is chosen to go along the neutral axis. If the FEM model contains only prismatic beams, this \(\mathbf {K}^e\) is nodally exact, and consequently statically optimal (it can also be derived from the equilibrium equations). This stiffness is kept fixed throughout the Fourier analysis.

6.2 BE Beam Template Fourier Analysis

The Fourier analysis procedure should be by now familiar to the reader. An infinite lattice of identical beam elements of length \(\ell \) is set up. This will look like Fig. 2b–d, except that the member is now a plane beam. Plane waves of wavenumber \(k\) and frequency \(\omega \) propagating over the lattice are represented by

$$\begin{aligned} v(x,t)&= B_v\,\exp \big (j(kx -\omega t\big ), \quad \theta (x,t)\nonumber \\&= B_\theta \,\exp \big (j(kx-\omega t\big ), \quad j=\sqrt{-1}. \end{aligned}$$
(94)

At a typical lattice node \(j\) there are two DOF: \(v_j\) and \(\theta _j\). Two patch equations are extracted, and converted to dimensioneless form on defining \(\kappa =k\ell \) and \(\Omega =\omega c_0/\ell \), in which \(c_0=EI/(\rho A\ell ^4)\) is a reference phase velocity. The condition for wave propagation gives the characteristic matrix equation

$$\begin{aligned} \det \left[ \begin{array}{cc} C_{vv} &{} \quad C_{v\theta } \\ C_{\theta v} &{} \quad C_{\theta \theta } \end{array}\right] = C_{vv}C_{\theta \theta }-C_{v\theta }C_{\theta v}=0, \end{aligned}$$
(95)

in which

$$\begin{aligned} C_{vv}&=\big (840 {-} 2(13 {+} 35\mu _{11})\Omega ^2\nonumber \\&\quad \,- (840 {+} (9 {-} 70\mu _{11})\Omega ^2)\cos \kappa \big )/35, \nonumber \\ -C_{\theta v}&=C_{v\theta }= j\big (2520 + (13 {+} 420\mu _{23})\Omega ^2\big ) \sin \kappa /210, \nonumber \\ C_{\theta \theta }&=\big (1680 - 4(1 {+} 105\mu _{22})\Omega ^2\nonumber \\&\quad \,+(840 + 3(1 {+} 140\mu _{24})\Omega ^2)\cos \kappa \big )/210. \end{aligned}$$
(96)

The condition (95) gives a quadratic equation in \(\Omega ^2\) that provides two dispersion solutions: AB \(\Omega ^2_a(\kappa )\) and OB \(\Omega ^2_o(\kappa )\). The AB represent genuine flexural modes, whereas the OB is a spurious byproduct of the FEM discretization. The small-\(\kappa \) (low frequency, long wavelength) expansions of these roots are

$$\begin{aligned} \Omega _a^2&= \kappa ^4 + C_6\kappa ^6 + C_8\kappa ^8 + C_{10}\kappa ^{10} + C_{12}\kappa ^{12} +\cdots , \nonumber \\ \Omega _o^2&= D_0 + D_2\kappa ^2 + D_4\kappa ^4 +\cdots , \end{aligned}$$
(97)

in which \(C_6= -\mu _{11} - 2\mu _{22} - 4\mu _{23} + 2\mu _{24}\), \(C_8= 1/720 + \mu _{11}^2 + 4\mu _{22}^2 + 2\mu _{23}/3 + 16\mu _{22}\mu _{23} + 16\mu _{23}^2 + \mu _{11}(1/12 + 4\mu _{22} + 8\mu _{23} - 4\mu _{24}) - \mu _{24} - 8\mu _{22}\mu _{24} - 16\mu _{23}\mu _{24} + 4\mu _{24}^2\), etc.; and \(D_0=2520/(1 + 420\mu _{22} - 420\mu _{24})\), etc. Mathematica calculated these series up to \(C_{14}\) and \(D_4\).

The continuum dispersion curve is \(\Omega ^2=\kappa ^4\), which automatically matches \(\Omega _a^2\) as \(\kappa \rightarrow 0\). Thus four free parameters offer the opportunity to match coefficients of four powers: \(\{\kappa ^{6},\kappa ^8,\kappa ^{10},\kappa ^{12}\}\). But it will be seen that the last match is unfeasible if \(\mathbf {M}^e\) is to stay nonnegative. We settle for a scheme that agrees up to \(\kappa ^{10}\). Setting \(C_6=C_8=C_{10}=0\) while keeping \(\mu _{22}\) free yields two sets of solutions, of which the most useful one is

$$\begin{aligned} \mu _{11}&= 4\mu _{22}-67/540 - (4/27)\sqrt{38/35 - 108\mu _{22}},\nonumber \\ \mu _{23}&= 43/1080- 2\mu _{22} + \sqrt{95/14 - 675\mu _{22}}/54,\nonumber \\ \mu _{24}&= 19/1080 - \mu _{22} + \sqrt{19/70 - 27\mu _{22}}/27. \end{aligned}$$
(98)

The positivity behavior of \(\mathbf {M}^e_\mu \) as \(\mu _{22}\) is varied is shown in Fig. 13a. \(\mathbf {M}^{(e)}\) is indefinite for \(\mu _{22}<\mu _{22}^{min}= (27-4\sqrt{35})/5040=0.0006618414419844316\). At the other extreme the solutions of (98) become complex if \(\mu _{22}>\mu _{22}^{max}=19/1890=0.010052910052910053\).

Fig. 13
figure 13

Behavior of \(\mathbf {M}^e_\mu \) as function of \(\mu _{22}\) with other parameters given by (98): a determinants \(d_k\) of principal minors of order \(k\) of \(\mathbf {M}^e_\mu \), showing legal positivity range \(\{\mu _{22}^{min},\mu _{22}^{max}\}\), b coefficient \(C_{12}\) of \(\kappa ^{12}\) in ABTS series

Figure 13b plots \(C_{12}(\mu _{22})=(-111545-3008\psi +15120(525+4\psi )\mu _{22})/685843200\), with \(\psi =\sqrt{70}\sqrt{19-1890\mu _{22}}\). This has one real root \(\mu _{22}^z=-0.02830257472322391\), but that gives an indefinite mass matrix. For \(\mu _{22}\) in the legal range \([\mu _{22}^{min},\mu _{22}^{max}]\), \(C_{12}\) is minimized for \(\mu _{22}^b= (25\sqrt{105}-171)/30240= 0.0028165928951385567\), which substituted gives the optimal mass matrix:

$$\begin{aligned} \mathbf {M}^e_{LFFO}&={m^e\over 30240} \left[ \begin{array}{cccc} a_{11} &{} \quad 1788\ell &{} \quad a_{13} &{} \quad -732\ell \\ &{} \quad a_{22}\ell ^2 &{} \quad 732\ell &{} \quad a_{24}\ell ^2 \\ &{} \quad &{} \quad a_{33} &{} \quad 1788\ell \\ \hbox {symm} &{} \quad &{} \quad &{} \quad a_{44}\ell ^2 \\ \end{array}\right] \nonumber \\&= m^e \!\left[ \!\begin{array}{cccc} 0.389589 &{} \quad 0.059127\ell &{} \quad 0.110410 &{} \quad -0.024206\ell \\ &{} \quad 0.012340\ell ^2 &{} \quad 0.024206\ell &{} \quad -0.005548\ell ^2 \\ &{} \quad &{} \quad 0.389589 &{} \quad -0.059127\ell \\ &{} \quad &{} \quad &{} \quad 0.012340\ell ^2 \end{array}\!\right] \!. \end{aligned}$$
(99)

