Conjugate gradient method

Last updated
A comparison of the convergence of gradient descent with optimal step size (in green) and conjugate vector (in red) for minimizing a quadratic function associated with a given linear system. Conjugate gradient, assuming exact arithmetic, converges in at most n steps, where n is the size of the matrix of the system (here n = 2). Conjugate gradient illustration.svg
A comparison of the convergence of gradient descent with optimal step size (in green) and conjugate vector (in red) for minimizing a quadratic function associated with a given linear system. Conjugate gradient, assuming exact arithmetic, converges in at most n steps, where n is the size of the matrix of the system (here n = 2).

In mathematics, the conjugate gradient method is an algorithm for the numerical solution of particular systems of linear equations, namely those whose matrix is positive-definite. The conjugate gradient method is often implemented as an iterative algorithm, applicable to sparse systems that are too large to be handled by a direct implementation or other direct methods such as the Cholesky decomposition. Large sparse systems often arise when numerically solving partial differential equations or optimization problems.

Contents

The conjugate gradient method can also be used to solve unconstrained optimization problems such as energy minimization. It is commonly attributed to Magnus Hestenes and Eduard Stiefel, [1] [2] who programmed it on the Z4, [3] and extensively researched it. [4] [5]

The biconjugate gradient method provides a generalization to non-symmetric matrices. Various nonlinear conjugate gradient methods seek minima of nonlinear optimization problems.

Description of the problem addressed by conjugate gradients

Suppose we want to solve the system of linear equations

for the vector , where the known matrix is symmetric (i.e., AT = A), positive-definite (i.e. xTAx > 0 for all non-zero vectors in Rn), and real, and is known as well. We denote the unique solution of this system by .

Derivation as a direct method

The conjugate gradient method can be derived from several different perspectives, including specialization of the conjugate direction method for optimization, and variation of the Arnoldi/Lanczos iteration for eigenvalue problems. Despite differences in their approaches, these derivations share a common topic—proving the orthogonality of the residuals and conjugacy of the search directions. These two properties are crucial to developing the well-known succinct formulation of the method.

We say that two non-zero vectors u and v are conjugate (with respect to ) if

Since is symmetric and positive-definite, the left-hand side defines an inner product

Two vectors are conjugate if and only if they are orthogonal with respect to this inner product. Being conjugate is a symmetric relation: if is conjugate to , then is conjugate to . Suppose that

is a set of mutually conjugate vectors with respect to , i.e. for all . Then forms a basis for , and we may express the solution of in this basis:

Left-multiplying the problem with the vector yields

and so

This gives the following method [4] for solving the equation Ax = b: find a sequence of conjugate directions, and then compute the coefficients .

As an iterative method

If we choose the conjugate vectors carefully, then we may not need all of them to obtain a good approximation to the solution . So, we want to regard the conjugate gradient method as an iterative method. This also allows us to approximately solve systems where n is so large that the direct method would take too much time.

We denote the initial guess for x by x0 (we can assume without loss of generality that x0 = 0, otherwise consider the system Az = bAx0 instead). Starting with x0 we search for the solution and in each iteration we need a metric to tell us whether we are closer to the solution x (that is unknown to us). This metric comes from the fact that the solution x is also the unique minimizer of the following quadratic function

The existence of a unique minimizer is apparent as its Hessian matrix of second derivatives is symmetric positive-definite

and that the minimizer (use Df(x)=0) solves the initial problem follows from its first derivative

This suggests taking the first basis vector p0 to be the negative of the gradient of f at x = x0. The gradient of f equals Axb. Starting with an initial guess x0, this means we take p0 = bAx0. The other vectors in the basis will be conjugate to the gradient, hence the name conjugate gradient method. Note that p0 is also the residual provided by this initial step of the algorithm.

