Skip to main content
Log in

Operator Scaling: Theory and Applications

  • Published:
Foundations of Computational Mathematics Aims and scope Submit manuscript

Abstract

In this paper, we present a deterministic polynomial time algorithm for testing whether a symbolic matrix in non-commuting variables over \({\mathbb {Q}}\) is invertible or not. The analogous question for commuting variables is the celebrated polynomial identity testing (PIT) for symbolic determinants. In contrast to the commutative case, which has an efficient probabilistic algorithm, the best previous algorithm for the non-commutative setting required exponential time (Ivanyos et al. in Comput Complex 26(3):717–763, 2017) (whether or not randomization is allowed). The algorithm efficiently solves the “word problem” for the free skew field, and the identity testing problem for arithmetic formulae with division over non-commuting variables, two problems which had only exponential time algorithms prior to this work. The main contribution of this paper is a complexity analysis of an existing algorithm due to Gurvits (J Comput Syst Sci 69(3):448–484, 2004), who proved it was polynomial time for certain classes of inputs. We prove it always runs in polynomial time. The main component of our analysis is a simple (given the necessary known tools) lower bound on central notion of capacity of operators (introduced by Gurvits 2004). We extend the algorithm to actually approximate capacity to any accuracy in polynomial time, and use this analysis to give quantitative bounds on the continuity of capacity (the latter is used in a subsequent paper on Brascamp–Lieb inequalities). We also extend the algorithm to compute not only singularity, but actually the (non-commutative) rank of a symbolic matrix, yielding a factor 2 approximation of the commutative rank. This naturally raises a relaxation of the commutative PIT problem to achieving better deterministic approximation of the commutative rank. Symbolic matrices in non-commuting variables, and the related structural and algorithmic questions, have a remarkable number of diverse origins and motivations. They arise independently in (commutative) invariant theory and representation theory, linear algebra, optimization, linear system theory, quantum information theory, approximation of the permanent and naturally in non-commutative algebra. We provide a detailed account of some of these sources and their interconnections. In particular, we explain how some of these sources played an important role in the development of Gurvits’ algorithm and in our analysis of it here.

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.

Institutional subscriptions

Similar content being viewed by others

Notes

  1. Our main results will be for the rationals \({\mathbb {Q}}\) (and will hold for \({\mathbb {R}}\) and \({\mathbb {C}}\) as well) but not for finite fields. However, many of the questions are interesting for any field.

  2. For all purposes, we may assume that the matrices \(A_i\) are linearly independent, namely span a space of matrices of dimension exactly m.

  3. For now, the reader may think of the elements of this “free skew field” simply as containing all expressions (formulas) built from the variables and constants using the arithmetic operations of addition, multiplication and division (we define it more formally a bit later). We note that while this is syntactically the same definition one can use for commuting variables, the skew field is vastly more complex, and in particular its elements cannot be described canonically as ratios of polynomials.

  4. Actually there are many, but only one “universal field of fractions”.

  5. Moreover, the polynomial entries of KM in such a minimal decomposition can actually be taken to be polynomials of degree at most 1, namely affine combinations of the variables.

  6. Here \(A^{\dagger }\) denotes the conjugate-transpose of a complex matrix A.

  7. The left–right action and its invariant polynomials are defined as follows. Consider \(mn^2\) commuting variables which are arranged in m matrices \((Y_1, Y_2, \ldots Y_m)\), and consider polynomials in these variables. Every pair BC of determinant 1 matrices over \({\mathbb {F}}\) defines a linear map of these variables by sending this tuple to \((BY_1C, BY_2C, \ldots BY_mC)\). A polynomial in these variables is invariant if it remains unchanged by this action for every such pair BC.

  8. This is a technical term which we will not define here.

  9. In general inversion height, the minimum amount of nesting needed, can be arbitrarily high. This is an important theorem of Reutenauer [73]. However, in the example above, the nested inversion can be eliminated, and in fact, the two expressions are equal (a simple fact which the reader might try to prove)! This equality is called Hua’s identity [48], underlying the fundamental theorem of projective geometry.

  10. when the expression attempts to invert a singular matrix it is undefined on that input. The domain of an expression is simply all input tuples on which it is defined.

  11. For example, deciding whether two knots diagrams describe the same knot was proved decidable by Haken [42], and deciding whether two presentations with generators and relations describe the same group were proved undecidable by Rabin [69].

  12. While this paper focuses on identity testing, we note that our interest is partly (and indirectly) motivated by the more basic problem of proving lower bounds for non-commutative circuits. We refer the reader to the papers [46, 47, 59, 66] and their references for existing lower bounds on weaker models, some completeness results, and possible approaches to proving stronger lower bounds.

  13. Replacing formulas by circuits there is no contrast—in both the commutative and non-commutative setting matrix inverse has a polynomial size circuit (with division of course) [45].

  14. It is interesting to note that most recent progress on deterministic PIT algorithms (e.g., [25, 31, 55, 57, 75] among many others) are for polynomials computed by a variety of restricted classes of arithmetic circuits. Algorithm G seems to differ from all of them in solving PIT for a very different class of polynomials, which we do not know how to classify in arithmetic complexity terms.

  15. For both the commutative and non-commutative definitions.

  16. It is an interesting question whether a compression space naturally arises from matroid parity duality of Lovasz [62, 63].

  17. A “non-triviality” assumption is that no row or column in A is all zero.

  18. Again using a “non-triviality” assumption these matrices are invertible.

  19. Arising in particular in the GCT program of Mulmuley and Sohoni.

  20. Note that for it to be a group action in the strict sense, one should study the action which takes Y to \(BYC^{-1}\) or \(BYC^{T}\) but for simplicity, we will avoid this distinction.

