1 Introduction

In this paper we study the behavior of quasi-Newton methods (QN) on an unconstrained quadratic problem of the form

$$\begin{aligned} {\mathop {\hbox {minimize}}\limits _{x \in \mathbb {R}^n}} \quad \frac{1}{2}x^T Hx+c^T x, \end{aligned}$$
(QP)

where \(H=H^T \succ 0\). Solving (QP) is equivalent to solving a symmetric system of linear equations \(Hx+c=0\). In particular, our concern is to give conditions under which a quasi-Newton method utilizing exact linesearch generates search directions that are parallel to those of the method of conjugate gradients (CG). As exact linesearch is considered, parallel search directions imply identical iterates. At iteration k, the x-iterate and the gradient \(Hx+c\) are denoted by \(x_k\) and \(g_k\) respectively. In a quasi-Newton method, the search direction \(p_k\) is computed from \(B_k p_k=-g_k\), where \(B_k\) is nonsingular.

We give necessary and sufficient conditions on a QN-method for this equivalence with CG on (QP). This is not the first time necessary and sufficient conditions are given. In [12, Theorem 2.2], a necessary and sufficient condition is given, which is based on projections from iterations 0, 1, ..., \(k-1\), allowing also the preconditioned setting to be considered. In contrast, we interpret the method of conjugate gradients as a particular quasi-Newton method and base the necessary and sufficient conditions on this observation. The result we give is thus directly based on the projection given by the method of conjugate gradients, i.e., based on quantities from iteration \(k-1\) and k involving one projection only.

If considering update matrices \(U_k\) defined by \(U_k=B_k-B_{k-1}\), it is well-known that, on (QP), QN using exact linesearch and an update scheme in the one-parameter Broyden family generates identical iterates to those generated by CG, see, e.g., [3, 11, 14]. The unique rank-1 update matrix in the Broyden family is usually referred to as the SR1 update matrix, and it is determined entirely by the so-called secant condition. As a result of our equivalence result, we show that the symmetric rank-1 update matrices that give parallel search directions to CG are given by the family of update matrices on the form

$$\begin{aligned} U_k = - \frac{1}{ (\gamma _k-1) p_{k-1}^T g_{k-1}} (\gamma _{k}g_k - g_{k-1})(\gamma _{k}g_k-g_{k-1})^T, \end{aligned}$$
(1)

where \(\gamma _k\) is a free parameter. The free parameter can be seen as a relaxation of the secant condition, as the SR1 update matrix is the only matrix in our parameterized rank-1 family which satisfies this condition. We show how to choose the parameter so that positive definiteness of the quasi-Newton matrix is preserved.

To simplify the exposition, we discuss equivalence to CG, which corresponds to the initial Hessian approximation being the identity matrix in the quasi-Newton method in our analysis. We then give the corresponding results in the preconditioned setting, which corresponds to an arbitrary positive definite and symmetric initial Hessian approximation. For the rank-1 case, the family of symmetric update matrices take the form (1) also in the preconditioned setting.

In Sect. 2, we make a brief introduction to CG and QN. In Sect. 3, we present our results which include necessary and sufficient conditions on QN such that CG and QN generate parallel search directions. These results are specialized to update matrices in Sect. 4. In particular, in Sect. 4.1, we give the results on symmetric rank-1 update matrices. Section 5 contains a discussion on how the results would apply if the inverse of the Hessian was updated instead of the Hessian itself. In Sect. 6, the corresponding results in the preconditioned setting are stated. Finally, in Sect. 7 we make some concluding remarks.

2 Background

For solving (QP), we consider linesearch methods on the following form. At iteration k, a search direction \(p_k\) is computed. The x-iterate and the gradient are updated as

$$\begin{aligned} x_{k+1}=x_k+\theta _k p_k, \quad g_{k+1}=g_k+\theta _k Hp_k,\; \text {for }\,\theta _k=-\frac{g_k^T p_k}{p_k^T Hp_k}. \end{aligned}$$

The choice of steplength \(\theta _k\) corresponds to exact linesearch, i.e., given a search direction \(p_k\) the steplength gives the exact minimizer along \(p_k\). This is a natural choice for (QP), as it can be done explicitly. For a given initial point \(x_0\), the iteration process is terminated at an iteration r if \(g_{r}=0\), in which case \(x_{r}\) is given as the optimal solution to (QP) or equivalently as the unique solution to \(Hx+c=0\). The method is summarized in Algorithm 1.

figure a

The particular linesearch method is defined by the way the search direction \(p_k\) is obtained in each iteration k. Our model method is the method of conjugate gradients, CG, by Hestenes and Stiefel [10]. There are different varieties of CG, which are equivalent on (QP). The variety we describe is referred to as the Fletcher-Reeves method of conjugate gradients, as stated in the following definition.

Definition 1

(The method of conjugate gradients (CG)) The method of conjugate gradients, CG, is the linesearch method of the form given by Algorithm 1 in which the search direction \(p_k\) is given by \(p_k^{CG}\), with

$$\begin{aligned} p_k^{CG}={\left\{ \begin{array}{ll} -g_0 &{} \text{ if } k=0, \\ -g_k+\frac{g_k^Tg_k}{g_{k-1}^Tg_{k-1}}p_{k-1}^{CG} &{} \text{ if } k\ge 1. \end{array}\right. } \end{aligned}$$
(CG)

For CG it holds that, for all k, \(g_k^Tg_i=0\), \(i=0, \ldots , k-1\), so the method terminates with \(g_r=0\) for some r, \(r\le n\), and \(x_r\) solves (QP). In addition, it holds that \(\{p_k^{CG}\}_{k=0}^{r-1}\) are mutually conjugate with respect to H. For an introduction to CG, see, e.g., [2, 15, 16]. In [5], CG is extended to general unconstrained problems. The reason for CG being our model method is that it requires one matrix-vector product \(Hp_k\) per iteration, and it terminates in r iterations, with \(r\le n\).

Next we define what we will refer to as a quasi-Newton method, QN.

Definition 2

(Quasi-Newton method (QN)) A quasi-Newton method, QN, is a linesearch method of the form given by Algorithm 1 in which the search direction \(p_k\) is given by

$$\begin{aligned} B_k p_k=-g_k, \end{aligned}$$
(QN)

where the matrix \(B_k\) is assumed nonsingular.

Quasi-Newton methods were first suggested by Davidon, see [1], and later modified and formalized by Fletcher and Powell, see [4]. For an introduction to QN-methods, see, e.g., [7, Chapter 4].

