Abstract

Two closed-form methods to solve the continuous-time algebraic Riccati equation (CARE) for second-order systems in terms of the mass, damping, and stiffness matrices are presented. One method utilizes the modal transformation of mass and stiffness matrices, and the other does not require this transformation. Hundreds of high-dimensional second-order systems are used to show that these methods achieve similar or better accuracy compared to the state-of-the-art, while significantly reducing the computation time. Furthermore, advantages of these methods are illustrated in vibration control problems.

Graphical Abstract Figure
Graphical Abstract Figure
Close modal

1 Introduction

Continuous-time algebraic Riccati equations (CAREs) often appear in control and optimization problems such as linear quadratic regulator (LQR) [1], linear quadratic Gaussian [2], Kalman filter [2], H2 and H methods [1], the gap (δ) and ν-gap (δν) [1] computation, etc. As a result, they are employed in applications such as control design for helicopters [3], unmanned aerial vehicles [4], spacecraft formation [5], space structures [6], induction motors [7], and others.

These CAREs are generally solved using the Schur eigenvector method [8], which is formulated for first-order systems. The first-order state-space form, with the state dimension n¯, is used to create a Hamiltonian matrix (2n¯×2n¯), whose rearranged eigenvectors are used to compute the CARE solution. Since the computational effort is proportional to n¯3 [9], the method in Ref. [8] works well only for moderate size systems (n¯100) [8]. It becomes inefficient and prohibitive for larger systems [911], particularly when the system is poorly conditioned. Thus, it becomes a limiting factor for applications that require solving CARE for large-size systems such as the control of large space structures (n¯=128; [12]), finite element techniques [13], etc. Solving CARE in real-time is challenging even for small-size systems. For example, Zhao et al. [14] mention that solving CARE in real-time for only six states is considerably difficult for a processor onboard aerial robot.

Furthermore, the linear dynamics of many systems can be expressed as second-order linear time-invariant (LTI) ordinary differential equations (ODEs), described by the mass (M), stiffness (K), damping (C), and input B matrices. The standard approach to solve CARE for such a system, of size n×n, is to convert it to an equivalent first-order system, of size n¯×n¯=2n×2n, which doubles the system size. Thus, the problem in solving CARE is exacerbated for second-order systems, such as in Refs. [6,12], reducing the limit of n to n50, beyond which the method in Ref. [8] becomes progressively less reliable and inefficient. Moreover, conversion of second-order systems to their first-order form leads to the loss of desirable properties such as symmetry and sparsity of the second-order matrices [15].

However, second-order systems have a special structure that can be used to solve CAREs in a closed-form manner, i.e., expressing the solution explicitly as an analytical formula. This special structure has been exploited in Refs. [15,16] to find simple, specialized conditions for the controllability and observability of second-order systems. In this article, we use a similar strategy to obtain simple closed-form solutions to CAREs for second-order systems. Prior works [1720] that attempted to find analytical or closed-form solutions to CARE for second-order systems are briefly described next, revealing their limitations.

1.1 Prior Works.

Prior works are evaluated based on whether they actually solve the CARE and if certain requirements are met. Specifically, the solution matrix P and design matrix Q must be positive semidefinite, and another design matrix R must be positive definite. Also, prior works that claim a unique positive semidefinite solution should ensure the satisfaction of the detectability and stabilizability conditions.

In Ref. [17], a method that could simplify solving CARE for systems where M and K are simultaneously congruently diagonalizable is suggested. To that end, the second-order system is transformed into a first-order one, and a modal transformation is applied to diagonalize M and K (see Ref. [21] for an extensive analysis of simultaneous diagonalizability of matrices). The obtained CARE is broken down into four equations involving the transformed matrices. One of these equations is treated as a CARE by restrictively assuming the off-diagonal blocks of P to be symmetric and solved using the eigenvector-based approach [8]. However, this assumption, while mathematically convenient, is not generally true and can lead to an incorrect solution. Also, the approach in Ref. [17] can lead to a non-symmetric, and hence, non-positive semidefinite P. Therefore, the method in Ref. [17] does not always solve the CARE.

In Ref. [18], a proposed closed-form solution for the CARE for undamped fully actuated second-order systems is expressed in terms of scalar weights α and β. The assumption of no damping significantly limits the application of Ref. [18].

Furthermore, Hanks and Skelton [19] propose a method to solve CARE for second-order systems but only if the sensors and actuators are collocated. The method includes the design of the matrix Q (or Q^), which is intrinsic to the CARE formulation, and the CARE solution P in terms of scalar constants a and b, and another matrix θ^. However, there is no discussion about how a and b should be selected to ensure that Q (or Q^) is positive semidefinite, as required by the CARE. Moreover, to prove that P is positive definite, certain constraints are imposed which, if solved, could lead to a and b such that P0. These constraints might be unsolvable, or be very expensive to solve, and levy structural impositions on the system limiting its application to very particular cases.

In Ref. [20], a closed-form solution to the CARE for second-order systems is presented in terms of scalars α and β. This solution satisfies the CARE and the authors discuss how to find α and β. However, the work in Ref. [20] also assumes collocated actuators and sensors. Under this assumption, the authors prove that the obtained P always exponentially stabilizes the system using LQR. By additionally proving the observability condition, they invoke the argument of a unique stabilizing solution of Ref. [8] and conclude that their solution is unique and positive semidefinite. However, the proof is valid only for systems with collocated actuators and sensors. The authors do not address the more realistic and practical non-collocated actuators and sensors systems [22].

