Dynamic substructuring

Last updated

Dynamic Substructuring (DS) is an engineering tool used to model and analyse the dynamics of mechanical systems by means of its components or substructures. Using the dynamic substructuring approach one is able to analyse the dynamic behaviour of substructures separately and to later on calculate the assembled dynamics using coupling procedures. Dynamic substructuring has several advantages over the analysis of the fully assembled system:

Contents

Dynamic substructuring is particularly tailored to simulation of mechanical vibrations, which has implications for many product aspects such as sound / acoustics, fatigue / durability, comfort and safety. Also, dynamic substructuring is applicable to any scale of size and frequency. It is therefore a widely used paradigm in industrial applications ranging from automotive and aerospace engineering to design of wind turbines and high-tech precision machinery.

History

Two levels of domain decomposition in dynamic substructuring. Two different levels of domain decomposition.svg
Two levels of domain decomposition in dynamic substructuring.

The roots of dynamic substructuring can be found in the field of domain decomposition. In 1890 the mathematician Hermann Schwarz came up with an iterative procedure for domain decomposition which allows to solve for continuous coupled subdomains. However, many of the analytical models of coupled continuous subdomains do not have closed-form solutions, which led to discretization and approximation techniques such as the Ritz method [1] (which is sometimes called the Rayleigh-Ritz method due to the similarity between Ritz's formulation and the Rayleigh ratio) the boundary element method (BEM) and the finite element method (FEM). These methods can be considered as "first level" domain decomposition techniques.

The finite element method proved to be the most efficient method and the invention of the microprocessor made it possible to easily solve a large variety of physical problems. [2] In order to analyse even larger and more complex problems, methods were invented to optimize the efficiency of the discretized calculations. The first step was replacing the direct solvers by iterative solvers such as the conjugate gradient method. [3] The lack of robustness and slow convergence of these solvers did not make them an interesting alternative in the beginning. The rise of parallel computing in the 1980s however sparked their popularity. Complex problems could now be solved by dividing the problem into subdomains, each processed by a separate processor, and solving for the interface coupling iteratively. This can be seen as a second level domain decomposition as is visualized in the figure.

The efficiency of dynamic modelling could be increased even further by reducing the complexity of the individual subdomains. This reduction of the subdomains (or substructures in the context of structural dynamics) is realized by representing substructures by means of their general responses. Expressing the separate substructures by means of their general response instead of their detailed discretization led to the so-called dynamic substructuring method. This reduction step also allowed for replacing the mathematical description of the domains by experimentally obtained information. This reduction step is also visualized by the reduction arrow in the figure.

The first dynamic substructuring methods were developed in the 1960s and were more commonly known under the name component mode synthesis (CMS). The benefits of dynamic substructuring were quickly discovered by the scientific and engineering communities and it became an important research topic in the field of structural dynamics and vibrations. Major developments followed, resulting in e.g. the classic Craig-Bampton method. [4]

Due to improvements in sensor and signal processing technology in the 1980s, substructuring techniques also became attractive for the experimental community. Methods dealing with structural dynamic modification were created in which coupling techniques were directly applied to measured frequency response functions (FRFs). Broad popularity of the method was gained when Jetmundsen et al. formulated the classical frequency-based substructuring (FBS) method, [5] which laid the ground work for frequency-based dynamic substructuring. In 2006 a systematic notation was introduced by De Klerk et al. [6] in order to simplify the difficult and elaborate notation that had been used prior. The simplification was done by means of two Boolean matrices that handle all the "bookkeeping" involved in the assembly of substructures [7]

Domains

Five domains typically used for dynamic substructuring. Ds domains v3.png
Five domains typically used for dynamic substructuring.