Our interest is now to set up conditions on \(B_k\) such that \(p_k\) and \(p_k^{CG}\) are parallel for all k, so that QN also terminates in r iterations. In [6], we derived such conditions based on a sufficient condition to obtain mutually conjugate search directions. Here, we give a direct necessary and sufficient condition based on \(p_k^{CG}\) only.

The results of the paper are derived with (CG) as the model method, which corresponds to \(B_0=I\) in (QN) giving \(p_0=p_0^{CG}\). It is also of interest to consider the case when a symmetric positive definite matrix M is given for which a preconditioned method of conjugate gradients is defined. This corresponds to \(B_0=M\) in (QN) giving the initial search directions identical. To simplify the exposition, we derive the results for the unpreconditioned case given by (CG) and give the corresponding results for the preconditioned setting in Sect. 6.

3 Necessary and sufficient conditions for QN

In this section we give precise conditions on \(B_k\) such that \(p_k\) is parallel to \(p_k^{CG}\). The main benefit of the conditions compared to previous work is that our result is based on the single iteration k. The dependence on the previous iterates is contained in the search direction \(p_{k-1}\), and there is no need to check any condition for all the previous iterates.

In the following proposition, we give a necessary and sufficient condition on \(B_k\) at a particular iteration k to give a search direction \(p_k\) such that \(p_k=\delta _k p_k^{CG}\) for a scalar \(\delta _k\). We assume that each previous search direction \(p_i\) has been parallel to the corresponding search direction of CG, \(p_i^{CG}\), so that QN and CG have generated the same iterate \(x_k\).

Proposition 1

Consider iteration k of the exact linesearch method of Algorithm 1, where \(1\le k < r\). Assume that \(p_i=\delta _i p_i^{CG}\) with \(\delta _i\ne 0\) for \(i=0,\ldots ,k-1\), where \(p_i^{CG}\), \(i=0,\ldots ,k-1\), are the search directions of the method of conjugate gradients, as stated in Definition 1. Let \(A_k\) be defined as

$$\begin{aligned} A_k = I-\frac{1}{g_{k-1}^T p_{k-1}} p_{k-1} g_k^T. \end{aligned}$$
(2)

Then,

$$\begin{aligned} A_k^{-1}= I+\frac{1}{g_{k-1}^T p_{k-1}} p_{k-1} g_k^T, \end{aligned}$$
(3)

and it holds that \(A_k p_k^{CG}=- g_k\). In addition, if \(p_k\) is given by \(B_k p_k=-g_k\) with \(B_k\) nonsingular, then, for any nonzero scalar \(\delta _k\), it holds that \(p_k=\delta _k p_k^{CG}\) if and only if

$$\begin{aligned} B_k A_k^{-1}g_k = \frac{1}{\delta _k} g_k, \end{aligned}$$
(4)

or equivalently if and only if

$$\begin{aligned} B_k = A_k^T W_k A_k, \text {with }W_k g_k = \frac{1}{\delta _k} g_k, \text {for } W_k \text { nonsingular.} \end{aligned}$$
(5)

Finally, it holds that \(B_k\succ 0\) if and only if \(W_k\succ 0\).

Proof

We have

$$\begin{aligned} p_k^{CG} =- g_k +\frac{g_k^T g_k}{g_{k-1}^Tg_{k-1}}p_{k-1}^{CG} = -\left( I+\frac{1}{g_{k-1}^T p_{k-1}^{CG}}p_{k-1}^{CG} g_k^T\right) g_k. \end{aligned}$$
(6)

Therefore, since \(p_{k-1}=\delta _{k-1}p_{k-1}^{CG}\), with \(\delta _{k-1}\ne 0\), (6) gives

$$\begin{aligned} p_k^{CG} = -\left( I+\frac{1}{g_{k-1}^T p_{k-1}}p_{k-1} g_k^T\right) g_k= -A_k^{-1}g_k, \end{aligned}$$
(7)

with \(A_k^{-1}\) given by (3). Since \(g_k^T p_{k-1}=0\), multiplication of \(A_k\) of (2) by \(A_k^{-1}\) of (3) gives \(A_k A_k^{-1}=I\), so that the stated \(A_k\) is nonsingular with corresponding inverse \(A_k^{-1}\). Therefore, (7) gives \(A_k p_k^{CG}=-g_k\).

Since \(B_k\) is assumed nonsingular and \(\delta _k\ne 0\), it holds that \(p_k=\delta _k p_k^{CG}\) if and only if \(B_k (-\delta _k A_k^{-1}g_k) = -g_k\), which is equivalent to (4). Since \(g_k^T p_{k-1}=0\), we obtain \(A_k^{-T}\!g_k=g_k\), so that (4) is equivalent to

$$\begin{aligned} A_k^{-T}\!B_k A_k^{-1}g_k=\frac{1}{\delta _k} g_k, \end{aligned}$$

which in turn is equivalent to (5). The final result on positive definiteness follows from the nonsingularity of \(A_k\) by Sylvester’s law of inertia, see, e.g., [8, Theorem 8.1.17]. \(\square \)

The necessary and sufficient conditions of Proposition 1 give a straightforward way to check if a matrix \(B_k\) is such that the corresponding QN-method and CG will generate parallel search directions. The scaling of \(p_k^{CG}\) has a special role in our analysis and we relate \(p_k\) to \(p_k^{CG}\) by a scalar \(\delta _k\). The observation that \(p_k^{CG}\) may be written as \(p_k^{CG}=-A_k^{-1}g_k\) for a nonsingular \(A_k\) has been made in [12, Example 2], but the equivalence result of [12] concerns \(p_k\) without relating to the scaling of \(p_k^{CG}\) explicitly. Therefore, the condition of [12] involves projections on all previous iterations \(0,1,\ldots ,k-1\), not one single projection as we obtain. In addition, since there is no relationship to a particular scaling, there is no parameter corresponding to our \(\delta _k\). Since such a parameter is vital for deriving later results in our paper, in particular when characterizing symmetric rank-one updates, we cannot apply the equivalence result of [12] directly. A difference in [12] is that they consider matrices \(N_k\) that approximate \(H^{-1}\) rather than matrices \(B_k\) that approximate H. This is not a major difference, we discuss these issues in Sect. 5.

Note that it is not necessary to make \(B_k-I\) increase in rank. In particular, \(B_k = A_k^T A_k\), corresponding to \(W_k=I\) in Proposition 1, is a positive-definite symmetric matrix for which \(B_k p_k=-g_k\) gives \(p_k=p_k^{CG}\).

We also note that the characterization of \(B_k\) does only depend on information from iteration k and \(k-1\), since it directly inherits the properties of the method of conjugate gradients. In addition, the characterization of Proposition 1 only depends directly on quantities computed by the quasi-Newton method, the scaling of the method of conjugate gradients is not needed.