in which \(a_{11}=a_{33}=12396 - 60 \sqrt{105},\, a_{13}=2724 + 60 \sqrt{105},\, a_{22}=a_{44}=117 + 25 \sqrt{105}\) and \(a_{24}=-219 + 5 \sqrt{105}\). For this set, \(C_{12}=(25\sqrt{105}-441)/91445760= -2.021\,10^{-6}\). Another interesting value is \(\mu _{22}=13/3150= 0.004126984126984127\), which substituted in (98) yields rational values for the other parameters: \(\mu _{11}=-\mu _{13}=23/2100, \, \mu _{12}=-\mu _{14}=-\mu _{23}=23/4200,\, \mu _{24}=23/4200\) and \(\mu _{24}=-17/12600\). Substitution into (91) gives

$$\begin{aligned} \mathbf {M}^e_{BLFM}&={m^e\over 12600} \left[ \begin{array}{cccc} 4818 &{} \quad 729\ell &{} \quad 1482 &{} \quad -321\ell \\ &{} \quad 172\ell ^2 &{} \quad 321\ell &{} \quad -73\ell ^2 \\ &{} \quad &{} \quad 4818 &{} \quad -729\ell \\ \hbox {symm} &{} \quad &{} \quad &{} \quad 172\ell ^2 \\ \end{array}\right] \nonumber \\&= m^e\!\left[ \begin{array}{cccc} 0.382381 &{} \quad 0.057857\ell &{} \quad 0.117619 &{} \quad -0.025476\ell \\ &{} \quad 0.013651\ell ^2 &{} \quad 0.025476\ell &{} \quad -0.005794\ell ^2 \\ &{} \quad &{} \quad 0.382381 &{} \quad -0.057857\ell \\ \hbox {symm} &{} \quad &{} \quad &{} \quad 0.013651\ell ^2 \\ \end{array}\!\right] \!. \end{aligned}$$
(100)

For this matrix, \(C_{12}=-41/18144000=-2.26\,10^{-6}\). Its magnitude is only about 10 % higher than for the truly LFF optimal (99). Since its entries are simpler, (100) is adopted as BLFM matrix for the BE element, and used as a baseline for the Timoshenko beam element investigated in Sect. 7.

7 The Timoshenko Plane Beam Element

This last example is far more elaborate than the previous ones. The goal is to construct a mass template for the prismatic, plane beam Timoshenko model, a name often abbreviated to Ti-beam. It includes the BE model as special case; consequently results can be crosschecked with those of Sect. 6.2. One interesting feature of this model is that the continuum dispersion diagram has two branches, both of which are physical.

The “acoustical-like branch”, which has zero frequency at zero wavenumber, corresponds to lower-frequency bending oscillations in which the beam displaces transversely. The “optical-like branch”, which exhibits a nonzero cutoff frequency, corresponds to higher frequency shear oscillations. Because of these interpretations, they are called the flexural frequency branch (FFB) and the shear frequency branch (SFB), respectively. They are identified by subscripts \(f\) and \(s\), respectively, for the continuum model, while \(a\) and \(o\) are reused for the FEM discretization. Both branches are dispersive, meaning that group velocity depends on wavenumber. Those velocities tend to finite values, except in the BE limit.

The upshoot of these complications is that LFCF customization, which was so clear-cut with Bar3, becomes ambiguous: do we want to fit the FFB or the SSB? For thin beams, (as well as in the BE limit, in which case the SSB moves to \(\infty \)) the FFB is dominant. But as the beam becomes progressively thicker (as measured by a slenderness coefficient introduced in Sect. 7.1) the situation is less clear: for an extremely thick beam the shear oscillations may well dominate (of course in that case the Timoshenko model is questionable).

The continuum model is first studied in some detail, since frequency expansion formulas applicable to template customization by characteristic root fitting are not available in the literature.

7.1 Ti-Beam Continuum Elastodynamic Analysis

Consider a structural beam member modeled as a shear-flexible Timoshenko plane beam (Ti-beam), as illustrated in Fig. 14. This figure provides the notation used below. Section properties \(\{\rho ,E,A,A_s,I,I_R\}\) are constant along \(x\). The beam is transversally loaded by line load \(q(x,t)\) (not shown in figure), with dimension of force per length. The primary kinematic variables are the transverse deflection \(v(x,t)\) and the total cross-section rotation \(\theta (x,t)=v'(x,t)+\gamma (x,t)\), where \(\gamma =V/(GA_s)\) is the mean shear rotation. The kinetic and potential energies in terms of those variables are

$$\begin{aligned} T[v,\theta ]&= {\textstyle {1\over 2}}\int _0^L \big (\rho A \, \dot{v}^2 + \rho I_R \,\,\dot{\theta }^2\big )\,dx, \,\, \Pi [v,\theta ]\nonumber \\&= \int _0^L \Big ( {\textstyle {1\over 2}}EI(v'')^2 + {\textstyle {1\over 2}}GA_s(\theta {-}v'\big )^2-qv\Big )\,dx.\nonumber \\ \end{aligned}$$
(101)

where superposed dots denote time derivatives. The equations of motion (EOM) follow on forming the Euler equations from the Lagrangian \(L=T-\Pi \):

$$\begin{aligned} {\delta L\over \delta v}&= 0\rightarrow GA_s\,(\theta '-v'')+\rho A\ddot{v} =q, \;\, {\delta L\over \delta \theta }=0\rightarrow EI\,\theta ''\nonumber \\&+ GA_s\,(v'-\theta )-\rho I_R\,\ddot{\theta }=0. \end{aligned}$$
(102)

An expedient way to eliminate \(\theta \) is to rewrite the coupled equations (102) in transform space:

$$\begin{aligned} \left[ \begin{array}{ll} \rho A s^2-GA_s\,p^2 &{} GA_s\,p \\ GA_s\,p &{} EI\,p^2-GA_s-\rho \,I_R\,s^2\\ \end{array}\right] \left[ \begin{array}{c} \tilde{v}\\ \tilde{\theta }\\ \end{array}\right] = \left[ \begin{array}{c} \tilde{q}\\ 0 \end{array}\right] , \end{aligned}$$
(103)

in which \(\{p,s,\tilde{v},\tilde{\theta },\tilde{q}\}\) denote transforms of \(\{d/dx,d/dt,v, \theta ,q\}\), respectively (Fourier in \(x\) and Laplace in \(t\)). Eliminating \(\tilde{\theta }\) and returning to the physical domain yields

$$\begin{aligned} EI \,v''''+\rho A \ddot{v}&- \left( {\rho I_R}+{\rho A EI\over GA_s}\right) \ddot{v}''+ {\rho ^2 A I_R\over GA_s}{\ddddot{v}}\nonumber \\&\quad = q-{EI\over G A_s}q''+{\rho I_R\over G A_s}\ddot{q}. \end{aligned}$$
(104)

This derivation does not preset \(I\equiv I_R\), as usually done in textbooks. For the unforced case \(q=0\), (104) has plane wave solutions \(v=B\,\exp \big (i\,(k_0\,x-\omega _0\,t)\big )\). The propagation condition yields a characteristic equation relating \(k_0\) and \(\omega _0\). To render it dimensionless, introduce a reference phase velocity \(c_0^2=EI/(\rho A L^4)\) so that \(\,k_0=\omega _0/c_0=2\pi /\lambda _0\), a dimensionless frequency \(\Omega =\omega _0 L/c_0\) and a dimensionless wavenumber \(\kappa =k_0 L\).