Let rk be the residual at the kth step:

As observed above, is the negative gradient of at , so the gradient descent method would require to move in the direction rk. Here, however, we insist that the directions must be conjugate to each other. A practical way to enforce this is by requiring that the next search direction be built out of the current residual and all previous search directions. The conjugation constraint is an orthonormal-type constraint and hence the algorithm can be viewed as an example of Gram-Schmidt orthonormalization. This gives the following expression:

(see the picture at the top of the article for the effect of the conjugacy constraint on convergence). Following this direction, the next optimal location is given by

with

where the last equality follows from the definition of . The expression for can be derived if one substitutes the expression for xk+1 into f and minimizing it with respect to

The resulting algorithm

The above algorithm gives the most straightforward explanation of the conjugate gradient method. Seemingly, the algorithm as stated requires storage of all previous searching directions and residue vectors, as well as many matrix–vector multiplications, and thus can be computationally expensive. However, a closer analysis of the algorithm shows that is orthogonal to , i.e. , for i ≠ j. And is -orthogonal to , i.e. , for . This can be regarded that as the algorithm progresses, and span the same Krylov subspace. Where form the orthogonal basis with respect to the standard inner product, and form the orthogonal basis with respect to the inner product induced by . Therefore, can be regarded as the projection of on the Krylov subspace.

That is, if the CG method starts with , then [6]

The algorithm is detailed below for solving where is a real, symmetric, positive-definite matrix. The input vector can be an approximate initial solution or 0. It is a different formulation of the exact procedure described above.

This is the most commonly used algorithm. The same formula for βk is also used in the Fletcher–Reeves nonlinear conjugate gradient method.

Restarts

We note that is computed by the gradient descent method applied to . Setting would similarly make computed by the gradient descent method from , i.e., can be used as a simple implementation of a restart of the conjugate gradient iterations. [4] Restarts could slow down convergence, but may improve stability if the conjugate gradient method misbehaves, e.g., due to round-off error.

Explicit residual calculation

The formulas and , which both hold in exact arithmetic, make the formulas and mathematically equivalent. The former is used in the algorithm to avoid an extra multiplication by since the vector is already computed to evaluate . The latter may be more accurate, substituting the explicit calculation for the implicit one by the recursion subject to round-off error accumulation, and is thus recommended for an occasional evaluation. [7]

A norm of the residual is typically used for stopping criteria. The norm of the explicit residual provides a guaranteed level of accuracy both in exact arithmetic and in the presence of the rounding errors, where convergence naturally stagnates. In contrast, the implicit residual is known to keep getting smaller in amplitude well below the level of rounding errors and thus cannot be used to determine the stagnation of convergence.

Computation of alpha and beta

In the algorithm, αk is chosen such that is orthogonal to . The denominator is simplified from

since . The βk is chosen such that is conjugate to . Initially, βk is

using

and equivalently

the numerator of βk is rewritten as

because and are orthogonal by design. The denominator is rewritten as

using that the search directions pk are conjugated and again that the residuals are orthogonal. This gives the β in the algorithm after cancelling αk.

Example code in Julia (programming language)

"""    conjugate_gradient!(A, b, x)Return the solution to `A * x = b` using the conjugate gradient method."""functionconjugate_gradient!(A::AbstractMatrix,b::AbstractVector,x::AbstractVector;tol=eps(eltype(b)))# Initialize residual vectorresidual=b-A*x# Initialize search direction vectorsearch_direction=residual# Compute initial squared residual normnorm(x)=sqrt(sum(x.^2))old_resid_norm=norm(residual)# Iterate until convergencewhileold_resid_norm>tolA_search_direction=A*search_directionstep_size=old_resid_norm^2/(search_direction'*A_search_direction)# Update solution@.x=x+step_size*search_direction# Update residual@.residual=residual-step_size*A_search_directionnew_resid_norm=norm(residual)# Update search direction vector@.search_direction=residual+(new_resid_norm/old_resid_norm)^2*search_direction# Update squared residual norm for next iterationold_resid_norm=new_resid_normendreturnxend

Numerical example

Consider the linear system Ax = b given by

we will perform two steps of the conjugate gradient method beginning with the initial guess

in order to find an approximate solution to the system.

