Skip to main content
Log in

A polynomial primal-dual affine scaling algorithm for symmetric conic optimization

  • Published:
Computational Optimization and Applications Aims and scope Submit manuscript

Abstract

The primal-dual Dikin-type affine scaling method was originally proposed for linear optimization and then extended to semidefinite optimization. Here, the method is generalized to symmetric conic optimization using the notion of Euclidean Jordan algebras. The method starts with an interior feasible but not necessarily centered primal-dual solution, and it features both centering and reducing the duality gap simultaneously. The method’s iteration complexity bound is analogous to the semidefinite optimization case. Numerical experiments demonstrate that the method is viable and robust when compared to SeDuMi, MOSEK and SDPT3.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. http://sedumi.ie.lehigh.edu/.

  2. http://www.math.nus.edu.sg/~mattohkc/sdpt3.html.

  3. https://www.mosek.com/.

References

  1. Alizadeh, F., Schmieta, S.H.: Optimization with semidefinite, quadratic and linear constraints. Technical report RRR 23-97, RUTCOR, Rutgers University, New Brunswick (1997)

  2. Alizadeh, F., Schmieta, S.: Symmetric cones, potential reduction methods and word-by-word extensions. In: Wolkowicz, H., Saigal, S., Vandenberghe, L. (eds.) Handbook on Semidefinite Programming, pp. 195–233. Kluwer Academic Publishers, Dordrecht (2000)

    Chapter  Google Scholar 

  3. Ben-Tal, A., Nemirovskii, A.: Lectures on Modern Convex Optimization. MPS/SIAM Series on Optimization. SIAM, Philadelphia (2001)

    Book  Google Scholar 

  4. Faraut, J., Korányi, A.: Analysis on Symmetric Cones. Oxford University Press, Oxford (1994)

    MATH  Google Scholar 

  5. Faybusovich, L.: Euclidean Jordan algebras and interior point algorithms. Positivity 1(4), 331–357 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  6. Faybusovich, L.: A Jordan algebraic approach to potential-reduction algorithms. Math. Z. 239(1), 117–129 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  7. Fiacco, A.V., McCormick, G.P.: Nonlinear Programming: Sequential Unconstrained Minimization Techniques. SIAM Classics in Applied Mathematics. SIAM Publications, Philadelphia (1990)

    Book  MATH  Google Scholar 

  8. Frisch, K.R.: The logarithmic potential method for solving linear programming problems. Memorandum. University Institute of Economics, Oslo (1955)

    Google Scholar 

  9. Goldfarb, D., Scheinberg, K.: Interior point trajectories in semidefinite programming. SIAM J. Optim. 8, 871–886 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  10. Gu, G., Zangiabadi, M., Roos, C.: Full Nesterov–Todd step infeasible interior-point method for symmetric optimization. Eur. J. Oper. Res. 214(3), 473–484 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  11. Güler, O.: Barrier functions in interior point methods. Math. Oper. Res. 21(4), 860–885 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  12. Jansen, B., Roos, C., Terlaky, T.: A Polynomial primal-dual Dikin-type algorithm for linear programming. Math. Oper. Res. 21(2), 341–353 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  13. Karmarkar, N.: A new polynomial-time algorithm for linear programming. Combinatorica 4(4), 373–395 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  14. de Klerk, E., Roos, C., Terlaky, T.: Polynomial primal-dual affine scaling algorithms in semidefinite programming. J. Comb. Optim. 2(1), 51–69 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  15. Mittelmann, H.D.: An independent benchmarking of SDP and SOCP solvers. Math. Program. 95(2), 407–430 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  16. Monteiro, R.D., Adler, I., Resende, M.G.: A polynomial-time primal-dual affine scaling algorithm for linear and convex quadratic programming and its power series extension. Math. Oper. Res. 15(2), 191–214 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  17. Nesterov, Y.E., Nemirovskii, A.: Interior Point Polynomial Methods in Convex Programming: Theory and Applications. SIAM, Philadelphia (1994)