  21. We note that this is part of the larger project of understanding quiver representations, started by the works of Procesi, Razmysolov, and Formanek [32, 68, 72].

  22. Note though that the roles of which matrices in the tensor product are variable, and which are constant, has switched!

  23. This also follows from [11, 56].

  24. This bound may a-priori depend on m, the number of matrices, but we already noted that \(m\le n^2\).

  25. We note that this notion of capacity seems to have nothing to do with the usual capacity of a quantum channel.

  26. Recall that log determinant is a concave function over the domain of positive definite matrices.

  27. Taking the dimension’s root of capacity.

  28. Notice that we can make the following assumption just to simplify notation. In actuality, we do not know where the full rank minor is located in M.

References

  1. B. Adsul, S. Nayak, and K. V. Subrahmanyam. A geometric approach to the Kronecker problem ii : rectangular shapes, invariants of n*n matrices, and a generalization of the Artin-Procesi theorem. Manuscript, available at http://www.cmi.ac.in/~kv/ANS10.pdf, 2010.

  2. N. Alon. Combinatorial Nullstellensatz. Combinatorics, Probability and Computing, 8(1-2):7–29, 1999.

    Article  MathSciNet  MATH  Google Scholar 

  3. S. A. Amistur and J. Levitzki. Minimal identities for algebras. Proceedings of the American Mathematical Society, 1:449–463, 1950.

    Article  MathSciNet  MATH  Google Scholar 

  4. S. Amitsur. Rational identities and applications to algebra and geometry. Journal of Algebra, 3:304–359, 1966.

    Article  MathSciNet  MATH  Google Scholar 

  5. M. D. Atkinson. Spaces of matrices with several zero eigenvalues. Bulletin of the London Mathematical Society, 12(89-95), 1980.

  6. M. D. Atkinson and S. Lloyd. Large spaces of matrices of bounded rank. Quarterly Journal of Math. Oxford, 31:253–262, 1980.

    Article  MathSciNet  MATH  Google Scholar 

  7. L. B. Beasley. Nullspaces of spaces of matrices of bounded rank. Current trends in matrix theory, 1987.

  8. S. J. Berkowitz. On computing the determinant in small parallel time using a small number of processors. Information Processing Letters, 18(3):147–150, 1984.

    Article  MathSciNet  MATH  Google Scholar 

  9. M. Bläser, G. Jindal, and A. Pandey. Greedy strikes again: A deterministic PTAS for commutative rank of matrix spaces. In LIPIcs-Leibniz International Proceedings in Informatics, volume 79. Schloss Dagstuhl-Leibniz-Zentrum fuer Informatik, 2017.

  10. A. Bogdanov and H. Wee. More on noncommutative polynomial identity testing. Computational Complexity, pages 92–99, 2005.

  11. M. Bürgin and J. Draisma. The hilbert null-cone on tuples of matrices and bilinear forms. Math Z, 254(4):785–809, 2006.

    Article  MathSciNet  MATH  Google Scholar 

  12. M. Choi. Completely positive linear maps on complex matrices. Linear Algebra and Its Applications, pages 285–290, 1975.

  13. P. M. Cohn. The embedding of firs in skew fields. Proceedings of the London Mathematical Society, 23:193–213, 1971.

    Article  MathSciNet  MATH  Google Scholar 

  14. P. M. Cohn. The word problem for free fields. The Journal of Symbolic Logic, 38(2):309–314, 1973.

    Article  MathSciNet  MATH  Google Scholar 

  15. P. M. Cohn. The word problem for free fields: A correction and an addendum. Journal of Symbolic Logic, 40(1):69–74, 1975.

    Article  MathSciNet  MATH  Google Scholar 

  16. P. M. Cohn. Skew Fields, Theory of General Division Rings. Cambridge University Press, 1995.

  17. P. M. Cohn and C. Reutenauer. On the construction of the free field. International journal of Algebra and Computation, 9(3):307–323, 1999.

    Article  MathSciNet  MATH  Google Scholar 

  18. D. Cox, J. Little, and D. O’Shea. Ideals, Varieties, and Algorithms. Undergraduate Texts in Mathematics. Springer, New York, third edition edition, 2007.

  19. H. Derksen. Polynomial bounds for rings of invariants. Proceedings of the American Mathematical Society, 129(4):955–964, 2001.

    Article  MathSciNet  MATH  Google Scholar 

  20. H. Derksen and G. Kemper. Computational Invariant Theory, volume 130. Springer-Verlag, Berlin, 2002.

    Book  MATH  Google Scholar 

  21. H. Derksen and V. Makam. Polynomial degree bounds for matrix semi-invariants. Advances in Mathematics, 310:44–63, 2017.

    Article  MathSciNet  MATH  Google Scholar 

  22. H. Derksen and J. Weyman. Semi-invariants of quivers and saturation for Littlewood-Richardson coefficients. Journal of the American Mathematical Society, 13(3):467–479, 2000.

    Article  MathSciNet  MATH  Google Scholar 

  23. J. Dieudonné. Sur une généralisation du groupe orthogonal à quatre variables. Arch. Math., 1:282–287, 1949.

    Article  MathSciNet  MATH  Google Scholar 

  24. M. Domokos and A. N. Zubkov. Semi-invariants of quivers as determinants. Transformation Groups, 6(1):9–24, 2001.

    Article  MathSciNet  MATH  Google Scholar 

  25. Z. Dvir and A. Shpilka. Locally decodable codes with 2 queries and polynomial identity testing for depth 3 circuits. SIAM J. Comput, 2006.