Finally, the icare method [23] in Matlab uses a modified version of the method of Ref. [8] for first-order systems, by employing a proprietary scaling routine. This extends the limit of n¯100 (or n50) without much loss in accuracy, albeit with a very significant increase in computation time. In comparison, the unscaled icare of Matlab produces poor results as n¯ increases beyond 100 (or n>50).

Main Contribution: We present two novel closed-form solutions, in Theorems 2 and 3, to CARE for second-order systems. The first applies when the mass and stiffness matrices {M,K} can be simultaneously congruently diagonalized. The second is valid for any positive definite M,C, and K. The solutions are unique, positive semidefinite, and optimal for the cost in (3). They are expressed in terms of the second-order system matrices M,C,K, and offer physical insights. The proposed methods work very well for n>>50 and achieve better accuracy than the Matlab’s unscaled icare [23] and similar accuracy to Matlab’s scaled icare, but are significantly faster than both of them.

Section 2 discusses the problem statement. The novel methods to solve CARE are presented in Secs. 3 and 4, followed by results in Sec. 5. Some applications are presented in Sec. 6 and the conclusions follow in Sec. 7.

2 Problem Statement

Consider a general second-order system:
(1)
where the generalized coordinates qRn, {M,C,K,B}Rn×n,andtheinputuRn. Matrices M,C,K are assumed positive definite, and thus, symmetric and invertible. The equivalent first-order system is
(2)
Matrices {A,B¯} describe a first-order LTI ODE system and can be used to design an optimal stabilizing controller, known as LQR [1]. It minimizes the cost (J) in (3), where Q and R are positive semidefinite and positive definite matrices, respectively:
(3)
The optimal LQR solution is obtained by solving the associated CARE (4) for P, and the optimal LQR control is given by u=R1B¯TPw.
(4)
A unique positive semidefinite solution (P) to (4) that is optimal for the LQR cost in (3) exists under certain conditions. These conditions are provided in Theorem 1 for a positive semidefinite Q and a positive definite R, where C~ is a full-rank factorization (FRF) of Q (Q=C~TC~) such that columnrank(C~)=rank(Q). This P is also stabilizing, i.e., if the open loop system (2) is unstable, then the closed-loop system obtained by LQR control is exponentially stable.

Theorem 1 [8] Stabilizability of{A,B¯}& detectability of{A,C~}are necessary and sufficient for(4)to have a unique non-negative solutionP.

Moreover, if the conditions of stabilizability and detectability are, respectively, strengthened to controllability and observability, then the solution becomes positive definite. Note that the conditions of stabilizability and detectability are critical. Thus, any method that claims to solve the CARE for a unique positive semidefinite, stabilizing solution that is optimal to the cost in (3) should prove these conditions.

Now, corresponding to the second-order structure in (2), P and Q can be partitioned as follows:
(5)
(6)
Using (2) with (5) and (6), (4) transforms to
(7)
Since M,C,K are symmetric, the transpose superscript on them can be dropped. This simplifies (7) to (8), where LRn×n is a symmetric matrix.
(8)
Our objective is to find the solution P in terms of the second-order matrices M,C,K instead of the first-order matrices A,B¯. This amounts to solving (8) by a method unlike [8], such that the Hamiltonian eigenvector calculation is avoided. To this end, we exploit the structure of (8) such that a closed-form expression of P is obtained. We break (8), which is 2n×2n, into four n×n Eqs. (9)(12), out of which three are independent (9), (11), (12), with two nonlinear (9), (12), and one linear (11):
(9)
(10)
(11)
(12)
The method in Ref. [17] converts the original-sized CARE problem into four equations similar to (9)(12), but with the assumption that {M,K} are congruently simultaneously diagonalizable. The system is then converted to its normalized modal form, with M transformed to an identity matrix and K to another diagonal matrix (Ω2). Matrices C and B do not necessarily transform to diagonal forms, i.e., they become C^ and B^, respectively, and are not necessarily diagonal. The transformed M, which is now an identity matrix, and the transformed B (i.e., B^) lead to transformed L, represented by L^. This transforms (9)(12) to (13)(16):
(13)
(14)
(15)
(16)
Note that (13) is not a CARE and cannot be solved using any solver of CARE. To solve it, Skelton [17] assumes that the non-diagonal block matrix of the solution, P12, is symmetric, thus converting (13) into a CARE of size n×n, which is solved using Ref. [8]. Thus, Skelton [17] claims that a 2n×2n CARE solution, which could be computationally expensive for a large n, can be obtained by solving two n×n nonlinear equations (13) and (16) and one linear equation (15). However, the assumption that P12 is symmetric is not generally true. It would be trivially satisfied when P12 is diagonal. However, because L^ (and B^) are not necessarily diagonal, (13) would not be satisfied for a general Q11. Instead, Q11 should have a special structure that addresses the non-diagonal parts of (13). Hence, the method in Ref. [8] cannot be directly applied to solve (13), as Ref. [17] implies.

