Accepted Manuscript SIMPLE-like preconditioners for saddle point problems from the steady Navier-Stokes equations Zhao-Zheng Liang, Guo-Feng Zhang PII: DOI: Reference:
S0377-0427(16)30059-0 http://dx.doi.org/10.1016/j.cam.2016.02.012 CAM 10509
To appear in:
Journal of Computational and Applied Mathematics
Received date: 9 May 2015 Revised date: 10 October 2015 Please cite this article as: Z.-Z. Liang, G.-F. Zhang, SIMPLE-like preconditioners for saddle point problems from the steady Navier-Stokes equations, Journal of Computational and Applied Mathematics (2016), http://dx.doi.org/10.1016/j.cam.2016.02.012 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Manuscript Click here to view linked References
SIMPLE-like preconditioners for saddle point problems from the steady Navier-Stokes equationsI Zhao-Zheng Liang, Guo-Feng Zhang∗ School of Mathematics and Statistics, Lanzhou University, Lanzhou 730000, P.R. China
Abstract The semi implicit method for pressure linked equations (SIMPLE) preconditioner is further generalized to a class of SIMPLE-like (SL) preconditioners for solving saddle point problems from the steady Navier-Stokes equations. The SL preconditioners can also be viewed as a generalization of the relaxed deteriorated PSS (RDPSS) preconditioner proposed by Cao et al. [J. Comput. Appl. Math., 273 (2015) 41-60]. Convergence analysis of the corresponding SL iteration is presented and the optimal iteration parameter is obtained by minimizing the spectral radius of the SL iteration matrix. Moreover, Krylov subspace acceleration of the SL preconditioning is studied. The SL preconditioned saddle point matrix is analyzed. Results about eigenvalue and eigenvector distributions and the minimal polynomial are derived. Numerical experiments from “leaky” two dimensional lid-driven cavity problems are given, which show the advantages of the SL preconditioned GMRES over the DPSS and RDPSS preconditioned ones for solving saddle point problems. Keywords: SIMPLE preconditioner, spectral analysis, saddle point problems, convergence, preconditioning, Navier-Stokes equation 2000 MSC: 65F10, 65F15, 65F50
1. Introduction Consider the incompressible Navier-Stokes equations of the form: −ν∆u + u · ∇u + ∇p = f, in Ω, ∇ · u = 0, in Ω, u = g, on ∂ Ω,
(1.1)
where Ω ⊂ R2 (or R3 ) is an open bounded domain with boundary ∂ Ω. Here, ν > 0 is the viscosity parameter, ∆ is the vector laplacian, ∇ is the gradient, ∇ · u is the divergence of u, f is a given external force field and g is the Dirichlet boundary data. The goal is to find the unknown velocity and pressure fields u and p . The convective term u · ∇u makes this system nonlinear. Linearization of the Navier-Stokes system (1.1) by Picard fixed-point iteration results in a sequence of Oseen problems of the form −ν∆u + w · ∇u + ∇p = f, in Ω, (1.2) ∇ · u = 0, in Ω, u = g, on ∂ Ω, I This work was supported
by the National Natural Science Foundation of China (No. 11271174, 11511130015). author. Email addresses:
[email protected] (Zhao-Zheng Liang),
[email protected] (Guo-Feng Zhang)
∗ Corresponding
Preprint submitted to Elsevier
October 10, 2015
where the divergence free field w is the velocity field obtained from the previous Picard iteration step. Spatial discretizations of the Oseen equations (1.2) using the Ladyzhenskaya-Babuša-Brezzi (LBB) stable finite element methods [23] result in the saddle point problems of the following structure: A BT u f A+ X = = ≡ c+ , (1.3) B 0 p g where now u and p represent discrete velocity and pressure, respectively. A is the discretization of the diffusion and convection terms, B T is the discrete gradient, B is the discrete divergence, and f and g contain forcing and boundary terms. In general, A ∈ Rn×n is a positive real matrix, i.e., its symmetric part is positive definite, B ∈ Rm×n (m < n) is a matrix of full rank, f ∈ Rn and g ∈ Rm . Negativing the second block row of (1.3) equivalently yields the nonsymmetric saddle point problems A BT u f AX = = ≡ c. (1.4) −B 0 p −g The nonsymmetric formulation (1.4) is especially natural when A is nonsymmetric, but positive real. In fact, A is positive stable, i.e., the eigenvalues of A have real and positive parts, see [8]. This can be advantageous when using certain Krylov subspace methods, like GMRES. A variety of scientific computing and engineering applications can derive the augmented linear system (1.3) or (1.4), for example, computational fluid dynamics, mixed finite element approximation of elliptic PDEs, weighted and equality constrained least squares estimation. See [9, 22, 32] and the references therein for a broad overview of applications and numerical solution techniques of saddle point problems. Due to the large and sparse structure of its coefficient matrix, iteration methods are more attractive for solving the saddle point problem (1.3) or (1.4). A large amount of efficient methods have been presented in the literature, such as Uzawa-type methods [6, 7, 13, 14, 20], Hermitian and skew-Hermitian splitting methods [1–4], SIMPLE-type methods [18, 28, 32, 37, 38] and so forth. Meanwhile, Krylov subspace methods [33, 34] are considered more efficient in general. However, they are not competitive without good preconditioners to speed up their convergence. Important and efficient preconditioners include block and approximate Schur complement preconditioners [27, 29, 30, 36], constraint preconditioners [5, 19, 26], augmented Lagrangian preconditioners [10, 12, 16, 24, 25], HSS preconditioners [8, 11, 15, 17, 31, 35] and so on. In [15], based on the factorized form of the deteriorated PSS (DPSS) preconditioner: 1 αI + A 0 αI B T PDPSS = (1.5) 0 αI −B αI 2α for the nonsymmetric saddle point problem (1.4), Cao et al. proposed the RDPSS preconditioner structured as 1 T 1 A 0 αI B T A α AB , (1.6) PRDPSS = = 0 −B 0 α 0 αI −B which is much closer to the saddle point matrix A compared with PDPSS . Here, α > 0 and I denotes identity matrix of proper size. Convergence of the corresponding RDPSS iteration is analyzed and optimal parameter which minimizes the spectral radius of the iteration matrix is derived in [15]. Besides, some spectrum properties −1 A is also described there. of the RDPSS preconditioned matrix PRDPSS In [28], the SIMPLE preconditioner for saddle point problem (1.3) was presented with the following structure A AD −1 B T PSIMPLE = , (1.7) B 0 −1 A+ has with D = diag(A) being the diagonal part of A. Eigenvalue analysis of the preconditioned matrix PSIMPLE been presented in [28]. Some eigenvalue bounds and the estimation for the spectral condition number are also given. We see that the RDPSS preconditioner (1.6) is somewhat similar to the SIMPLE preconditioner (1.7).