Solution

For reference, the exact solution is

Our first step is to calculate the residual vector r0 associated with x0. This residual is computed from the formula r0 = b - Ax0, and in our case is equal to

Since this is the first iteration, we will use the residual vector r0 as our initial search direction p0; the method of selecting pk will change in further iterations.

We now compute the scalar α0 using the relationship

We can now compute x1 using the formula

This result completes the first iteration, the result being an "improved" approximate solution to the system, x1. We may now move on and compute the next residual vector r1 using the formula

Our next step in the process is to compute the scalar β0 that will eventually be used to determine the next search direction p1.

Now, using this scalar β0, we can compute the next search direction p1 using the relationship

We now compute the scalar α1 using our newly acquired p1 using the same method as that used for α0.

Finally, we find x2 using the same method as that used to find x1.

The result, x2, is a "better" approximation to the system's solution than x1 and x0. If exact arithmetic were to be used in this example instead of limited-precision, then the exact solution would theoretically have been reached after n = 2 iterations (n being the order of the system).

Convergence properties

The conjugate gradient method can theoretically be viewed as a direct method, as in the absence of round-off error it produces the exact solution after a finite number of iterations, which is not larger than the size of the matrix. In practice, the exact solution is never obtained since the conjugate gradient method is unstable with respect to even small perturbations, e.g., most directions are not in practice conjugate, due to a degenerative nature of generating the Krylov subspaces.

As an iterative method, the conjugate gradient method monotonically (in the energy norm) improves approximations to the exact solution and may reach the required tolerance after a relatively small (compared to the problem size) number of iterations. The improvement is typically linear and its speed is determined by the condition number of the system matrix : the larger is, the slower the improvement. [8]

If is large, preconditioning is commonly used to replace the original system with such that is smaller than , see below.

Convergence theorem

Define a subset of polynomials as

where is the set of polynomials of maximal degree .

Let be the iterative approximations of the exact solution , and define the errors as . Now, the rate of convergence can be approximated as [4] [9]

where denotes the spectrum, and denotes the condition number.

This shows iterations suffices to reduce the error to for any .

Note, the important limit when tends to

This limit shows a faster convergence rate compared to the iterative methods of Jacobi or Gauss–Seidel which scale as .

No round-off error is assumed in the convergence theorem, but the convergence bound is commonly valid in practice as theoretically explained [5] by Anne Greenbaum.

Practical convergence

If initialized randomly, the first stage of iterations is often the fastest, as the error is eliminated within the Krylov subspace that initially reflects a smaller effective condition number. The second stage of convergence is typically well defined by the theoretical convergence bound with , but may be super-linear, depending on a distribution of the spectrum of the matrix and the spectral distribution of the error. [5] In the last stage, the smallest attainable accuracy is reached and the convergence stalls or the method may even start diverging. In typical scientific computing applications in double-precision floating-point format for matrices of large sizes, the conjugate gradient method uses a stopping criteria with a tolerance that terminates the iterations during the first or second stage.

The preconditioned conjugate gradient method

In most cases, preconditioning is necessary to ensure fast convergence of the conjugate gradient method. If is symmetric positive-definite and has a better condition number than , a preconditioned conjugate gradient method can be used. It takes the following form: [10]

repeat
ifrk+1 is sufficiently small then exit loop end if
end repeat
The result is xk+1

The above formulation is equivalent to applying the regular conjugate gradient method to the preconditioned system [11]

where

The Cholesky decomposition of the preconditioner must be used to keep the symmetry (and positive definiteness) of the system. However, this decomposition does not need to be computed, and it is sufficient to know . It can be shown that has the same spectrum as .