Moreover, P11, from (15), depends on terms that may be non-symmetric. For example, C^ is not necessarily diagonal, the term C^P12T might be non-symmetric because even if we assume P12 to be symmetric, the product of symmetric matrices is not generally symmetric. Similarly, P22L^P12T is not necessarily symmetric. This would result in P not being necessarily symmetric, in contradiction with a key requirement for the correct CARE solution. Furthermore, it should be shown that the solution (P11,P12,P22) and the design matrices (Q11,Q12,Q22) from (13) to (16) would result in positive semidefinite P and Q. The work in Ref. [17] does not address these issues in finding a positive semidefinite solution to CARE. Also, the proposed solution is not closed-form and still depends on the eigenvector method [8].

We now present novel closed-form methods that solve CARE for second-order systems.

3 Method-I

We start with rewriting Eqs. (9)(12) as (17)(19) such that they, respectively, resemble a standard CARE (19) and a standard quadratic matrix equation (17), for which the solutions may already be known. Here, X=M1K, P12=ZT, Y=M1C, and Y~=Y:
(17)
(18)
(19)
Equation (17) is a quadratic matrix equation in only one variable, Z. Once (17) is solved, then (19) can be solved using the solution Z. Then, once we know both P12 and P22, we can solve (18) to obtain P11. Thus, we can find the solution P for a CARE associated with a second-order system vectorized in the first-order form, subject to the constraints that P and Q are positive semidefinite, R is positive definite, and the conditions of stabilizability and detectability are met. However, this is only possible when (17) can be solved for Z. Since Z need not be symmetric, (17) is not a CARE, and cannot be solved using Ref. [8].
The method in Ref. [24] mentions that if L is positive definite and Q11 is symmetric, then a solution exists for (17) if and only if XTL1X+Q11 is positive semidefinite. Since Q11 is a leading principal minor of Q, it must be positive semidefinite. We now consider the sign-definiteness of L.
(20)
From (20), L is a positive definite matrix if and only if F is non-singular. Since the matrices M and R are invertible, F is non-singular whenever matrix B is invertible. Even though F can be non-singular when B has a full-column rank, we will only consider a full-rank B because it simplifies the stabilizability condition of {A,B¯}. This is discussed in more detail toward the end of method-I. Thus, with the assumption that B is full-rank, L is a positive definite matrix. Since matrix B is full-rank, the original second-order system is fully actuated. We remark that there are many systems that can be modeled as fully actuated, e.g., propeller tilting multirotors [25], underwater vehicles [26], etc.

Furthermore, since L and Q11 are positive definite and semidefinite matrices, respectively, and X is full-rank, the matrix J=XTL1X+Q11 is positive definite. Thus, all of the conditions for the existence of a solution to (17), as per Ref. [24], are satisfied.

Given that the solution exists, Ref. [24] mentions that Z satisfies (17) if and only if it is of the form:
(21)
where U is a column unitary matrix (UUT=I) whose columns span the column space of J, and V is a column unitary matrix. Based on Ref. [24], we now write a corollary for when X is a square matrix.
Corollary 1
Given a positive definite matrixLand a positive semidefinite matrixQ11, the solution to(17)is provided by(22)if the matrixXis square.
(22)
Proof

The matrices U and V (in (21)) become identity matrices when X is square. The rest follows the proof in Ref. [24] and can be simply verified by substituting (22) in (17).

Thus, (17) can now be solved for any positive semidefinite Q11, for example, Q11=XTL1X. Now, we design the matrix R as follows, with β>0:
(23)
This implies that
(24)
Substituting (23) and (24) in (22), we obtain
(25)
Here, we assume Z to be invertible, but this assumption is dropped toward the end of method-I.
We now focus on (19) rewritten below:
(26)
where Q~22=Q22+Z+ZT. Here, Q22 should be chosen such that Q~22 is also positive semidefinite (or definite) for any Z. One such choice of Q22 is
(27)
where SQ22 is any positive definite matrix. Thus, with Q~22 positive definite, (26) now represents a CARE, with size n×n. Since B is full-rank, the controllability condition is trivially satisfied. Furthermore, since Q22 is positive definite, there exists some FRF C~2 of Q~22 (Q~22=C~2TC~2) such that {Y~,C~2} is observable (hence detectable). Then, there always exists a unique positive definite solution (P22) to (26).
Now, based on Ref. [20], the following closed-form solution to (26) can be used:
(28)
where a>0 must be selected such that Q220. The selection of such an a is addressed later.
Finally, we focus on (18) rewritten as follows:
(29)
Given Z and P22 obtained, respectively, from (22) and (28), we can find P11, as shown in (29), for some choice of Q12. However, it should be noted from (29) that the resulting P11 is not inherently symmetric unless all the matrix products in it commute. Since P11 is the principal minor of the CARE solution P, it must be positive semidefinite and, hence, symmetric. The matrix Q12 needs to be specifically designed such that the resulting P11 is symmetric. We provide a structure of Q12 that ensures the positive semidefiniteness and symmetry of P11:
(30)
where SQ12 is a positive definite matrix. This leads to a positive definite P11 as shown in (31).
(31)
Thus, we have provided a matrix P that could be a solution for the CARE associated with a vectorized second-order system. If the pair {A,B¯} is stabilizable and the pair {A,C~} is detectable, then (8) will have only one non-negative solution [8].

Thus, if we can show that the proposed P and Q are positive semidefinite along with the stabilizability and detectability conditions, then this P would be the same as that obtained from Ref. [8] by virtue of uniqueness. Therefore, the proposed P would be the unique positive semidefinite and stabilizing solution to (4) that is optimal for the LQR cost in (3) if the choice of {Q11,Q12,Q22} ensures that matrices Q and P are positive semidefinite, and the conditions of stabilizability and detectability are met.