Fig. 14
figure 14

Plane beam member modeled as Ti-beam, illustrating notation followed in Sect. 7.1. Transverse load \(q(x)\) not shown to reduce clutter. Infinitesimal deflections and deformations grossly exaggerated for visibility

As dimensionless measures of relative bending-to-shear rigidities and rotary inertia take

$$\begin{aligned} \Phi _0=12EI/(GA_s L^2), \qquad r_R^2=I_R/A, \qquad \Psi _0=r_R/L. \end{aligned}$$
(105)

The resulting dimensionless characteristic equation is

$$\begin{aligned} \kappa ^4-\Omega ^2-({\textstyle {1\over 12}}\Phi _0+\Psi _0^2)\,\kappa ^2\,\Omega ^2+ {\textstyle {1\over 12}}\Phi _0 \,\Psi _0^2\,\Omega ^4=0. \end{aligned}$$
(106)

This is quadratic in \(\Omega ^2\). Its solution yields two kinds of squared-frequencies, which will be denoted by \(\Omega _f^2\) and \(\Omega _s^2\) because they are associated with flexural and shear modes, respectively. Their expressions are listed below along with their small-\(\kappa \) (long wavelength) Taylor series:

$$\begin{aligned} \Omega ^2_f&=6{P-\sqrt{Q}\over \Phi _0\Psi _0^2} = \kappa ^4 - ({\textstyle {1\over 12}}\Phi _0+\Psi _0^2)\, \kappa ^6\nonumber \\&+ \, \left( \frac{1}{144}\Phi _0^2 + {\textstyle {1\over 4}}\Phi _0\Psi _0^2 + \Psi _0^4\right) \, \kappa ^8 \nonumber \\&\quad -(\frac{1}{1728}\Phi _0^3+ \frac{1}{24}\Phi _0^2\Psi _0^2+{\textstyle {1\over 2}}\Phi _0\Psi _0^4+\Psi _0^6) \,\kappa ^{10}+\cdots \nonumber \\&\qquad = A_4\kappa ^4+A_6\kappa ^6+ A_8\kappa ^8 + \cdots .\end{aligned}$$
(107)
$$\begin{aligned} \Omega ^2_s&=6{P+\sqrt{Q}\over \Phi _0\Psi _0^2} = {12\over \Phi _0\Psi _0^2} + \left( {12\over \Phi _0} + {1\over \Psi _0^2}\right) \kappa ^2 -\kappa ^4\nonumber \\&\quad +\, ( {\textstyle {1\over 12}}\Phi _0 + \Psi _0^2)\kappa ^6 + \cdots = B_0 + B_2\kappa ^2 + \cdots , \end{aligned}$$
(108)

in which \( P=1+\kappa ^2(\Psi _0^2 +{\textstyle {1\over 12}}\Phi _0)\) and \(Q=P^2-{\textstyle {1\over 3}}\kappa ^4\Phi _0\Psi _0^2\). The dispersion relation \(\Omega ^2_f(\kappa )\) defines the flexural frequency branch (FFB) whereas \(\Omega ^2_s(\kappa )\) defines the shear frequency branch (SFB). If \(\Phi _0\rightarrow 0\) and \(\Psi _0\rightarrow 0\), which reduces the the Ti model to BE, (106) collapses to \(\Omega ^2=\kappa ^4\) or (in principal value) \(\Omega =\kappa ^2\). This surviving branch pertains to flexural motions whereas the shear branch disappears; more precisely, \(\Omega ^2_s(\kappa ) \rightarrow \infty \).

It is easily shown that the radicand \(Q\) in the exact expressions is strictly positive for any \(\{\Phi _0>0,\Psi _0>0,\kappa \ge 0\}\). Thus for any such triple, \(\Omega _f^2\) and \(\Omega _s^2\) are real, finite and distinct with \(\Omega _f^2(\kappa )<\Omega _s^2(\kappa )\). Further \(\{\Omega _f^2,\Omega _s^2\}\) increase indefinitely as \(\kappa \rightarrow \infty \). Following the dispersion-diagram nomenclature introduced in Fig. 8, the value \(\Omega _s\) at \(\kappa =0\) is called the cutoff frequency.

To see what branches look like, consider a beam of narrow rectangular cross section of width \(b\) and height \(h\), fabricated of isotropic material with Poisson’s ratio \(\nu \). Accordingly \(E/G=2(1+\nu )\) and \(A_s/A\approx 5/6\) (actually a more refined \(A_s/A\) ratio would be \(10(1+\nu )/(12+11\nu )\), but that makes little difference in the results). We have \(A=bh\), \(I=I_R= b h^3/12\), \(r_R^2=I_R/A=h^2/12\), \(\Psi _0^2=r_R^2/L^2={\textstyle {1\over 12}}h^2/L^2\) and \(\Phi _0=12 EI/(GA_sL^2)=12(1+\nu )h^2/(5L^2)\). Since \(\Phi _0/12=12(1+\nu )\,\Psi _0^2/5\), the first-order effect of shear on \(\Omega _f^2\), as measured by the \(\kappa ^6\) term in (107), is 2.4 to 3.6 times that from rotary inertia, depending on \(\nu \). Replacing into (107) and (108) yields

$$\begin{aligned} \left. \begin{array}{lll} \Omega ^2_f\\ \Omega ^2_s\\ \end{array}\right\}&={60+\kappa ^2(17+12\nu )\Lambda ^2 \mp \sqrt{\big (60+\kappa ^2(17+12\nu )\Lambda ^2\big )^2- 240\kappa ^4(1+\nu )\Lambda ^4}\over 2(1+\nu ) \Lambda ^4} \\&=\left\{ \begin{array}{l} \kappa ^4-\frac{1}{60}(17+12\nu )\Lambda ^2\,\kappa ^6+ \frac{1}{3600}(349+468\nu +144\nu ^2)\Lambda ^4\kappa ^8 +\cdots \\ {\displaystyle 60+ (17+12\nu )\Lambda ^2\kappa ^2 - (1+\nu )\Lambda ^4\,\kappa ^4 +\ldots \over \displaystyle (1+\nu )\Lambda ^4 } \end{array}\right. \end{aligned}$$
(109)

in which \(\Lambda =h/L\). Dispersion curves \(\Omega (\kappa )\) for \(\Lambda =h/L=\frac{1}{4}\) and \(\nu =\{0,{\textstyle {1\over 2}}\}\) are plotted in Fig. 15a. Phase velocities \(\Omega /\kappa \) are shown in Fig. 15b. The figure also shows the flexural branch of the BE model. The phase velocities of the Timoshenko model tend to finite values in the shortwave, high-frequency limit \(\kappa \rightarrow \infty \), which is physically correct. The BE model is physically wrong in that limit because it predicts an infinite propagation speed.

