Manifold Calculus in System Theory and Control—Second Order Structures and Systems
Abstract
:1. Introduction
- it provides a clear and well-motivated introduction to advanced concepts in manifold calculus, the basis of system and control theory on manifolds, with special emphasis to its computational and application aspects. The present contribution provides practical formulas to cope with those real-valued manifolds which, in the author’s experience, are most accessed in engineering and applied science problems. As a matter of fact, complex-valued manifolds are not treated at all;
- it clearly states and illustrates the idea that, when dealing with a computer-based implementation of dynamical systems, it is necessary to discretize the differential equations that describe the system in a suitable way. In order to perform such discretization, standard discretization methods (such as the ones based on Euler forward-backward discretization) are not advisable since these do not work as they stand on curved manifolds. It is therefore advisable to resort to more sophisticated integration techniques such as the one based on geodesics.
2. Notation and Recapitulation of Fundamentals
- Hypercube: The simplest manifold of interest is the hypercube , which is essentially the set spanned by p real-valued variables (or p-tuples). A generalization of the hypercube is the set of real-valued matrices.
- Hypersphere: A hypersphere is represented as and is the subset of points of the hypercube with unit Euclidean distance from the point 0. This is a smooth manifold of dimension .
- General linear group and special linear group: The general linear group is defined as }. The special linear group is defined as }. It represents the set of invertible linear operators of given dimension.
- Special orthogonal group: The manifold of special orthogonal matrices is defined as . It represents the set of hyper-rotations in a hypercube of given dimension. In the definition, the symbol denotes a identity matrix.
- Stiefel manifold: The (compact) Stiefel manifold is defined as:
- Real symplectic group: The real symplectic group is defined as
- Manifold of symmetric, positive-definite (SPD) matrices: The space of symmetric, positive-definite matrices is defined as
- Grassmann manifold: A Grassmann manifold is a set of subspaces of spanned by p independent vectors, namely
3. Covariant Derivative of a Vector Field
3.1. Brief Review of Directional Derivative and of Its Properties
- Since f is a multivariable function, because its domain is , there are plenty of ways to ‘move away’ from the foot-point p. The rate of change of the function f depends on the direction towards which one moves away from point x, hence a direction needs to be specified. For example, if one takes , the function f changes little along the x axis, changes more along the y axis, increases along the positive direction of the z axis and decreases along the negative direction of the z axis, not to mention the behavior corresponding to every possible combinations of the triple .
- Since f is arbitrary, its rate of change depends also on ‘how far’ one moves away from the foot-point p. Taking a large leap is, however, not very useful, as one may always replace the foot-point p with a new foot-point to know how the function behaves far from p. Therefore, it is understood that the rate of change is evaluated in close proximity of the point p.
- () Assume that the direction v along which the rate of change is sought is given by the superposition of two partial directions, namely , with being two scalar fields, namely and . How the rate of change relates to rates of change and along partial directions is easily understood:
- () Assume the field f be given as the superposition of two concurrent fields, namely . The way the rate of change relates to rates of change and is also easily understood:
- () In addition, assume the field f is given as the dilation of another field, namely , with being a scalar field. How the rate of change relates to the rate of change is readily understood:
3.2. Coordinate-Free Conceptualization of Covariant Derivation along a Curve
- Top-down from parallel transport: Assuming that a notion of parallel transport is available, which enables us to compare the values of a vector field at different points of a manifold, covariant derivation stems quite naturally by adapting the classical definition of directional derivative (14). In this sense, covariant derivation represents a ‘local’ version of parallel transport.
- Direct axiomatization: In this instance, covariant derivative is defined to be an operator that meets certain requirements, which basically retrace properties , , noted in Section 3.1. Such definition is quite general and leads to a family of covariant derivatives. Moreover, such definition is given independently from that of parallel transport, which may, in turn, be defined as the solution of a covariant-derivative-based differential equation on the tangent bundle . In this sense, parallel transport represents a ‘global’ version of covariant derivation.