Now, we examine how the choice of {Q11,Q22,Q12} influences the positive semidefinite constraint on matrices Q and P. By using Schur’s complement of Q, we have
(32)
Any choice of Q that satisfies these constraints is a valid candidate if its FRF C~ is such that {A,C~} is detectable, which happens trivially for any positive definite Q. Substituting R and Q11 from (23) and (24), respectively, (32) transforms to the following:
(33)
where Q22(β,SQ22) indicates that Q22 is a function of β and SQ22. Thus, Q is positive semidefinite if (33) can be satisfied for a real positive constant β and a positive definite matrix SQ22.
Similarly, we can write the Schur’s complement conditions for P, noting that matrices P11=SQ12 and P22 are positive definite:
(34)
Furthermore, let SQ12=(1/γ)P12P221P12T0, where γ>0. Then, (34) transforms to the following:
(35)
Thus, this selection of Q and R results in a positive definite P for 0<γ<1.
Remark 1

Equation (33) may be difficult to solve analytically for a general second-order system. However, the problem simplifies greatly when the system is transformed to a modal form where matrices {M,K} transform to their diagonal forms. Moreover, a block diagonal solution can be obtained in this case for the CARE in (4).

We now modify the presented method by transforming the system to its modal form. Furthermore, the assumption made on Z being invertible, after (25), is no longer needed in the modified method.

3.1 Modal Form—Block Diagonal Solution.

Since the matrices M and K are positive definite, they can be simultaneously diagonalized using congruence transformation [21]. Then, there exists a transformation matrix U (q=Ur) such that MD=UTMU and KD=UTKU are diagonal. Note that C^=UTCU is not required to be diagonal. Then, (36)(38) show the transformation of the system to the form where MD and KD are diagonal:
(36)
(37)
(38)
Now, we can define the following matrices:
(39)
Similarly, matrices (X,Y,L) are rewritten with the subscript D. Then, (2) can be rewritten to obtain the CARE, similar to (4), in terms of AD and B^. Taking R=B^TKD2B^, Q11=αI with α>0, Z can be obtained using Corollary 1, as shown in (40)(42):
(40)
(41)
(42)
In this case, Z is a positive definite matrix, which implies that Z+ZT is also positive definite. Now, from (28), P22=aMD becomes diagonal, where a>0 is such that Q22, as in (43), is positive semidefinite:
(43)
Finally, P11 also becomes diagonal for Q12=(1+α1)C^KD1, per (29), as shown in (44):
(44)
To show that the matrix Q is positive semidefinite, we use (32) again noting that Q11=αI0. Then, (45)(49) show the conditions under which Q0:
(45)
(46)
An a>0 can be chosen such that (45) is true even if we drop the a2KD2 term. Then, (45) is equivalent to
(47)
(48)
(49)

Since C^ is a real symmetric matrix, C^1/2 can be directly obtained by orthogonal diagonalization. Note that all the matrix products inside the square bracket in (48) are symmetric. Thus, their positively weighted sum is also symmetric with real eigenvalues. Hence, the inequality in (49) is valid. Thus, the matrix Q is positive semidefinite for a given by (49). This choice of a also ensures that Q22, from (43), is positive semidefinite.

Now, for P to be positive semidefinite, the following must hold besides P22=aMD being positive definite:
(50)
(51)
(52)
(53)
Thus, the matrix P is positive semidefinite if a is constrained from (53).
For matrices P and Q to be simultaneously positive semidefinite, the scalar a can be greater than or equal to the maximum of the values obtained from (49) and (53). Moreover, if the inequalities in (49) and (53) are strict, Q and P become positive definite. Since Q becomes full-rank, the detectability, in fact observability, condition is trivially satisfied. For the pair {A,B¯} to be controllable, the controllability matrix OC=[B¯AB¯A2n1B¯] must have full-row rank, i.e., 2n. This is true if the matrix OCr=[B¯AB¯], of size 2n×2n, has a full-rank. Note that
(54)
(55)
Since B is assumed full-rank, the matrix II is invertible. Furthermore, since M1 and C are positive definite, the matrix I is invertible. Therefore, both OCr and OC are full-rank, which implies that both {AD,B^} and {A,B¯} are controllable. Thus, since the controllability and observability conditions are satisfied, there exists only one positive semidefinite solution to (4) [8]. Furthermore, since our closed-form solution is also positive semidefinite, it is the same solution as that from Ref. [8], but obtained in a closed-form expression. The results are summarized in Theorem 2.
Theorem 2
Given a second-order system defined by(38), the followingQ,R,Psolve the corresponding CARE(4)in a closed-form manner such thatPis the unique positive semidefinite solution. ThisPis stabilizing and also optimal for the LQR cost in(3).
(56)
(57)
(58)
Here, α>0andais greater than or equal to the maximum of the values from(49)and(53).

4 Method-II

Now, we present another novel closed-form solution to (4). Here, we do not require to transform the system to its modal form, and only require the matrices M,C,K to be positive definite. The solution is in terms of M,C,K, as shown in Theorem 3.