4 Results on update matrices

In the previous section we gave results on \(B_k\) for a particular iteration k without directly relating to any other \(B_i\), \(i\ne k\). It is often the case that \(B_k\) is defined in terms of the previous matrix \(B_{k-1}\) and an update matrix \(U_k\) such that \(B_k=B_{k-1}+U_k\), and that conditions are put on \(U_k\). We have in mind a setting where information from the generated gradients is used, so that \(B_{k-1}\) may be expressed as \(B_{k-1}=I+V_k\), with \(\mathcal {R}(V_k) \subseteq span\{g_0, \ldots ,g_{k-1}\}\). As we have in mind such a setting where in addition \(B_k\) is symmetric, we make the assumption \(B_{k-1}g_k=g_k\).

Proposition 1 can then be applied in a straightforward manner to give conditions on \(U_k\) such that \(p_k=\delta _k p_k^{CG}\). Note that there is a one-to-one correspondence between \(U_k\) and \(B_k\) given \(B_{k-1}\).

Proposition 2

Consider iteration k of the exact linesearch method of Algorithm 1, where \(1\le k < r\). Assume that \(p_i=\delta _i p_i^{CG}\) with \(\delta _i\ne 0\) for \(i=0,\ldots ,k-1\), where \(p_i^{CG}\), \(i=0,\ldots ,k-1\), are the search directions of the method of conjugate gradients, as stated in Definition 1. Let \(B_{k-1}\) be a nonsingular matrix such that \(B_{k-1} p_{k-1}=-g_{k-1}\) and \(B_{k-1}g_k = g_k\). Let \(U_{k}=B_k - B_{k-1}\) and assume that \(B_k\) and \(p_k\) satisfy \(B_kp_k=-g_k\), with \(B_k\) nonsingular. Then, for any nonzero scalar \(\delta _k\), it holds that \(p_k=\delta _k p_k^{CG}\) if and only if

$$\begin{aligned} U_{k} \left( g_k+\frac{g_k^Tg_k}{p_{k-1}^T g_{k-1}}p_{k-1} \right) = \left( \frac{1}{\delta _k} -1\right) g_k + \frac{ g_k^T g_k}{p_{k-1}^T g_{k-1}} g_{k-1}. \end{aligned}$$
(8)

Proof

By assumption, \(B_k\) is nonsingular so for \(B_k=B_{k-1}+U_k\), Proposition 1 gives \(p_k=\delta _k p_k^{CG}\) if and only if

$$\begin{aligned} U_k \left( g_k+\frac{g_k^Tg_k}{p_{k-1}^T g_{k-1}}p_{k-1} \right)&=\frac{1}{\delta _k}g_k-B_{k-1}\left( g_k+\frac{g_k^Tg_k}{p_{k-1}^T g_{k-1}}p_{k-1} \right) \\&=\frac{1}{\delta _k}g_k- g_k+\frac{g_k^Tg_k}{p_{k-1}^T g_{k-1}}g_{k-1}, \end{aligned}$$

since \(p_{k-1}\) is computed from \(B_{k-1}p_{k-1}=-g_{k-1}\) and it is assumed that \(B_{k-1} g_k = g_k\), so the statement of the proposition follows. \(\square \)

Note that in the right-hand side of (8) in Proposition 2, the component along \(g_{k-1}\) is nonzero and independent of \(\delta _k\). The component along \(g_k\) is zero for \(\delta _k=1\), i.e., when \(p_k=p_k^{CG}\).

4.1 Results on symmetric rank-one update matrices

Next we consider the case when \(U_k\) is a symmetric matrix of rank one. It is well known that the secant condition gives a unique update referred to as SR1, see, e.g., [13, Chapter 9]. The secant condition and SR1 will be discussed later in this section. Using Proposition 2 we can give a different result concerning the case when the update matrix \(U_k\) is a symmetric matrix of rank one. In particular, we show that the family of rank-1 update matrices can be parameterized by a free parameter and that the matrix is unique for a fixed value of the parameter. This parametrization allows positive definiteness of the quasi-Newton matrix to be preserved.

The situation can be considered in two ways. First, for any given value of the scalar \(\gamma _k\), except for three distinct values, there is a symmetric rank-1 update matrix \(U_k\) of the form

$$\begin{aligned} U_k = - \frac{1}{ (\gamma _k-1) p_{k-1}^T g_{k-1}} (\gamma _{k}g_k - g_{k-1})(\gamma _{k}g_k-g_{k-1})^T, \end{aligned}$$
(9)

for which \(p_k=\delta _k(\gamma _k) p_k^{CG}\), where \(\delta _k(\cdot )\) is a real-valued function. Second, if \(p_k=\delta _k p_k^{CG}\) is required for any given value of the scalar \(\delta _k\), except for three distinct values, and \(U_k\) is symmetric and of rank one, \(U_k\) must take the form (9), with \(\gamma _k=\gamma _k(\delta _k)\), where \(\gamma _k(\cdot )\) is the inverse function of \(\delta _k(\cdot )\). Consequently, except for three distinct values, there is a one-to-one correspondence between \(\delta _k\) such that \(p_k=\delta _k p_k^{CG}\) and \(\gamma _k\) of the symmetric rank-1 update matrix \(U_k\) of (9).

The functions \(\delta _k(\cdot )\) and \(\gamma _k(\cdot )\) are defined in the following lemma.

Lemma 1

Consider iteration k of the exact linesearch method of Algorithm 1, where \(1\le k < r\). Assume that \(p_i=\delta _i p_i^{CG}\) with \(\delta _i\ne 0\) for \(i=0,\ldots ,k-1\), where \(p_i^{CG}\), \(i=0,\ldots ,k-1\), are the search directions of the method of conjugate gradients, as stated in Definition 1. Let \(\hat{\gamma }_k={p_{k-1}^T g_{k-1}}/{g_k^T g_k}\). For \(\delta _k\ne 0\) and \(\gamma _k\ne \hat{\gamma }_k\), let the functions \(\gamma _k(\delta _k)\) and \(\delta _k(\gamma _k)\) be defined by

$$\begin{aligned} \gamma _k(\delta _k) =- \frac{p_{k-1}^T g_{k-1}}{g_k^T g_k} \left( \frac{1}{\delta _k}-1\right) , \quad \delta _k(\gamma _k) =\frac{1}{1 - \gamma _k\frac{g_k^T g_k}{p_{k-1}^T g_{k-1}}}. \end{aligned}$$

Then, the functions \(\gamma _k(\cdot )\) and \(\delta _k(\cdot )\) are inverses to each other.