Dynamic substructuring can best be seen as a domain-independent toolset for assembly of component models, rather than a modelling method of its own. Generally, dynamic substructuring can be used for all domains that are well suited to simulate multiple input/multiple output behaviour. [7] Five domains that are well suited for substructuring are:

The physical domain concerns methods that are based on (linearised) mass, damping and stiffness matrices, typically obtained from numerical FEM modelling. Popular solutions to solve the associated system of second order differential equations are the time integration schemes of Newmark [8] and the Hilbert-Hughes-Taylor scheme. [9] The modal domain concerns component mode synthesis (CMS) techniques such as the Craig-Bampton, Rubin and McNeal method. These methods provide efficient modal reduction bases and assembly techniques for numerical models in the physical domain. The frequency domain is more popularly known as Frequency Based Substructuring (FBS). Based on the classic formulation of Jetmundsen et al. [5] and the reformulation of De Klerk et al., [9] it has become the most commonly used domain for substructuring, because of the ease of expressing the differential equations of a dynamical system (by means of Frequency Response Functions, FRFs) and the convenience of implementing experimentally obtained models. The time domain refers to the recently proposed concept of Impulse Based Substructuring (IBS), [10] which expresses the behaviour of a dynamic system using a set of Impulse Response Functions (IRFs). The state-space domain, finally, refers to methods proposed by Sjövall et al. [11] that employ system identification techniques common to control theory.

An overview of the governing equations of the five herementioned domains is presented in the table below.

Dynamic equations for five domains
DomainDynamic equationAdditional information
Physical domain represent the linear(ised) mass, damping and stiffness matrix of the system.
Modal Domain represent the modally reduced mass, damping and stiffness matrix; is the set of modal amplitudes.
Frequency Domain is the impedance FRF matrix; is the admittance FRF matrix.
Time domain is the IRF matrix.
State-space domain are the state-space matrices; , and represent the state, input and output vector.

As dynamic substructuring is a domain-independent toolset, it is applicable to the dynamic equations of all domains. In order to establish substructure assembly in a particular domain, two interface conditions need to be implemented. This is explained next, followed by a few common substructuring techniques.

Interface Conditions

To establish substructuring coupling / decoupling in each of the above-mentioned domains, two conditions should be met:

These are the two essential conditions that keep substructures together, hence allow to construct an assembly of multiple components. Note that the conditions are comparable with Kirchhoff's laws for electric circuits, in which case similar conditions apply to currents and voltages though/over electric components in a network; see also Mechanical–electrical analogies.

Substructure connectivity

Assembly of two substructures A and B, connected by the DoFs
u
2
{\displaystyle \mathbf {u} _{2}}
and interface forces
g
2
{\displaystyle \mathbf {g} _{2}}
of the coupling nodes. Assembly of 2 substructures consisting of nodes.png
Assembly of two substructures A and B, connected by the DoFs and interface forces of the coupling nodes.

Consider two substructures A and B as depicted in the figure. The two substructures comprise a total of six nodes; the displacements of the nodes are described by a set of Degrees of Freedom (DoFs). The DoFs of the six nodes are partitioned as follows:

  1. DoFs of the internal nodes of substructure A;
  2. DoFs of the coupling nodes of substructures A and B, i.e. interface DoFs;
  3. DoFs of the internal nodes of substructure B.

Note that the denotation 1, 2 and 3 indicates the function of the nodes/DoFs rather than the total amount. Let us define the sets of DoFs for the two substructures A and B in concatenated form. The displacements and applied forces are represented by the sets and . For the purpose of substructuring, a set of interface forces is introduced which only contains non-zero entries on the interface DoFs:

The relation between dynamic displacements and applied forces of the uncoupled problem is governed by a particular dynamic equation, such as presented in the table above. The uncoupled equations of motion are augmented by extra terms/equations for compatibility and equilibrium, as discussed next.

Compatibility

The compatibility condition requires that the interface DoFs have the same sign and value at both sides of the interface: . This condition can be expressed using a so-called signed Boolean matrix, [6] denoted by . For the given example this can be expressed as:

In some cases the interface nodes of the substructures are non-conforming, e.g. when two substructures are meshed separately. In such cases a non-Boolean matrix has to be used in order to enforce a weak interface compatibility. [12] [13]

A second form in which the compatibility condition can be expressed is by means of coordinate substitution by a set of generalised coordinates . The set contains the unique coordinates that remain after assembly of the substructures. Every matching pair of interface DoFs is described by a single generalised coordinate, which means that the compatibility condition is automatically enforced. Expressing using gives:

Matrix is referred to as the Boolean localisation matrix. A useful relation between matrix and can be exposed by noting that compatibility should hold for any set of physical coordinates expressed by . Indeed, substituting in the equation :

Hence represents the nullspace of :

This means in practice that one only needs to define or ; the other Boolean matrix is calculated using the nullspace property.