2
In this paper, a new generalized variant of the RDPSS preconditioner (1.6), which resembles the SIMPLE preconditioner (1.7), is presented for the nonsymmetric saddle point problem (1.4). We call it SIMPLE-like (SL) preconditioner, which is given by: 1 −1 T A B α AQ PSL = , (1.8) −B 0 with α > 0 and Q being an approximation of A. It is obvious that this preconditioner reduces to the RDPSS preconditioner when Q = I . Convergence properties of the corresponding SL iteration is analyzed and the optimal −1 iteration parameter is obtained. Moreover, spectral properties of the SL preconditioned matrix PSL A is studied. Compared with the RDPSS preconditioner, the SL preconditioner can result in more rapid convergence rate and denser eigenvalue distributions with suitable choices of the preconditioning matrices Q . The remainder of this paper is organized as follows. In Section 2, convergence of the SL iteration for solving saddle point problem (1.4) is analyzed and the optimal iteration parameter is described. In Section 3, spectrum properties of the SL preconditioned matrix are analyzed. In Section 4, the SL preconditioner is considered to accelerate GMRES and its implementation is given. In Section 5, numerical experiments from “leaky” lid-driven cavity problems are presented to show the effectiveness of the SL preconditioned GMRES. Finally in Section 6, we end this paper with some brief conclusions. Throughout this paper, we denote x ∗ as the conjugate transpose of the vector x . I is the identity matrix of proper order. The minimum and maximum eigenvalues of a symmetric matrix H are denoted by λmin (H ) and λmax (H ), respectively. For a matrix A, ρ(A) and σ(A) denote its spectral radius and spectrum, respectively. The real part of a complex number c is denoted as Re(c ). The imaginary unit is denoted as ı . 2. Analysis of the SL iteration In this section, we analyze convergence properties of the SL iteration method. The choice of the optimal iteration parameter is also studied. For the SL preconditioner PSL defined in (1.8), it can also be induced by the following matrix splitting 1 −1 T B 0 ( α1 A − Q )Q −1 B T A α AQ − , A = PSL − RSL = −B 0 0 0 which results in the following SL iteration method. Method 2.1. (The SL iteration method) Given initial vectors x (0) ∈ Rn , y (0) ∈ Rm and positive parameter α. For k = 0, 1, 2, · · · until the iteration sequence {((x (k ) )T , (y (k ) )T )T } converges to the exact solution of the augmented linear system (1.4), compute 1 −1 T f x (k +1) x (k ) 0 ( α1 A − Q )Q −1 B T B A α AQ = + . (2.1) −g y (k +1) y (k ) −B 0 0 0 −1 −1 The SL iteration (2.1) can be rewritten in a fixed-point form: X (k +1) = TSL X (k ) + PSL c , where TSL = PSL RSL is the iteration matrix. Now, we discuss convergence of the SL iteration method.
Lemma 2.2. Let A ∈ Rn×n be a nonsymmetric positive definite matrix, B ∈ Rm×n be a matrix of full rank and Q ∈ Rn×n be a symmetric positive definite matrix. If µ is any eigenvalue of (BQ −1 B T )−1 B A −1 B T , then R e (µ) > 0. Proof. Let µ be an eigenvalue of the matrix (BQ −1 B T )−1 B A −1 B T and u ∈ Cn be the corresponding eigenvector. Then we have B A −1 B T u = µBQ −1 B T u. Since u , 0, taking the scalar product of the above equation by u and dividing by u ∗ u, we obtain α1 + ı β1 = µα2 , with α1 + ı β1 =
u ∗ B A −1 B T u , u∗u
α2 =
u ∗ BQ −1 B T u . u∗u 3
Since B is of full rank and u , 0, we have B T u , 0. So αi > 0 (i = 1, 2), as A and Q are, respectively, nonsymmetric and symmetric positive definite. Thus, it is obvious that R e (µ) > 0. Now, we present convergence results about the SL iteration method. Theorem 2.3. Assume the conditions of Lemma 2.2 hold. Denote the upper and lower bounds of the real parts of the eigenvalues of (BQ −1 B T )−1 B A −1 B T as γmax and γmin , respectively. The upper bound of the absolute values of the imaginary parts of these eigenvalues is denoted as ηmax . Then the SL iteration (2.1) is convergent if α satisfies:
0<α<
2γ min γ2min +η2max ,
2γmax γ2max +η2max ,
if if
ηmax ¾
p
ηmax <
p
γmax γmin , (2.2) γmax γmin .
Proof. From (1.8), by simple calculations, we have A −1 − Q −1 B T (BQ −1 B T )−1 B A −1 −Q −1 B T (BQ −1 B T )−1 −1 PSL = . α(BQ −1 B T )−1 B A −1 α(BQ −1 B T )−1 Then the iteration matrix A −1 − Q −1 B T (BQ −1 B T )−1 B A −1 −Q −1 B T (BQ −1 B T )−1 0 TSL = α(BQ −1 B T )−1 B A −1 α(BQ −1 B T )−1 0 0 −A −1 B T + Q −1 B T (BQ −1 B T )−1 B A −1 B T . = 0 I − α(BQ −1 B T )−1 B A −1 B T
1 −1 T B α AQ
0
− BT
(2.3)
From (2.3), we know that the spectral radius of the SL iteration matrix is ρ(TSL ) = max
1¶i ¶m
q
(1 − αγi )2 + α2 η2i ,
where µ = γi + ı ηi (i = 1, 2, · · · , m ) are eigenvalues of (BQ −1 B T )−1 B A −1 B T . To ensure the convergence of the SL iteration, it must hold that ρ(TSL ) < 1. By using Lemma 2.2, we have γi > 0. Simple calculations give the conditions in (2.2). Based on the convergence condition (2.2), we have the following result about the optimal α that minimizes the spectral radius of the SL iteration matrix (2.3) by similar discussions with Theorem 2.2 in [15]. Theorem 2.4. Assume the conditions of Theorem 2.3 hold. Then the optimal α, which minimizes the spectral radius of the SL iteration matrix (2.3), is γ p min γ2min +η2max , if ηmax ¾ γmax γmin , αopt = p 2 if ηmax < γmax γmin . γmax +γmin , The corresponding optimal spectral radius is
ρ(TSL ) =
ηmax q , γ2min +η2max
p(γmax −γmin )2 +4η2max γmax +γmin
,
if
ηmax ¾
p
γmax γmin ,
if
ηmax <
p
γmax γmin .