The preconditioner matrix M has to be symmetric positive-definite and fixed, i.e., cannot change from iteration to iteration. If any of these assumptions on the preconditioner is violated, the behavior of the preconditioned conjugate gradient method may become unpredictable.

An example of a commonly used preconditioner is the incomplete Cholesky factorization. [12]

Using the preconditioner in practice

It is importart to keep in mind that we don't want to invert the matrix explicitly in order to get for use it in the process, since inverting would take more time/computational resources than solving the conjugate gradient algorithm itself. As an example, let's say that we are using a preconditioner coming from incomplete Cholesky factorization. The resulting matrix is the lower triangular matrix , and the preconditioner matrix is:

Then we have to solve:

But:

Then:

Let's take an intermediary vector :

Since and and known, and is lower triangular, solving for is easy and computationally cheap by using forward substitution. Then, we substitute in the original equation:

Since and are known, and is upper triangular, solving for is easy and computationally cheap by using backward substitution.

Using this method, there is no need to invert or explicitly at all, and we still obtain .

The flexible preconditioned conjugate gradient method

In numerically challenging applications, sophisticated preconditioners are used, which may lead to variable preconditioning, changing between iterations. Even if the preconditioner is symmetric positive-definite on every iteration, the fact that it may change makes the arguments above invalid, and in practical tests leads to a significant slow down of the convergence of the algorithm presented above. Using the Polak–Ribière formula

instead of the Fletcher–Reeves formula

may dramatically improve the convergence in this case. [13] This version of the preconditioned conjugate gradient method can be called [14] flexible, as it allows for variable preconditioning. The flexible version is also shown [15] to be robust even if the preconditioner is not symmetric positive definite (SPD).

The implementation of the flexible version requires storing an extra vector. For a fixed SPD preconditioner, so both formulas for βk are equivalent in exact arithmetic, i.e., without the round-off error.

The mathematical explanation of the better convergence behavior of the method with the Polak–Ribière formula is that the method is locally optimal in this case, in particular, it does not converge slower than the locally optimal steepest descent method. [16]

Vs. the locally optimal steepest descent method

In both the original and the preconditioned conjugate gradient methods one only needs to set in order to make them locally optimal, using the line search, steepest descent methods. With this substitution, vectors p are always the same as vectors z, so there is no need to store vectors p. Thus, every iteration of these steepest descent methods is a bit cheaper compared to that for the conjugate gradient methods. However, the latter converge faster, unless a (highly) variable and/or non-SPD preconditioner is used, see above.

Conjugate gradient method as optimal feedback controller for double integrator

The conjugate gradient method can also be derived using optimal control theory. [17] In this approach, the conjugate gradient method falls out as an optimal feedback controller,

for the double integrator system,

The quantities and are variable feedback gains. [17]

Conjugate gradient on the normal equations

The conjugate gradient method can be applied to an arbitrary n-by-m matrix by applying it to normal equations ATA and right-hand side vector ATb, since ATA is a symmetric positive-semidefinite matrix for any A. The result is conjugate gradient on the normal equations (CGN or CGNR).

ATAx = ATb

As an iterative method, it is not necessary to form ATA explicitly in memory but only to perform the matrix–vector and transpose matrix–vector multiplications. Therefore, CGNR is particularly useful when A is a sparse matrix since these operations are usually extremely efficient. However the downside of forming the normal equations is that the condition number κ(ATA) is equal to κ2(A) and so the rate of convergence of CGNR may be slow and the quality of the approximate solution may be sensitive to roundoff errors. Finding a good preconditioner is often an important part of using the CGNR method.

Several algorithms have been proposed (e.g., CGLS, LSQR). The LSQR algorithm purportedly has the best numerical stability when A is ill-conditioned, i.e., A has a large condition number.

Conjugate gradient method for complex Hermitian matrices

The conjugate gradient method with a trivial modification is extendable to solving, given complex-valued matrix A and vector b, the system of linear equations for the complex-valued vector x, where A is Hermitian (i.e., A' = A) and positive-definite matrix, and the symbol ' denotes the conjugate transpose. The trivial modification is simply substituting the conjugate transpose for the real transpose everywhere.