  26. J. Edmonds. Systems of distinct representatives and linear algebra. Journal of research of the National Bureau of Standards, 71(241-245), 1967.

  27. J. Edmonds. Submodular functions, matroids, and certain polyhedra. Lectures, Calgary International Symposium on Combinatorial Structures, 1969.

  28. D. Eisenbud and J. Harris. Vector spaces of matrices of low rank. Advances in Math, 70:135–155, 1988.

    Article  MathSciNet  MATH  Google Scholar 

  29. S. A. Fenner, R. Gurjar, and T. Thierauf. Bipartite perfect matching is in quasi-NC. STOC, 2016.

  30. M. Forbes and A. Shpilka. Explicit noether normalization for simultaneous conjugation via polynomial identity testing. RANDOM, 2013.

  31. M. Forbes and A. Shpilka. Quasipolynomial-time identity testing of non-commutative and read-once oblivious algebraic branching programs. FOCS, pages 243–252, 2013.

  32. E. Formanek. Generating the ring of matrix invariants. Ring Theory, pages 73–82, 1986.

  33. M. Fortin and C. Reutenauer. Commutative/noncommutative rank of linear matrices and subspaces of matrices of low rank. 2004.

  34. A. Garg, L. Gurvits, R. Oliveira, and A. Wigderson. Algorithmic and optimization aspects of Brascamp-Lieb inequalities, via operator scaling. Geometric and Functional Analysis, 28(1):100–145, 2018.

    Article  MathSciNet  MATH  Google Scholar 

  35. B. Gelbord and R. Meshulam. Spaces of singular matrices and matroid parity. European Journal of Combinatorics, 23(4):389–397, 2002.

    Article  MathSciNet  MATH  Google Scholar 

  36. I. Gelfand, S. Gelfand, V. Retakh, and R. Wilson. Quasideterminants. arXiv:math/0208146, 2002.

  37. L. Gurvits. Classical complexity and quantum entanglement. Journal of Computer and System Sciences, 69(3):448–484, 2004.

    Article  MathSciNet  MATH  Google Scholar 

  38. L. Gurvits. Hyperbolic polynomials approach to Van der Waerden/Schrijver-Valiant like conjectures: sharper bounds, simpler proofs and algorithmic applications. STOC, pages 417–426, 2006.

  39. L. Gurvits and A. Samorodnitsky. A deterministic polynomial-time algorithm for approximating mixed discriminant and mixed volume. STOC, 2000.

  40. L. Gurvits and A. Samorodnitsky. A deterministic algorithm approximating the mixed discriminant and mixed volume, and a combinatorial corollary. Discrete Computational Geometry, 27(531-550), 2002.

  41. L. Gurvits and P. N. Yianilos. The deflation-inflation method for certain semidefinite programming and maximum determinant completion problems. Technical Report, NECI, 1998.

  42. W. Haken. Theorie der Normalflächen. Acta Math, 105:245–375, 1961.

    Article  MathSciNet  MATH  Google Scholar 

  43. G. Higman. Units in group rings. PhD thesis, Balliol College, 1940.

  44. D. Hilbert. Uber die vollen Invariantensysteme. Math. Ann., 42:313–370, 1893.

    Article  MathSciNet  MATH  Google Scholar 

  45. P. Hrubes and A. Wigderson. Non-commutative arithmetic circuits with division. ITCS, 2014.

  46. P. Hrubes, A. Wigderson, and A. Yehudayoff. Relationless completeness and separations. In Computational Complexity (CCC), 2010 IEEE 25th Annual Conference on, pages 280–290. IEEE, 2010.

  47. P. Hrubes, A. Wigderson, and A. Yehudayoff. Non-commutative circuits and the sum-of-squares problem. Journal of the American Mathematical Society, 24(3):871–898, 2011.

    Article  MathSciNet  MATH  Google Scholar 

  48. L.-K. Hua. Some properties of a sfield. Proceedings of National Academy of Sciences USA, 35:533–537, 1949.

    Article  MathSciNet  MATH  Google Scholar 

  49. L. Hyafil. On the parallel evaluation of multivariate polynomials. SIAM Journal on Computing, 8(2):120–123, 1979.

    Article  MathSciNet  MATH  Google Scholar 

  50. G. Ivanyos, M. Karpinski, Y. Qiao, and M. Santha. Generalized Wong sequences and their applications to Edmonds’ problems. Journal of Computer and System Sciences, 81(7):1373–1386, 2015.

    Article  MathSciNet  MATH  Google Scholar 

  51. G. Ivanyos, Y. Qiao, and K. Subrahmanyam. Non-commutative Edmonds’ problem and matrix semi-invariants. Computational Complexity, 26(3):717–763, 2017.

    Article  MathSciNet  MATH  Google Scholar 

  52. G. Ivanyos, Y. Qiao, and K. V. Subrahmanyam. Constructive noncommutative rank computation in deterministic polynomial time over fields of arbitrary characteristics. Computational Complexity, 27(4):561–593, December 2018.

    Article  MathSciNet  MATH  Google Scholar 

  53. V. Kabanets and R. Impagliazzo. Derandomizing polynomial identity tests means proving circuit lower bounds. Computational Complexity, 13:1–46, 2004.

    Article  MathSciNet  MATH  Google Scholar 

  54. D. S. Kaliuzhnyi-Verbovetskyi and V. Vinnikov. Noncommutative rational functions, their difference-differential calculus and realizations. Multidimensional Systems and Signal Processing, 23(1-2):49–77, 2010.