Equilibrium

The second condition that has to be satisfied for substructure assembly is the force equilibrium for matching interface forces . For the current example, this condition can be written as . Similar to the compatibility equation, the force equilibrium condition can be expressed using a Boolean matrix. Use is made of the transpose of the Boolean localisation matrix that was introduced to write compatibility:

The equations for and state that the interface forces on internal nodes are zero, hence not present. The equation for correctly establishes the force equilibrium between a matching pair of interface DoFs according to Newton's third law.

A second notation in which the equilibrium condition can be expressed is by introducing a set of Lagrange multipliers . The substitution of these Lagrange multipliers is possible as and differ only in sign, not in value. Using again the signed Boolean matrix :

The set defines the intensity of the interface forces . Each Lagrange multiplier represents the magnitude of two matching interface forces in the assembly. By defining the interface forces using Lagrange multipliers , force equilibrium is automatically satisfied. This can be seen by substituting into the first equilibrium equation:

Again, the nullspace property of the Boolean matrices is used here, namely: .

The two conditions as presented above can be applied to establish coupling / decoupling in a myriad of domains and are thus independent of variables such as time, frequency, mode, etc. Some implementations of the interface conditions for the most common domains of substructuring are presented below.

Substructuring in the physical domain

The physical domain is the domain that has the most straightforward physical interpretation. For each discrete linearised dynamic system one is able to write an equilibrium between the externally applied forces and the internal forces originating from intrinsic inertia, viscous damping and elasticity. This relation is governed by one of the most elementary formulas in structural vibrations:

represent the mass, damping and stiffness matrix of the system. These matrices are often obtained from finite element modelling (FEM), and are referred to as the numerical model of the structure. Furthermore, represents the DoFs and the force vector which are dependent on time . This dependency is omitted in the following equations in order to improve readability.

Coupling in the physical domain

Coupling of substructures in the physical domain first requires writing the uncoupled equations of motion of the substructures in block diagonal form:

Next, two assembly approaches can be distinguished: primal and dual assembly.

Primal assembly

For primal assembly, a unique set of degrees of freedom is defined in order to satisfy compatibility, . Furthermore, a second equation is added to enforce interface force equilibrium. This results in the following coupled dynamic equilibrium equations:

Pre-multiplying the first equation by and noting that , the primal assembly reduces to:

The primally assembled system matrices can be used for a transient simulation by any standard time stepping algorithm. Note that the primal assembly technique is analogue to assembly of super-elements in finite element methods.

Dual assembly

In the dual assembly formulation the global set of DoFs is retained and an assembly is made by a priori satisfying the equilibrium condition . Again, the Lagrange multipliers represent the interface forces connecting the DoFs at the interface. As these are unknowns, they are moved to the left-hand side of the equation. In order to satisfy compatibility, a second equation is added to the system, now operating on the displacements:

The dually assembled system can be written in matrix form as:

This dually assembled system can also be used in a transient simulation by means of a standard time stepping algorithm. [9]

Substructuring in the frequency domain

In order to write out the equations for frequency based substructuring (FBS), the dynamic equilibrium first has to be put in the frequency domain. Starting with the dynamic equilibrium in the physical domain:

Taking the Fourier transform of this equation gives the dynamic equilibrium in the frequency domain:

Matrix is referred to as the dynamic stiffness matrix. This matrix consists of the complex-valued frequency-dependent functions that describe the force required to generate a unit harmonic displacement at a certain DoF. The inverse of the matrix is defined as and yields the more intuitive admittance notation:

The receptance matrix contains the frequency response functions (FRFs) of the structure which describe the displacement response to a unit input force. Other variants of the receptance matrix are the mobility and accelerance matrix, which respectively describe the velocity and acceleration response. The elements of the dynamic stiffness (or impedance in general) and receptance (or admittance in general) matrix are defined as follows:

Coupling in the frequency domain

In order to couple two substructures in the frequency domain, use is made of the admittance and impedance matrices of both substructures. Using the definition of substructures A and B as introduced previously, the following impedance and admittance matrices are defined (note that the frequency dependency is omitted from the terms to improve readability):

The two admittance and impedance matrices can be put in block diagonal form in order to align with the global set of DoFs :

The off-diagonal zero terms show that at this point no coupling is present between the two substructures. In order to create this coupling, use can be made of the primal or dual assembly method. Both assembly methods make use of the dynamic equations as was defined before:

In these equations is again used to define the set of interface forces, which are yet unknown.

Primal assembly

In order to obtain the primal system of equations, a unique set of coordinates is defined: . By definition of an appropriate Boolean localisation matrix , a unique set of DoFs remains for which the compatibility condition is satisfied a priori (compatibility condition). In order to satisfy the equilibrium condition a second equation is added to the equations of motion:

Pre-multiplying the first equation with yields the notation of the assembled equations of motion for the generalised coordinates :

This result can be rewritten in admittance form as:

This last result gives access to the generalised responses as a result of the generalised applied forces , namely by inverting the primally assembled impedance matrix.

The primal assembly procedure is mainly of interest when one has access to the dynamics in impedance form, e.g. from finite element modelling. When one only has access to the dynamics in admittance notation, [14] the dual formulation is a more suitable approach.

Dual assembly

A dually assembled system starts with the system written in the admittance notation. For a dually assembled system the force equilibrium condition is satisfied a priori by substituting Lagrange multipliers for the interface forces: . The compatibility condition is enforced by adding an additional equation:

Substituting the first line in the second and solving for gives:

The term represents the incompatibility caused by the uncoupled responses of the substructures to the applied forces . By multiplying the incompatibility with the combined interface stiffness, i.e. , the forces that keep the substructures together are determined. The coupled response is obtained by substituting the calculated back into the original equation:

This coupling method is referred to as the Lagrange-multiplier frequency-based substructuring (LM-FBS) method. [6] The LM-FBS method allows for quick and easy assembling of an arbitrary number of substructures in a systematic fashion. Note that the result is theoretically the same as was obtained above by application of primal assembly.

Decoupling in the frequency domain

Decoupling of substructure B from assembly AB Decoupling of substruces.png
Decoupling of substructure B from assembly AB

In addition to coupling of substructures, one is also able to decouple substructures from assemblies. [15] [16] [17] [18] Using the plus sign as a substructure coupling operator, the coupling procedure could simply be described as AB = A + B. Using a similar notation, decoupling could be formulated as AB - B = A. Decoupling procedures are often required to remove substructures that were added for measurement purposes, e.g. to fix the structure. Similar to coupling, a primal and dual formulation exists for decoupling procedures.

Primal disassembly

As a result of the primal coupling, the impedance matrix of the assembled system can be written as follows:

Using this relation, the following trivial subtraction operation would suffice for the decoupling of the substructure B from assembly AB:

By placing the impedance of AB and B in block-diagonal form, with a minus sign for the impedance of B to account for the subtraction operation, the same equation that was used for primal coupling can now be used to perform the primal decoupling procedures.

with:

The primal disassembly can thus be understood as the assembly of structure AB with the negative impedance of substructure B. A limitation of the primal disassembly is that all DoF of the substructure that is to be decoupled have to be exactly represented in the assembled situation. For numerical decoupling situations this should not pose any problems, however for experimental cases this can be troublesome. A solution to this problem can be found in the dual disassembly.