Fig. 15
figure 15

Spectral behavior of continuum Ti-beam model for a narrow \(b\times h\) rectangular cross section. a Dispersion curves \(\Omega (\kappa )\) for \(\Lambda =h/\ell =1/4\) and two Poisson’s ratios; Timoshenko flexural and shear branches in red and blue, respectively; BE curve \(\Omega =\kappa ^2\) in black, b Wavespeed \(\partial \Omega /\partial \kappa \). (Color figure online)

7.2 Ti-Beam Element

The shear-flexible plane beam member of Fig. 14 is discretized by two-node elements. An individual element of this type is shown in Fig. 16, which illustrates its kinematics. The element has four nodal freedoms arranged as

$$\begin{aligned} \mathbf {u}^e=\left[ \begin{array}{lllll}v_1&{}\quad \theta _1&{}\quad v_2&{}\quad \theta _2\\ \end{array}\right] ^T \end{aligned}$$
(110)

Here \(\theta _1=v_1+\gamma _1\) and \(\theta _2=v_2+\gamma _2\) are the total cross section rotations evaluated at the end nodes. The dimensionless properties (105) that characterize relative shear rigidity and rotary inertia are redefined using the element length:

$$\begin{aligned} \Phi =12EI/(GA_s \ell ^2), \qquad r_R^2=I_R/A, \qquad \Psi =r_R/\ell . \end{aligned}$$
(111)

If the beam member is divided into \(N_e\) elements of equal length, \(\ell =L/N_e\) whence \(\Phi =\Phi _0N_e^2\) and \(\Psi =\Psi _0 N_e\). Thus even if \(\Phi _0\) and \(\Psi _0\) are small with respect to one, they can grow without bound as the mesh is refined. For example if \(\Phi _0=1/4\) and \(\Psi _0^2=1/100\), which are typical values for a moderately thick beam, and we take \(N_e=32\), then \(\Phi \approx 250\) and \(\Psi ^2\approx 10\). Those are no longer small numbers, a fact that will impact performance as \(N_e\) increases. The stiffness matrix to be paired with the mass template is taken to be that of the equilibrium element:

$$\begin{aligned} \mathbf {K}^e={EI\over \ell ^3(1+\Phi )} \left[ \begin{array}{cccc} 12 &{} \quad 6 \ell &{} \quad -12 &{} \quad 6\ell \\ 6\ell &{} \quad \ell ^2 (4 + \Phi ) &{} \quad -6\ell &{} \quad \ell ^2(2-\Phi ) \\ -12 &{} \quad -6 \ell &{} \quad 12 &{} \quad -6 \ell \\ 6\ell &{} \quad \ell ^2(2-\Phi ) &{} \quad -6\ell &{} \quad \ell ^2(4 +\Phi ) \end{array}\right] \!. \end{aligned}$$
(112)

This is known to be optimal in static analysis for a prismatic beam member. It will be kept fixed in the ensuing derivations. It reduces to the stiffness matrix (93) of the BE model if \(\Phi =0\).

Fig. 16
figure 16

Two-node element for Timoshenko plane beam, illustrating kinematics

7.3 The Ti-Beam Mass Template

FEM derivations usually split the \(4\times 4\) mass matrix of this element into \(\mathbf {M}^e=\mathbf {M}_v^e+\mathbf {M}_\theta ^e\), where \(\mathbf {M}_v^e\) and \(\mathbf {M}_\theta ^e\) come from the translational inertia and rotary inertia terms, respectively, of the kinetic energy functional \(T[v,\theta ]\) of (101). The most general mass template would result from applying a EW parametrization of those two matrices. This would require a set of 20 parameters (10 in each matrix), reducible to 9 through 11 on account of invariance and conservation conditions. Attacking the problem this way, however, leads to unwieldy algebraic equations even with the help of a CAS, while concealing the underlying physics. A divide and conquer approach works better. This is briefly outlined next and covered in more detail in the next subsections.

(I) Express \(\mathbf {M}^e\) as the one-parameter matrix-weighted form \(\mathbf {M}^e=(1-\mu _0)\,\mathbf {M}_F^e + \mu _0\,\mathbf {M}_D^e\). Here \(\mathbf {M}_F^e\) is full and includes the CMM as instance, whereas \(\mathbf {M}_D^e\) is \(2\times 2\) block diagonal and includes the DLMM as instance. This is plainly a generalization of the LC linear combination (2).

(II) Decompose the foregoing mass components as \(\mathbf {M}_F^e=\mathbf {M}_{FT}^e+\mathbf {M}_{FR}\) and \(\mathbf {M}_D^e=\mathbf {M}_{DT}^e+\mathbf {M}_{DR}^e\), where T and R subscripts identify their source in the kinetic energy functional: \(T\) if coming from the translational inertia term \({\textstyle {1\over 2}}\rho A \,\dot{v}^2\) and \(R\) from the rotary inertia term \({\textstyle {1\over 2}}\rho I_R \,\dot{\theta }^2\).

(III) Both components of \(\mathbf {M}_F^e\) are expressed as parametrized spectral forms, whereas those of \(\mathbf {M}_D^e\) are expressed as EW. The main reasons for choosing spectral forms for the full matrix are reduction of parameters and physical transparency. No such concerns apply to \(\mathbf {M}_D^e\).

The analysis follows a “bottom up” sequence, in order (III)–(II)–(I). This has the advantage that if a satisfactory custom mass matrix for a target application emerges during (III), stages (II) and (I) need not be carried out, and that matrix directly used by setting the remaining parameters to zero.

7.4 Ti-Beam Full Mass Parametrization

As noted above, one starts with full-matrix spectral forms. Let \(\xi \) denote the natural iso-P coordinate that varies from \(-1\) at node 1 to \(+1\) at node 2. Two element transverse displacement expansions in generalized coordinates are introduced:

$$\begin{aligned} v_T(\xi )&=L_{1}(\xi )\,c_{T1}+L_{2}(\xi )\, c_{T2}+L_{3}(\xi )\,c_{T3}\nonumber \\&\quad +\, L_{4}(\xi )\,c_{T4}=\mathbf {L}_T\,\mathbf {c}_T,\nonumber \\ v_R(\xi )&=L_{1}(\xi )\,c_{R1}+L_{2}(\xi )\, c_{R2}+L_{3}(\xi )\,c_{R3}\nonumber \\&\quad +\, \widehat{L}_{4}(\xi )\,c_{R4}=\mathbf {L}_R\,\mathbf {c}_R,\nonumber \\ L_{1}(\xi )&= 1, \ L_{2}(\xi ) = \xi , \, L_{3}(\xi ) = {\textstyle {1\over 2}}(3\xi ^2-1), \nonumber \\ L_{4}(\xi )&= {\textstyle {1\over 2}}(5\xi ^3-3\xi ), \nonumber \\ \widehat{L}_{4}(\xi )&= {\textstyle {1\over 2}}\big (5\xi ^3-(5+10\Phi )\xi \big ) = L_{4}(\xi )-(1+5\Phi )\xi . \end{aligned}$$
(113)

The \(v_T\) and \(v_R\) expansions are used for the translational and rotational parts of the kinetic energy, respectively. The interpolation function set \(\{L_{i}\}\) used for \(v_T\) is formed by the first four Legendre polynomials over \(\xi =[-1,1]\). The set used for \(v_R\) is the same except that \(L_{4}\) is adjusted to \(\widehat{L}_4\) to produce a diagonal rotational mass matrix. All amplitudes \(c_{Ti}\) and \(c_{Ri}\) have dimension of length.