Based upon the proof of Theorem 2.3, when A and Q are symmetric, we can derive the following much simpler convergence condition for the SL iteration method, which can be used easily in practical applications. 4
Corollary 2.5. Let A, Q ∈ Rn×n be symmetric positive definite matrices, B ∈ Rm×n be a matrix of full rank. Then we know that (BQ −1 B T )−1 B A −1 B T has real and positive spectrum. Let µmax and µmin be its largest and smallest eigenvalues. Then the SL iteration method is convergent if 0<α<
2 . µmax
The optimal α, which minimizes the spectral radius ρ(TSL ) is αopt =
2 , µmax + µmin
and the corresponding optimal spectral radius is ρ(TSL ) =
µmax − µmin . µmax + µmin
3. Spectral properties of the SL preconditioned matrix −1 A are analyzed. Bounds on its eigenvalIn this section, spectral properties of the preconditioned matrix PSL ues and the corresponding eigenvector distributions are established. Besides, properties of the minimal polyno−1 mial of PSL A is also discussed, which is instructive for the implementation of the Krylov subspace acceleration. From (2.3), we have I A −1 B T − Q −1 B T (BQ −1 B T )−1 B A −1 B T −1 . (3.1) PSL A = I − TSL = −1 T −1 −1 T 0 α(BQ B ) B A B
Then it immediately leads to the following results. Theorem 3.1. Let A be a positive real matrix, B ∈ Rm×n be a matrix of full rank and Q ∈ Rn×n be an approximation −1 of A. Then the eigenvalues of PSL A defined in (3.1) are 1 with multiplicity at least n. The remaining eigenvalues are of the form αµi (i = 1, 2, · · · , m ), where µi are eigenvalues of the matrix (BQ −1 B T )−1 B A −1 B T . Remark 3.2. From Lemma 2.2, we know that the eigenvalues of (BQ −1 B T )−1 B A −1 B T all have positive real parts −1 when Q is symmetric positive definite. Thus the eigenvalues of PSL A lie in a positive box, which may result in faster convergence of Krylov subspace acceleration. Especially, when A and Q are both symmetric positive definite, −1 the eigenvalues of PSL A will locate in a positive real interval. Remark 3.3. Assume that A and Q are both symmetric positive definite. Consider the generalized eigenvalue problem B A −1 B T u = λBQ −1 B T u. Denote u˜ = B T u. Then it holds that λmin ((BQ −1 B T )−1 B A −1 B T ) = min u ,0
u˜ T A −1 u˜ u T B A −1 B T u = λmin (Q A −1 ). = min T −1 T u˜ ,0 u ˜ T Q −1 u˜ u BQ B u
Similarly, we have λmax ((BQ −1 B T )−1 B A −1 B T ) = λmax (Q A −1 ). Thus the eigenvalues of (BQ −1 B T )−1 B A −1 B T are located in the interval [λmin (Q A −1 ), λmax (Q A −1 )]. In the following, the eigenvector distributions and an upper bound of the degree of the minimal polynomial −1 A are presented, which are instructive for the Krylov subspace acceleration. of PSL −1 Theorem 3.4. Assume the conditions of Theorem 3.1 hold. Then the preconditioned matrix PSL A has n + i + j linearly independent eigenvectors. There are
5
(1) n eigenvectors of the form [u T 0T ]T , that correspond to the eigenvalue 1; (2) i (1 ¶ i ¶ m ) eigenvectors of the form [u T v T ]T , with ( α1 A − Q )Q −1 B T v = 0, that correspond to the eigenvalue 1; (3) j (1 ¶ j ¶ m ) eigenvectors of the form [u T v T ]T , that correspond to the nonunit eigenvalues. Proof. Let θ and [u T v T ]T be an eigenpair of PS−1 L A . Then 1 −1 T u A B u B A α AQ . =θ v −B T 0 v −B 0 It is equivalent to θ (1 − θ )Au = ( A − Q )Q −1 B T v α
(3.2)
and (1 − θ )B u = 0, which implies that either θ = 1 or B u = 0. If θ = 1, (3.2) reduces to θ ( A − Q )Q −1 B T v = 0, α
(3.3)
which is satisfied when v = 0. So there are n linearly independent eigenvectors of the form [u T 0T ]T that correspond to the eigenvalue 1. If v , 0, there are i (1 ¶ i ¶ m ) linearly independent eigenvectors of the form [u T v T ]T with v satisfying (3.3). 1 If θ , 1, from (3.2), we have u = 1−θ A −1 ( θα A − Q )Q −1 B T v . Substituting it into (1 − θ )B u = 0, we get αB A −1 B T v = θ BQ −1 B T v. 1 A −1 ( θα A − Q )Q −1 B T v = 0, which contradicts that [u T v T ]T is an It is easy to see that v , 0. Otherwise, u = 1−θ eigenvector. Then there are j (1 ¶ j ¶ m ) linearly independent eigenvectors of the form [u T v T ]T that correspond to nonunit eigenvalues. Thus the eigenvectors are obtained. Next, we will prove they are linearly independent. Let (1)
a (1) = [a 1 , · · · , a n(1) ],
(2)
(2)
a (2) = [a 1 , · · · , a i ] and
(3)
(3)
a (3) = [a 1 , · · · , a j ]
be three vectors and (1) (2) (3) a 1(3) 0 a (1) (2) a1 (3) (2) (1) 1 u1 · · · u j . u · · · ui . u1 · · · un .. . . . + = .. + 1(2) (3) (3) (2) . . . , 0 ··· 0 v · · · v v1 · · · vi 1 j (3) (2) (1) 0 an aj ai
(3.4)
where the first matrix arises from the case θk = 1 (k = 1, · · · , n), the second matrix from the case θk = 1 (k = 1, · · · , i ) and the third matrix from the case θk , 1 (k = 1, · · · , j ). To show that the m + i + j eigenvectors are linearly −1 independent, a (k ) (k = 1, 2, 3) should be all zero vectors. Multiplying both sides of (3.4) from left with PSL A , we have (2) (1) θ1 a 1(3) (3) 0 a1 (1) (2) a1 (3) (2) u1 · · · u j . u1 · · · ui . u 1 · · · u n(1) . . . . (3.5) .. + (2) (2) . + (3) (3) . = .. . 0 ··· 0 v1 · · · vi v1 · · · v j (3) (2) (1) 0 an θj a j ai Subtracting (3.4) from (3.5), gives (3) (θ1 − 1)a 1(3) 0 (3) u1 · · · u j . . . = .. . (3) (3) . v1 · · · v j (3) 0 (θ j − 1)a j
6
(3)
Since θk , 1 and the coefficient matrix are i linearly independent vectors, we have a k = 0 (k = 1, · · · , j ). Then (2) ak
we can easy to obtain = 0 (k = 1, · · · , i ) as (1) 0 a1 (1) u 1 · · · u n(1) . .. .. = . . 0 ··· 0 0 a n(1)
(2) vk
(1)