Theorem 3
Consider a second-order system defined by(1). Letγ, a1, anda2be any scalars such that
Letamax{a1,a2}. Then, the followingP,Q,Rsolve the CARE(4)in a closed-form manner such thatPis the unique positive semidefinite and stabilizing solution that is optimal for the LQR cost in(3).
(59)
(60)
(61)

The formulas for Q,P were designed by choosing candidate elements of Q (Q11,Q12,Q22) and P (P11,P12,P22) such that Eqs. (17)(19) were satisfied. Additionally, these elements should also satisfy P0 and Q0. It can be easily verified that the given choice of P,Q,R exactly satisfy the associated CARE. In the following, we prove that the matrices P and Q are positive semidefinite.

Matrix Q will be positive semidefinite, using (32), if (62) and (63) are satisfied:
(62)
It will be shown later (in (68)) that γ(,2) is not feasible. Thus, γ(0,). Consider (63):
(63)
(64)
An a>0 can be chosen such that (64) is true even without the a2K2 term. Indeed, (64) is equivalent to
(65)
(66)
(67)

Note that the terms I and II are both symmetric and so is their positively weighted sum. Hence, the combined matrix on the right-hand side of (67) has only real eigenvalues. Thus, we have shown that Q0 for any γ>0 and for a as per (67).

Matrix P will be positive definite if (68) and (69) are satisfied:
(68)
(69)
(70)
(71)
Since M and K3 are positive definite, their product has real positive eigenvalues. Thus, P is positive definite for γ>1, and for a as given by (71).

For P and Q to be simultaneously positive semidefinite, the scalar a can be chosen as greater than or equal to the maximum of the values given by (67) and (71). If the inequality in (67) is considered strict, then Q0. Then, based on the arguments given toward the end of method-I (after (54)), P is the unique positive definite and stabilizing solution to (4) that is optimal for the LQR cost in (3).

5 Results and Discussion

Now, we present simulation results that numerically verify the proposed method. Since method-II works for any positive definite M,C,K, it is more general than method-I, which requires modal decomposition. Thus, we only present results from method-II. We compare method-II’s results with the scaled and unscaled versions of matlab’s icare on the basis of the residual and computation time.

The residual is obtained by first finding the matrix residual, given by (72), left by a candidate solution P. Then, its induced-norm (2-norm) is taken to obtain a scalar metric, given by (73)
(72)
(73)
The matrices M,C,K used for simulations are randomized positive definite matrices, whose elements draw values from [15] and then added to nI. This is done to avoid the issue of ill-conditioning or poor-conditioning of the Hamiltonian matrix. Matlab’s icare is susceptible to these issues, which significantly deteriorate its results. This happens when the eigenvalues of the Hamiltonian matrix are either close to or are at the origin. The chosen structure of matrices M,C,K avoided the issue of ill-conditioning and extremely reduced the issue of poor-conditioning of the Hamiltonian matrix. The matrix B is a full-rank matrix whose elements draw values from [01] and are then added to an identity matrix to ensure full rank.

Figure 1 shows the 2-norm of the residual left behind, using scattered points, by different methods when solving the CARE. The median of the residuals is also shown for an interval of 50 by solid lines. It is clear that method-II and the scaled and unscaled icare work almost equally well when n50 for second-order systems. However, as n increases, Matlab’s unscaled icare starts giving progressively large residuals. The poor performance of the unscaled icare could be attributed to the errors generated while computing the eigenvectors for large-size systems (n>50). In contrast, method-II avoids eigenvector calculations and its associated errors.

Fig. 1
Residual with respect to the system size. Systems have their dimension equal to the next immediate x-axis tickmark. The scatter points show the actual residual norm. The solid lines show the median for an interval of 50. The size of the corresponding first-order matrix A is 2n×2n.
Fig. 1
Residual with respect to the system size. Systems have their dimension equal to the next immediate x-axis tickmark. The scatter points show the actual residual norm. The solid lines show the median for an interval of 50. The size of the corresponding first-order matrix A is 2n×2n.
Close modal

The scaled icare produces much better results than the unscaled one because of the proprietary scaling routine. However, it does so at the cost of significant time and computation penalty, as shown in Fig. 2. The proposed method-II achieves similar accuracy to the scaled icare, and at a cost even smaller than the unscaled icare, see Fig. 2.

Fig. 2
Computation time with respect to the system size. Figure shows the variation in time taken to compute the CARE solution as a function of the system size. Systems have their size equal to the next immediate tickmark on the x-axis.
Fig. 2
Computation time with respect to the system size. Figure shows the variation in time taken to compute the CARE solution as a function of the system size. Systems have their size equal to the next immediate tickmark on the x-axis.
Close modal

Figure 2 shows the time taken to compute the CARE solution with respect to the size of the second-order system, clearly illustrating the advantage of method-II from the computation cost perspective. Since Matlab’s unscaled icare requires the computation and reordering of the Hamiltonian matrix’s eigenvectors, it becomes computationally expensive for large-size systems. In addition, the scaled icare scales the input matrices according to a proprietary scaling routine, which further increases the computation time and cost. In contrast, method-II is closed-form and avoids eigenvector computation. Consequently, the time taken to compute the solution is significantly less for method-II, as evident from Fig. 2. Thus, method-II would perform much better than other methods in solving CARE for large-size systems in real-time.

Remark 2

The proposed methods enforce a fixed structure for design matrices Q and R. Though the structure is fixed, matrix Q can still be varied by a designer by varying the design variables γ and a. The variable γ can be selected as any number between (0,), and the variable “a” can be selected as any number greater or equal to max(a1,a2), where a1,a2 are as provided in Theorem 3.