Advantages and disadvantages

The advantages and disadvantages of the conjugate gradient methods are summarized in the lecture notes by Nemirovsky and BenTal. [18] :Sec.7.3

A pathological example

This example is from [19]

Let , and define

Since is invertible, there exists a unique solution to . Solving it by conjugate gradient descent gives us rather bad convergence:

In words, during the CG process, the error grows exponentially, until it suddenly becomes zero as the unique solution is found.

See also

Related Research Articles

<span class="mw-page-title-main">Gradient</span> Multivariate derivative (mathematics)

In vector calculus, the gradient of a scalar-valued differentiable function of several variables is the vector field whose value at a point gives the direction and the rate of fastest increase. The gradient transforms like a vector under change of basis of the space of variables of . If the gradient of a function is non-zero at a point , the direction of the gradient is the direction in which the function increases most quickly from , and the magnitude of the gradient is the rate of increase in that direction, the greatest absolute directional derivative. Further, a point where the gradient is the zero vector is known as a stationary point. The gradient thus plays a fundamental role in optimization theory, where it is used to minimize a function by gradient descent. In coordinate-free terms, the gradient of a function may be defined by:

In mathematics, a symmetric matrix with real entries is positive-definite if the real number is positive for every nonzero real column vector where is the transpose of . More generally, a Hermitian matrix is positive-definite if the real number is positive for every nonzero complex column vector where denotes the conjugate transpose of

<span class="mw-page-title-main">Matrix multiplication</span> Mathematical operation in linear algebra

In mathematics, particularly in linear algebra, matrix multiplication is a binary operation that produces a matrix from two matrices. For matrix multiplication, the number of columns in the first matrix must be equal to the number of rows in the second matrix. The resulting matrix, known as the matrix product, has the number of rows of the first and the number of columns of the second matrix. The product of matrices A and B is denoted as AB.

Unit quaternions, known as versors, provide a convenient mathematical notation for representing spatial orientations and rotations of elements in three dimensional space. Specifically, they encode information about an axis-angle rotation about an arbitrary axis. Rotation and orientation quaternions have applications in computer graphics, computer vision, robotics, navigation, molecular dynamics, flight dynamics, orbital mechanics of satellites, and crystallographic texture analysis.

In the mathematical field of differential geometry, a metric tensor is an additional structure on a manifold M that allows defining distances and angles, just as the inner product on a Euclidean space allows defining distances and angles there. More precisely, a metric tensor at a point p of M is a bilinear form defined on the tangent space at p, and a metric field on M consists of a metric tensor at each point p of M that varies smoothly with p.

<span class="mw-page-title-main">Gradient descent</span> Optimization algorithm

Gradient descent is a method for unconstrained mathematical optimization. It is a first-order iterative algorithm for finding a local minimum of a differentiable multivariate function.

In linear algebra, the adjugate of a square matrix A is the transpose of its cofactor matrix and is denoted by adj(A). It is also occasionally known as adjunct matrix, or "adjoint", though the latter term today normally refers to a different concept, the adjoint operator which for a matrix is the conjugate transpose.

In numerical analysis, one of the most important problems is designing efficient and stable algorithms for finding the eigenvalues of a matrix. These eigenvalue algorithms may also find eigenvectors.

<span class="mw-page-title-main">Projection (linear algebra)</span> Idempotent linear transformation from a vector space to itself

In linear algebra and functional analysis, a projection is a linear transformation from a vector space to itself such that . That is, whenever is applied twice to any vector, it gives the same result as if it were applied once. It leaves its image unchanged. This definition of "projection" formalizes and generalizes the idea of graphical projection. One can also consider the effect of a projection on a geometrical object by examining the effect of the projection on points in the object.