are linearly independent. Thus, (3.4) can be simplified to
(1)
As u k (k = 1, . . . , n) are linearly independent, we have a k = 0, k = 1, . . ., n. Theorem 3.5. Assume the conditions of Theorem 2.3 hold. Let l (1 ¶ l ¶ m ) be the degree of the minimal polyno−1 A is mial of the matrix α(BQ −1 B T )−1 B A −1 B T − I . Then the degree of the minimal polynomial of the matrix PSL at most l + 1. −1 Proof. We know that PSL A takes the form 0 Θ2 −1 PSL A =I + ≡ I +Υ, 0 Θ1
where Θ1 = α(BQ −1 B T )−1 B A −1 B T − I and Θ2 = A −1 B T + Q −1 B (BQ −1 B T )−1 B A −1 B T . Note that, for any positive integer k , we have k 0 Θ2 0 Θ2 Θ1k −1 Υk = . = k −1 0 Θ1 0 Θ1 It follows that if a polynomial p (t ) is of the form pˆ (t ) = α0 + α1 t + · · · + αk t k , for p (t ) = t pˆ (t ), it holds that 0 Θ2 pˆ (Θ1 ) p (Υ ) = . 0 Θ1 pˆ (Θ1 ) Therefore, if p ∗ (t ) is the minimal polynomial of Θ1 , p (t ) = t p ∗ (t ) is an annihilating polynomial of Υ . Because −1 −1 A is at most l + 1. A = I + Υ , it immediately results in the degree of the minimal polynomial of PSL PSL Remark 3.6. Based on Theorem 3.5 and the results in [33], we know that the dimension of the Krylov subspace −1 A , c ) is at most l + 1. Thus GMRES with SL preconditioner for solving the saddle point problem (1.4) will K (PSL terminate in at most l + 1 steps. 4. Implementation aspects From the discussions in Section 2, we notice that when the condition number of (BQ −1 B T )−1 B A −1 B T is too large, even with optimal choice of α, the convergence of the stationary iteration (3.4) is typically too slow for it to be competitive. In practical implementations, we use the SL splitting as a preconditioner for the Krylov subspace method, e.g., GMRES. The linear system (1.4) is equivalent to the linear system PS−1 A X = PS−1 c. L L It is easy to verify that 1 A αI 0 PS L = 0 α −B BQ −1 B T
(4.1)
Q −1 B T . I
Since the pre-factor α1 has no effect on the preconditioned system (4.1), at each step of the application of the preconditioned GMRES (PGMRES) with SL preconditioner, we need to solve a linear system of the form αI Q −1 B T A 0 z = r, 0 I −B BQ −1 B T for a given r . Then we have the following algorithm for the implementation of the SL iteration. 7
Algorithm 4.1. For a given residual vector r = (r1T , r2T )T , we can compute the update vector z = (z 1T , z 2T )T from the following steps: (1) solve Au 1 = r1 ; (2) solve (BQ −1 B T )z 2 = B u 1 + r2 ; (3) solve Q u 2 = B T z 2 ; (4) z 1 = α1 (u 1 − u 2 ). From the above Algorithm 4.1, we know that three subsystems with coefficient matrices A, BQ −1 B T and Q have to be solved at each iteration steps. If we choose Q to be diagonal or tridiagonal approximations of A, BQ −1 B T will also have sparse structure. Then direct method such as LU factorization or inexact solves such as GMRES can be used to solve the corresponding linear system. Especially, when A and Q are symmetric positive definite, Cholesky factorization or conjugate gradient (CG) method can be used. Now we give analysis about the numerical stability and convergence behavior of GMRES. From the characterization of Krylov subspace method, it follows that the k th residual for GMRES can be expressed in polynomial formed as rk = c − A X k = pk (A )r0 , with pk (0) = 1 and pk ∈ Πk , the set of polynomials of degree not greater than k . GMRES generates implicitly the polynomial pk for which ||rk || is minimal. It is well known that if A is diagonalizable with the eigenvector matrix W , then it holds ||rk || ¶ κ2 (W ) min max |pk (λ)|, pk ∈Πk ,pk (0)=1 λ∈σ(A ) ||r0 || where κ2 (W ) = ||W || · ||W −1 || is the spectral condition number of W . In some cases, the upper bound may be used to predict a true reduction for the GMRES residuals. When the eigenvalues of A are in an ellipse in the right half plane, which are satisfied for results in this paper, then rate of convergence can be bounded. If the eigenvalues of A appear in complex conjugate pairs, the ellipse then has its center d at the real axis, say d > 0. We denoted the focal points by d − c and d + c and the intersection of the ellipse with the real axis by d − c and d + c . For this situation, the asymptotic convergence factor is defined as p a + a2 − c 2 . (4.2) p d + d2 −c2 See [33] for more detailed discussions. When preconditioners are used in GMRES, asymptotic convergence factor (4.2) will depend on the distribution of the eigenvalues of the preconditioned matrix. For the SL preconditioner (1.8), in the case of symmetric positive definite A and Q , the distribution can be measured by the spectral condition number, which is the quotient of the largest and smallest eigenvalues in absolute value. However, in the case that A is nonsymmetric, the condition number appears not to be a good estimate due to the large number of unit eigenvalue. Note that the number of iterations of GMRES is at most equal to the number of different eigenvalues if they are all simple. Next, we will use an effective spectral condition number κ, instead of the real spectral condition number to measure the distribution of the eigenvalues. The number κ is the quotient of the largest and smallest eigenvalue in the absolute value not equal to one. This technique is also used in [18]. Let µi (i = 1, 2, · · · , m ) be the eigenvalues of (BQ −1 B T )−1 B A −1 B T . Then from Theorem 3.1, we have −1 κ(PSL A)=
α max |µi | max |µi | = . α min |µi | min |µi |
−1 This analysis shows that κ(PSL A ) will become much better when max |µi | is close to min|µi |, i.e., Q is a good approximation to A.