The matrix R is fixed. However, this is the case with many control theory applications such as in obtaining the normalized coprime factors [1], ν-gap metric [1], and in other applications such as Youla-Kucera parameterization [27]. These applications fix R as an identity matrix.

The freedom of selecting any structure of Q and R (though Q can still be varied) is traded to obtain a closed-form solution to CARE for second-order systems that can now be directly written given a second-order system.

6 Applications

In this section, we discuss the advantages of our work in the context of some specific engineering problems such as vibration control.

6.1 Vibration Control of Large Space Structures.

Many large space structures require solving CARE for second-order systems to achieve vibration control [6,12]. Consider Ref. [6], where the authors improve the active vibration control of large space structures through structural modifications. To that end, the authors of Ref. [6] first solve a CARE to find the LQR solution. Then, with this solution, the authors obtain the closed-loop system, compute its eigenvalues, and the damping factors corresponding to these eigenvalues. Since these damping factors (and the eigenvalues) are dependent on the system’s matrices M,C,andK, any structural modification will change these matrices as well as the damping factors. Thus, required changes in the damping factors can be obtained through modification of these matrices by changing properties such as the area, length, etc. of the constituent structural elements.

Both methods proposed in this article to solve CARE can be used to improve upon the work in Ref. [6]; however, we shall focus on method-II. Particularly, we replace steps 4 and 5 in the optimum design routine of Ref. [6] which involve solving a CARE, finding the closed-loop matrix and its eigenvalues, and the damping factors. These steps become computationally expensive and inaccurate for large-size systems, and, due to their numerical nature, offer no insight into how matrices M,C,K affect the closed-loop system. In contrast, due to the closed-form nature of our solution, there is no need to use the computationally expensive standard CARE solution. Matrix P is now directly known in terms of scalars γ and a, which can be easily computed as per the steps 25 in Theorem 3. Using matrix P, we can now obtain a closed-form expression of the closed-loop matrix (ACL) for LQR control, as shown through (74) and (75):
(74)
(75)
Thus, the closed-loop matrix can now directly be written in terms of M,C,andK, which was previously done numerically in Ref. [6]. The eigenvalues (λi=σi±jρi) of ACL can then be found using (76), which transforms to an easier to solve (77) after some algebraic manipulations:
(76)
(77)
For the particular case when matrices M,C,K are simultaneously diagonalizable, the system can be converted to its modal form, as in Ref. [6], and the eigenvalues can be directly obtained in a closed-form manner. In that case, M becomes I, C becomes 2[ζ][ω], and K becomes [ω2]. Here, [ζ] is the damping factor matrix consisting of damping factor ζi for i th state, and [ω] is the matrix of natural frequencies of the system consisting of natural frequency ωi for i th state of the second-order system [6]. Then, the 2n eigenvalues of the closed-loop system, two for each state i, are given by (78)
(78)
Thus, due to the closed-form solution P, the closed-loop eigenvalues can now be directly obtained in terms of the modal second-order system, with significant computational savings.
Now, the right (νi of size 2n×1) and the left (μi of size 1×2n) eigenvectors of ACL can be obtained by solving (79) and (80), respectively:
(79)
(80)
Let tk be a structural parameter such as the length or area of the k th structural element of a system like a tetrahedral truss with multiple edges [6]. Any change in tk modifies the matrices M,C,K and, hence, the closed-loop eigenvalues. This variation of the closed-loop eigenvalues with respect to tk can be obtained in terms of the eigenvectors and the sensitivity of ACL with respect to tk. This variation given in Ref. [6] is shown in (81):
(81)
In Ref. [6], δACL/δtk was obtained by differentiating (74), which in turn required δP/δtk, δA/δtk, and δB¯/δtk. The partial derivative δP/δtk was obtained by differentiating (4) with respect to tk, which again required δA/δtk and δB¯/δtk. This process becomes tedious for large-size systems. However, since we have expressed ACL in a closed-form expression (75), δACL/δtk can now simply be obtained directly using (75).
Now, the damping factors, ζi=(σi/(σi2+ρ2)), also change with tk, and the change is given by (82) [6]
(82)
where ρi/tk and σi/tk can be obtained from (81). Thus, the sensitivity of the damping factors to a modification in structural parameter (tk) can now be obtained in a simplified manner due to the closed-form expression of P and ACL. Furthermore, the closed-loop eigenvalues can also be obtained in a closed-form manner if matrices M,C,K are simultaneously diagonalizable, as in Ref. [6]. These improvements offer computational savings and a simplified process.

6.2 Optimal Control of Vibrating Structures.

In this subsection, we show that a closed-form expression for the control gain of LQR can be directly obtained given a second-order system.

Consider a spring-mass-damper system with n units, shown in Fig. 3, fixed from one side. The dynamics of such a system can be described by (1). This system can be used to describe simplified models of structures such as buildings [28], offshore wind turbines [29], etc. Each i th unit has a lumped mass mi, with the spring and damping constants of ki and ci, respectively. An actuator force, Fi, acts on each lumped mass, mi, to counter oscillations. These forces are modeled in terms of the control input ui, Fi=biui, with bi constant. The displacement of mi relative to its rest position is given by qi along the direction shown.