Unlike the usual Hermite cubic shape functions, the polynomials in (113) have a direct physical interpretation. \(L_1\): translational rigid mode; \(L_2\): rotational rigid mode; \(L_3\): pure-bending mode symmetric about \(\xi =0\); \(L_4\) and \(\widehat{L}_4\): bending-with-shear mode antisymmetric about \(\xi =0\).

With the abbreviation \((.)'\equiv d(.)/dx=(2/\ell )d(.)/d\xi \), the associated cross section rotations are compactly expressed as

$$\begin{aligned} \theta _T=v_T'+\gamma _T=\mathbf {L}_T'\mathbf {c}_T+\gamma _T, \quad \theta _R=v_R'+\gamma _R=\mathbf {L}_R'\,\mathbf {c}_R+\gamma _R, \end{aligned}$$
(114)

in which the mean shear distortions are constant over the element:

$$\begin{aligned} \gamma _T={\Phi \ell ^2\over 12} v_T''' = {10\Phi \over \ell } c_{T4}, \quad \gamma _R={\Phi \ell ^2\over 12} v_R''' = {10\Phi \over \ell } c_{R4}. \end{aligned}$$
(115)

The kinetic energy of the element in generalized coordinates is

$$\begin{aligned} T^e&= {\textstyle {1\over 2}}\int _0^\ell \big (\rho A \, \dot{v}_T^2 + \rho I_R \,\,\dot{\theta }_R^2\big )\,dx\nonumber \\&= {\ell \over 4} \int _{-1}^1 \big (\rho A \, \dot{v}_T^2 + \rho I_R \,\,\dot{\theta }_R^2\big )\,d\xi \nonumber \\&= {\textstyle {1\over 2}}\dot{\mathbf {c}}_T^T\,\mathbf {D}_T\,\dot{\mathbf {c}}_T + {\textstyle {1\over 2}}\dot{\mathbf {c}}_R^T\,\mathbf {D}_R\,\dot{\mathbf {c}}_R, \end{aligned}$$
(116)

Both generalized mass matrices turn out to be diagonal as expected:

$$\begin{aligned} \mathbf {D}_T&= m^e\,\mathbf{diag}\left[ \begin{array}{cccc} 1&{} \quad {\textstyle {1\over 3}}&{} \quad {\textstyle {1\over 5}}&{} \quad {\textstyle {1\over 7}}\\ \end{array}\right] ,\nonumber \\ \mathbf {D}_R&= 4m^e\,\Psi ^2 \,\mathbf{diag}\left[ \begin{array}{cccc} 0&{} \quad 1&{} \quad 3&{} \quad 5\\ \end{array}\right] , \end{aligned}$$
(117)

in which as usual \(m^e=\rho \,A\,\ell \). To convert \(\mathbf {D}_T\) and \(\mathbf {D}_R\) to physical coordinates (110), \(v_T,\, v_R,\, \theta _T\) and \(\theta _R\) are evaluated at the nodes by setting \(\xi =\pm 1\). This establishes the transformations \(\mathbf {u}^e=\mathbf {G}_T\,\mathbf {c}_T\) and \(\mathbf {u}^e=\mathbf {G}_R\,\mathbf {c}_R\). Inverting: \(\mathbf {c}_T=\mathbf {H}_T\,\mathbf {u}^e\) and \(\mathbf {c}_R=\mathbf {H}_R\,\mathbf {u}^e\) with \(\mathbf {H}_T=\mathbf {G}_T^{-1}\) and \(\mathbf {H}_R=\mathbf {G}_R^{-1}\). A symbolic calculation yields for \(\mathbf {H}_T\):

$$\begin{aligned} \mathbf {H}_T = {1\over 60 (1+\Phi )} \left[ \begin{array}{llll} 30(1+\Phi ) &{} \quad 5\ell (1+\Phi ) &{} \quad 30(1+\Phi ) &{} \quad -5\ell (1+\Phi ) \\ -36-30\Phi &{} \quad -3\ell &{} \quad 36+30\Phi &{} \quad -3\ell \\ 0 &{} \quad -5\ell (1+\Phi ) &{} \quad 0 &{} \quad 5\ell (1+\Phi ) \\ 6 &{} \quad 3\ell &{} \quad -6 &{} \quad 3\ell \\ \end{array}\right] . \end{aligned}$$
(118)

Matrix \(\mathbf {H}_R\) differs only in the second row:

$$\begin{aligned} \mathbf {H}_R={1\over 60(1+\Phi )} \left[ \begin{array}{llll} 30(1+\Phi ) &{} \quad 5\ell (1+\Phi ) &{} \quad 30(1+\Phi ) &{} \quad -5\ell (1+\Phi ) \\ -30 &{} \quad 15\ell \Phi &{} \quad 30 &{} \quad 15\ell \Phi \\ 0 &{} \quad -5\ell (1+\Phi ) &{} \quad 0 &{} \quad 5\ell (1+\Phi ) \\ 6 &{} \quad 3\ell &{} \quad -6 &{} \quad 3\ell \\ \end{array}\right] . \end{aligned}$$
(119)

The difference comes from adjusting \(L_4\) to \(\widehat{L}_4\) in (113). To map this into a spectral template, inject six free parameters in the generalized masses while moving \(4\Psi ^2\) inside \(\mathbf {D}_{R\mu }\):

$$\begin{aligned} \mathbf {D}_{T\mu }&= m^e\,\mathbf{diag}\left[ \begin{array}{cccc} 1&{}\quad {\textstyle {1\over 3}}\mu _{T1}&{}\quad {\textstyle {1\over 5}}\mu _{T2}&{}\quad {\textstyle {1\over 7}}\mu _{T3}\\ \end{array}\right] , \quad \mathbf {D}_{R\mu }\nonumber \\&=m^e\,\mathbf{diag}\left[ \begin{array}{cccc} 0&{}\quad \mu _{R1}&{}\quad 3 \mu _{R2} &{}\quad 5\mu _{R3} \\ \end{array}\right] . \end{aligned}$$
(120)