    Book  MATH  Google Scholar 

  18. Nesterov, Y.E., Todd, M.J.: Self-scaled barriers and interior point methods for convex programming. Math. Oper. Res. 22(1), 1–42 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  19. Nesterov, Y.E., Todd, M.J.: Primal-dual interior point methods for self-scaled cones. SIAM J. Optim. 8(2), 324–364 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  20. Peng, J., Roos, C., Terlaky, T.: Self-Regularity: A New Paradigm for Primal-Dual Interior Point Algorithms. Princeton Series in Applied Mathematics. Princeton University Press, Princeton (2002)

    MATH  Google Scholar 

  21. Rangarajan, B.K.: Polynomial convergence of infeasible-interior point methods over symmetric cones. SIAM J. Optim. 16(4), 1211–1229 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  22. Schmieta, S.H.: Application of Jordan Algebras to the design and analysis of interior-point algorithms for linear, quadratically constrained quadratic, and semidefinite programming. PhD Thesis, RUTCOR, Rutgers University, New Brunswick (1999)

  23. Schmieta, S.H., Alizadeh, F.: Extension of commutative class of primal-dual interior-point algorithms to symmetric cones. Technical Report RRR 13-99, RUTCOR, Rutgers University, New Brunswick (1999)

  24. Schmieta, S.H., Alizadeh, F.: Associative and Jordan algebras, and polynomial time interior-point algorithms for symmetric cones. Math. Oper. Res. 26(3), 543–564 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  25. Schmieta, S.H., Alizadeh, F.: Extension of polynomial time interior-point algorithms to symmetric cones. Math. Program. 96(3), 409–438 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  26. Sturm, J.F.: Using SeDuMi 1.02, a MATLAB toolbox for optimization over symmetric cones. Optim. Methods Softw. 11(1–4), 625–653 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  27. Sturm, J.F.: Similarity and other spectral relations for symmetric cones. Linear Algebra and Its Applications 312(1), 135–154 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  28. Terlaky, T., Wang, Z.: On the identification of the optimal partition of second order cone optimization problems. SIAM J. Optim. 24, 385–414 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  29. Toh, K.C., Todd, M.J., Tütüncü, R.H.: SDPT3 - A Matlab software package for semidefinite programming, version 1.3. Optim. Methods Softw. 11(1–4), 545–581 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  30. Tütüncü, R.H., Toh, K.C., Todd, M.J.: Solving semidefinite-quadratic-linear programs using SDPT3. Math. Program. 95(2), 189–217 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  31. Zhang, Y.: On extending primal-dual interior-point algorithms from linear programming to semidefinite programming. SIAM J. Optim. 8(2), 356–386 (1998)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgments

This work is supported by Airforce Office of Scientific Research (AFOSR) Grant # FA9550-15-1-0222.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ali Mohammad-Nezhad.

Appendix: Euclidean Jordan algebras

Appendix: Euclidean Jordan algebras

This section gives a brief review of the basic properties of the Euclidean Jordan algebras. For the sake of simplicity, we only provide the necessary concepts which will be required in this paper. For detailed studies of Euclidean Jordan algebras the reader can consult [4] and [25].

Definition 7.1

Let \(\mathcal {J}\) be an n-dimensional vector space over the field of real numbers with a bilinear map \((x,s) \rightarrow x \circ s\). Then, \((\mathcal {J},\circ )\) is referred to as the Euclidean Jordan algebra if for all \(x,s,z \in \mathcal {J}\)

  1. 1.

    \(x \circ s=s \circ x\),

  2. 2.

    \(x \circ (x^2 \circ s)=x^2 \circ (x \circ s)\),

  3. 3.

    \(\langle x,x \rangle > 0\) for all \(x \ne 0\),

where \(x^2=x \circ x\), and \(\langle . , . \rangle \) denotes an inner product defined over \(\mathcal {J}\). An identity element is defined for a Euclidean Jordan algebra \(\mathcal {J}\), if there exists a unique element e, such that \(x \circ e=e \circ x=x\) for all \(x \in \mathcal {J}\).

Roughly speaking, a Euclidean Jordan algebra is a commutative algebra over the field of real numbers which is not necessarily associative. Nevertheless, the Euclidean Jordan algebra is power associative; that is \(x^{p+q}:=x^p \circ x^q\).