- Case of normal Christoffel form. Whenever the manifold is endowed with a normal Christoffel form , namely whenever , the general expression of the covariant derivative may be simplified as
- Case of geodesics. When , namely the vector field under derivation coincides with the velocity field of a curve, it holds that . If is a geodesic curve, since in every point of a geodesic, then it turns out that, on every geodesic line,In fact, in a ‘direct axiomatization’ setting, the above relationship is taken as one defining a geodesic!
3.3. Covariant Derivation along a Direction, Connections
3.4. Relationship of Gradient to Hessian by Covariant Derivation
3.5. Axiomatization of Covariant Derivative and Relationship to Parallellism
- Proof of property : The covariant derivative may be written explicitly asSince both the Gateaux derivative and the Christoffel form are linear in their vectorial arguments, the property follows.
- Proof of property : The vector field may be expressed explicitly asBy the linearity of the terms in parentheses with respect to their vectorial arguments, the proof follows.
- Proof of property : The covariant derivative may be re-expressed asNoticing that leads to the sought result.
- Parallel transport preserves the length of transported vectors: In fact, if we take , we get . This means that parallel transport realizes an isometry.
- Parallel transport preserves the angle between transported vectors: In fact, if we define the cosine of the angle between two tangent vectors as their inner product normalized by their length, we get
3.6. Coordinate-Prone Covariant Derivation*
3.7. Commutator of Vector Fields and Torsion Field
3.8. A Remarkable Relationship Linking the Operator to the Christoffel Form
3.9. Lie Derivation*
4. Iterated Derivatives and Riemannian Curvature
4.1. Commutativity of Directional Derivatives in
4.2. Iterated Covariant Derivatives and Second Covariant Derivative
- A special case arises in the presence of a connection whose Christoffel form is normal, namely , in which case we may access the simplified relation (36) to compute covariant derivative, namely . In such a case, the covariant jolt may be computed as
- A even more particular case is given by a geodesic curve, for which . In this case, the acceleration , hence all covariant derivatives (from acceleration to pounce) are identically zero (no matter what is the structure of the Cristoffel form of the second kind).
4.3. Manifestation of Manifold Curvature in Parallel Transport along a Loop
4.4. Riemannian Curvature Endomorphism
- four points that denote the vertexes of the parallelogram in around the point x,
- two tangent vectors that define the ‘sides’ of the parallelogram,
- a smooth curve that joins the point x to the point a, such that and ; the curve is parametrized through the parameter s, namely a point on such curve is denoted by ,
- a smooth curve that joins the point a to the point c, such that and ; the curve is parametrized through the parameter t, namely a point on such curve is denoted by ,
- a smooth curve that joins the point x to the point b, such that and ; the curve is parametrized through the parameter t, namely a point on such curve is denoted by ,
- a smooth curve that joins the point b to the point b, such that and ; the curve is parametrized through the parameter s, namely a point on such curve is denoted by .
4.5. Tidal Effects, Geodesic Deviation, Jacobi Fields, Sectional Curvature
- for a fixed value of the index , the curve represents a geodesic parametrized by the parameter ; the geodesic curve represents the fiducial geodesic; each tangent vector denotes the speed of the geodesic (we shall omit the values of the parameters for the sake of notation conciseness); all geodesics depart from the same point , namely for any s;
- for a fixed value of the parameter , the function represents a smooth curve (but not a geodesic); each tangent vector represents the separation between two nearby geodesics (even in this case, we shall omit the values of the parameters for the sake of notation conciseness); more intuitively, the infinitesimal separation between nearby geodesics whose index s differ by an infinitesimal is . Since all geodesics depart from the same point, it holds that for every value of the index s.
- whenever , the tidal force is directed in the same direction of the current separation, hence nearby geodesics tend to spread even further from the fiducial one;
- whenever , the tidal force is directed in the opposite direction of the current separation vector, hence nearby geodesics tend to get closer to the fiducial one;
- as opposed to the previous cases, whenever , the tide does not produce any effects and the geodesics get neither closer nor farther apart (in fact, this case corresponds to zero curvature).
- in a region of positive sectional curvature, two nearby parallel geodesics tend to converge towards each other,
- in a region of negative curvature, two nearby parallel geodesics tend to deviate from one another,
- in a region of zero curvature, two nearby parallel geodesic curves neither converge nor diverge.