The transformation matrices (118) and (119) can be reused without change to produce \(\mathbf {M}^e_F=\mathbf {H}_T^T\mathbf {D}_{T\mu }\mathbf {H}_T+\mathbf {H}_R^T\mathbf {D}_{R\mu }\mathbf {H}_R\). If \(\mu _{T1}=\mu _{T2}=\mu _{T3}=1\) and \(\mu _{R1}=\mu _{R2}=\mu _{R3}=4\Psi ^2\) one obtains the well known CMM as a valuable check. The configuration (120) already accounts for linear momentum conservation, which is why the upper diagonal entries are not parametrized. Enforcing also angular momentum conservation requires \(\mu _{T1}=1\) and \(\mu _{R1}=4\Psi ^2\), whence the template is reduced to four parameters:

$$\begin{aligned} \mathbf {M}_F^e&= m^e\,\mathbf {H}_T^T\, \left[ \begin{array}{llll} 1&{} \quad 0&{} \quad 0&{} \quad 0\\ 0&{} \quad {\textstyle {1\over 3}}&{} \quad 0&{} \quad 0\\ 0&{} \quad 0&{} \quad {\textstyle {1\over 5}}\mu _{T2}&{} \quad 0\\ 0&{} \quad 0&{} \quad 0&{} \quad {\textstyle {1\over 7}}\mu _{T3}\\ \end{array}\right] \mathbf {H}_T\nonumber \\&+\,m^e \,\mathbf {H}_R^T\, \left[ \begin{array}{llll} 0&{} \quad 0&{} \quad 0&{} \quad 0\\ 0&{} \quad 4\Psi ^2 &{} \quad 0&{} \quad 0\\ 0&{} \quad 0&{} \quad 3\mu _{R2}&{} \quad 0\\ 0&{} \quad 0&{} \quad 0&{} \quad 5\mu _{R3}\\ \end{array}\right] \mathbf {H}_R. \end{aligned}$$
(121)

Since both \(\mathbf {H}_T\) and \(\mathbf {H}_R\) are nonsingular, choosing all parameters in (121) to be nonnegative guarantees that \(\mathbf {M}_F^e\) is nonnegative. This property eliminates lengthy a posteriori checks.

Setting \(\mu _{T2}^{}=\mu _{T3}^{}=\mu _{R2}^{}=\mu _{R3}^{}=0\) and \(\Phi =0\) yields the correct mass matrix for a rigid beam, including rotary inertia. This simple result highlights the physical transparency of spectral forms.

7.5 Ti-Beam Block-Diagonal Mass Parametrization

Template (121) has a flaw: it does not include the DLMM. To remedy the omission, a block diagonal form, with four free parameters: \(\nu _{T1},\, \nu _{T2}\), \(\nu _{R1}\), and \(\nu _{R2}\) is separately constructed:

$$\begin{aligned} \mathbf {M}_D^e&= m^e \,\left[ \begin{array}{lllll} {\textstyle {1\over 2}}&{} \quad \nu _{T1}\ell &{} \quad 0&{} \quad 0\\ \nu _{T1}\ell &{} \quad \nu _{T2}\ell ^2&{} \quad 0&{} \quad 0\\ 0&{} \quad 0&{} \quad {\textstyle {1\over 2}}&{} \quad -\nu _{T1}\ell \\ 0&{} \quad 0&{} \quad -\nu _{T1}\ell &{} \quad \nu _{T2}\ell ^2\\ \\ \end{array}\right] \nonumber \\&+\, m^e \,\left[ \begin{array}{lllll} 0&{} \quad \nu _{R1}\ell &{} \quad 0&{} \quad 0\\ \nu _{R1}\ell &{} \quad \nu _{R2}\ell ^2&{} \quad 0&{} \quad 0\\ 0&{} \quad 0&{} \quad 0&{} \quad -\nu _{R1}\ell \\ 0&{} \quad 0&{} \quad -\nu _{R1}\ell &{} \quad \nu _{R2}\ell ^2\\ \\ \end{array}\right] \!. \end{aligned}$$
(122)

Four parameters can be merged into two by adding the foregoing matrices:

$$\begin{aligned} \mathbf {M}_D^e= m^e\, \left[ \begin{array}{cccc} {\textstyle {1\over 2}}&{} \quad \nu _{1}\ell &{} \quad 0&{} \quad 0\\ \nu _{1}\ell &{} \quad \nu _{2}\ell ^2&{} \quad 0&{} \quad 0\\ 0&{} \quad 0&{} \quad {\textstyle {1\over 2}}&{} \quad -\nu _{1}\ell \\ 0&{} \quad 0&{} \quad -\nu _{1}\ell &{} \quad \nu _{2}\ell ^2\\ \end{array}\right] \!. \end{aligned}$$
(123)

in which \(\nu _1=\nu _{T1}+\nu _{R1}\) and \(\nu _2=\nu _{T2}+\nu _{R2}\). Sometimes it is convenient to use the split form (122), for example in lattices with varying beam properties or lengths, a topic not considered there. Otherwise (123) suffices. If \(\nu _{1}=0,\, \mathbf {M}_D^e\) is diagonal. However for computational purposes a block diagonal form is just as good and provides additional customization power. Terms in the (1,1) and (3,3) positions must be as shown to satisfy linear momentum conservation. If angular momentum conservation is imposed a priori it is necessary to set \(\nu _2={\textstyle {1\over 2}}\Psi ^2\), and only one parameter: \(\nu _1\), remains.

The general template is obtained as a linear combination of \(\mathbf {M}_F^e\) and \(\mathbf {M}_D^e\):

$$\begin{aligned} \mathbf {M}^e=(1-\mu _0) \mathbf {M}_F^e + \mu _0\mathbf {M}_D^e \end{aligned}$$
(124)

In summary, there is a total of 7 parameters to play with: 4 in \(\mathbf {M}_F^e\), 2 in \(\mathbf {M}_D^e\), plus \(\mu _0\). This is less that the 9-to-11 count that would result from a full EW parametrization, so not all possible mass matrices for this element are included by (124).

7.6 Ti-Beam Fourier Analysis

An infinite lattice of identical Ti-beam elements of length \(\ell \) is set up in th usual manner. As in Sect. 6.2, plane waves of wavenumber \(k\) and frequency \(\omega \) propagating over the lattice are represented by

$$\begin{aligned} v(x,t) = B_v\,\exp \big (i(kx -\omega t\big ), \quad \theta (x,t) = B_\theta \,\exp \big (i(kx-\omega t\big ). \end{aligned}$$
(125)

At each typical lattice node \(j\) there are two freedoms: \(v_j\) and \(\theta _j\). Two patch equations are extracted, and converted to dimensionless form on defining \(\kappa =k\ell \) and \(\Omega =\omega \,c/\ell \), in which \(c =EI/(\rho A\ell ^4)\) is a reference phase velocity (these should not be confused with \(c_0\)). The condition for wave propagation gives the characteristic matrix equation

$$\begin{aligned} \det \left[ \begin{array}{cc} C_{vv} &{} \quad C_{v\theta } \\ C_{\theta v} &{} \quad C_{\theta \theta } \end{array}\right] = C_{vv}C_{\theta \theta }-C_{v\theta }C_{\theta v}=0, \end{aligned}$$
(126)

in which the coefficients are complicated functions computed by Mathematica and omitted for brevity. Solving the equation provides two equations: \(\Omega _a^2\) and \(\Omega _o^2\), where \(a\) and \(o\) denote acoustic and OB, respectively. These are expanded in powers of \(\kappa \) for matching to the continuum. For the full mass matrix one obtains

$$\begin{aligned} \mathop {\Omega }\limits _a^2{=} \kappa ^4+ C_6\kappa ^6+ C_8\kappa ^8+C_{10}\kappa ^{10}+\cdots , \quad \Omega _o^2{=}D_0+ D_2\kappa ^2+ \cdots \end{aligned}$$
(127)