We now characterize the symmetric rank-one update matrices that give search directions which are parallel to those of the method of conjugate gradients. In addition, we give conditions for preserving positive definiteness and a hereditary result.

Proposition 3

Consider iteration k of the exact linesearch method of Algorithm 1, where \(1\le k < r\). Assume that \(p_i=\delta _i p_i^{CG}\) with \(\delta _i\ne 0\) for \(i=0,\ldots ,k-1\), where \(p_i^{CG}\), \(i=0,\ldots ,k-1\), are the search directions of the method of conjugate gradients, as stated in Definition 1. Let \(B_k\) and \(p_k\) satisfy \(B_kp_k=-g_k\), and let \(B_{k-1}\) be a nonsingular matrix such that \(B_{k-1} p_{k-1}=-g_{k-1}\) and \(B_{k-1}g_k = g_k\). In addition, let \(\gamma _k(\cdot )\), \(\delta _k(\cdot )\) and \(\hat{\gamma }_k\) be given by Lemma 1.

For any scalar \(\gamma _k\), except \(\gamma _k=0\), \(\gamma _k=\hat{\gamma }_k\) and \(\gamma _k=1\), let \(B_k\) be defined by

$$\begin{aligned} B_{k} = B_{k-1} - \frac{1}{ (\gamma _k-1) p_{k-1}^T g_{k-1}} (\gamma _{k}g_k - g_{k-1})(\gamma _{k}g_k-g_{k-1})^T. \end{aligned}$$
(10)

Then, \(B_k\) is nonsingular and \(p_k=\delta _k p_k^{CG}\) for \(\delta _k=\delta _k(\gamma _k)\).

Conversely, for any scalar \(\delta _k\), except \(\delta _k=0\), \(\delta _k=\delta _k(1)\) and \(\delta _k=1\), assume that \(p_k=\delta _k p_k^{CG}\) and assume that \(B_k-B_{k-1}\) is symmetric and of rank one. Then, \(B_k\) is a nonsingular matrix given by (10) for \(\gamma _k=\gamma _k(\delta _k)\).

If, in addition, \(B_{k-1}=B_{k-1}^T\succ 0\), then \(B_k\) defined by (10) satisfies \(B_k\succ 0\) if and only if \(\gamma _k>1\) or \(\hat{\gamma }_k<\gamma _k<0\), or equivalently if and only if \(\gamma _k=\gamma _k(\delta _k)\) for \(0<\delta _k<\delta _k(1)\) or \(\delta _k>1\).

Finally, if \(B_i p_i=-g_i\), \(i=0,\ldots ,k\), with \(B_0=I\) and if, for \(i=1,\ldots ,k\), \(B_{i-1}\) is updated to \(B_i\) according to (10) for \(\gamma _i\) such that \(\gamma _i\ne 0\), \(\gamma _i\ne \hat{\gamma }_i\) and \(\gamma _i\ne 1\), then

$$\begin{aligned} B_k p_i = \frac{\gamma _{i+1}\theta _i}{\gamma _{i+1}-1}H p_i, \quad i=0,\ldots ,k-1. \end{aligned}$$
(11)

Proof

Let \(U_k=B_k-B_{k-1}\). If \(U_k\) is symmetric and of rank one, we may write \(U_k=\beta _k u_k u_k^T\), where \(\beta _k\) is a scalar and \(u_k\) is a vector in \(\mathbb {R}^n\), both to be determined. If \(B_k\) is nonsingular and \(\delta _k\ne 0\), Proposition 2 shows that \(p_k=\delta _k p_k^{CG}\) if and only if

$$\begin{aligned} \beta _k u_k u_k^TA_k^{-1} g_k = \left( \frac{1}{\delta _k} -1 \right) g_k + \frac{ g_k^T g_k}{p_{k-1}^T g_{k-1}} g_{k-1}= - \frac{g_k^T g_k}{p_{k-1}^T g_{k-1}}(\gamma _{k} g_k-g_{k-1}), \end{aligned}$$
(12)

with \(\gamma _k=\gamma _k(\delta _k)\) given by Lemma 1. Throughout the proof, assume that \(\delta _k\not \in \{0,\delta _k(1),1\}\) and \(\gamma _k\not \in \{0,\hat{\gamma }_k,1\}\), which is assumed in the statement of the Proposition. Then, Lemma 1 shows that there is a one-to-one correspondence between \(\delta _k\) and \(\gamma _k\). Hence, (12) may be considered for either \(\delta _k\) or \(\gamma _k\). We choose \(\gamma _k\) for ease of notation.

We first assume that \(B_k\) is nonsingular, and verify that this is the case later in the proof. For \(B_k\) nonsingular, it follows from (12) that \(u_k\) will be equal to the right-hand side vector up to some arbitrary non-zero scaling. Let

$$\begin{aligned} u_k= \gamma _kg_k-g_{k-1}. \end{aligned}$$
(13)

The scaling of \(u_k\) will be reflected in \(\beta _k\) by insertion into (12) as

$$\begin{aligned} \beta _k (\gamma _kg_k-g_{k-1})^T\left( g_k + \frac{g_k^T g_k}{p_{k-1}^T g_{k-1}}p_{k-1}\right) = \beta _k (\gamma _k-1) g_k^T g_k = - \frac{g_k^T g_k}{p_{k-1}^T g_{k-1}}, \end{aligned}$$

so that

$$\begin{aligned} \beta _k=-\frac{1}{ (\gamma _k-1) p_{k-1}^T g_{k-1}}. \end{aligned}$$
(14)

Note that (14) is well defined as \(\gamma _k\ne 1\) is assumed. A combination of (12), (13) and (14) gives \(B_k\) expressed as in (10).

It remains to show that \(B_k\) is nonsingular. It follows from (10) that

$$\begin{aligned} B_{k} = B_{k-1} \left( I - \frac{1}{ (\gamma _k-1) p_{k-1}^T g_{k-1}} (\gamma _kB_{k-1}^{-1}g_k - B_{k-1}^{-1}g_{k-1})(\gamma _kg_k-g_{k-1})^T\right) , \end{aligned}$$

so that

$$\begin{aligned} \det (B_k)=\det (B_{k-1})\eta _k, \end{aligned}$$
(15)

with

$$\begin{aligned} \eta _k&= 1 - \frac{(\gamma _kB_{k-1}^{-1}g_k - B_{k-1}^{-1}g_{k-1})^T(\gamma _kg_k-g_{k-1})}{(\gamma _k-1) p_{k-1}^T g_{k-1}} \nonumber \\&= 1 - \frac{\gamma _k^2 g_k^T g_k - p_{k-1}^T g_{k-1}}{(\gamma _k-1) p_{k-1}^T g_{k-1}} = \frac{-\gamma _k ( \gamma _k g_k^T g_k - p_{k-1}^T g_{k-1} )}{(\gamma _k-1) p_{k-1}^T g_{k-1}} \nonumber \\&= \frac{-\gamma _k ( \gamma _k - \hat{\gamma }_k )}{(\gamma _k-1) \hat{\gamma }_k}, \end{aligned}$$
(16)

since \(B_{k-1}g_k=g_k\), \(B_{k-1}p_{k-1}=-g_{k-1}\) and \(g_k^T p_{k-1}=0\), with \(\hat{\gamma }_k\) given by Lemma 1. Hence, since \(B_{k-1}\) is assumed nonsingular, a combination of (15) and (16) shows that nonsingularity of \(B_k\) is equivalent to \(\eta _k\ne 0\), i.e., \(\gamma _k\ne 0\) and \(\gamma _k\ne \hat{\gamma }_k\), which is exactly what is assumed.