    MathSciNet  MATH  Google Scholar 

  55. N. Kayal and N. Saxena. Polynomial identity testing for depth 3 circuits. Computational Complexity, 2007.

  56. A. D. King. Moduli of representations of finite dimensional algebras. The Quarterly Journal of Mathematics, 45(4):515–530, 1994.

    Article  MathSciNet  MATH  Google Scholar 

  57. A. Klivans and D. Spielman. Randomness efficient identity testing of multivariate polynomials. In Proceedings of the 33rd Annual STOC, 2001.

  58. H. Kraft and C. Procesi. Classical invariant theory, a primer. https://math.unibas.ch/uploads/x4epersdb/files/primernew.pdf, 1996.

  59. N. Limaye, G. Malod, and S. Srinivasan. Lower bounds for non-commutative skew circuits. In Electronic Colloquium on Computational Complexity (ECCC), volume 22, page 22, 2015.

  60. N. Linial, A. Samorodnitsky, and A. Wigderson. A deterministic strongly polynomial algorithm for matrix scaling and approximate permanents. STOC, pages 644–652, 1998.

  61. L. Lovasz. On determinants, matchings, and random algorithms. Fundamentals of Computation Theory, pages 565–574, 1979.

  62. L. Lovasz. Selecting independent lines from a family of lines in a space. Acta Sci. Math., 42(121-131), 1980.

  63. L. Lovasz. Singular spaces of matrices and their application in combinatorics. Bulletin of the Brazilian Mathematical Society, 20:87–99, 1989.

    Article  MathSciNet  MATH  Google Scholar 

  64. P. Malcolmson. A prime matrix ideal yields a skew field. Journal of the London Mathematical Society, 18(221-233), 1978.

  65. K. Mulmuley. Geometric complexity theory V: Equivalence between blackbox derandomization of polynomial identity testing and derandomization of noether’s normalization lemma. FOCS, pages 629–638, 2012.

  66. N. Nisan. Lower bounds for non-commutative computation. In Proceedings of the twenty-third annual ACM symposium on Theory of computing, pages 410–418. ACM, 1991.

  67. V. L. Popov. The constructive theory of invariants. Izvestiya: Mathematics, 19(2):359–376, 1982.

    Article  MATH  Google Scholar 

  68. C. Procesi. The invariant theory of nn matrices. Advances in Mathematics, 19:306–381, 1976.

    Article  MathSciNet  MATH  Google Scholar 

  69. M. O. Rabin. Recursive unsolvability of group theoretic problems. Annals of Mathematics, 67(172-194), 1958.

  70. R. Rado. A theorem on independence relations. Quarterly Journal of Math. Oxford, 13:83–89, 1942.

    Article  MathSciNet  MATH  Google Scholar 

  71. R. Raz and A. Shpilka. Deterministic polynomial identity testing in non commutative models. Computational Complexity, 14:1–19, 2005.

    Article  MathSciNet  MATH  Google Scholar 

  72. J. P. Razmyslov. Trace identities of full matrix algebras over a field of characteristic zero. Mathematics of the USSR-Izvestiya, 8(4):727, 1974.

    Article  MATH  Google Scholar 

  73. C. Reutenauer. Inversion height in free fields. Selecta Mathematica, 2(1):93–109, 1996.

    Article  MathSciNet  MATH  Google Scholar 

  74. L. H. Rowen. Polynomial identities in ring theory. Academic Press, New York, 1980.

    MATH  Google Scholar 

  75. S. Saraf and I. Volkovich. Black-box identity testing of depth 4 multilinear circuits. In Proceedings of the 43rd annual STOC, 2011.

  76. A. Schofield and M. V. den Bergh. Semi-invariants of quivers for arbitrary dimension vectors. Indagationes Mathematicae, 12(1):125–138, 2001.

    Article  MathSciNet  MATH  Google Scholar 

  77. A. Shpilka and A. Yehudayoff. Arithmetic Circuits: A Survey of Recent Results and Open Questions, volume 5. NOW, Foundations and Trends in Theoretical Computer Science, 2010.

  78. R. Sinkhorn. A relationship between arbitrary positive matrices and doubly stochastic matrices. The Annals of Mathematical Statistics, 35:876–879, 1964.

    Article  MathSciNet  MATH  Google Scholar 

  79. V. Strassen. Vermeidung von Divisionen. Journal für Reine Angew. Math, 264:182–202, 1973.

    MathSciNet  MATH  Google Scholar 

  80. L. Valiant. The complexity of computing the permanent. Theoretical Computer Science, 8:189–201, 1979.

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank Harm Derksen, Pavel Hrubes, Louis Rowen and K. V. Subrahmanyam for helpful discussions. We would also like to thank Oded Regev for suggesting us that operator scaling could be used for approximating capacity. Finally, we thank the anonymous reviewers for a comprehensive reading of the paper and pointing several typographical errors and minor bugs.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ankit Garg.

Additional information

Communicated by Peter Bürgisser.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Ankit Garg: This research was done when the author was a student at Princeton University and his research was partially supported by Mark Braverman’s NSF Grant CCF-1149888, Simons Collaboration on Algorithms and Geometry, Simons Fellowship in Theoretical Computer Science and a Siebel Scholarship. Rafael Oliveira: This research was done when the author was a student at Princeton University and the research supported by NSF Career award (1451191) and by CCF-1523816 award. Avi Wigderson: This research was partially supported by NSF Grant CCF-1412958.

Symbolic Matrices with Polynomial Entries and Non-commutative Rank

Symbolic Matrices with Polynomial Entries and Non-commutative Rank

In this section, we show how to compute the non-commutative rank of any (not necessarily square) matrix with linear entries over the free skew field . This will be achieved in two ways: the first, in Sect. A.2, by reducing this problem to testing singularity of a certain square matrix with linear entries, and the second, in Sect. A.3, by a purely quantum approach which in a sense mimics the reduction from maximum matching to perfect matching.