5. Continuous-Time Dynamical Systems
5.1. Calculus of Variations on Manifold
- the index s determines which curve in the family is referred to; keeping the index s fixed and varying the affine parameter t, namely considering , corresponds to tracing the curve in the family;
- the affine parameter t determines which point on a curve is referred to; keeping t fixed and letting s vary, namely considering , corresponds to traversing the family of curves transversally in correspondence of homologous points; notice that still traces out a curve, although not necessarily endowed with the properties of (for example, while curves may be geodesics, transversal curves might not be geodesics).
5.2. Second-Order Dynamical Systems on Manifold
5.3. Coordinate-Prone Lagrangian Formulation of Dynamical Systems*
- the standard notation for covariant and contravariant tensors indexes as well as Einstein’s convention on summation indexes are in force;
- for any pair of tangent vectors , it holds that , where denote the components of the metric tensor associated to the inner product , namely, based on the canonical basis , ; the functions depend on the coordinates and the symmetry property holds;
- in coordinates, the Riemannian gradient of a regular function at a point is given by:
- the Christoffel symbols of the first kind associated to the Levi-Civita connection corresponding to the metric tensor of components are computed as:
- the kinetic energy functional is expressed by the symmetric bilinear form , namely , where the constant plays the role of a mass term; on a Riemannian manifold, the metric tensor is positive-definite, hence on every trajectory ; in the context of abstract manifold calculus, the mass term is quite immaterial since it does not represent any actual physical quantity, nevertheless, we shall leave it in the equation to help readers although, as already mentioned, it might well be incorporated into the metric tensor g;
- the potential energy function depends on the coordinates only;
- the variations may be chosen arbitrarily except at the boundaries of the trajectory where they vanish, namely, .
- the terms equal the Christoffel symbols of the second kind ;
- the terms denote the components of the Riemannian gradient ;
- the terms denote the components of the contravariant (tangent) dissipation force field .
5.4. Cotangent Bundle Derivation of Non-Linear Oscillators on Manifolds*
5.4.1. Classical Oscillators on Euclidean Spaces
- Case of a hard Duffing oscillator: In this case, the potential function is defined as ,
- Case of a double-well Duffing oscillator: In this instance, the potential is defined as ,
- Case of a soft Duffing oscillator: In this case, the potential is defined as ,
5.4.2. Cotangent Bundle Notation
5.4.3. Second-Order Dynamical Systems Formulation by Cotangent Bundles
- Friction-type damping: This kind of damping generalizes the Rayleigh damping and is expressed by the forcing term , with being a damping coefficient and being a viscosity coefficient. By computing the derivatives, it is found that the friction-type damping force equals .
- Non-linear damping: It generalizes the nonlinear damping term that appears in the van der Pol system and assumes the expression , with being a vector field that depends on a parameter, such that, for each , is a linear endomorphism.
- Sinusoidal driving force: It generalizes the notion of external sinusoidal forcing term from mono-dimensional dynamical systems. It takes the expression , with and .
5.4.4. Potential Energy Functions
6. Control Systems on Manifolds and Numerical Implementation
- A left translation on a Lie group is denoted by and is defined as . Since for the quadcopter model we deal with a matrix Lie group, we may take , obtaining .
- The Lie algebra associated to a Lie group is a vector space endowed with Lie brackets and an adjoint endomorphism . The pushforward map will be denoted as for brevity.
- Given a smooth function , for a matrix Lie group we can define the fiber derivative of ℓ, , with , as the unique element on the algebra such that for each , where the symbol denotes an inner product on the vector space and the symbol denotes the Gateaux derivative (or Jacobian matrix) of the function ℓ with respect to the matrix-variable .
- On a matrix Lie algebra, it holds that . In order to ease the notation in the following sections, let us define a matrix anti-commutator, which is a symmetric form, namely , no matter what is the structure of the arguments. Conversely, the matrix commutator in the algebra is an anti-symmetric bilinear form, namely .