To prove the result on positive definiteness, assume that \(B_{k-1}=B_{k-1}^T\succ 0\). In this case, since \(B_{k-1}\) and \(B_k\) differ by a symmetric rank-1 matrix, \(B_k\) can have at most one nonpositive eigenvalue, see, e.g., [8, Theorem 8.1.8]. Therefore, (15) shows that positive definiteness of \(B_k\) is equivalent to \(\eta _k>0\). Note that \(B_{k-1}\succ 0\) implies \(p_{k-1}^T g_{k-1}=-p_{k-1}^T B_{k-1}^{-1}p_{k-1}<0\), which in turn gives \(\hat{\gamma }_k<0\). We may now examine (16) to see what values of \(\gamma _k\) that give \(\eta _k>0\). The numerator of (16) is positive for \(\hat{\gamma }_k<\gamma _k<0\) and negative for \(\gamma _k<\hat{\gamma }_k\) and \(\gamma _k>0\). The denominator of (16) is positive for \(\gamma _k<1\) and negative for \(\gamma _k>1\). We conclude that \(\eta _k>0\) if and only if \(\hat{\gamma }_k<\gamma _k<0\) or \(\gamma _k>1\), which by Lemma 1 is equivalent to \(0<\delta _k<\delta _k(1)\) or \(\delta _k>1\).

To prove the final hereditary result, assume that \(B_i p_i=-g_i\), \(i=0,\ldots ,k\), with \(B_0=I\) and assume that \(B_{i-1}\) is updated to \(B_i\) according to (10) for \(\gamma _i\) such that \(\gamma _i\ne 0\), \(\gamma _i\ne \hat{\gamma }_i\) and \(\gamma _i\ne 1\). Then, for a given i, \(0< i < k\), k may be replaced by \(i+1\) in (10), which gives

$$\begin{aligned} B_{i+1}p_i= & {} B_{i}p_i - \frac{1}{ (\gamma _{i+1}-1) p_{i}^T g_{i}} (\gamma _{i+1}g_{i+1} - g_{i})(\gamma _{i+1}g_{i+1}-g_{i})^Tp_i \nonumber \\= & {} -g_i + \frac{1}{\gamma _{i+1}-1}(\gamma _{i+1}g_{i+1} - g_{i}) = \frac{\gamma _{i+1}}{\gamma _{i+1}-1}(g_{i+1} - g_{i}) \nonumber \\= & {} \frac{\gamma _{i+1}\theta _i}{\gamma _{i+1}-1} H p_i, \end{aligned}$$
(17)

where the identities \(B_i p_i = - g_i\), \(g_{i+1}^T p_i=0\) and \(g_{i+1}-g_i=\theta _i H p_i\) have been used. Finally, (10) gives \(B_j p_i = B_{i+1} p_i\) for \(j=i+2,\ldots ,k\), since \(g_j^T p_i=0\) for \(j\ge i+1\). Consequently, \(B_k p_i = B_{i+1}p_i\), with \(B_{i+1} p_i\) given by (17), proving (11). \(\square \)

Note that there are two ways in which positive definiteness of a symmetric \(B_{k-1}\) may be preserved in a symmetric rank-one update. The first one, \(\gamma _k>1\), or equivalently \(0<\delta _k<\delta _k(1)\), is straightforward, since it corresponds to \(U_k\succeq 0\). The second one, \(\hat{\gamma }_k<\gamma _k<0\), or equivalently \(\delta _k>1\), is less straightforward. The corresponding \(U_k\) is negative semidefinite, but still the resulting \(B_k\) is positive definite.

Proposition 3 gives precise conditions for which rank-one matrices that give a corresponding update matrix that preserves positive definiteness and gives search directions parallel to the method of conjugate gradients. We have the freedom to choose \(\gamma _k\) or \(\delta _{k}\) appropriately. This can be compared to SR1, the symmetric rank-one update scheme uniquely defined by the secant condition

$$\begin{aligned} B_k s_{k-1}=y_k, \text {for } s_{k-1}=\theta _{k-1} p_{k-1}, \quad y_k=g_k-g_{k-1}. \end{aligned}$$
(18)

By writing \(U_k=B_k-B_{k-1}\), the secant condition gives a requirement on \(U_k\) as

$$\begin{aligned} U_{k} s_{k-1}= y_k - B_{k-1}s_{k-1}, \end{aligned}$$
(19)

which for \(U_k\) symmetric and of rank one gives the SR1 update matrix \(U_k^{SR1}\) on the form

$$\begin{aligned} U_k^{SR1} = \frac{1}{s_{k-1}^T (y_k-B_ks_{k-1})}(y_k-B_ks_{k-1})(y_k-B_ks_{k-1})^T, \end{aligned}$$
(20)

see, e.g., [13, Chapter 9]. Since \(B_{k-1}p_{k-1}=-g_{k-1}\) holds by the definition of the quasi-Newton method, we may use the definitions of \(s_{k-1}\) and \(y_k\) of (18) to rewrite \(U_k^{SR1}\) of (20) as

$$\begin{aligned} U_k^{SR1} =\frac{1}{\theta _{k-1} p_{k-1}^T ( g_k - (1-\theta _{k-1}) g_{k-1}} (g_k - (1-\theta _{k-1}) g_{k-1})(g_k - (1-\theta _{k-1}) g_{k-1})^T. \end{aligned}$$