8
5. Numerical experiments In this section, we provide numerical experiments on linear systems from Oseen problem (1.2). Our aim is to compare the performance of the SL preconditioner with the DPSS and RDPSS preconditioners. Experiments have been performed by using GMRES with these preconditioners. All the sub-linear systems arising in the application of the preconditioners are solved by Cholesky or LU factorization in combination with AMD reordering. As for the preconditioning matrix Q in the SL preconditioner, we consider the following two choices: • Case I: Q = diag(A); • Case II: Q = tridiag(H ) with H = 12 (A + A T ). The test problem is a “leaky” two dimensional lid-driven cavity problem. We discretize the Oseen problem (1.2) with Q2 − Q1 finite element method on uniform or stretched l × l grids. In computations, we employ the function driver “navier_textproblem” of the IFISS software package [21] with default options except meshsize. The derived linear systems are structured as (1.4) with n = 2(l + 1)2 and m = 41 (l + 2)2 , for a total 94 l 2 + 5l + 3 unknowns. All computations are performed in MATLAB (version 8.3.0.532 (R2014a)) on a personal computer with 2.60 GHz central processing unit (Intel(R) Core(TM) i5-4210) and 4.00G memory. Starting with zero initial guesses, we compare the tested methods in terms of the number of iteration steps (denoted by ‘IT’) and elapsed CPU times in seconds (denoted by ‘CPU’). All iteration processes are terminated when the current residuals satisfy RES :=
Æ
k f − A x (k ) − B T y (k ) k22 + k g − B x (k ) k22 < 10−6 Æ k f k22 + k g k22
with {((x (k ) )T , (y (k ) )T )T } being the current approximation solutions. The numerical results of the tested methods on the uniform and stretched grids are depicted in Tables 1-4 and 5-8, respectively. Here, ‘DPSS’ and ‘RDPSS’ denote GMRES with DPSS and RDPSS preconditioners, respectively. Analogously, the SL preconditioned GMRES with the preconditioning matrices Q chosen as Cases I and II are, respectively, denoted as ‘SL-I’ and ‘SL-II’. GMRES without preconditioning is denoted by ‘Non’. In Tables 1-3 and 5-7, problems of different sizes and iteration parameters ν are used for each tested method with varied viscosity parameters α = 0.01, 0.1 and 0.01. Meanwhile, in Tables 4 and 8, the parameters αopt are the theoretically found optimal ones based on Theorem 2.4. Here, ‘–’ means the tested method is not convergent within the prescribed iteration number k max = 1000. From these tables, we can see significant improvements for the performance of GMRES with all preconditioners. Meanwhile, the SL and RDPSS preconditioners lead to much better numerical results than the DPSS preconditioner. We have the observations from Tables 1-3 and 5-7 that the SL and RDPSS preconditioned GMRES need less iteration steps and CPU times comparing with the DPSS preconditioned ones. Besides, we also observe that their performances are not sensitive to the iteration parameter α though the GMRES with DPSS preconditioners are unduly parameter dependent. Comparing the performance of the SL and RDPSS preconditioned GMRES, we see that the SL preconditioned GMRES with Q chosen as Case I leads to better performance than the RDPSS preconditoined GMRES. It appears to offer advantages in terms of both iteration steps and runtime. As for the SL preconditioned GMRES with Q chosen as Case II, it may outperform all other methods in terms of iteration steps but it becomes more time consuming with the incresing of problem size. The reason is that the involved sub-linear system BQ −1 B T y = b is much harder to solve when Q is chosen as Case II as it leads to denser coefficient matrix. Moreover, its poor performance for solving problems on large stretched grids is owing to the losing of symmetric dominance for A. In order to further investigate the dependence of the RDPSS and SL preconditioned GMRES on the iteration parameter α, we illustrate the changing of their iteration steps with respect to α from 0.001 to 1 in Figure 1. Here, we choose the case with ν = 0.1 for 32 × 32 grids. From this figure, we see that the SL preconditioned GMRES needs less iteration steps than the RDPSS preconditioned GMRES. What is more, their iteration counts do not change dramatically with the changing of α. Thus, it may exist a fairly wide range of values of α that produce similar fast convergence results. 9
The eigenvalue distributions of the the original matrix A and the preconditioned matrices of all the tested methods are shown in Figures 2-4. The extreme eigenvalues and condition numbers of these matrices are listed in Table 9. Meanwhile, the convergence plots for the corresponding PGMRES methods are shown in Figure 5. Here, we also choose the case with ν = 0.1 on 32 × 32 grids. RDPSS denotes the RDPSS preconditioned matrices, and the SL preconditioned ones with the preconditioning matrix Q chosen as Cases I and II are denoted by ‘SL-I’ and ‘SL-II’, respectively. For all the preconditioners, the parameters α is the optimal ones chosen as Table 4 or 8. In Table 9, γmax and γmin denote, respectively, the upper and lower bounds of the real part of the eigenvalues of the related matrix, ηmax denotes the upper bound of the absolute value of the imaginary part of the eigenvalues of the related matrix. κ is the spectral condition number defined as in Section 4. From Figures 2-4, we see that all the preconditioned matrices have much denser spectrum distributions comparing with the original matrices. Note that the spectrum of the original matrices are also bounded, but they are much closer to the origin, which might delay the convergence. See the numerical results of GMRES without preconditioning in Tables 4 and 8. Moreover, spectrum distribution of the SL preconditioned matrices are more clustered than the RDPSS preconditioned ones. The formers are more separate from zero which in turn lead to faster convergence of PGMRES, as we can see the numerical results in Tables 4 and 8. The results of Table 9 agree with the numerical results in Tables 4 and 8 and Figures 2-5. The SL preconditioned matrices have tighter extreme eigenvalue distributions and smaller condition numbers than the original matrices and the RDPSS preconditioned matrices. Thus the SL preconditioned GMRES is well defined and it has better numerical performance than the GMRES method without preconditioning and the RDPSS preconditioned GMRES. Table 1: Numerical results of PGMRES for uniform grids (ν = 0.01).