Dual disassembly

Similar to the dual assembly, the dual disassembly approaches the decoupling problem using the admittance matrices. Decoupling in the dual domain means finding a force that ensures compatibility, yet acts in the opposite direction. This newly found force would then counteract the force that is applied to the assembly due to the dynamics of substructure B. Writing this out in equations of motion:

In order to write the dynamics of both systems in one equation, using the LM-FBS assembly notation, the following matrices are defined:

In order to enforce compatibility, a similar approach is used as for the assembly task. Defining a -matrix to enforce compatibility:

Using this notation, the disassembly procedure can be performed using exactly the same equation as was used for the dual assembly:

This means that coupling and decoupling procedures using LM-FBS require identical steps, the only difference being the manner in which the global admittance matrix is defined. Indeed, the substructures to couple appear with a plus sign, whereas decoupled structures carry a minus sign:

More advanced decoupling techniques use the fact that internal points of substructure B appear in both the admittances of AB and B, hence can be used to enhance the decoupling process. Such techniques are described in. [17] [18]

See also

Related Research Articles

<span class="mw-page-title-main">Lorentz transformation</span> Family of linear transformations

In physics, the Lorentz transformations are a six-parameter family of linear transformations from a coordinate frame in spacetime to another frame that moves at a constant velocity relative to the former. The respective inverse transformation is then parameterized by the negative of this velocity. The transformations are named after the Dutch physicist Hendrik Lorentz.

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances. They were named after French engineer and physicist Claude-Louis Navier and the Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

<span class="mw-page-title-main">Singular value decomposition</span> Matrix decomposition

In linear algebra, the singular value decomposition (SVD) is a factorization of a real or complex matrix into a rotation, followed by a rescaling followed by another rotation. It generalizes the eigendecomposition of a square normal matrix with an orthonormal eigenbasis to any matrix. It is related to the polar decomposition.

<span class="mw-page-title-main">Ellipsoid</span> Quadric surface that looks like a deformed sphere

An ellipsoid is a surface that can be obtained from a sphere by deforming it by means of directional scalings, or more generally, of an affine transformation.

<span class="mw-page-title-main">Moment of inertia</span> Scalar measure of the rotational inertia with respect to a fixed axis of rotation

The moment of inertia, otherwise known as the mass moment of inertia, angular mass, second moment of mass, or most accurately, rotational inertia, of a rigid body is a quantity that determines the torque needed for a desired angular acceleration about a rotational axis, akin to how mass determines the force needed for a desired acceleration. It depends on the body's mass distribution and the axis chosen, with larger moments requiring more torque to change the body's rate of rotation by a given amount.

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

In continuum mechanics, the infinitesimal strain theory is a mathematical approach to the description of the deformation of a solid body in which the displacements of the material particles are assumed to be much smaller than any relevant dimension of the body; so that its geometry and the constitutive properties of the material at each point of space can be assumed to be unchanged by the deformation.

<span class="mw-page-title-main">Hooke's law</span> Physical law: force needed to deform a spring scales linearly with distance

In physics, Hooke's law is an empirical law which states that the force needed to extend or compress a spring by some distance scales linearly with respect to that distance—that is, Fs = kx, where k is a constant factor characteristic of the spring, and x is small compared to the total possible deformation of the spring. The law is named after 17th-century British physicist Robert Hooke. He first stated the law in 1676 as a Latin anagram. He published the solution of his anagram in 1678 as: ut tensio, sic vis. Hooke states in the 1678 work that he was aware of the law since 1660.

Linear elasticity is a mathematical model of how solid objects deform and become internally stressed due to prescribed loading conditions. It is a simplification of the more general nonlinear theory of elasticity and a branch of continuum mechanics.

In mathematics, the matrix exponential is a matrix function on square matrices analogous to the ordinary exponential function. It is used to solve systems of linear differential equations. In the theory of Lie groups, the matrix exponential gives the exponential map between a matrix Lie algebra and the corresponding Lie group.

In differential geometry, the four-gradient is the four-vector analogue of the gradient from vector calculus.

In geometry, various formalisms exist to express a rotation in three dimensions as a mathematical transformation. In physics, this concept is applied to classical mechanics where rotational kinematics is the science of quantitative description of a purely rotational motion. The orientation of an object at a given instant is described with the same tools, as it is defined as an imaginary rotation from a reference placement in space, rather than an actually observed rotation from a previous placement in space.

The derivatives of scalars, vectors, and second-order tensors with respect to second-order tensors are of considerable use in continuum mechanics. These derivatives are used in the theories of nonlinear elasticity and plasticity, particularly in the design of algorithms for numerical simulations.