Coefficients of terms up to \(\kappa ^{12}\) were computed by Mathematica. Those relevant for parameter selection are

$$\begin{aligned} C_6&= -\Phi /12-\Psi ^2,\nonumber \\ C_8&= \big [2 - 15 \mu _{R2} - \mu _{T2} + 5 \Phi (1 + \Phi )\nonumber \\&+\,60 (1 + 3 \Phi ) \Psi ^2 + 720 \Psi ^4\,\big ]\big /720,\nonumber \\ C_{10}&= \big [-44 + 35 \mu _{T2} - 3 \mu _{T3} - 282 \Phi + 525 \mu _{R2} (1 + \Phi )\nonumber \\&-\, 105 \mu _{R3} (1 + \Phi ) \nonumber \\&+\, 1575 \mu _{R2} \Phi (1 + \Phi ) - \Phi (3 \mu _{T3} - 35 \mu _{T2} (4 + 3 \Phi )\nonumber \\&+\,35 \Phi (17 + 5 \Phi (3 + \Phi ))) \nonumber \\&+\, (-2940 + 12600 \mu _{R2} (1 + \Phi ) + 420 (2 \mu _{T2} (1 + \Phi )\nonumber \\&-\,5 \Phi (7 + 6 \Phi (2 + \Phi )))) \Psi ^2 \nonumber \\&-\, 25200 (2 + \Phi (7 + 6 \Phi )) \Psi ^4\nonumber \\&-\, 302400 (1 + \Phi ) \Psi ^6\,\big ]\big /\big [302400 (1+\Phi )\big ],\nonumber \\ D_0&= 25200 (1 + \Phi )/ \big [7+105\mu _{R2}+3\mu _{T3}+2100\Phi ^2\Psi ^2\big ],\nonumber \\ D_2&= \big [ 2100 (1 + \Phi ) (-56 - 35 \mu _{T2} + 3 \mu _{T3} - 63 \Phi \nonumber \\&+\,3 \mu _{T3} \Phi + 105 \mu _{R3} (1 + \Phi ) \nonumber \\&-\, 525 \mu _{R2} (1 + \Phi )^2 - 35 \mu _{T2} \Phi (2 + \Phi ))\nonumber \\&+ 2100 (1 + \Phi ) (3360 \Phi + 6300 \Phi ^2 \nonumber \\&+\, 2100 \Phi ^3) \Psi ^2 + 52920000 \Phi ^2 (1 + \Phi ) \Psi ^4\big ]\big / \big [7\nonumber \\&+\,105\mu _{R2}+3\mu _{T3}+2100\Phi ^2\Psi ^2\big ]^2. \end{aligned}$$
(128)

For the block-diagonal template (123):

$$\begin{aligned}&\!\!\!\Omega _a^2= \kappa ^4+ F_6\kappa ^6+ F_8\kappa ^8+F_{10}\kappa ^{10}+\cdots , \nonumber \\&\!\!\! \Omega _o^2=G_0+ G_2\kappa ^2+ \cdots \end{aligned}$$
(129)

in which

$$\begin{aligned}&F_6 =-24\nu _2-\Phi , \nonumber \\&F_8 ={2880\nu _2-5\Phi +360\nu _2\Phi -1-5\Phi +5\Phi ^2 \over 720} \nonumber \\&G_0 ={6\over \nu _2(1+\Phi )}, \quad G_2 ={24\nu _2+\Phi -2\over 2\nu _2(1+\Phi )}. \end{aligned}$$
(130)

Expansions for the 7-parameter template (124) are considerably more involved than the above ones, and are omitted for brevity.

7.7 Ti-Beam Selected Template Instances

Seven useful instances of the foregoing templates are identified and described in Table 11. Table 12 lists the template signatures that produce those instances. These tables include two well known mass matrices (CMM and DLMM) re-expressed in the template context, and five new ones. The latter were primarily obtained by matching series such as (128) and (129) to the continuum ones (107) and (108), up to a certain number of terms as indicated in Table 12.

For the spectral template it is possible to match the flexure branch up to \(O(\kappa ^{10})\). Trying to match \(O(\kappa ^{12})\) leads to complex solutions. For the diagonal template the choice is more restrictive. It is only possible to match flexure up to \(O(\kappa ^6)\), which leads to the instance called FLMM. Trying to go further gives imaginary solutions. For the 7-parameter template (124) it is again possible to match up to \(O(\kappa ^{10})\) but no further. The instance that exhibits least truncation error while retaining positivity is FBMG. This is globally optimal for the BE limit \(\Phi =\Psi =0\), but the results are only slightly better for the reasons discussed below. Matching both flexure and shear branches leads to instances SBM0 and SBM2, which have the disadvantages noted in Table 11.

Table 11 Ti-beam selected template instances
Table 12 Signatures of selected Ti-beam template Instances

The exact dispersion curves of these instances are shown in Fig. 17 for \(\Phi =48/125\) and \(\Psi ^2=1/75\), which pertains to a thick beam. On examining Fig. 17 it is obvious that trying to match the shear branch is quite difficult; the fit only works well over a tiny range near \(\kappa =0\).

Fig. 17
figure 17

DDD for selected Ti-beam mass template instances with \(\Phi =48/125\) and \(\Psi ^2=1/75\)

7.8 Ti-Beam Vibration Analysis Example

The vibration analysis performance of the seven Ti-beam template instances listed in Tables 11 and 12 is evaluated on a simply supported (SS) prismatic plane beam. The beam has length \(L\) and is divided into \(N_e\) identical elements. The cross section is rectangular with width \(b\) and height \(h\). The material is isotropic with Poisson’s ratio \(\nu =0\). Three different height-to-span ratios \(h/L\) that characterize a thin, moderately thick and thick beam, respectively, are considered. Results for the three configurations are collected in Figs. 18, 19 and 20, respectively, for the first three vibration frequencies. All calculations are rendered dimensionless using appropriate scaling.