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

In linear algebra, a rotation matrix is a transformation matrix that is used to perform a rotation in Euclidean space. For example, using the convention below, the matrix

<span class="mw-page-title-main">Euler's rotation theorem</span> Movement with a fixed point is rotation

In geometry, Euler's rotation theorem states that, in three-dimensional space, any displacement of a rigid body such that a point on the rigid body remains fixed, is equivalent to a single rotation about some axis that runs through the fixed point. It also means that the composition of two rotations is also a rotation. Therefore the set of rotations has a group structure, known as a rotation group.

<span class="mw-page-title-main">Gauss–Newton algorithm</span> Mathematical algorithm

The Gauss–Newton algorithm is used to solve non-linear least squares problems, which is equivalent to minimizing a sum of squared function values. It is an extension of Newton's method for finding a minimum of a non-linear function. Since a sum of squares must be nonnegative, the algorithm can be viewed as using Newton's method to iteratively approximate zeroes of the components of the sum, and thus minimizing the sum. In this sense, the algorithm is also an effective method for solving overdetermined systems of equations. It has the advantage that second derivatives, which can be challenging to compute, are not required.

In numerical linear algebra, the method of successive over-relaxation (SOR) is a variant of the Gauss–Seidel method for solving a linear system of equations, resulting in faster convergence. A similar method can be used for any slowly converging iterative process.

In mathematics, preconditioning is the application of a transformation, called the preconditioner, that conditions a given problem into a form that is more suitable for numerical solving methods. Preconditioning is typically related to reducing a condition number of the problem. The preconditioned problem is then usually solved by an iterative method.

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

<span class="mw-page-title-main">Dual quaternion</span> Eight-dimensional algebra over the real numbers

In mathematics, the dual quaternions are an 8-dimensional real algebra isomorphic to the tensor product of the quaternions and the dual numbers. Thus, they may be constructed in the same way as the quaternions, except using dual numbers instead of real numbers as coefficients. A dual quaternion can be represented in the form A + εB, where A and B are ordinary quaternions and ε is the dual unit, which satisfies ε2 = 0 and commutes with every element of the algebra. Unlike quaternions, the dual quaternions do not form a division algebra.

Non-linear least squares is the form of least squares analysis used to fit a set of m observations with a model that is non-linear in n unknown parameters (m ≥ n). It is used in some forms of nonlinear regression. The basis of the method is to approximate the model by a linear one and to refine the parameters by successive iterations. There are many similarities to linear least squares, but also some significant differences. In economic theory, the non-linear least squares method is applied in (i) the probit regression, (ii) threshold regression, (iii) smooth regression, (iv) logistic link regression, (v) Box–Cox transformed regressors ().

In numerical linear algebra, the conjugate gradient method is an iterative method for numerically solving the linear system

The conjugate residual method is an iterative numeric method used for solving systems of linear equations. It's a Krylov subspace method very similar to the much more popular conjugate gradient method, with similar construction and convergence properties.