In fluid dynamics, Airy wave theory gives a linearised description of the propagation of gravity waves on the surface of a homogeneous fluid layer. The theory assumes that the fluid layer has a uniform mean depth, and that the fluid flow is inviscid, incompressible and irrotational. This theory was first published, in correct form, by George Biddell Airy in the 19th century.

<span class="mw-page-title-main">Stokes' theorem</span> Theorem in vector calculus

Stokes' theorem, also known as the Kelvin–Stokes theorem after Lord Kelvin and George Stokes, the fundamental theorem for curls or simply the curl theorem, is a theorem in vector calculus on . Given a vector field, the theorem relates the integral of the curl of the vector field over some surface, to the line integral of the vector field around the boundary of the surface. The classical theorem of Stokes can be stated in one sentence: The line integral of a vector field over a loop is equal to the surface integral of its curl over the enclosed surface. It is illustrated in the figure, where the direction of positive circulation of the bounding contour ∂Σ, and the direction n of positive flux through the surface Σ, are related by a right-hand-rule. For the right hand the fingers circulate along ∂Σ and the thumb is directed along n.

<span class="mw-page-title-main">Objective stress rate</span>

In continuum mechanics, objective stress rates are time derivatives of stress that do not depend on the frame of reference. Many constitutive equations are designed in the form of a relation between a stress-rate and a strain-rate. The mechanical response of a material should not depend on the frame of reference. In other words, material constitutive equations should be frame-indifferent (objective). If the stress and strain measures are material quantities then objectivity is automatically satisfied. However, if the quantities are spatial, then the objectivity of the stress-rate is not guaranteed even if the strain-rate is objective.

In fluid dynamics, Beltrami flows are flows in which the vorticity vector and the velocity vector are parallel to each other. In other words, Beltrami flow is a flow where Lamb vector is zero. It is named after the Italian mathematician Eugenio Beltrami due to his derivation of the Beltrami vector field, while initial developments in fluid dynamics were done by the Russian scientist Ippolit S. Gromeka in 1881.

The variational multiscale method (VMS) is a technique used for deriving models and numerical methods for multiscale phenomena. The VMS framework has been mainly applied to design stabilized finite element methods in which stability of the standard Galerkin method is not ensured both in terms of singular perturbation and of compatibility conditions with the finite element spaces.

The streamline upwind Petrov–Galerkin pressure-stabilizing Petrov–Galerkin formulation for incompressible Navier–Stokes equations can be used for finite element computations of high Reynolds number incompressible flow using equal order of finite element space by introducing additional stabilization terms in the Navier–Stokes Galerkin formulation.

<span class="mw-page-title-main">Forward problem of electrocardiology</span>

The forward problem of electrocardiology is a computational and mathematical approach to study the electrical activity of the heart through the body surface. The principal aim of this study is to computationally reproduce an electrocardiogram (ECG), which has important clinical relevance to define cardiac pathologies such as ischemia and infarction, or to test pharmaceutical intervention. Given their important functionalities and the relative small invasiveness, the electrocardiography techniques are used quite often as clinical diagnostic tests. Thus, it is natural to proceed to computationally reproduce an ECG, which means to mathematically model the cardiac behaviour inside the body.