In fact, we solve a more general problem. Section A.1 starts with a reduction of computing \({\text{ nc-rank }}\) of a matrix with polynomial entries (given by formulae), to the problem of computing the \({\text{ nc-rank }}\) of a matrix with linear entries, via the so-called “Higman’s trick” (Proposition A.2). We give the simple quantitative analysis of this reduction, which as far as we know does not appear in the literature and may be useful elsewhere. This reduction, with the two above, allow computing the non-commutative rank of any matrix in time polynomial in the description of its entries.

1.1 Higman’s Trick

Before stating the full version of the effective Higman trick, we need to define the bit complexity of a formula computing a non-commutative polynomial.

Definition A.1

(Bit Complexity of a Formula) Let \(\Phi \) be a non-commutative formula without divisions such that each of its gates computes a polynomial in \( \mathbb {Q}\langle {\mathbf {x}}\rangle \) (i.e., the inputs to the formula are either rational numbers or non-commutative variables). The bit complexity of \(\Phi \) is the maximum bit complexity of any rational input appearing in the formula \(\Phi \).

With this definition in hand, we can state and prove Higman’s trick, which first appeared in [43]. In the proposition below, it will be useful to have the following notation to denote the direct sum of two matrices A and B:

$$\begin{aligned} A \oplus B = \begin{pmatrix} A &{}\quad 0 \\ 0 &{}\quad B \end{pmatrix}, \end{aligned}$$

where the zero matrices in the top right and bottom left corners are of appropriate dimensions. Before stating and proving Higman’s trick, let us work through a small example which showcases the essence of the trick.

Suppose we want to know the \({\text{ nc-rank }}\) of matrix \(\begin{pmatrix} 1 &{} x\\ y &{} z + xy \end{pmatrix}\). The problem here is that this matrix is not linear, and we need to have a linear matrix. How can we convert this matrix into a linear matrix while preserving the rank, or the complement of the rank? To do this, we need to remove the multiplication happening in \(z + xy\).

Notice that the complement of its rank does not change after the following transformation:

$$\begin{aligned} \begin{pmatrix} 1 &{}\quad x \\ y &{}\quad z + xy \end{pmatrix} \mapsto \begin{pmatrix} 1 &{}\quad x &{}\quad 0 \\ y &{}\quad z + xy &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 1 \end{pmatrix}. \end{aligned}$$

Since the complement of the rank does not change after we perform elementary row or column operations, we can first add \(x \cdot \text {(third row)}\) to the second row, and then subtract \(\text {(third column)} \cdot y\) to the second column, to obtain:

$$\begin{aligned} \begin{pmatrix} 1 &{}\quad x &{}\quad 0 \\ y &{}\quad z + xy &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 1 \end{pmatrix} \mapsto \begin{pmatrix} 1 &{}\quad x &{}\quad 0 \\ y &{}\quad z + xy &{}\quad x \\ 0 &{}\quad 0 &{}\quad 1 \end{pmatrix} \mapsto \begin{pmatrix} 1 &{}\quad x &{}\quad 0 \\ y &{}\quad z &{}\quad x \\ 0 &{}\quad -y &{}\quad 1 \end{pmatrix} \end{aligned}$$

The complement of the rank of this last matrix is the same as the complement of the rank of our original matrix! In particular, if this last matrix is full rank, it implies that our original matrix is also full rank. This is the essence of Higman’s trick. We now proceed to its full version.

Proposition A.2

(Effective Higman’s Trick) Let \(A \in \mathbb {Q}\langle {\mathbf {x}}\rangle ^{m \times n}\) be a matrix where each entry \(a_{ij}\) is computed by a non-commutative formula of size \(\le s\) and bit complexity \(\le b\) without divisions. Let k be the total number of multiplication gates used in the computation of the entries of A. There exist matrices \(P \in \mathbb {GL}_{m+k}( \mathbb {Q}\langle {\mathbf {x}}\rangle )\), \(Q \in \mathbb {GL}_{n+k}( \mathbb {Q}\langle {\mathbf {x}}\rangle )\) such that \(P (A \oplus I_k) Q\) is a matrix with linear entries and coefficients with bit complexity bounded by b. Moreover, given access to the formulas computing the entries, one can construct P and Q efficiently in time \({\text{ poly }}(m, n, s, b)\). Since P and Q are non-singular matrices, the co-rank and the co-nc-rank of \(P (A \oplus I_k) Q\) are the same as the co-rank and the co-nc-rank of A.

Proof

Let \({\textsf {Mult}}(a_{ij})\) be the number of multiplication gates in the formula computing entry \(a_{ij}\) and

$$\begin{aligned} T = \displaystyle \sum _{\begin{array}{c} 1 \le i \le m \\ 1 \le j \le n \end{array}} {\textsf {Mult}}(a_{ij}). \end{aligned}$$

That is, T is the total number of multiplication gates used to compute all entries of the matrix A.

We prove this proposition by induction on T, for matrices of all dimensions. The base case, when \(T = 0\), is trivial, as in this case A itself has linear entries. Suppose now that the proposition is true for all matrices (regardless of their dimensions) which can be computed by formulas using \(< T\) multiplication gates.