Case α
0.01
DPSS
IT CPU
RDPSS
32 × 32 0.1
1
0.01
42 0.1623
137 0.9569
399 8.0655
IT CPU
52 0.2216
60 0.2645
SL-I
IT CPU
36 0.1421
SL-II
IT CPU
33 0.1955
64 × 64
128 × 128
0.1
1
0.01
0.1
1
93 2.7633
368 15.6052
856 94.1069
206 23.2811
914 320.0132
– –
64 0.2871
76 1.3757
86 1.6204
93 1.7928
103 9.2892
120 11.4096
139 13.6078
44 0.1865
49 0.2121
46 0.7994
61 1.0737
63 1.1122
61 5.2938
79 6.9173
77 6.7923
41 0.2363
45 0.2601
42 1.3627
54 1.5103
57 1.6121
54 16.8545
70 16.5465
68 15.1968
6. Conclusions We have proposed and analyzed a class of SIMPLE-like (SL) preconditioners for solving saddle point problems from Navier-Stokes problems. The new preconditioners can include the RDPSS preconditioner proposed in [15] as a special case. Convergence analysis of the corresponding SL iteration method and the selection of the optimal iteration parameter are given. Spectral analysis of the SL preconditioned matrix are presented, including bounds of the eigenvalues, distribution of the corresponding eigenvectors and its minimal polynomial. Numerical experiments from “leaky” two-dimensional lid-driven cavity problem are used to show its advantages over the DPSS and RDPSS preconditioners. References [1] Z.-Z. Bai, Optimal parameters in the HSS-like methods for saddle-point problems, Numer. Linear Algebra Appl. 16(6) (2009) 447–479.
10
Table 2: Numerical results of PGMRES for uniform grids (ν = 0.1).
Case α
0.01
DPSS
IT CPU
RDPSS
32 × 32 0.1
1
0.01
40 0.1618
70 0.3464
203 1.9703
IT CPU
26 0.1008
31 0.1181
SL-I
IT CPU
24 0.1002
SL-II
IT CPU
21 0.1505
64 × 64
128 × 128
0.1
1
0.01
0.1
1
55 0.9839
160 3.7812
467 25.3184
98 8.7591
236 8.6679
989 373.8501
34 0.1283
40 0.7112
46 0.8198
51 0.8814
59 5.1424
63 5.4762
77 6.6272
28 0.1092
30 0.1179
34 0.5973
41 0.7114
42 0.7263
50 4.3202
55 4.7845
59 5.1354
24 0.1545
27 0.1772
30 1.1793
36 1.1998
30 1.1931
44 15.5499
50 15.7271
51 15.9122
Table 3: Numerical results of PGMRES for uniform grids (ν = 1).
Case α
0.01
DPSS
IT CPU
RDPSS
32 × 32 0.1
1
0.01
44 0.2856
50 0.2212
132 0.8995
IT CPU
23 0.1605
26 0.1004
SL-I
IT CPU
25 0.1145
SL-II
IT CPU
21 0.1695
64 × 64 0.1
1
0.01
65 1.0978
89 1.6814
265 8.4163
28 0.1099
36 0.6239
38 0.6493
26 0.1061
28 0.1105
36 0.6292
22 0.1447
24 0.1549
33 1.1408
0.1
1
76 6.5216
112 10.1362
504 99.3014
40 0.6807
52 4.4792
52 4.4327
56 4.6849
38 0.6475
43 0.7528
49 4.1803
51 4.3627
65 5.5702
34 1.1578
37 1.1889
44 14.9657
46 15.1361
56 16.3128
Uniform grids
Stretched grids 40
SL−I SL−II RDPSS
36 34
128 × 128
SL−I SL−II RDPSS
35
32 30
30
Iter
Iter
28 26
25
24 20 22 20
15
18 16
0
0.2
0.4
0.6
0.8
10
1
α
0
0.2
0.4
0.6
α
Figure 1: Iteration steps for SL and RDPSS methods with varying α for 32 × 32 grids (ν = 0.1).
11
0.8
1
Table 4: Numerical results of GMRES and PGMRES with optimal parameters for uniform grids.
Case Non
ν
0.01
IT CPU
248 0.0243
16 × 16 0.1
1
124 157 0.5103 1.0213
4.10e-03 0.0563 0.4614 29 19 17 0.0243 0.0322 0.0302
0.01 521 8.7303
32 × 32 0.1
1
0.01
0.1
1
636 39.5201
700 40.8375
3.91e-05 6.19e-04 22 30 0.4125 0.5062
0.0302 38 0.7187
284 349 3.4619 4.7825
2.86e-04 0.0146 0.1198 23 27 26 0.0903 0.1009 0.1071
64 × 64
– –
RDPSS
αopt IT CPU
SL-I
αopt IT CPU
0.0779 31 0.0294
0.1075 0.0980 18 16 0.0287 0.0351
0.0053 23 0.0853
0.0295 0.0253 25 25 0.0955 0.1109
4.57e-04 22 0.4321
0.0013 28 0.5105
0.0064 37 0.7457
SL-II
αopt IT CPU
0.1139 28 0.0341
0.1537 0.1509 16 15 0.0376 0.0352
0.0085 20 0.1413
0.0455 0.0407 23 22 0.1497 0.1678
9.68e-04 27 0.9961
0.0023 27 0.9848
0.0104 33 1.1902
Table 5: Numerical results of PGMRES for stretched grids (ν = 0.01).