For all \(x,s \in \mathcal {J}\), the bilinear map \((x,s) \rightarrow x \circ s\) is characterized by

$$\begin{aligned} L(x)s=x \circ s, \end{aligned}$$

where L(x) denotes a symmetric matrix. In particular, \(L(x)e=x\) and \(L(x)x=x^2\). The quadratic representation of x is defined as

$$\begin{aligned} P(x):=2L^2(x)-L(x^2). \end{aligned}$$

Definition 7.2

The cone of squares of a Euclidean Jordan algebra \(\mathcal {J}\) is defined as

$$\begin{aligned} \mathcal {K}(\mathcal {J}):=\{x^2:x \in \mathcal {J}\}, \end{aligned}$$

where \(x^2=x \circ x\), and \(\mathcal {K}(\mathcal {J})\) is a closed pointed convex cone with nonempty interior.

Example 7.3

Let \(\mathcal {J}\) be an n-dimensional vector space over the field of real numbers, where the identity element e and L(x) are defined as

$$\begin{aligned} e_{n \times 1}:=\begin{bmatrix} 1 \\ 0 \\ . \\ . \\ . \\ 0 \\ \end{bmatrix}, \ \ \ L(x):=\begin{bmatrix}&x_0 \ \&x_{1:n-1}^T \\&x_{1:n-1} \ \&x_0 I_{n-1} \end{bmatrix}. \end{aligned}$$

The vector space \(\mathcal {J}\) endowed with the bilinear map characterized by L(x) is a Euclidean Jordan algebra. Further, \(\mathcal {K}(\mathcal {J}) \equiv \mathcal {L}^n\), where

$$\begin{aligned} \mathcal {L}^n:=\left\{ x \in \mathbb {R}^n: \ x_0 \ge \Vert {x_{1:n-1}}\Vert \right\} , \end{aligned}$$

where \(\Vert {.}\Vert \) denotes the \(l^2\)-norm. The cone \(\mathcal {L}^n\) is referred to as Lorentz cone or the second-order cone. In this algebra, the quadratic representation of \(x \in \mathcal {K}(\mathcal {J})\) is given by

$$\begin{aligned} P(x):=\begin{bmatrix}&\Vert {x}\Vert ^2 \ \ \ \&2x_0x_{1:n-1}^T \\&2x_0x_{1:n-1} \ \ \ \&\gamma (x) I_{n-1}+2x_{1:n-1}x_{1:n-1}^T \end{bmatrix}, \end{aligned}$$

where \(\gamma (x):=x_0^2-\Vert {x_{1:n-1}}\Vert ^2\), and \(I_{n-1}\) denotes an identity matrix of size \(n-1\).

The following lemma specifies some important properties of the quadratic representation. The interested reader is referred to [4] for more details.

Lemma 7.4

(Proposition II.3.1 in [4], Lemma 8 in [25]) For an invertible \(x \in \mathcal {J}\) and integer value t, we have

  1. 1.

    \(P(x^{-1})=P(x)^{-1}\) and in general \(P(x^t)=P(x)^t\),

  2. 2.

    \(P(x)x^{-1}=x\),

  3. 3.

    \(P(x)e=x^2\).

We now introduce the concept of eigenvalue and spectral decomposition in Euclidean Jordan algebras. Let r be the smallest integer such that the set \(\{e,x,x^2,\ldots ,x^r\}\) is linearly dependent for \(x \in \mathcal {J}\). Then, r is denoted as the degree of x, \(\deg (x)\). The rank of \(\mathcal {J}\) is defined as the largest value of \(\deg (x)\) over \(x \in \mathcal {J}\).

A nonzero element \(q \in \mathcal {J}\) is called idempotent if \(q^2=q\). Furthermore, an idempotent is primitive if it is not the sum of two other idempotents. In light of these definition, a Jordan frame is defined as a set of primitive idempotents \(\{q_1,\ldots ,q_r\}\), where \(q_i \circ q_j=0\) for all \(i \ne j\) and \(q_1+\ldots +q_r=e\).