References

  1. Hestenes, Magnus R.; Stiefel, Eduard (December 1952). "Methods of Conjugate Gradients for Solving Linear Systems" (PDF). Journal of Research of the National Bureau of Standards. 49 (6): 409. doi: 10.6028/jres.049.044 .
  2. Straeter, T. A. (1971). On the Extension of the Davidon–Broyden Class of Rank One, Quasi-Newton Minimization Methods to an Infinite Dimensional Hilbert Space with Applications to Optimal Control Problems (PhD thesis). North Carolina State University. hdl: 2060/19710026200 via NASA Technical Reports Server.
  3. Speiser, Ambros (2004). "Konrad Zuse und die ERMETH: Ein weltweiter Architektur-Vergleich" [Konrad Zuse and the ERMETH: A worldwide comparison of architectures]. In Hellige, Hans Dieter (ed.). Geschichten der Informatik. Visionen, Paradigmen, Leitmotive (in German). Berlin: Springer. p. 185. ISBN   3-540-00217-0.
  4. 1 2 3 4 Polyak, Boris (1987). Introduction to Optimization.
  5. 1 2 3 Greenbaum, Anne (1997). Iterative Methods for Solving Linear Systems. doi:10.1137/1.9781611970937. ISBN   978-0-89871-396-1.
  6. Paquette, Elliot; Trogdon, Thomas (March 2023). "Universality for the Conjugate Gradient and MINRES Algorithms on Sample Covariance Matrices". Communications on Pure and Applied Mathematics. 76 (5): 1085–1136. arXiv: 2007.00640 . doi:10.1002/cpa.22081. ISSN   0010-3640.
  7. Shewchuk, Jonathan R (1994). An Introduction to the Conjugate Gradient Method Without the Agonizing Pain (PDF).
  8. Saad, Yousef (2003). Iterative methods for sparse linear systems (2nd ed.). Philadelphia, Pa.: Society for Industrial and Applied Mathematics. pp.  195. ISBN   978-0-89871-534-7.
  9. Hackbusch, W. (2016-06-21). Iterative solution of large sparse systems of equations (2nd ed.). Switzerland: Springer. ISBN   978-3-319-28483-5. OCLC   952572240.
  10. Barrett, Richard; Berry, Michael; Chan, Tony F.; Demmel, James; Donato, June; Dongarra, Jack; Eijkhout, Victor; Pozo, Roldan; Romine, Charles; van der Vorst, Henk. Templates for the Solution of Linear Systems: Building Blocks for Iterative Methods (PDF) (2nd ed.). Philadelphia, PA: SIAM. p. 13. Retrieved 2020-03-31.
  11. Golub, Gene H.; Van Loan, Charles F. (2013). Matrix Computations (4th ed.). Johns Hopkins University Press. sec. 11.5.2. ISBN   978-1-4214-0794-4.
  12. Concus, P.; Golub, G. H.; Meurant, G. (1985). "Block Preconditioning for the Conjugate Gradient Method". SIAM Journal on Scientific and Statistical Computing. 6 (1): 220–252. doi:10.1137/0906018.
  13. Golub, Gene H.; Ye, Qiang (1999). "Inexact Preconditioned Conjugate Gradient Method with Inner-Outer Iteration". SIAM Journal on Scientific Computing. 21 (4): 1305. CiteSeerX   10.1.1.56.1755 . doi:10.1137/S1064827597323415.
  14. Notay, Yvan (2000). "Flexible Conjugate Gradients". SIAM Journal on Scientific Computing. 22 (4): 1444–1460. CiteSeerX   10.1.1.35.7473 . doi:10.1137/S1064827599362314.
  15. Bouwmeester, Henricus; Dougherty, Andrew; Knyazev, Andrew V. (2015). "Nonsymmetric Preconditioning for Conjugate Gradient and Steepest Descent Methods 1". Procedia Computer Science. 51: 276–285. arXiv: 1212.6680 . doi: 10.1016/j.procs.2015.05.241 . S2CID   51978658.
  16. Knyazev, Andrew V.; Lashuk, Ilya (2008). "Steepest Descent and Conjugate Gradient Methods with Variable Preconditioning". SIAM Journal on Matrix Analysis and Applications. 29 (4): 1267. arXiv: math/0605767 . doi:10.1137/060675290. S2CID   17614913.
  17. 1 2 Ross, I. M., "An Optimal Control Theory for Accelerated Optimization," arXiv : 1902.09004, 2019.
  18. Nemirovsky and Ben-Tal (2023). "Optimization III: Convex Optimization" (PDF).
  19. Pennington, Fabian Pedregosa, Courtney Paquette, Tom Trogdon, Jeffrey. "Random Matrix Theory and Machine Learning Tutorial". random-matrix-learning.github.io. Retrieved 2023-12-05.{{cite web}}: CS1 maint: multiple names: authors list (link)

Further reading