6.1. Design and Analysis of a Lie-Group Synchronization Theory
6.1.1. Velocity Synchronization
6.1.2. Attitude Synchronization
6.2. Application to Quadcopters Synchronization
6.2.1. Lie-Group Model Formulation of a Quadrotor Drone
- Left-invariance of the kinetic energy: The kinetic energy associated to a trajectory g is defined as . We shall assume the inner product representing the metric to be left-invariant, which may be expressed analytically by the condition that
- Left-invariance of the potential energy: There is no notion of left-invariance for the potential function, since it does not depend on the velocity . We shall henceforth assume (and we shall see that it is, in fact, the case) that the potential energy function for the system is simply constant with respect to the configuration g.
- Left-invariance of the generalized force field: In general, the product depends explicitly on the quantities g and . Under the assumption that the inner product is left-invariant, we may however write thatTo this, we add the assumption that the force field F be left-invariant, namely that , where denotes a ‘reduced force field’.
6.2.2. Physical Realizability of the Synchronizing Controller
- the values taken by the scalar control component , and do not differ too much from one another and are absolutely bounded,
- the control field components are small compared to the mass-related term .
6.2.3. Numerical Recipes to Implement Synchronization of Quadrotors
Algorithm 1 Pseudo-code to implement the (uncontrolled) mathematical model of a quadrotor on a computing platform |
Initialize the attitude and the rotation velocity |
Set the step-size h and the number of numerical steps K to complete a simulation |
for to do |
Update the attitude |
Update the angular velocity |
end for |
Algorithm 2 Pseudo-code to implement numerically a leader-follower pair and velocity synchronization by a proportional-type control algorithm |
Initialize the attitudes , and the rotation velocities , |
Set the step-size h, the number of numerical steps K and the proportional control gain |
for to do |
Update |
Update |
Evaluate the control field |
Update |
Update |
end for |
Algorithm 3 Pseudo-code to implement numerically a leader-follower pair and an attitude synchronization control by a proportioanal-derivative-type controller |
Initialize the attitudes , and the rotation velocities , |
Set the discrete step-size h, the number of numerical steps K, the proportional control |
gain and the derivative control gain |
for to do |
Update |
Update |
Evaluate the synchronization error |
Evaluate the synchronization error velocity |
Evaluate the control field (from Equation (341)) |
Update |
Update |
end for |
6.3. Synchronization of Second-Order Systems on Manifolds with Vanishing Control Effort
- in undamped systems, the function f depends only on the system’s state, therefore, the Lipschitz continuity condition is automatically verified and the differential inequation (398) becomes a differential equation, hence the control effort magnitude peters out as fast as , as long as .
- in linearly damped systems, the transition function is of the type + a term that depends on the state ‘x’ only. In this case, it is trivial to verify that the function f is globally -Lipschitz with .
- the term clearly equals because of the identity (440);
- the term may be majorized by means of the Cauchy-Schwarz inequality, by the assumption that the function f is continuous in the first argument and Lipschitz-continuous in the second argument, and by the properties (447). Namely, , such that with , it holds that
- the further scalar component may be majorized by invoking the Cauchy-Schwarz inequality and the continuity of the parallel transport and of its tangent maps. Namely, , such that with , it holds that
- to end with, the term may be majorized by the help of the Cauchy-Schwarz inequality, by the assumptions on the continuity of the pushforward maps of the exponential map and by the inequalities (447), namely, , such that for which then
6.4. Feedback Error Velocity, Principal Pushforward Map and Relation to Curvature
6.4.1. Algebraic Properties of the Principal Pushforward Map of Parallel Transport
- Equivalence of matrix representations: Given a linear operator on a vector space , two of its matrix representations and are equivalent on the vector space if for any , although the matrices and may appear different.
- Representation of the adjoint: If is a matrix representation of a linear operator on a vector space with respect to an orthonormal basis of , then the matrix representation of its adjoint, namely coincides with the transpose (this is a reason why the adjoint is sometimes also referred to as transpose operator).
6.4.2. Relationship between the Principal Pushforward Map and the Curvature Endomorphism
7. Discrete-Time Dynamical Systems on Manifolds
7.1. Auto-Regressive Moving-Average (ARMA) Systems on Riemannian Manifolds
- Moving average (MA) subsystem: The moving-average subsystem consists of a linear combination of the current value and of previous samples of the input signal;
- Auto-regressive (AR) subsystem: The auto-regressive subsystem consists of a linear combination of previous values of the internal state of the system.