Let A be our matrix, which can be computed using T multiplications. W.l.o.g., we can assume that \({\textsf {Mult}}(a_{mn}) \ge 1\). Then, by finding a multiplication gate in the formula for \(a_{mn}\) that has no other multiplication gate as an ancestor, we can write \(a_{mn}\) in the form \(a_{mn} = a + b \cdot c\), where

$$\begin{aligned} {\textsf {Mult}}(a_{mn}) = {\textsf {Mult}}(a) + {\textsf {Mult}}(b) + {\textsf {Mult}}(c) + 1. \end{aligned}$$

Hence, the matrix

$$\begin{aligned} B = \left( I_{m-1} \oplus \begin{pmatrix} 1 &{}\quad b \\ 0 &{}\quad 1 \end{pmatrix} \right) (A \oplus 1) \left( I_{n-1} \oplus \begin{pmatrix} 1 &{}\quad 0 \\ -c &{}\quad 1 \end{pmatrix} \right) \end{aligned}$$

is such that

$$\begin{aligned} b_{ij} = {\left\{ \begin{array}{ll} a_{ij}, \text { if } i \le m, j \le n \text { and } (i,j) \ne (m,n) \\ a, \text { if } (i,j) = (m,n) \\ b, \text { if } (i,j) = (m, n+1) \\ -c, \text { if } (i,j) = (m+1, n) \\ 1, \text { if } (i,j) = (m+1, n+1) \\ 0, \text { otherwise} \end{array}\right. } \end{aligned}$$

Therefore, the number of multiplications needed to compute B is given by

$$\begin{aligned} \displaystyle \sum _{\begin{array}{c} 1 \le i \le m+1 \\ 1 \le j \le n+1 \end{array}} {\textsf {Mult}}(b_{ij})&= \left( \displaystyle \sum _{\begin{array}{c} 1 \le i \le m \\ 1 \le j \le n \end{array}} {\textsf {Mult}}(a_{ij}) \right) - {\textsf {Mult}}(a_{mn}) \\&\quad + {\textsf {Mult}}(a) + {\textsf {Mult}}(b) + {\textsf {Mult}}(c) \\&= T - {\textsf {Mult}}(a_{mn}) + {\textsf {Mult}}(a) + {\textsf {Mult}}(b) + {\textsf {Mult}}(c) \\&= T -1 \end{aligned}$$

Since B is an \((m+1) \times (n+1)\) matrix which can be computed by using a total of \(T-1\) multiplication gates, by the induction hypothesis, there exist \(P' \in \mathbb {GL}_{m+1 + (T-1)}( \mathbb {Q}\langle {\mathbf {x}}\rangle ) = \mathbb {GL}_{m+T}( \mathbb {Q}\langle {\mathbf {x}}\rangle )\) and \(Q' \in \mathbb {GL}_{n+1+(T-1)}( \mathbb {Q}\langle {\mathbf {x}}\rangle ) = \mathbb {GL}_{n+T}( \mathbb {Q}\langle {\mathbf {x}}\rangle )\) such that \(P'(B \oplus I_{T-1})Q'\) is a linear matrix. Since

$$\begin{aligned} B \oplus I_{T-1}&= \left( I_{m-1} \oplus \begin{pmatrix} 1 &{}\quad b \\ 0 &{}\quad 1 \end{pmatrix} \oplus I_{T-1} \right) (A \oplus I_T) \left( I_{n-1} \oplus \begin{pmatrix} 1 &{}\quad 0 \\ -c &{}\quad 1 \end{pmatrix} \oplus I_{T-1} \right) \\&= R(A \oplus I_T)S, \end{aligned}$$

where \(R = \left( I_{m-1} \oplus \begin{pmatrix} 1 &{} b \\ 0 &{} 1 \end{pmatrix} \oplus I_{T-1} \right) \in \mathbb {GL}_{m+T}( \mathbb {Q}\langle {\mathbf {x}}\rangle )\) and \(S = \left( I_{n-1} \oplus \begin{pmatrix} 1 &{} 0 \\ -c &{} 1 \end{pmatrix} \oplus I_{T-1} \right) \in \mathbb {GL}_{n+T}( \mathbb {Q}\langle {\mathbf {x}}\rangle )\), we have that

$$\begin{aligned} P'(B \oplus I_{T-1})Q' = (P'R) (A \oplus I_T) (SQ'). \end{aligned}$$

Setting \(P = P'R\) and \(Q = SQ'\) proves the inductive step and completes the proof. Since we only use subformulas of the formulas computing the entries of A, the bound on the bit complexity does not change. \(\square \)

1.2 Classical Reduction

Having shown the effective version of Higman’s trick, we can now compute the \({\text{ nc-rank }}\) of a matrix over \( \mathbb {Q}\langle {\mathbf {x}}\rangle \). We begin with a lemma which will tell us that we can reduce the problem of computing the \({\text{ nc-rank }}\) of a matrix by testing fullness of a smaller matrix with polynomial entries.

Lemma A.3

(Reduction to Fullness Testing) Let \(M \in \mathbb {F}\langle {\mathbf {x}}\rangle ^{m \times n}\) be any matrix. In addition, let \(U = (u_{ij})\) and \(V = (v_{ij})\) be generic matrices in new, non-commuting variables \(u_{ij}\) and \(v_{ij}\), of dimensions \(r \times m\) and \(n \times r\), respectively. Then, \({\text{ nc-rank }}(M) \ge r\) iff the matrix UMV is full.

Proof

Since \({\text{ nc-rank }}(M) \ge r\), there exists an \(r \times r\) minor of M of full rank. Let Q be such a minor of M. W.l.o.g.,Footnote 28 we can assume that Q is the \([r] \times [r]\) principal minor of M. Hence, we have that

$$\begin{aligned} UMV = \begin{pmatrix} U_1&\quad U_2 \end{pmatrix} \begin{pmatrix} Q &{}\quad M_2 \\ M_3 &{}\quad M_4 \end{pmatrix} \begin{pmatrix} V_1 \\ V_2 \end{pmatrix}, \end{aligned}$$

where \(U_1\) and \(V_1\) are \(r \times r\) matrices and the others are matrices with the proper dimensions.