Case α
0.01
DPSS
IT CPU
RDPSS
32 × 32 0.1
1
0.01
32 0.1221
101 0.6107
401 8.2552
IT CPU
37 0.1516
49 0.1982
SL-I
IT CPU
24 0.0962
SL-II
IT CPU
24 0.1635
64 × 64 0.1
1
0.01
57 1.0019
302 11.1828
– –
56 0.2336
52 0.9036
72 1.3651
30 0.1193
34 0.1193
32 0.5699
39 0.2118
46 0.2934
77 2.7475
128 × 128 0.1
1
136 11.0712
867 152.2219
– –
81 1.4584
63 10.5969
83 13.7179
103 16.2173
42 0.7384
48 0.8521
43 7.3003
57 9.6872
61 9.9483
142 4.7715
158 5.2693
154 46.3043
364 115.9819
436 148.8143
[2] Z.-Z. Bai, G.H. Golub, J.-Y. Pan, Preconditioned Hermitian and skew-Hermitian splitting methods for non-Hermitian positive semidefinite linear systems, Numer. Math. 98(1) (2004) 1–32. [3] Z.-Z. Bai, G.H. Golub, Accelerated Hermitian and skew-Hermitian splitting iteration methods for saddle-point problems, IMA J. Numer. Anal. 27(1) (2007) 1–23. [4] Z.-Z. Bai, G.H. Golub, L.-Z. Lu, J.-F. Yin, Block triangular and skew-Hermitian splitting methods for positive-definite linear systems, SIAM J. Sci. Comput. 26(3) (2005) 844–863. [5] Z.-Z. Bai, M.K. Ng, Z.-Q. Wang, Constraint preconditioners for symmetric indefinite matrices, SIAM J. Matrix Anal. Appl. 31(2) (2009) 410–433. [6] Z.-Z. Bai, B.N. Parlett, Z.-Q. Wang, On generalized successive overrelaxation methods for augmented linear systems, Numer. Math. 102(1) (2005) 1–38. [7] Z.-Z. Bai, Z.-Q. Wang, On parameterized inexact Uzawa methods for generalized saddle point problems, Linear Algebra Appl. 428(11) (2008) 2900–2932. [8] M. Benzi, G.H. Golub, A preconditioner for generalized saddle point problems, SIAM J. Matrix Anal. Appl. 26(1) (2004) 20–41. [9] M. Benzi, G.H. Golub, J. Liesen. Numerical solution of saddle point problems, Acta Numer. 14 (2005) 1–137. [10] M. Benzi, A. Olshanskii, An augmented lageangian-based approach to the Ossen problem, SIAM J. Sci. Comput. 28 (2006) 2095–2113. [11] M. Benzi, V. Simoncini, On the eigenvalues of a class of saddle point matrices, Numer. Math. 103(2) (2006) 173–196. [12] M. Benzi, Z. Wang, Analysis of augmented Lagrangian-based preconditioners for the steady incompressible Navier-Stokes equations, SIAM J. Sci. Comput. 33 (2011) 2761–2784. [13] J.H. Bramble, J.E. Pasciak, A.T. Vassilev, Analysis of the inexact Uzawa algorithm for saddle point problems, SIAM J. Numer. Anal. 34(3)
12
Table 6: Numerical results of PGMRES for stretched grids (ν = 0.1).
Case α
0.01
DPSS
IT CPU
RDPSS
32 × 32 0.1
1
0.01
41 0.1975
79 0.4132
333 5.3926
IT CPU
36 0.1396
32 0.1217
SL-I
IT CPU
21 0.0843
SL-II
IT CPU
18 0.1435
64 × 64 0.1
1
0.01
67 1.1572
255 8.0346
990 133.0228
36 0.1364
51 0.8602
56 0.9732
25 0.0977
29 0.1138
37 0.6447
23 0.1471
26 0.1741
80 2.5266
128 × 128 0.1
1
146 27.2852
573 131.4725
– –
59 0.9945
75 12.7767
79 13.6293
80 13.6665
38 0.6742
37 0.6443
50 8.4356
53 9.1227
62 10.7012
99 3.1742
106 3.4213
228 69.2017
340 104.0383
354 110.4306
Table 7: Numerical results of PGMRES for stretched grids (ν = 1).
Case α
0.01
DPSS
IT CPU
RDPSS
32 × 32 0.1
1
0.01
65 0.3072
91 0.4953
272 3.2254
IT CPU
36 0.1378
39 0.1485
SL-I
IT CPU
23 0.0909
SL-II
IT CPU
28 0.1575
64 × 64 0.1
1
0.01
124 2.5416
258 8.0959
737 65.8169
38 0.1496
70 1.2213
70 1.2505
24 0.1222
28 0.1072
34 0.5927
25 0.1678
27 0.1745
81 2.6792
0.1
1
181 35.0339
612 192.7699
– –
69 1.2018
117 20.9962
110 19.2573
110 19.3117
39 0.6704
45 0.7725
51 8.5301
52 8.7742
76 13.1077
93 3.1211
97 3.2647
305 67.1044
307 71.2677
307 82.7895
Stretched grids
Uniform grids 0.08
0.2
0.06
0.15 0.1
imaginary
imaginary
0.04 0.02 0 −0.02
0.05 0 −0.05
−0.04
−0.1
−0.06
−0.15
−0.08
128 × 128
0
0.2
0.4
0.6
0.8
1
1.2
−0.2
1.4
0
0.5
1
1.5
2
2.5
real
real
Figure 2: The eigenvalue distributions of the original matrices A for 32 × 32 grids (ν = 0.1).
13
3
3.5
4
Table 8: Numerical results of GMRES and PGMRES with optimal parameters for stretched grids.
Case 0.01
ν Non
16 × 16 0.1
1
32 × 32
0.01
0.1
1
0.01
535 866 10.2099 22.1571
64 × 64 0.1
1
– –
– –
IT CPU
207 151 255 1.3174 0.6468 2.0578
522 9.8754
– –
RDPSS
αopt IT CPU
0.0016 0.0953 0.8165 23 17 20 0.0304 0.0234 0.0381
1.76e-03 28 0.1037
0.0311 31 0.1132
0.2680 38 0.1535
SL-I
αopt IT CPU
0.1188 0.1537 0.1385 22 16 16 0.0287 0.0205 0.0255
0.0436 27 0.1016
0.0464 24 0.0946
0.0405 24 0.1031
0.0020 22 0.4892
0.0059 0.0113 31 34 0.5632 0.7023
SL-II
αopt IT CPU
0.1629 0.2297 0.2442 25 15 14 0.0475 0.0277 0.0373
0.0264 31 0.1822
0.0722 22 0.1472
0.0722 25 0.1599
0.0016 33 0.8621
0.0190 0.0189 82 83 2.7549 2.7219
1.55e-04 0.0015 0.0826 22 30 70 0.4526 0.5136 2.2162
Table 9: The extreme eigenvalues and condition numbers of the original and preconditioned matrices for 32 × 32 grids (ν = 0.1).