- At each step k, it is necessary to compute a linear combination of several tangent vectors in via the transition operators for the auto-regressive subsystem and for the moving-average subsystem, where the operators and are linear in the second (namely, the vectorial) argument.
- In order to compute the linear combination of previous and current values of the state variable, at each step k it is necessary to parallel transport each tangent vector , …, , , …, to the tangent space .
- In order to propagate the constant bias, at each step k it is necessary to parallel transport it to the tangent space .
Algorithm 4 Pseudocode to implement the M-ARMA() systems (503) and (505). | |
1: | Choose a metric for the manifold and make sure the exponential map and the parallel transport operator or the vector transport operator , with , are available |
2: | Choose the initial seed and the drift constant |
3: | Choose 2 transition operators and q transition operators |
4: | Set and set up the zero-padding , , |
5: | for to K do |
6: | Get a tangent vector from an input source |
7: | Compute tangent vectors and , or and , for and and the actualized drift term or |
8: | Compute the next state-velocity term |
9: | Compute the next term in the output sequence |
10: | end for |
7.2. Examples of M-ARMA-Type Dynamical Systems
7.2.1. Generation of Pseudo-Random Paths on the Unit Hyper-Sphere
- Generate a random vector by stacking the outputs of a pseudo-random scalar generator;
- Project the vector into the tangent space by .
7.2.2. Generation of Pseudo-Random Paths on the Manifold of Symmetric, Positive-Definite Matrices
- Generate a pseudo-random matrix by means of a pseudo-random scalar generator;
- Project the matrix into the tangent space by the rule .
7.2.3. Generation of Pseudo-Random Paths on the Compact Stiefel Manifold
- Generate a pseudo-random matrix by a pseudo-random scalar generator;
- Project the matrix into the tangent space by the rule .
8. Conclusions
Funding
Institutional Review Board Statement
Informed Consent Statement
Conflicts of Interest
References
- Bloch, A. An introduction to aspects of geometric control theory. In Nonholonomic Mechanics and Control; Krishnaprasad, P., Murray, R., Eds.; Interdisciplinary Applied Mathematics; Springer: New York, NY, USA, 2015; Volume 24. [Google Scholar]
- Bullo, F.; Lewis, A. Geometric control of mechanical systems. In Texts in Applied Mathematics; Springer: New York, NY, USA; Heidelberg/Berlin, Germany, 2004; Volume 49. [Google Scholar]
- Návrat, A.; Vašík, P. On geometric control models of a robotic snake. Note Mat. 2017, 37, 120–129. [Google Scholar]
- Nijmeijer, H.; van der Schaft, A. Nonlinear Dynamical Control Systems; Springer: New York, NY, USA, 1990. [Google Scholar]
- Sreenath, K.; Lee, T.; Kumar, V. Geometric control and differential flatness of a quadrotor UAV with a cable-suspended load. In Proceedings of the 52nd IEEE Conference on Decision and Control, Firenze, Italy, 10–13 December 2013; pp. 2269–2274. [Google Scholar] [CrossRef]
- Agrachev, A.; Sachkov, Y. Control Theory from the Geometric Viewpoint; Springer: Berlin/Heidelberg, Germnay, 2004; Volume 87. [Google Scholar]
- Fiori, S. Extension of PID regulators to dynamical systems on smooth manifolds (M-PID). SIAM J. Control Optim. 2021, 59, 78–102. [Google Scholar] [CrossRef]
- Brockett, R. Lie algebras and lie groups in control theory. In Geometric Methods in System Theory; Mayne, D., Brockett, R., Eds.; Springer Netherlands: Dordrecht, The Netherlands, 1973; pp. 43–82. [Google Scholar]