Fig. 3
Spring-mass-damper model for a vibrating structure
Fig. 3
Spring-mass-damper model for a vibrating structure
Close modal
Consider that the structure in Fig. 3 is under vibration, with its dynamics described by (1) in second-order form and by (2) in first-order form. Then, forces Fi can be designed such that the displacements qi are driven to zero in an optimal way. This can be done by designing an LQR, which in turn requires solving the CARE (4). Now, the LQR solution that optimally drives qi to zero can be directly obtained using the closed-form solution of the associated CARE. The standard LQR gain (K) is given by (83), which converts to (85) after using method-II, where γ and a are as given by steps 3 and 4 of Theorem 3.
(83)
(84)
(85)
Thus, given a second-order system described by (1), with B invertible, the LQR control gain can be directly expressed in a closed-form manner in terms of the second-order system’s matrices.

7 Conclusions

This article discussed the pertinent literature that attempts to find an analytical and closed-form solution to CARE for second-order systems. These methods have various shortcomings, for example, they do not solve the CARE accurately, do not always satisfy the positive semi-definiteness property of P and/or Q, or are applicable only to a particular class of systems (collocated actuators and sensors systems). To address these shortcomings, we presented two novel methods that solve the CARE accurately and are applicable for both collocated and non-collocated control. The solution we designed is positive definite, stabilizing, and optimal to the LQR cost. Moreover, this solution always exists because the stabilizability and detectability conditions necessary for the existence and uniqueness of such a solution are satisfied.

Numerical simulations of 500 randomized systems, with sizes ranging from 50×50 to 500×500, were used to illustrate the advantages of the proposed methods. Results showed that method-II achieved a much better accuracy compared to the Hamiltonian eigenvector-based standard method, and similar accuracy with the state-of-the-art (Matlab’s scaled icare). However, it did so with a huge reduction in computation time compared to Matlab’s scaled and unscaled icare routine. Hence, method-II can significantly extend the limit on the system size, n50, to compute the CARE solution for large-size second-order systems.

Furthermore, the solution is a simple and direct formulation in terms of physically meaningful second-order matrices, M,C,K, which have the property of being symmetric positive definite and can be sparse too. Thus, these closed-form expressions offer physical insight into the structure of the solution, which is lost otherwise. We showed that engineering problems such as optimal control of vibrating structures and vibration control of large space structures can be addressed in a simpler and computationally inexpensive way with our solution, given the second-order system’s matrices M,C,K.

These methods, however, have restricted structures for Q and R. We traded some freedom in selecting structures for Q and R to obtain explicit closed-form solutions to CARE for second-order systems. Moreover, since method-I requires simultaneous diagonalization of matrices M and K, it could add to the computational cost. However, since method-II does not require simultaneous diagonalization, this additional cost is avoided.

Acknowledgment

The continuous support of the Office of Naval Research (ONR) under the grant number N00014-20-1-2080 is gratefully acknowledged.

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request.