Case
Uniform grids γmax
γmin
Stretched grids
ηmax
κ
γmax
γmin
ηmax
κ
Non
1.0000 0.0017 0.0724 575.6935
3.9831 4.04e-04 0.1742
RDPSS
1.9830 0.0146 0.1480 136.2006
1.9827
0.0185
0.1359 107.1463
SL-I
1.9709 0.0276 0.2298
71.8912
1.9567
0.0417
0.1538
46.9040
SL-II
1.9567 0.0437 0.2814
45.2066
1.9619
0.0368
0.1138
53.3220
RDPSS
0.15
SL-I
0.25
9.85e-03
SL-II
0.3
0.2 0.1
0.2 0.15 0.1
0
−0.05
imaginary
0.1
imaginary
imaginary
0.05
0.05 0 −0.05 −0.1
−0.1
0
−0.1
−0.2
−0.15 −0.15
−0.3 −0.2
−0.2
0
0.5
1
real
1.5
2
−0.25
0
0.5
1
real
1.5
2
−0.4
0
0.5
1
1.5
2
real
Figure 3: The eigenvalue distributions of the preconditioned matrices for 32 × 32 uniform grids (ν = 0.1).
(1997) 1072–1092. [14] J.H. Bramble, J.E. Pasciak, A.T. Vassilev, Uzawa type algorithms for nonsymmetric saddle point problems. Math. Comput. 69(230) (2000) 667–689. [15] Y. Cao, J.-L. Dong, Y.-M. Wang, A relaxed deteriorated PSS preconditioner for nonsymmetric saddle point problems from the steady Navier-Stokes equation, J. Comput. Appl. Math. 273 (2015) 41–60. [16] Y. Cao, M.-Q. Jiang, Y.-L. Zheng, A splitting preconditioner for saddle point problems, Numer. Linear Algebra Appl. 18(5) (2011) 875–895. [17] Y. Cao, L.-Q. Yao, M.-Q. Jiang, A relaxed HSS preconditioner for saddle point problems from meshfree discretization, J. Comput. Math.
14
RDPSS
0.15
−0.05
−0.1
0.05
imaginary
imaginary
imaginary
0.1
0.1
0
0.05 0 −0.05
−0.15
−0.15
0
0.5
1
1.5
−0.2
2
0
−0.05
−0.1
−0.1
−0.15
SL-II
0.15
0.15
0.1
0.05
−0.2
SL-I
0.2
0
0.5
real
1
1.5
−0.2
2
0
0.5
real
1
1.5
2
real
Figure 4: The eigenvalue distributions of the preconditioned matrices for 32 × 32 stretched grids (ν = 0.1).
Uniform grids
−1
10
SL−I SL−II RDPSS
−2
10
−2
−3
−3
||rk ||2 /||r0 ||2
||rk ||2 /||r0 ||2
SL−I SL−II RDPSS
10
10
−4
10
−5
10
−6
10
−4
10
−5
10
−6
10
10
−7
10
Stretched grids
−1
10
−7
0
5
10
15
20
25
10
30
0
5
10
15
20
25
30
35
k
k
Figure 5: The convergence curves of PGMRES for 32 × 32 grids (ν = 0.1).
31(4) (2013) 398–421. [18] A.C. de Niet, F.W. Wubs, Two preconditioners for saddle point problems in fluid flows. Int. J. Numer. Meth. Fluids 54(4) (2007) 355–377. [19] H.C. Elman, Preconditioning for the steady-state Navier-Stokes equations with low viscosity, SIAM J. Sci. Comput. 20(4) (1999) 1299– 1316. [20] H.C. Elman, G.H. Golub, Inexact and preconditioned Uzawa algorithms for saddle point problems, SIAM J. Numer. Anal. 31(6) (1994) 1645–1661. [21] H.C. Elman, A. Ramage, D.J. Silvester, Algorithm 866: IFISS, a Matlab toolbox for modelling incompressible flow, ACM Trans. Math. Software, 33(2) (2007) Article 14. [22] J.H. Ferziger, M. Peri´c, Computational Methods for Fluid Dynamics, Springer Berlin, 1996. [23] R. Glowinski, O. Pironneau, Finite Element Methods for Navier-Stokes Equations, Springer, Berlin, 1986. [24] G.H. Golub, C. Greif, On solving block-structured indefinite linear systems, SIAM J. Sci. Comput. 24 (2003) 2076–2092. [25] X. He, M. Neytcheva, S.S. Capizzano, On an augmented Lagrangian-based preconditioning of Oseen type problems. BIT Numer. Math. 51(4) (2011) 865–888. [26] C. Keller, N. I. Gould, A.J. Wathen, Constraint preconditioning for indefinite linear systems, SIAM J. Matrix Anal. Appl. 21(4) (2000) 1300–1317. [27] P. Krzyzanowski, On block preconditioners for nonsymmetric saddle point problems, SIAM J. Sci. Comput. 23(1) (2001) 157–169. [28] C.-G. Li, C. Vuik, Eigenvalue analysis of the SIMPLE preconditioning for incompressible flow, Numer. Linear Algebra Appl. 11 (2004)511–523. [29] M.F. Murphy, G.H. Golub, A.J. Wathen, A note on preconditioning for indefinite linear systems, SIAM J. Sci. Comput. 21(6) (2000) 1969–1972. [30] Y. Notay, A new analysis of block preconditioners for saddle point problems, SIAM J. Matrix Anal. Appl. 35(1) (2014) 143–173. [31] J.-Y. Pan, M.K. Ng, Z.-Z. Bai, New preconditioners for saddle point problems, Appl. Math. Comput. 172(2) (2006) 762–771. [32] S.V. Patankar. Numerical Heat Transfer and Fluid Flow, McGraw-Hill, New York, 1980. [33] Y. Saad, Iterative Methods for Sparse Linear Systems, SIAM, Philadelphia, PA, 2003.
15
[34] Y. Saad, M.H. Schultz, GMRES: A generalized minimal residual algorithm for solving nonsymmetric linear systems, SIAM J. Sci. Stat. comput. 7(3) (1986) 856–869. [35] V. Simoncin, M. Benzi, Specrtal properities of the Hermitian and skew-Hermitian splitting preconditioner for saddle point problems, SIAM J. Matrix Anal. Appl. 26(2) (2004) 377–389. [36] V. Simoncini, Block triangular preconditioners for symmetric saddle-point problems, Appl. Numer. Math. 49(1) (2004) 63–80. [37] C. Vuik, G. Segal, A comparision of preconditioners for incompressible Navier-Stokes solvers, Int. J. Numer. Meth. Fluids 57(12) (2008) 1731–1751. [38] C. Vuik, G. Segal, G.P. Boerstoel, The Krylov accelerated SIMPLE(R) method for flow problems in industrial furnaces. Int. J. Numer. Meth. Fluids 33(7) (2000) 1027–1040.
16