- Jakubzyk, B. Introduction to geometric nonlinear control; controllability and Lie bracket. In Mathematical Control Theory, Lectures Notes of a Minicourse; Polish Academy of Sciences: Warsaw, Poland, 2002; pp. 107–168. [Google Scholar]
- Jurdjevic, V. Optimal control on Lie groups and integrable Hamiltonian systems. Regul. Chaotic Dyn. 2011, 16, 514–535. [Google Scholar] [CrossRef]
- Sastry, S. Geometric nonlinear control. In Nonlinear Systems—Analysis, Stability, and Control; Sastry, S., Ed.; Springer Science + Business Media: New York, NY, USA, 1999; pp. 510–573. [Google Scholar]
- Fiori, S. Non-delayed synchronization of non-autonomous dynamical systems on Riemannian manifolds and its applications. Nonlinear Dyn. 2018, 94, 3077–3100. [Google Scholar] [CrossRef]
- Fiori, S. Manifold calculus in system theory and control—Fundamentals and first-order systems. Symmetry 2021, 13, 2092. [Google Scholar] [CrossRef]
- Poisson, E.; Pound, A.; Vega, I. The motion of point particles in curved spacetime. Living Rev. Relativ. 2011, 14, 7. [Google Scholar] [CrossRef] [Green Version]
- Ferreira, R.; Xavier, J. Hessian of the Riemannian squared distance function on connected locally symmetric spaces with applications. In Proceedings of the 7th Portuguese conference on automatic control (Controlo 2006), Lisbon, Portugal, 11–13 September 2006. [Google Scholar]
- Pennec, X. Hessian of the Riemannian Squared Distance; Technical report; Université Côte d’Azur and Inria Sophia-Antipolis Méditerranée: Valbonne, France, 2017. [Google Scholar]
- Fiori, S. Gyroscopic signals smoothness assessment by geometric jolt estimation. Math. Methods Appl. Sci. 2017, 40, 5893–5905. [Google Scholar] [CrossRef]
- Fiori, S. Error-based control systems on Riemannian state manifolds: Properties of the principal pushforward map associated to parallel transport. Math. Control Relat. Fields 2021, 11, 143–167. [Google Scholar] [CrossRef]
- Röbenack, K. Computation of multiple Lie derivatives by algorithmic differentiation. J. Comput. Appl. Math. 2008, 213, 454–464. [Google Scholar] [CrossRef] [Green Version]
- Civita, A.; Fiori, S.; Romani, G. A mobile acquisition system and a method for hips sway fluency assessment. Information 2018, 9, 321. [Google Scholar] [CrossRef] [Green Version]
- Fiori, S. A closed-form expression of the instantaneous rotational lurch index to evaluate its numerical approximation. Symmetry 2019, 11, 1208. [Google Scholar] [CrossRef] [Green Version]
- Ferreira, R.; Xavier, J.; Costeira, J.; Barroso, V. Newton method for Riemannian centroid computation in naturally reductive homogeneous spaces. In Proceedings of the 2006 IEEE International Conference on Acoustics Speech and Signal Processing, Toulouse, France, 14–19 May 2006; Volume 3. [Google Scholar] [CrossRef]
- Isvoranu, D.; Udrişte, C. Fluid flow versus geometric dynamics. In Proceedings of the 5th Conference on Differential Geometry, Mangalia, Romania, 29 August–2 September 2005; Geometry Balkan Press: Bucharest, Romania, 2006; pp. 70–82. [Google Scholar]
- van der Pol, B. On “relaxation-oscillations”. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1926, 2, 978–992. [Google Scholar] [CrossRef]
- Tôrres, L.; Aguirre, L. Transmitting information by controlling nonlinear oscillators. Physica D 2004, 196, 387–406. [Google Scholar] [CrossRef]
- Cao, T.; Yi, H. On the complex oscillation of higher order linear differential equations with meromorphic coefficients. J. Syst. Sci. Complex. 2007, 20, 135–148. [Google Scholar] [CrossRef]