Theorem 7.5

(Theorem III.1.2 in [4]) Let \(\mathcal {J}\) be a Euclidean Jordan algebra with rank r. Then each \(x \in \mathcal {J}\) can be represented as

$$\begin{aligned} x=\lambda _1 q_1+\ldots +\lambda _r q_r, \end{aligned}$$

where \(\{q_1, \ldots , q_r\}\) denotes a Jordan frame, and \(\lambda _i\) stands for the eigenvalues of x.

Example 7.6

Let \(x \in \mathcal {L}^n\). It can be easily shown that

$$\begin{aligned} x^2-2x_0x+(x_0^2-\Vert {x_{1:n-1}}\Vert ^2)e=0. \end{aligned}$$

This implies that \(r=2\) for this Euclidean Jordan algebra. Hence, the spectral decomposition for an element \(x \in \mathcal {L}^n\) is given by

$$\begin{aligned} x=\lambda _1 q_1+\lambda _2 q_2, \end{aligned}$$

where

$$\begin{aligned} \lambda _1=x_0-\Vert {x_{1:n-1}}\Vert , \ \text {and} \ \ \lambda _2=x_0+\Vert {x_{1:n-1}}\Vert ,\\ q_1=\tfrac{1}{2} \begin{pmatrix} 1\\ -\frac{x_{1:n-1}}{\Vert {x_{1:n-1}}\Vert } \end{pmatrix}, \ \text {and} \ \ q_2=\tfrac{1}{2} \begin{pmatrix} 1\\ \frac{x_{1:n-1}}{\Vert {x_{1:n-1}}\Vert } \end{pmatrix}. \end{aligned}$$

We can now extend the definition of any real valued function f(.) to elements of the Euclidean Jordan algebras by

$$\begin{aligned} f(x):=f(\lambda _1)q_1+\cdots +f(\lambda _r)q_r. \end{aligned}$$

In particular, we have

$$\begin{aligned} x^{\tfrac{1}{2}}&:=\lambda _1^{\tfrac{1}{2}} q_1+\cdots +\lambda _r^{\tfrac{1}{2}} q_r,\\ x^{-1}&:=\lambda _1^{-1}q_1+\cdots +\lambda _r^{-1}q_r. \end{aligned}$$

Note that \(x^{-1} \circ x=e\). Further, x is called invertible if all the eigenvalues of x are nonzeros.

In light of the definitions given so far, trace, determinant and norms of x are formally defined as follows.

Definition 7.7

Let \(x \in \mathcal {J}\) and \(\lambda _1,\cdots ,\lambda _r\) be the eigenvalues of x. Then,

  1. 1.

    \({{\mathrm{tr}}}(x):=\lambda _1+\lambda _2+\cdots +\lambda _r\),

  2. 2.

    \(\det (x):=\lambda _1 \lambda _2 \cdots \lambda _r\),

  3. 3.

    \(\langle x,y \rangle :={{\mathrm{tr}}}(x \circ y)\),

  4. 4.

    \(\Vert {x}\Vert _F:=\sqrt{\lambda _1^2+\lambda _2^2+\cdots +\lambda _r^2}\),

  5. 5.

    \(\Vert {x}\Vert _2:=\max _{i} |\lambda _i|\).

We now state an important theorem from [2] which is adopted from [4].

Theorem 7.8

(Theorem 8.3.6 in [2, 4]) Let \(\mathcal{J}\) be an Euclidean Jordan algebra. Then \(\mathcal{J}\) falls into one of the following categories:

  1. 1.

    An n-dimensional vector space over the field of real numbers, where

    $$\begin{aligned} x \circ s := \begin{bmatrix} x^Ts\\x_0s_{1:n-1}+s_0x_{1:n-1}\end{bmatrix}. \end{aligned}$$
  2. 2.

    The space of \(n \times n\) real symmetric matrices, where

    $$\begin{aligned} X \circ S := \frac{XS+SX}{2}. \end{aligned}$$
    (18)

    for symmetric matrices X and S.

  3. 3.

    The space of \(n \times n\) complex Hermitian matrices, where the Jordan product is defined as in (18) for Hermitian matrices X and S.

  4. 4.

    The space of \(n \times n\) Hermitian matrices with quaternion entries, where the Jordan product is defined as in (18) for quaternion Hermitian matrices X and S.