Letting \(U' = \begin{pmatrix} I_r&0 \end{pmatrix}\) and \(V' = \begin{pmatrix} I_r \\ 0 \end{pmatrix}\), the equality above becomes:

$$\begin{aligned} U'MV' = Q. \end{aligned}$$

As

$$\begin{aligned} r \ge {\text{ nc-rank }}(UMV) \ge {\text{ nc-rank }}(U'MV') = {\text{ nc-rank }}(Q) = r, \end{aligned}$$

we obtain that UMV is full, as we wanted. Notice that the second inequality comes from the fact that rank does not increase after restrictions of the new variables. \(\square \)

Notice that we do not know the rank \({\text{ nc-rank }}(M)\) a priori. Therefore, our algorithm will try all possible values of \(r \in [n]\) and output the maximum value of r for which we find a full matrix.

For each \(r \times r\) matrix UMV, we can use the effective Higman’s trick to convert UMV into a \(s \times s\) matrix with linear entries. With this matrix, we can use the truncated Gurvits’ algorithm to check whether the matrix we just obtained is full. Since we have this test, we will be able to output the correct rank. Algorithm 5 is the precise formulation of the procedure just described.

figure f

Theorem A.4

Let \(M \in \mathbb {Q}\langle {\mathbf {x}}\rangle ^{m \times n}\) be s.t. each entry of M is a polynomial computed by a formula of size bounded by s and bit complexity bounded by b. There exists a deterministic algorithm that finds the non-commutative rank of M in time \({\text{ poly }}(m, n, s, b)\).

Proof

To prove this theorem, it is enough to show that Algorithm 5 is correct and it runs with the desired runtime.

Without loss of generality, we can assume that \(n \le m\). Therefore we have that \({\text{ nc-rank }}(M) \le n\). By Lemma A.3, if \(r \le {\text{ nc-rank }}(M)\), then matrix \(M_r\) will be of full rank (and therefore will not have a shrunk subspace, by Theorem 1.4). Since \(M_r = U_r M V_r\), from the formulas computing the entries of M, we obtain formulas of size at most 2smn computing the entries of \(M_r\). Moreover, the bit complexities of these formulas will still be bounded by b, as multiplication by generic matrices do not mix any of the polynomials of M.

By Proposition A.2 and the fact that the size of the formulas computing the entries of \(M_r\) are bounded by 2smn, we have that \(N_r\) is a linear matrix of dimensions \((k+r) \times (k+r)\), where \(k \le 2s(mn)^2\) and the bit complexity of the coefficients bounded by b. Moreover, \(N_r = P (M_r \oplus I_{k}) Q\) implies that \(N_r\) is full if, and only if, \(M_r\) is full, which is true if, and only if, \({\text{ nc-rank }}(M) \ge r\).

Now, by Theorem 1.1, we have a deterministic polynomial time algorithm to determine whether \(N_r\) is full rank. If \(r \le {\text{ nc-rank }}(M)\), \(N_r\) will be full, and the maximum such r will be exactly when \(r = {\text{ nc-rank }}(M)\). Therefore, by outputting the maximum r for which \(N_r\) we compute \({\text{ nc-rank }}(M)\). This proves that our algorithm is correct. Notice that the runtime is polynomial in the input size, as we perform at most n applications of the Higman trick and of Algorithm \(G'\). This completes the proof. \(\square \)

1.3 The Quantum Reduction

Here we present a different reduction from computing non-commutative rank to fullness testing from a quantum viewpoint. We will only work with square matrices though. As we saw, by Higman’s trick, we can assume the matrices to be linear. So we are given a matrix \(L = \sum _{i=1}^m x_i A_i \in M_n( \mathbb {F}\langle {\mathbf {x}}\rangle )\). A combination of Theorems 1.4 and 1.17 shows that \({\text{ nc-rank }}(L) \le r\) iff the operator defined by \(A_1,\ldots ,A_m\) is \(n-r\)-rank-decreasing. So we just want to check whether a completely positive operator is c-rank-decreasing, and we will do this by using an algorithm for checking if an operator is rank-decreasing as a black box using the following lemma:

Lemma A.5

Let \(T : M_n(\mathbb {C}) \rightarrow M_n(\mathbb {C})\) be a completely positive operator. Define an operator \(\overline{T} : M_{n+c-1}(\mathbb {C}) \rightarrow M_{n+c-1}(\mathbb {C})\) as follows:

$$\begin{aligned} \overline{T} \left( \begin{bmatrix} X_{1,1}&\quad X_{1,2} \\ X_{2,1}&\quad X_{2,2} \end{bmatrix} \right) = \begin{bmatrix} T(X_{1,1}) + \text {tr}(X_{2,2})I_n&\quad 0 \\ 0&\quad \text {tr}(X_{1,1})I_{c-1} \end{bmatrix} \end{aligned}$$

Here \(X_{1,1}\), \(X_{1,2}\), \(X_{2,1}\), \(X_{2,2}\) are \(n \times n\), \(n \times c-1\), \(c-1 \times n\), \(c-1 \times c-1\) matrices, respectively. Then \(\overline{T}\) is completely positive and T is c-rank-decreasing iff \(\overline{T}\) is rank-decreasing. Note that we are considering \(c \le n\).