Fig. 18
figure 18

Accuracy of first 3 natural vibration frequencies of SS prismatic beam using mass matrices of instances listed in Tables 11 and 12. BE model with \(\Phi _0=\Psi _0=0\). Exact (12-decimal) frequencies \(\Omega _1=\pi ^2=9.869604401089\), \(\Omega _2=4\pi ^2= 39.478417604357\) and \(\Omega _3=9\pi ^2=88.826439609804\). Cutoff frequency \(+\infty \)

Fig. 19
figure 19

Accuracy of first 3 natural vibration frequencies of SS prismatic beam using mass matrices of instances listed in Tables 11 and 12. Timoshenko model with \(\Phi _0=3/80=0.0375\) and \(\Psi _0^2=1/768=0.00130\), pertaining to a rectangular x-section with \(h/L=1/8\) and \(\nu =0\). Exact (12-decimal) frequencies \(\Omega _1=9.662562122511\), \(\Omega _2=36.507937703548\) and \(\Omega _3=75.894968024537\). Cutoff frequency \(\Omega _{cut}=12/(\Phi _0\Psi _0^2)=495.741868314549\)

Fig. 20
figure 20

Accuracy of first 3 natural vibration frequencies of SS prismatic beam using mass matrices of instances listed in Tables 11 and 12. Timoshenko model with \(\Phi _0=24/625=0.384\) and \(\Psi _0^2=1/75=0.0133\), pertaining to a rectangular x-section with \(h/L=2/5\) and \(\nu =0\). Exact (12-decimal) frequencies \(\Omega _1=8.287891683498\), \(\Omega _2=24.837128591729\) and \(\Omega _3= 43.182948411234\). Cutoff frequency \(\Omega _{cut}=12/(\Phi _0\Psi _0^2)=48.412291827593\)

The accuracy of the computed frequencies is depicted using log-log plots of dimensionless natural frequency error versus \(N_e\). The error is displayed as \(d=\log _{10}(|\Omega _{comp}-\Omega _{exact}|\), which gives at a glance the number of correct digits \(d\), versus \(\log _2 N_e\) for \(N_e=1\) through 32. If the LF error is approximately controlled by a truncation term of the form \(\propto \kappa ^m\), the log-log plot should be roughly a straight line of slope \(\propto m\), inasmuch as \(\kappa =k\ell =kL/N_e\) (note that the accuracy curves for CMM and FLMM are virtually on top of each other for \(N_e\ge 4\), although errors have opposite signs; that is why their average CDLA does much better).

The results for the BE model, shown in Fig. 18, agree perfectly with the truncation error in the \(\Omega ^2_f\) branch as listed in Table 11. For example, the top performers FBMG and FBMS gain digits twice as fast as CMM, DLMM and SBM2, since the formers match \(\Omega ^2_f\) to \(O(\kappa ^{10})\) whereas the latter do that only to \(O(\kappa ^6)\). Instances CDLA and SMB0, which agree through \(O(\kappa ^8)\), come in between. The highly complicated FBMG is only slightly better than the much simpler FBMS. The case for their high accuracy should be emphasized. For example, four FBMS elements give \(\Omega _1\) to six figures: \(9.86960281\ldots \) versus \(\pi ^2=9.86960440\ldots \), whereas CMM gives less than three: \(9.87216716\ldots \). The “accuracy ceiling” of about 11 digits for FBMS and FBMG observable for \(N_e>16\) is due to the eigensolver working in double precision (\(\approx \) 16 digits). Rerunning with higher (quad) floating point precision, the plots continues marching up as straight lines before leveling off at approximately 25 digits.

On passing to the Timoshenko model, the well ordered BE world of Fig. 18 unravels. The culprits are \(\Phi \) and \(\Psi \). These figure prominently in the branch series and grow without bound as \(N_e\) increases, as discussed in Sect. 7.2. Figure 19 collects results for a moderately thick beam with \(h/L=1/8\), which corresponds to \(\Phi _0=3/80\) and \(\Psi _0^2=1/768\). The BE top performers, FBMS and FBMG, gradually slow down and are caught by CDLA by \(N_e=32\). All other instances trail, with the standard ones: CMM and FLMM, becoming the worst performers. Note that for \(N_e=32\), CMM and FLMM provide only 1 digit of accuracy in \(\Omega _3\), although there are \(32/1.5\approx 21\) elements per wavelength.

Figure 20 collects results for a thick beam with \(h/L=2/5\), corresponding to \(\Phi _0=24/625\) and \(\Psi _0^2=1/75\). The trends of Fig. 19 are exacerbated, with FBMS and FBMG running out of steam by \(N_e=4\) and CDLA clearly emerging as best for \(N_e\ge 8\). Again CMM and FLMM trail badly.

The reason for the performance degradation of FBMS and FBMG as the Ti-beam gets thicker is unclear as of this writing. Eigensolver accuracy is not responsible since rerunning the cases of Figs. 19 and 20 in quad precision did not change the plots. A numerical study of the \(\Omega _f^2\) truncation error shows that FBMS and FBMG fit the continuum branch better than CDLA even for very thick beams. Possible contamination of vibration mode shapes with the shear branch was not investigated.

8 Conclusions and Future Work

It is clear from the previous studies that mass matrix customization by templates can be effective in structural dynamics. The examples of Sects. 4.9, 5.14 and 7.8 illustrates two typical advantages:

  • Orders of magnitude improvements in frequency accuracy can be achieved for the same computational effort.

  • The space discretization need not be changed at all. Only the template free parameters need to be adjusted by supplying the appropiate signature.

These should be attractive to engineers for practical FEM computations. The last one is particularly important, since redoing a structural dynamics model not amenable to mesh generation may take a significant portion of a design and analysis process.

Would availability of customized templates eliminate the need for \(h\) and \(p\) adaptivity? Certainly not. Elements have performance limits, so such refinement schemes cannot be ruled out. It should be noted, however, that mesh adaptivity is less effective in dynamics, particularly in problems with rapid transients and shocks. Irregular meshes and high order elements are notorious sources of HF pollution, and adaptivity can make things worse by exacerbating nonphysical dispersion.

As regards future work, the most ambitious plan is extension to multiple space dimensions. Additional challenges emerge there:

  • Directionality. This means that the dynamic accuracy of the FEM model, as compared to the continuum, depends on the direction of plane wave propagation. This is a phenomenon missing in 1D. Integration averaging over propagation angles necessarily appears as another component of the Fourier analysis.

  • Material property dependence. Optimal free parameters become dependent on additional elastic material properties missing from the 1D treatment. For example if the material is isotropic Poisson’s ratio appears. This is not actually unique to mass templates, but affects stiffness templates as well.

  • Multiple plane wave types. In isotropic 2D and 3D continua, one needs to consider two types: pressure (P-waves) and shear (S-waves). Plainly this impacts customization. If the medium is non-isotropic, more complicated wave types may have to be considered.

  • Parameter explosion. This can be expected to hinder symbolic calculations. At first sight it seems inevitable given the rapidly increasing size of the element matrices. Growth can be controlled, however, by making use of a priori reduction techniques such as those mentioned in Sects. 3.1, 3.2, and 3.4.

These challenges are illustrated for a simple 2D element (three-node linear triangle) in Appendix 5.

Less ambitious research thrusts may focus on extending the results presented here for 1D elements. For example:

  • Assess the performance of different template variant construction approaches to reduce high-frequency pollution caused in DTI. Three were described for the Bar3 element: singular mass, SMS and constant OB. As of this writing, no comparative rating based on numerical experiments is available.

  • Find out whether the impressive gains in accuracy observed in LFF-customized templates survive in irregular meshes and/or heterogeneous element mixtures.

  • Unfinished business remains for the Timoshenko beam. First, the unexpected deterioration in vibration accuracy as the coefficients \(\varvec{\Phi }\) and \(\varvec{\Psi }\) increase is presently unexplained. Could the degradation be arrested using MS template pairs? Prior experience with the BE-beam, referenced in Appendix Sect. 1.7, shows only modest improvements, but such continuum model is comparatively well behaved. Second, the relative performnace of the various template instances listed in Tables 11 and 12 for direct time integration (DTI) remains to be assessed.