References

  1. Ritz, W. (1909). "Über eine neue Methode zur Lösung gewisser Variations Probleme der Mathematishen Physik". Journal für die Reine und Angewandte Mathematik. 1909 (135): 1–61. doi:10.1515/crll.1909.135.1. S2CID   116143760.
  2. Huebner, Dewhirst; Smith, Byrom (2001). The Finite Element Method for Engineers. New York: Wiley. ISBN   978-0471370789.
  3. Hestnes, Stiefel (1952). "Method of Conjugate Gradients for Solving Linear Systems" (PDF). Journal of Engineering Mechanics. 86 (4): 51–69.
  4. Craig, Bampton (1968). "Coupling of Substructures for Dynamic Analysis" (PDF). AIAA Journal. 6 (7): 1313–1319. Bibcode:1968AIAAJ...6.1313B. doi:10.2514/3.4741.
  5. 1 2 Jetmundsen, Bjorn; Bielawa, Richard L.; Flannelly, William G. (1988-01-01). "Generalized Frequency Domain Substructure Synthesis". Journal of the American Helicopter Society. 33 (1): 55–64. doi:10.4050/JAHS.33.55.
  6. 1 2 3 D. de Klerk; D. Rixen; J. de Jong (2006). "The Frequency Based Substructuring method reformulated according to the dual domain decomposition method". Proceedings of the XXIV International Modal Analysis Conference (IMAC), St. Louis. Archived from the original on 2016-07-01.
  7. 1 2 Klerk, D. De; Rixen, D. J.; Voormeeren, S. N. (2008-01-01). "General Framework for Dynamic Substructuring: History, Review and Classification of Techniques". AIAA Journal. 46 (5): 1169–1181. Bibcode:2008AIAAJ..46.1169D. doi:10.2514/1.33274. ISSN   0001-1452.
  8. Newmark, N.M. (1959). "A Method of Computation for Structural Dynamics". Journal of the Engineering Mechanics Division. 85 (3): 67–94. doi:10.1061/JMCEA3.0000098.
  9. 1 2 3 Geradin, Michel; Rixen, Daniel J. (2014). Mechanical Vibrations: Theory and Application to Structural Dynamics, 3rd Edition. John Wiley & Sons. ISBN   978-1-118-90020-8.
  10. Rixen, Daniel J.; van der Valk, Paul L. C. (2013-12-23). "An Impulse Based Substructuring approach for impact analysis and load case simulations". Journal of Sound and Vibration. 332 (26): 7174–7190. Bibcode:2013JSV...332.7174R. doi:10.1016/j.jsv.2013.08.004.
  11. Sjövall, Per; Abrahamsson, Thomas (2007-10-01). "Component system identification and state-space model synthesis". Mechanical Systems and Signal Processing. 21 (7): 2697–2714. Bibcode:2007MSSP...21.2697S. doi:10.1016/j.ymssp.2007.03.002.
  12. Bernardi, C.; Maday, Y.; Patera, A. T. (1994). "New Nonconforming Approach to Domain Decomposition: The Mortar Element Method". Nonlinear Partial Differential Equations and Their Applications.
  13. Voormeeren, S.N. (7 November 2012). Dynamic Substructuring Methodologies for Integrated Dynamic Analysis of Wind Turbines (PhD). Delft University of Technology. doi:10.4233/uuid:f45f0548-d5ec-46aa-be7e-7f1c2b57590d.
  14. Allen, M.; Mayes, R (2007). "Comparison of FRF and Modal Methods for Combining Experimental and Analytical Substructures". Proceedings of the Twenty Fifth International Modal Analysis Conference.
  15. D'Ambrogio, Walter; Fregolent, Annalisa (2010-05-19). "The role of interface DoFs in decoupling of substructures based on the dual domain decomposition". Mechanical Systems and Signal Processing. 24 (7): 2035–2048 via Elsevier Science Direct.
  16. D’Ambrogio, Walter; Fregolent, Annalisa (2011-01-01). Proulx, Tom (ed.). Direct decoupling of substructures using primal and dual formulation. Conference Proceedings of the Society for Experimental Mechanics Series. Springer New York. pp. 47–76. doi:10.1007/978-1-4419-9305-2_5. ISBN   9781441993045.
  17. 1 2 Voormeeren, S. N.; Rixen, D. J. (2012-02-01). "A family of substructure decoupling techniques based on a dual assembly approach". Mechanical Systems and Signal Processing. 27: 379–396. Bibcode:2012MSSP...27..379V. doi:10.1016/j.ymssp.2011.07.028.
  18. 1 2 D'Ambrogio, Walter; Fregolent, Annalisa (2014-04-04). "Inverse dynamic substructuring using the direct hybrid assembly in the frequency domain". Mechanical Systems and Signal Processing. 45 (2): 360–377. Bibcode:2014MSSP...45..360D. doi:10.1016/j.ymssp.2013.11.007.