- Chen, J.; Lu, J.; Wu, X. Bidirectionally coupled synchronization of the generalized Lorenz systems. J. Syst. Sci. Complex. 2011, 24, 433–448. [Google Scholar] [CrossRef]
- Huang, J.; Zhang, H. Bifurcations of periodic orbits in three-well Duffing system with a phase shift. J. Syst. Sci. Complex. 2011, 24, 519–531. [Google Scholar] [CrossRef]
- Mo, J.; Lin, W. Generalized variation iteration solution of an atmosphere-ocean oscillator model for global climate. J. Syst. Sci. Complex. 2011, 24, 271–276. [Google Scholar] [CrossRef]
- Fiori, S. Nonlinear damped oscillators on Riemannian manifolds: Fundamentals. J. Syst. Sci. Complex. 2016, 29, 22–40. [Google Scholar] [CrossRef]
- FitzHugh, R. Mathematical models of excitation and propagation in nerve. In Biological Engineering; H.P. Schwan, Ed.; McGraw-Hill: New York, NY, USA, 1969. [Google Scholar]
- Barbosa, R.; Tenreiro Machado, J.; Vinagre, B.; Calderón, A. Analysis of the van der Pol oscillator containing derivatives of fractional order. J. Vib. Control 2007, 13, 1291–1301. [Google Scholar] [CrossRef] [Green Version]
- Duffing, G. Erzwungene Schwingungen bei Veränderlicher Eigenfrequenz und Ihre Technische Bedeutung; Braunschweig, F., Ed.; Vieweg & Sohn: Berlin, Germany, 1918. [Google Scholar]
- Trueba, J.; Rams, J.; Sanjuán, M. Analytical estimates of the effect of nonlinear damping in some nonlinear oscillators. Int. J. Bifurc. Chaos 2000, 10, 2257–2267. [Google Scholar] [CrossRef]
- Sanjuán, M. The effect of nonlinear damping on the universal escape oscillator. Int. J. Bifurc. Chaos 1999, 9, 735–744. [Google Scholar] [CrossRef]
- Lorenz, E. Deterministic nonperiodic flow. J. Atmos. Sci. 1963, 20, 130–141. [Google Scholar] [CrossRef] [Green Version]
- Greer, M.; Saha, R.; Gogliettino, A.; Yu, C.; Zollo-Venecek, K. Emergence of oscillations in a simple epidemic model with demographic data. R. Soc. Open Sci. 2020, 7, 191187. [Google Scholar] [CrossRef] [Green Version]
- Georgiou, I.; Corless, M.; Bajaj, A. Dynamics of nonlinear structures with multiple equilibria: A singular perturbation-invariant manifold approach. Z. Angew. Math. Phys. 1999, 50, 892–924. [Google Scholar] [CrossRef]
- Shorek, S. A stationarity principle for non-conservative systems. Adv. Water Resour. 1984, 7, 85–88. [Google Scholar] [CrossRef]
- Molero, F.; Lara, M.; Ferrer, S.; Cèspedes, F. 2-D Duffing oscillator: Elliptic functions from a dynamical systems point of view. Qual. Theory Dyn. Syst. 2013, 12, 115–139. [Google Scholar] [CrossRef]
- Cucker, F.; Smale, S. Emergent behavior in flocks. IEEE Trans. Autom. Control 2007, 52, 852–862. [Google Scholar] [CrossRef] [Green Version]
- Ha, S.Y.; Kim, D.; Schlöder, F.W. Emergent behaviors of Cucker–Smale flocks on Riemannian manifolds. IEEE Trans. Autom. Control 2021, 66, 3020–3035. [Google Scholar] [CrossRef]
- Fiori, S.; Del Rossi, L. Minimal control effort and time Lie-group synchronisation design based on proportional-derivative control. Int. J. Control 2022, 95, 138–150. [Google Scholar] [CrossRef]
- Fiori, S. Synchronization of first-order autonomous oscillators on Riemannian manifolds. Discret. Contin. Dyn. Syst. Ser. B 2019, 24, 1725–1741. [Google Scholar] [CrossRef] [Green Version]
- Fiori, S. Model formulation over Lie groups and numerical methods to simulate the motion of gyrostats and quadrotors. Mathematics 2019, 7, 935. [Google Scholar] [CrossRef] [Green Version]