  5. 5.

    The space of \(3 \times 3\) Hermitian matrices with octonion entries, known as Albert algebra, where the Jordan product is defined as in (18).

In the rest of this section, we review some technical lemmas (without their proofs) which are necessary for the complexity analysis of the Dikin-type algorithm. As stated at the begining part of this paper, it is assumed that \(\mathcal {K}\) is a symmetric cone with nonempty interior \(\mathcal {K}_+\).

Lemma 7.9

(Lemma 2.13 in [10]) Assume that \(x,s \in \mathcal {J}\) and \({{\mathrm{tr}}}(x \circ s)=0\). Then, we have

$$\begin{aligned} -\tfrac{1}{4} \Vert {x+s}\Vert _F^2 e \preceq _{\mathcal {K}} x \circ s \preceq _{\mathcal {K}} \tfrac{1}{4} \Vert {x+s}\Vert ^2_F e. \end{aligned}$$

Lemma 7.10

(Lemma 2.15 in [10]) Assume that \(0 \prec _{\mathcal {K}} x \circ s\), where \(x,s \in \mathcal {J}\). Then, \(\det (x) \ne 0\).

Lemma 7.11

(Lemma 2.17 in [10]) Let \(x \in \mathcal {J}\) and \(0 \prec _{\mathcal {K}} s\). Then, we have

$$\begin{aligned} \lambda _{\min }(x) {{\mathrm{tr}}}(s) \le {{\mathrm{tr}}}(x \circ s) \le \lambda _{\max }(x) {{\mathrm{tr}}}(s). \end{aligned}$$

The following lemma points out a nice property of the quadratic representation.

Lemma 7.12

(Theorem III.2.1 and Proposition III.2.2 in [4]) Let \(x \in \mathcal {J}\). Then, L(x) is positive definite (semidefinite) if and only if \(0 \prec _{\mathcal {K}} x\) (\(0 \preceq _{\mathcal {K}} x\)). Further, \(P(x)\mathcal {K}_+=\mathcal {K}_+\) if x is invertible.

In fact, Lemma 7.12 states that for each interior solution \(0 \prec _{\mathcal {K}} x\) and \(0 \prec _{\mathcal {K}} s\), P(x)s is an invertible linear map from \(\mathcal {K}_+\) to \(\mathcal {K}_+\). All this hints that the NT search directions obtained from (13) are well-defined.

Lemma 7.13

(Proposition 21 in [25]) Let \(0 \prec _{\mathcal {K}} x\) and \(0 \prec _{\mathcal {K}} s\), and w be the scaling point of x and s as defined in (5). Then,

  1. 1.

    \({{\mathrm{tr}}}(P(w)^{-\tfrac{1}{2}}x \circ P(w)^{\tfrac{1}{2}}s)={{\mathrm{tr}}}(x \circ s)\),

  2. 2.

    \(P(x^{\tfrac{1}{2}})s \sim P(s^{\tfrac{1}{2}})x\),

  3. 3.

    \(P(\tilde{x}^{\tfrac{1}{2}})\tilde{s} \sim P(x^{\tfrac{1}{2}})s\),

where \(\tilde{x}:=P(w)^{-\tfrac{1}{2}}x\) and \(\tilde{s}:=P(w)^{\tfrac{1}{2}}s\).

Lemma 7.14

(Lemma 30 in [25]) Let \(0 \prec _{\mathcal {K}} x\) and \(0 \prec _{\mathcal {K}} s\). Then, we have

  1. 1.

    \(\lambda _{\min }(P(x)^{\tfrac{1}{2}}s) \ge \lambda _{\min } (x \circ s)\),

  2. 2.

    \(\lambda _{\max }(P(x)^{\tfrac{1}{2}}s) \le \lambda _{\max } (x \circ s)\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mohammad-Nezhad, A., Terlaky, T. A polynomial primal-dual affine scaling algorithm for symmetric conic optimization. Comput Optim Appl 66, 577–600 (2017). https://doi.org/10.1007/s10589-016-9874-5

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10589-016-9874-5

Keywords

Mathematics Subject Classification

Navigation