Since exact linesearch is performed in our case, it holds that \(p_{k-1}^T g_k=0\), so that \(U_k^{SR1}\) takes the form

$$\begin{aligned} U_k^{SR1}&=\frac{-1}{\theta _{k-1} (1-\theta _{k-1}) p_{k-1}^Tg_{k-1}} (g_k - (1-\theta _{k-1}) g_{k-1})(g_k - (1-\theta _{k-1}) g_{k-1})^T \nonumber \\&= \frac{-(1-\theta _{k-1})}{\theta _{k-1} p_{k-1}^T g_{k-1}} \left( \frac{1}{1-\theta _{k-1}}g_k - g_{k-1}\right) \left( \frac{1}{1-\theta _{k-1}}g_k-g_{k-1}\right) ^T, \end{aligned}$$
(21)

where in the last step, a scaling of the rank-1 vector by a factor \(1/(1-\theta _{k-1})\) has been made. A comparison of (10) and (21) shows that the SR1 update is the particular member of the family of symmetric rank-1 updates given by Proposition 3 for which \(\gamma _k=1/(1-\theta _{k-1})\). In particular, for \(\theta _{k-1}=1\), SR1 is not well defined. In addition, as there is no freedom in choosing the rank-one matrix for SR1, there is no way to ensure \(B_k \succ 0\) even if \(B_{k-1}=B_{k-1}^T\succ 0\). Note that the condition on \(B_k\) of (18) giving a condition on \(U_k\) of (19) and a unique symmetric rank-1 \(U_k\) of (20) is analogous to our condition on \(B_k\) of Proposition 1 for a fixed \(\delta _k\) giving a condition on \(U_k\) of Proposition 2 and a unique rank-1 \(U_k\) of Proposition 3.

Example 1 illustrates the SR1 update and another rank-1 update of Proposition 3 which preserves positive definiteness. The H and c of the example are parameterized by a positive scalar \(\phi \). We obtain \(\theta _0=2/(3\phi )\), so by selecting \(\phi =2/3\), it follows that \(\theta _0=1\) and the SR1 update becomes undefined. By selecting \(\phi \) slightly smaller than 2/3, for example 0.65, we obtain \(\theta _0\) slightly larger than one (\(\theta _0=40/39\)), so that \(\gamma _1=-39\) and the corresponding \(\delta _1\) is negative (\(\delta _1=-3/10\)). Consequently, \(B_1^{SR1}\) is indefinite and the corresponding \(p_1\) is an ascent direction. For comparison, the rank-1 update of Proposition 3 is given for \(\delta _1=2\), which preserves positive definiteness. As can be seen from (10), the rank-1 update of Proposition 3 is independent of \(\phi \).

Example 1

For a positive parameter \(\phi \), consider the example

$$\begin{aligned} H= \phi \left( \! \begin{array}{cc}2 &{} \ 0 \\ 0 &{} \ 1\end{array}\! \right) , \quad c= \phi \left( \! \begin{array}{c}-1 \\ -1\end{array}\! \right) , \quad B_0= \left( \! \begin{array}{cc}1 &{} \ 0 \\ 0 &{} \ 1\end{array}\! \right) , \quad x_0=\left( \! \begin{array}{c}0 \\ 0\end{array}\! \right) \end{aligned}$$

for which

$$\begin{aligned} \quad x_1=\left( \! \begin{array}{r}\frac{2}{3} \\ \frac{2}{3}\end{array}\! \right) , \quad x_2=\left( \! \begin{array}{r}\frac{1}{2} \\ 1\end{array}\! \right) . \end{aligned}$$

Then

$$\begin{aligned} \phi&=\frac{2}{3}&\implies&B_1^{SR1}&\ \text{ undefined }&B_1^{\delta _1=2}&= \frac{1}{44}\left( \! \begin{array}{rr} 43 &{} 5 \\ 5 &{} \ 19 \end{array}\! \right) \\ \phi&=\frac{65}{100}&\implies&B_1^{SR1}&=\frac{1}{20} \left( \! \begin{array}{rr} -16 &{} 42 \\ 42 &{} \ -29 \end{array}\! \right)&B_1^{\delta _1=2}&= \frac{1}{44}\left( \! \begin{array}{rr} 43 &{} 5 \\ 5 &{} \ 19 \end{array}\! \right) \end{aligned}$$

Note that the numerical values and the dimension of Example 1 are not important. For a given quadratic problem, there will always exist a particular positive scaling such that the resulting \(B_1^{SR1}\) is undefined.

5 On the approximation of the Hessian

The results of the present manuscript have been written based on the search directions of the method of conjugate gradients. The reason for doing so is that it allows a direct treatment of \(B_k\), and there is no need to focus on the update matrix \(B_k-B_{k-1}\). This is the choice of the authors, but other choices are of course possible.

The results are stated for a matrix \(B_k\) that approximates the Hessian H. We prefer to think of the quasi-Newton method in this way, but there would be little difference if one instead stated the results for a matrix \(N_k\) that approximates \(H^{-1}\), which is done for example in [12]. The search direction \(p_k\) would then be defined by \(p_k=-N_k g_k\) rather than by \(B_k p_k=-g_k\) and conditions would be imposed on \(N_k\) rather than on \(B_k\). Proposition 1 could be equivalently stated using \(N_k\) as the approximation of \(H^{-1}\). Then, the counterparts of (4) and (5) would read

$$\begin{aligned} N_k g_k = \delta _k A_k^{-1}g_k \text { and } A_k N_k A_k^T g_k = \delta _k g_k. \end{aligned}$$

When considering update matrices, with \(V_k=N_k-N_{k-1}\), the counterpart of the update formula (8) of Proposition 2 would read

$$\begin{aligned} V_k g_k&=\delta _k A_k^{-1}g_k - N_{k-1} g_k =(\delta _k-1) g_k + \delta _k \frac{g_k^T g_k}{g_{k-1}^T p_{k-1}}p_{k-1}. \end{aligned}$$
(22)

For the rank-one case, the update matrix is unique for a given \(\delta _k\), and (22) gives

$$\begin{aligned} N_k = N_{k-1} -\frac{1}{\gamma _k ( \gamma _k g_k^T g_k - p_{k-1}^T g_{k-1} )} (\gamma _k g_k + p_{k-1})(\gamma _kg_k + p_{k-1})^T, \end{aligned}$$
(23)

with \(\gamma _k=\gamma _k(\delta _k)\) of Lemma 1. The uniqueness of the update implies that if \(N_{k-1}=B_{k-1}^{-1}\), then \(N_{k}=B_{k}^{-1}\) and (23) follows from (10) by the Sherman–Morrison formula.