- Gaponov, I.; Razinkova, A. Quadcopter design and implementation as a multidisciplinary engineering course. In Proceedings of the IEEE International Conference on Teaching, Assessment, and Learning for Engineering (TALE) 2012, Hong Kong, China, 20–23 August 2012; pp. H2B-16–H2B-19. [Google Scholar] [CrossRef]
- Bloch, A.; Krishnaprasad, P.; Marsden, J.; Ratiu, T. The Euler-Poincaré equations and double bracket dissipation. Commun. Math. Phys. 1996, 175, 1–42. [Google Scholar] [CrossRef] [Green Version]
- Gajbhiye, S.; Banavar, R. The Euler-Poincaré equations for a spherical robot actuated by a pendulum. IFAC Proc. Vol. 2012, 45, 72–77. [Google Scholar] [CrossRef] [Green Version]
- Becker, M.; Sampaio, R.; Bouabdallah, S.; de Perrot, V.; Siegwart, R. In-flight collision avoidance controller based only on OS4 embedded sensors. J. Braz. Soc. Mech. Sci. Eng. 2012, 34, 295–297. [Google Scholar] [CrossRef] [Green Version]
- Fiori, S. A control-theoretic approach to the synchronization of second-order continuous-time dynamical systems on real connected Riemannian manifolds. SIAM J. Control Optim. 2020, 58, 787–813. [Google Scholar] [CrossRef]
- Magdy, M.; Ng, T. Regulation and control effort in self-tuning controllers. IEE Proc. D Control Theory Appl. 1986, 133, 289–292. [Google Scholar] [CrossRef]
- Grönwall, T. Note on the derivatives with respect to a parameter of the solutions of a system of differential equations. Ann. Math. 1919, 20, 292–296. [Google Scholar] [CrossRef]
- Henderson, H.; Searle, S. On deriving the inverse of a sum of matrices. SIAM Rev. 1981, 23, 53–60. [Google Scholar] [CrossRef]
- Gavrilov, A. The Taylor series related to the differential of the exponential map. arXiv 2012, arXiv:1205:2868v1. [Google Scholar]
- Bullo, F.; Lewis, A. Geometric Control of Mechanical Systems: Modeling, Analysis, and Design for Mechanical Control Systems; Springer: Berlin/Heidelberg, Germany, 2005. [Google Scholar]
- Osborne, J.; Hicks, G. The geodesic spring on the Euclidean sphere with parallel-transport-based damping. Not. AMS 2013, 60, 544–556. [Google Scholar] [CrossRef]
- Ambrose, W.; Singer, I.M. A theorem on holonomy. Trans. Am. Math. Soc. 1953, 75, 428–443. [Google Scholar] [CrossRef]
- Fiori, S. Auto-regressive moving-average discrete-time dynamical systems and autocorrelation functions on real-valued Riemannian matrix manifolds. Discret. Contin. Dyn. Syst. Ser. B 2014, 19, 2785–2808. [Google Scholar] [CrossRef]
- Escobar, J.; Poznyak, A. Robust parametric identification for ARMAX models with non-Gaussian and coloured noise: A survey. Mathematics 2022, 10, 1291. [Google Scholar] [CrossRef]
Description | Parameter | Value |
---|---|---|
Overall quadrotor mass | ||
Inertia on x axis | ||
Inertia along y axis | ||
Inertia along z axis | ||
Thrust coefficient | b | |
Drag coefficient | ||
Rotational inertia | ||
Arm length | r |
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. |
© 2022 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Fiori, S. Manifold Calculus in System Theory and Control—Second Order Structures and Systems. Symmetry 2022, 14, 1144. https://doi.org/10.3390/sym14061144
Fiori S. Manifold Calculus in System Theory and Control—Second Order Structures and Systems. Symmetry. 2022; 14(6):1144. https://doi.org/10.3390/sym14061144
Chicago/Turabian StyleFiori, Simone. 2022. "Manifold Calculus in System Theory and Control—Second Order Structures and Systems" Symmetry 14, no. 6: 1144. https://doi.org/10.3390/sym14061144
APA StyleFiori, S. (2022). Manifold Calculus in System Theory and Control—Second Order Structures and Systems. Symmetry, 14(6), 1144. https://doi.org/10.3390/sym14061144