Proof

A well known characterization due to Choi [12] states that \(\overline{T}\) is completely positive iff \(\sum _{i,j = 1}^{n+c-1} E_{i,j} \otimes \overline{T}(E_{i,j})\) is psd. Here \(E_{i,j}\) is the matrix with 1 at ij position and 0 everywhere else. Now

$$\begin{aligned}&\overline{T}(E_{i,j}) \\&\quad = {\left\{ \begin{array}{ll} \begin{bmatrix} T(E_{i,j}) &{}\quad 0 \\ 0 &{}\quad I_{c-1} \end{bmatrix} &{}\quad 1 \le i=j \le n \\ \\ \begin{bmatrix} T(E_{i,j}) &{}\quad 0 \\ 0 &{}\quad 0 \end{bmatrix} &{} 1 \le i, j \le n, i \ne j \\ \\ \begin{bmatrix} 0 &{}\quad 0 \\ 0 &{}\quad 0 \end{bmatrix} &{}\quad 1 \le i \le n, n+1 \le j \le n+c-1 \, \text {or} \, n+1 \le i \le n+c-1, 1 \le j \le n \\ \\ \begin{bmatrix} I_n &{}\quad 0 \\ 0 &{}\quad 0 \end{bmatrix} &{} n+1 \le i=j \le n+c-1 \\ \\ \begin{bmatrix} 0 &{}\quad 0 \\ 0 &{}\quad 0 \end{bmatrix} &{} n+1 \le i, j \le n+c-1, i \ne j \\ \end{array}\right. } \end{aligned}$$

From here, it is easy to verify that \(\sum _{i,j = 1}^{n+c-1} E_{i,j} \otimes \overline{T}(E_{i,j})\) is psd given that \(\sum _{i,j=1}^n E_{i,j} \otimes T(E_{i,j})\) is psd. Now suppose that \(\overline{T}\) is rank-decreasing. This can only happen if \(X_{1,1} = 0\) or \(X_{2,2} = 0\), otherwise

$$\begin{aligned} \overline{T} \left( \begin{bmatrix} X_{1,1}&\quad X_{1,2} \\ X_{2,1}&\quad X_{2,2} \end{bmatrix} \right) = \begin{bmatrix} T(X_{1,1}) + \text {tr}(X_{2,2})I_n&0 \\ 0&\quad \text {tr}(X_{1,1})I_{c-1} \end{bmatrix} \end{aligned}$$

is full rank. If \(X_{1,1} = 0\), then

$$\begin{aligned} \begin{bmatrix} 0&X_{1,2} \\ X_{2,1}&X_{2,2} \end{bmatrix} \end{aligned}$$

can be psd (and Hermitian) only if \(X_{1,2} = X_{2,1} = 0\). In this case a \(c-1\) ranked matrix is mapped to rank n matrix. So \(X_{2,2}\) has to be zero. Then again by the psd condition \(X_{1,2} = X_{2,1} = 0\). So

$$\begin{aligned} \overline{T} \left( \begin{bmatrix} X_{1,1}&\quad 0 \\ 0&\quad 0 \end{bmatrix} \right) = \begin{bmatrix} T(X_{1,1})&\quad 0 \\ 0&\quad \text {tr}(X_{1,1})I_{c-1} \end{bmatrix} \end{aligned}$$

and \(X_{1,1} \ne 0\) and

$$\begin{aligned} \text {Rank} \left( \begin{bmatrix} X_{1,1}&\quad 0 \\ 0&\quad 0 \end{bmatrix} \right) > \text {Rank} \left( \begin{bmatrix} T(X_{1,1})&\quad 0 \\ 0&\quad \text {tr}(X_{1,1})I_{c-1} \end{bmatrix} \right) \end{aligned}$$

Hence \(\text {Rank}(T(X_{1,1})) \le \text {Rank}(X_{1,1}) - c\). This proves one direction. Now suppose that T is c-rank-decreasing and \(\text {Rank}(T(X)) \le \text {Rank}(X)-c\), then

$$\begin{aligned} \text {Rank} \left( \overline{T} \left( \begin{bmatrix} X&\quad 0 \\ 0&\quad 0 \end{bmatrix} \right) \right) = \text {Rank} \left( \begin{bmatrix} T(X)&\quad 0 \\ 0&\quad \text {tr}(X) I_{c-1} \end{bmatrix} \right) < \text {Rank} \left( \begin{bmatrix} X&\quad 0 \\ 0&\quad 0 \end{bmatrix} \right) \end{aligned}$$

This proves the lemma. \(\square \)

Remark A.6

This seems to be the “quantum” analogue of obtaining a maximum matching oracle based on a perfect matching oracle: add c-1 dummy vertices to both sides of the bipartite graph and connect them to everything. Then the new graph has a perfect matching iff the original graph had a matching of size \(\ge n - c + 1\).

Remark A.7

Here we didn’t specify a set of Kraus operators for the operator \(\overline{T}\) which seem to be needed to run Algorithms 1 and 2 but Kraus operators can be obtained by looking at the eigenvectors of \(\sum _{i,j=1}^{n+c-1} E_{i,j} \otimes \overline{T}(E_{i,j})\). Alternatively Algorithms 1 and 2 can also be interpreted as acting directly on the Choi-Jamiolkowski state of \(\overline{T}\), i.e., \(\sum _{i,j=1}^{n+c-1} E_{i,j} \otimes \overline{T}(E_{i,j})\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Garg, A., Gurvits, L., Oliveira, R. et al. Operator Scaling: Theory and Applications. Found Comput Math 20, 223–290 (2020). https://doi.org/10.1007/s10208-019-09417-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10208-019-09417-z

Keywords

Mathematics Subject Classification

Navigation