As for the rank-one case, in light of the results of Sect. 4.1, one of the referees has pointed out that the parametrization given by \(\delta _k\) can be replaced by a different parametrization. The secant condition \(\theta _{k-1} B_k p_{k-1} = g_k-g_{k-1}\) and its counterpart on the inverse \(\theta _{k-1} p_{k-1} = N_k (g_k-g_{k-1})\) may be relaxed by a parameter \(\lambda _k\) so that \(\lambda _k\theta _{k-1} B_k p_{k-1} = g_k-g_{k-1}\) and \(\lambda _k\theta _{k-1} p_{k-1} = N_k (g_k-g_{k-1})\) respectively. For the updates, we obtain

$$\begin{aligned} \lambda _k \theta _{k-1} U_k p_k&= g_k -(1-\lambda _k\theta _{k-1})g_{k-1} \text { and} \end{aligned}$$
(24a)
$$\begin{aligned} V_k (g_k-g_{k-1})&= -g_k-(1-\lambda _k\theta _{k-1})p_{k-1}, \end{aligned}$$
(24b)

since \(B_{k-1}p_{k-1}=-g_{k-1}\), \(N_{k-1} g_k=g_k\) and \(N_{k-1} g_{k-1}=-p_{k-1}\). If having read the previous sections of this paper, we would see that \(p_k=\delta _k p_k^{CG}\), where we can relate \(\gamma _k\) to \(\lambda _k\) by \(\gamma _k = 1/(1-\lambda _k\theta _{k-1})\), by comparing the right-hand side vector of (24a) to the rank-one vector of Proposition 3 or comparing the right-hand side vector of (24b) to the rank-one vector of (23). The corresponding relationship to \(\delta _k\) is given by Lemma 1. An alternative to reading the previous sections of this paper, however, would be to say that \(\lambda _k=1\) corresponds to the SR1 update, and \(\lambda _k=0\) corresponds to the conjugate projection update [3, Eq. (4.1.10)]. They are considered in the update of the inverse and are both known to give \(p_k\) parallel to \(p_k^{CG}\) if \(N_0=I\). By replacing \(\gamma _k\) by \(1/(1-\lambda _k\theta _{k-1})\), one could show that a rank-1 matrix of the form (9) would give \(p_k\) parallel to \(p_k^{CG}\) using induction similar to what is done in [3, Theorem 3.4.1] and give conditions on preserving positive definiteness on \(\lambda _k\). This would, however, not show that there is no other family of rank-1 updates giving \(p_k\) parallel to \(p_k^{CG}\). We prefer to give a direct proof based on our result of Proposition 1, as we from there get both necessary and sufficient conditions.

6 Preconditioning

Our results have been derived in the setting of CG, which corresponds to \(B_0=I\) in QN giving the initial search directions identical. In this section, we give the analogous results in a preconditioned setting. In the preconditioned method of conjugate gradients, there is a positive definite symmetric matrix M, providing an estimate of H. For the quasi-Newton method, this will correspond to \(B_0=M\) giving the initial search directions identical.

The preconditioned method of conjugate gradient takes the following form. If the Cholesky factor of M is denoted by L, so that \(M=LL^T\), then the method of conjugate gradients is applied to

(25)

for , see, e.g., [15, Chapter 9.2]. Letting “hat” be associated with quantities of (25), we obtain and . Since is associated with a “usual” unpreconditioned system, we write , and since p is associated with a preconditioned system, we write \(p^{PCG}\), so that . It is straightforward to use these relations to derive the result analogous to those given in the the previous sections also for the preconditioned system.

Definition 3

(The preconditioned method of conjugate gradients (PCG)) For a positive definite symmetric \(n\times n\) matrix M, the preconditioned method of conjugate gradients, PCG, is the linesearch method of the form given by Algorithm 1 in which the search direction \(p_k\) is given by \(p_k^{PCG}\), with

$$\begin{aligned} p_k^{PCG}={\left\{ \begin{array}{ll} -M^{-1}g_0 &{} \text{ if } k=0, \\ -M^{-1}g_k+\frac{g_k^TM^{-1}g_k}{g_{k-1}^TM^{-1}g_{k-1}} p_{k-1}^{PCG} &{} \text{ if } k\ge 1. \end{array}\right. } \end{aligned}$$
(PCG)

For PCG it holds that, for all k, \(g_k^T M^{-1}g_i=0\), \(i=0, \ldots , k-1\), so the method terminates with \(g_r=0\) for some r, \(r\le n\), and \(x_r\) solves (QP). In addition, it holds that \(\{p_k^{PCG}\}_{k=0}^{r-1}\) are mutually conjugate with respect to H.

Proposition 4

Consider iteration k of the exact linesearch method of Algorithm 1, where \(1\le k < r\). Assume that \(p_i=\delta _i p_i^{PCG}\) with \(\delta _i\ne 0\) for \(i=0,\ldots ,k-1\), where \(p_i^{PCG}\), \(i=0,\ldots ,k-1\), are the search directions of the preconditioned method of conjugate gradients, as stated in Definition 3. Let \(A_k\) be defined as

$$\begin{aligned} A_k = I-\frac{1}{g_{k-1}^T p_{k-1}} p_{k-1} g_k^T. \end{aligned}$$

Then,

$$\begin{aligned} A_k^{-1}= I+\frac{1}{g_{k-1}^T p_{k-1}} p_{k-1} g_k^T, \end{aligned}$$

and it holds that \(M A_k p_k^{PCG}=- g_k\). In addition, if \(p_k\) is given by \(B_k p_k=-g_k\) with \(B_k\) nonsingular, then, for any nonzero scalar \(\delta _k\), it holds that \(p_k=\delta _k p_k^{PCG}\) if and only if

$$\begin{aligned} B_k A_k^{-1}M^{-1}g_k = \frac{1}{\delta _k} g_k, \end{aligned}$$

or equivalently if and only if

$$\begin{aligned} B_k = A_k^T W_k A_k, \text { with } W_k M^{-1}g_k = \frac{1}{\delta _k} g_k, \text { for } W_k \text { nonsingular.} \end{aligned}$$

Finally, it holds that \(B_k\succ 0\) if and only if \(W_k\succ 0\).

In particular, \(B_k = A_k^T M A_k\), corresponding to \(W_k=M\) in Proposition 4, is a positive-definite symmetric matrix for which \(B_k p_k=-g_k\) gives \(p_k=p_k^{PCG}\).