References

1.
Zhou
,
K.
, and
Doyle
,
J. C.
,
1998
,
Essentials of Robust Control
, Vol. 104,
Prentice Hall
,
Upper Saddle River, NJ
.
2.
Skogestad
,
S.
, and
Postlethwaite
,
I.
,
2005
,
Multivariable Feedback Control: Analysis and Design
,
John Wiley & Sons
,
Hoboken, NJ
.
3.
Bogdanov
,
A.
, and
Wan
,
E.
,
2007
, “
State-Dependent Riccati Equation Control for Small Autonomous Helicopters
,”
J. Guidance Control Dyn.
,
30
(
1
), pp.
47
60
.
4.
Rinaldi
,
F.
,
Chiesa
,
S.
, and
Quagliotti
,
F.
,
2013
, “
Linear Quadratic Control for Quadrotors UAVs Dynamics and Formation Flight
,”
J. Intell. Robot. Syst.
,
70
(
1
), pp.
203
220
.
5.
Starin
,
S.
,
Sparks
,
A.
, and
Yedavalli
,
R.
,
2001
, “
Spacecraft Formation Flying Maneuvers Using Linear-Quadratic Regulation With No Radial Axis Inputs
,”
AIAA Guidance, Navigation, and Control Conference and Exhibit
,
Montreal, Canada
,
Aug. 6–9
, p.
4029
.
6.
Eastep
,
F.
,
Khot
,
N.
, and
Grandhi
,
R.
,
1987
, “
Improving the Active Vibrational Control of Large Space Structures Through Structural Modifications
,”
Acta Astronaut.
,
15
(
6–7
), pp.
383
389
.
7.
Ebrahim
,
O.
,
Salem
,
M.
,
Jain
,
P.
, and
Badr
,
M.
,
2010
, “
Application of Linear Quadratic Regulator Theory to the Stator Field-Oriented Control of Induction Motors
,”
IET Electric Power Appl.
,
4
(
8
), pp.
637
646
.
8.
Laub
,
A.
,
1979
, “
A Schur Method for Solving Algebraic Riccati Equations
,”
IEEE Trans. Automat. Contr.
,
24
(
6
), pp.
913
921
.
9.
Morris
,
K.
, and
Navasca
,
C.
,
2004
, “Solution of Algebraic Riccati Equations Arising in Control of Partial Differential Equations,”
Control and Boundary Analysis
,
J. P.
Zolesio
, and
J.
Cagnol
, eds.,
CRC Press
,
Boca Raton, FL
.
10.
Banks
,
H.
, and
Ito
,
K.
,
1991
, “
A Numerical Algorithm for Optimal Feedback Gains in High Dimensional Linear Quadratic Regulator Problems
,”
SIAM J. Control Optim.
,
29
(
3
), pp.
499
515
.
11.
Meirovitch
,
L.
,
Baruh
,
H.
, and
Oz
,
H.
,
1983
, “
A Comparison of Control Techniques for Large Flexible Systems
,”
J. Guidance Control Dyn.
,
6
(
4
), pp.
302
310
.
12.
O’Donoghue
,
P.
, and
Atluri
,
S.
,
1986
, “
Control of Dynamic Response of a Continuum Model of a Large Space Structure
,”
Comput. Struct.
,
23
(
2
), pp.
199
209
.
13.
Chahlaoui
,
Y.
,
Lemonnier
,
D.
,
Vandendorpe
,
A.
, and
Van Dooren
,
P.
,
2006
, “
Second-Order Balanced Truncation
,”
Linear Algebra Appl.
,
415
(
2–3
), pp.
373
384
.
14.
Zhao
,
M.
,
Anzai
,
T.
,
Shi
,
F.
,
Chen
,
X.
,
Okada
,
K.
, and
Inaba
,
M.
,
2018
, “
Design, Modeling, and Control of an Aerial Robot Dragon: A Dual-Rotor-Embedded Multilink Robot with the Ability of Multi-Degree-of-Freedom Aerial Transformation
,”
IEEE Rob. Autom. Lett.
,
3
(
2
), pp.
1176
1183
.
15.
Hamdan
,
A.
, and
Nayfeh
,
A.
,
1989
, “
Measures of Modal Controllability and Observability for First- and Second-Order Linear Systems
,”
J. Guidance Control Dyn.
,
12
(
3
), pp.
421
428
.
16.
Hughes
,
P. C.
, and
Skelton
,
R. E.
,
1980
, “
Controllability and Observability of Linear Matrix-Second-Order Systems
,”
J. Appl. Mech.
,
47
(
2
), pp.
415
420
.
17.
Skelton
,
R. E.
,
1988
,
Dynamic Systems Control: Linear Systems Analysis and Synthesis
,
John Wiley & Sons, Inc.
,
Hoboken, NJ
.
18.
Belvin
,
W. K.
, and
Park
,
K. C.
,
1990
, “
Structural Tailoring and Feedback Control Synthesis—An Interdisciplinary Approach
,”
J. Guidance Control Dyn.
,
13
(
3
), pp.
424
429
.
19.
Hanks
,
B.
, and
Skelton
,
R.
,
1991
, “
Closed-Form Solutions for Linear Regulator-Design of Mechanical Systems Including Optimal Weighting Matrix Selection
,”
Structures, Structural Dynamics, and Materials Conference
,
Baltimore, MD
,
Apr. 8–10
, p.
1117
.
20.
Koujitani
,
K.
,
Ikeda
,
M.
, and
Kida
,
T.
,
1989
, “
Optimal Control of Large Space Structures by Collocated Feedback
,”
Trans. Soc. Instrum. Control Eng.
,
25
(
8
), pp.
882
888
.
21.
Sultan
,
C.
,
2022
, “
Decoupling of Second Order Systems Via Linear Time-Invariant Transformations
,”
Mech. Syst. Signal Process.
,
169
, p.
108295
.
22.
Liu
,
H.
,
Zhang
,
J.
, and
Zhao
,
W.
,
2017
, “
An Intelligent Non-collocated Control Strategy for Ball-Screw Feed Drives With Dynamic Variations
,”
Engineering
,
3
(
5
), pp.
641
647
.
23.
MATLAB
,
R2019
, “Implicit Solver for Continuous-Time Algebraic Riccati Equations (ICARE),” Version. https://www.mathworks.com/help/control/ref/icare.html.
24.
Crone
,
L.
,
1981
, “
Second Order Adjoint Matrix Equations
,”
Linear Algebra Appl.
,
39
, pp.
61
71
.
25.
Invernizzi
,
D.
,
Giurato
,
M.
,
Gattazzo
,
P.
, and
Lovera
,
M.
,
2021
, “
Comparison of Control Methods for Trajectory Tracking in Fully Actuated Unmanned Aerial Vehicles
,”
IEEE Trans. Control Syst. Technol.
,
29
(
3
), pp.
1147
1160
.
26.
Xiang
,
X.
,
Lapierre
,
L.
, and
Jouvencel
,
B.
,
2015
, “
Smooth Transition of AUV Motion Control: From Fully-Actuated to Under-Actuated Configuration
,”
Rob. Auton. Syst.
,
67
, pp.
14
22
.
27.
Mahtout
,
I.
,
Navas
,
F.
,
Milanés
,
V.
, and
Nashashibi
,
F.
,
2020
, “
Advances in Youla-Kucera Parametrization: A Review
,”
Ann. Rev. Control
,
49
, pp.
81
94
.
28.
Suresh
,
L.
, and
Mini
,
K.
,
2019
, “
Effect of Multiple Tuned Mass Dampers for Vibration Control in High-Rise Buildings
,” Practice Periodical on Structural Design and Construction, Vol.
24
, No.
4
.
29.
Zuo
,
H.
,
Bi
,
K.
, and
Hao
,
H.
,
2017
, “
Using Multiple Tuned Mass Dampers to Control Offshore Wind Turbine Vibrations Under Multiple Hazards
,”
Eng. Struct.
,
141
, pp.
303
315
.