Proposition 5

Consider iteration k of the exact linesearch method of Algorithm 1, where \(1\le k < r\). Assume that \(p_i=\delta _i p_i^{PCG}\) with \(\delta _i\ne 0\) for \(i=0,\ldots ,k-1\), where \(p_i^{PCG}\), \(i=0,\ldots ,k-1\), are the search directions of the preconditioned method of conjugate gradients using a positive definite symmetric preconditioning matrix M, as stated in Definition 3. Let \(B_{k-1}\) be a nonsingular matrix such that \(B_{k-1} p_{k-1}=-g_{k-1}\) and \(B_{k-1}M^{-1}g_k = g_k\). Let \(U_{k}=B_k - B_{k-1}\) and assume that \(B_k\) and \(p_k\) satisfy \(B_kp_k=-g_k\), with \(B_k\) nonsingular. Then, for any nonzero scalar \(\delta _k\), it holds that \(p_k=\delta _k p_k^{PCG}\) if and only if

$$\begin{aligned} U_{k} \left( M^{-1}g_k+\frac{g_k^TM^{-1}g_k}{p_{k-1}^T g_{k-1}}p_{k-1} \right) = \left( \frac{1}{\delta _k} -1\right) g_k + \frac{ g_k^T M^{-1}g_k}{p_{k-1}^T g_{k-1}} g_{k-1}. \end{aligned}$$

Lemma 2

Consider iteration k of the exact linesearch method of Algorithm 1, where \(1\le k < r\). Assume that \(p_i=\delta _i p_i^{PCG}\) with \(\delta _i\ne 0\) for \(i=0,\ldots ,k-1\), where \(p_i^{PCG}\), \(i=0,\ldots ,k-1\), are the search directions of the preconditioned method of conjugate gradients using a positive definite symmetric preconditioning matrix M, as stated in Definition 3. Let \(\hat{\gamma }_k={p_{k-1}^T g_{k-1}}/{g_k^T M^{-1}g_k}\). For \(\delta _k\ne 0\) and \(\gamma _k\ne \hat{\gamma }_k\), let the functions \(\gamma _k(\delta _k)\) and \(\delta _k(\gamma _k)\) be defined by

$$\begin{aligned} \gamma _k(\delta _k) =- \frac{p_{k-1}^T g_{k-1}}{g_k^T M^{-1}g_k} \left( \frac{1}{\delta _k}-1\right) , \quad \delta _k(\gamma _k) =\frac{1}{1 - \gamma _k\frac{g_k^T M^{-1}g_k}{p_{k-1}^T g_{k-1}}}. \end{aligned}$$

Then, the functions \(\gamma _k(\cdot )\) and \(\delta _k(\cdot )\) are inverses to each other.

Proposition 6

Consider iteration k of the exact linesearch method of Algorithm 1, where \(1\le k < r\). Assume that \(p_i=\delta _i p_i^{PCG}\) with \(\delta _i\ne 0\) for \(i=0,\ldots ,k-1\), where \(p_i^{PCG}\), \(i=0,\ldots ,k-1\), are the search directions of the preconditioned method of conjugate gradients using a positive definite symmetric preconditioning matrix M, as stated in Definition 3. Let \(B_k\) and \(p_k\) satisfy \(B_kp_k=-g_k\), and let \(B_{k-1}\) be a nonsingular matrix such that \(B_{k-1} p_{k-1}=-g_{k-1}\) and \(B_{k-1}M^{-1}g_k = g_k\). In addition, let \(\gamma _k(\cdot )\), \(\delta _k(\cdot )\) and \(\hat{\gamma }_k\) be given by Lemma 2.

For any scalar \(\gamma _k\), except \(\gamma _k=0\), \(\gamma _k=\hat{\gamma }_k\) and \(\gamma _k=1\), let \(B_k\) be defined by

$$\begin{aligned} B_{k} = B_{k-1} - \frac{1}{ (\gamma _k-1) p_{k-1}^T g_{k-1}} (\gamma _{k}g_k - g_{k-1})(\gamma _{k}g_k-g_{k-1})^T. \end{aligned}$$
(26)

Then, \(B_k\) is nonsingular and \(p_k=\delta _k p_k^{PCG}\) for \(\delta _k=\delta _k(\gamma _k)\).

Conversely, for any scalar \(\delta _k\), except \(\delta _k=0\), \(\delta _k=\delta _k(1)\) and \(\delta _k=1\), assume that \(p_k=\delta _k p_k^{PCG}\) and assume that \(B_k-B_{k-1}\) is symmetric and of rank one. Then, \(B_k\) is a nonsingular matrix given by (26) for \(\gamma _k=\gamma _k(\delta _k)\).

If, in addition, \(B_{k-1}=B_{k-1}^T\succ 0\), then \(B_k\) defined by (26) satisfies \(B_k\succ 0\) if and only if \(\gamma _k>1\) or \(\hat{\gamma }_k<\gamma _k<0\), or equivalently if and only if \(\gamma _k=\gamma _k(\delta _k)\) for \(0<\delta _k<\delta _k(1)\) or \(\delta _k>1\).

Finally, if \(B_i p_i=-g_i\), \(i=0,\ldots ,k\), with \(B_0=M\) and if, for \(i=1,\ldots ,k\), \(B_{i-1}\) is updated to \(B_i\) according to (26) for \(\gamma _i\) such that \(\gamma _i\ne 0\), \(\gamma _i\ne \hat{\gamma }_i\) and \(\gamma _i\ne 1\), then

$$\begin{aligned} B_k p_i = \frac{\gamma _{i+1}\theta _i}{\gamma _{i+1}-1}H p_i, \quad i=0,\ldots ,k-1. \end{aligned}$$

7 Conclusion

In this paper we have derived necessary and sufficient conditions on the matrix \(B_k\) in a QN-method such that \(p_k\), obtained by solving \(B_k p_k=-g_k\), satisfies \(p_k=\delta _k p_k^{PCG}\) for some \(\delta _k \ne 0\), where \(p_k^{PCG}\) is the search direction of the preconditioned method of conjugate gradients. These conditions are stated in Proposition 4. The results have been derived for the case of CG and then extended to PCG for a symmetric positive definite preconditioning matrix M.

Further, we have characterized the symmetric rank-one update matrices for QN that give parallel search directions to those of PCG. In Proposition 6, we show that the rank-one matrix must be a linear combination of \(g_k\) and \(g_{k-1}\), and also that almost any linear combination will do. In addition, we characterize the family of symmetric rank-one updates that preserve symmetry and positive definiteness of \(B_{k-1}\).

Our focus is on the mathematical properties of PCG and QN in exact arithmetic. We want to stress that considering the numerical properties in finite precision is of utmost importance, but such an analysis is beyond the scope of this paper. See, e.g., [9] for an illustration of a case where PCG and QN generate identical iterates in exact arithmetic but the difference between numerically computed iterates for the two methods is large.

The results of the paper are meant to be useful as such, for understanding the behavior of exact linesearch quasi-Newton methods for minimizing a quadratic function. In addition, we hope that they can lead to further research on methods for unconstrained minimization. In particular, understanding the behavior of quasi-Newton methods on near-quadratic functions would be a subject of future research.