Skip to main content
Log in

Approximating the little Grothendieck problem over the orthogonal and unitary groups

  • Full Length Paper
  • Series A
  • Published:
Mathematical Programming Submit manuscript

Abstract

The little Grothendieck problem consists of maximizing \(\sum _{ij}C_{ij}x_ix_j\) for a positive semidefinite matrix C, over binary variables \(x_i\in \{\pm 1\}\). In this paper we focus on a natural generalization of this problem, the little Grothendieck problem over the orthogonal group. Given \(C\in \mathbb {R}^{dn\times dn}\) a positive semidefinite matrix, the objective is to maximize \(\sum _{ij}{\text {tr}}\left( C^T_{ij}O_iO_j^T\right) \) restricting \(O_i\) to take values in the group of orthogonal matrices \(\mathcal {O}_d\), where \(C_{ij}\) denotes the (ij)-th \(d\times d\) block of C. We propose an approximation algorithm, which we refer to as Orthogonal-Cut, to solve the little Grothendieck problem over the group of orthogonal matrices \(\mathcal {O}_d\) and show a constant approximation ratio. Our method is based on semidefinite programming. For a given \(d\ge 1\), we show a constant approximation ratio of \(\alpha _{\mathbb {R}}(d)^2\), where \(\alpha _{\mathbb {R}}(d)\) is the expected average singular value of a \(d\times d\) matrix with random Gaussian \(\mathcal {N}\left( 0,\frac{1}{d}\right) \) i.i.d. entries. For \(d=1\) we recover the known \(\alpha _{\mathbb {R}}(1)^2=2/\pi \) approximation guarantee for the classical little Grothendieck problem. Our algorithm and analysis naturally extends to the complex valued case also providing a constant approximation ratio for the analogous little Grothendieck problem over the Unitary Group \(\mathcal {U}_d\). Orthogonal-Cut also serves as an approximation algorithm for several applications, including the Procrustes problem where it improves over the best previously known approximation ratio of \(\frac{1}{2\sqrt{2}}\). The little Grothendieck problem falls under the larger class of problems approximated by a recent algorithm proposed in the context of the non-commutative Grothendieck inequality. Nonetheless, our approach is simpler and provides better approximation with matching integrality gaps. Finally, we also provide an improved approximation algorithm for the more general little Grothendieck problem over the orthogonal (or unitary) group with rank constraints, recovering, when \(d=1\), the sharp, known ratios.

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.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. We also note that these semidefinite programs satisfy Slater’s condition as the identity matrix is a feasible point. This ensures strong duality, which can be exploited by many semidefinite programming solvers.

  2. These ideas also play a major role in the unidimensional complex case treated by Man-Cho So et al. [25].

  3. The additional constraint that forces a matrix to be in the special orthogonal or unitary group is having determinant equal to 1 which is not quadratic.

References

  1. Alizadeh, F., Haeberly, J.-P.A., Overton, M.L.: Primal-dual interior-point methods for semidefinite programming: convergence rates, stability and numerical results. SIAM J. Optim. 8(3), 746–768 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  2. Alon, N., Makarychev, K., Makarychev, Y., Naor, A.: Quadratic forms on graphs. Invent. Math. 163, 486–493 (2005)

    MATH  Google Scholar 

  3. Alon, N., Naor, A.: Approximating the cut-norm via Grothendieck’s inequality. In: Proceedings of the 36 th ACM STOC, pp. 72–80. ACM Press (2004)

  4. Abramowitz, M., Stegun, I.A.: Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. Dover, New York (1964)

    MATH  Google Scholar 

  5. Bandeira, A.S.: Convex relaxations for certain inverse problems on graphs. Ph.D. thesis, Program in Applied and Computational Mathematics, Princeton University (2015)

  6. Briet, J., Buhrman, H., Toner, B.: A generalized Grothendieck inequality and nonlocal correlations that require high entanglement. Commun. Math. Phys. 305(3), 827–843 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  7. Briet, J., Filho, F.M.O., Vallentin, F.: The positive semidefinite Grothendieck problem with rank constraint. In: Automata, Languages and Programming, vol. 6198 of Lecture Notes in Computer Science, pp. 31–42. Springer, Berlin (2010)

  8. Briet, J., Regev, O., Saket, R.: Tight hardness of the non-commutative Grothendieck problem. In: 2015 IEEE 56th Annual Symposium on Foundations of Computer Science (FOCS), pp. 1108–1122. doi:10.1109/FOCS.2015.72 (2015)

  9. Bandeira, A.S., Singer, A., Spielman, D.A.: A Cheeger inequality for the graph connection Laplacian. SIAM J. Matrix Anal. Appl. 34(4), 1611–1630 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  10. Ben-Tal, A., Nemirovski, A.: On tractable approximations of uncertain linear matrix inequalities affected by interval uncertainty. SIAM J. Optim. 12, 811–833 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  11. Carlen, E.A.: Trace inequalities and quantum entropy: an introductory course. http://www.ueltschi.org/azschool/notes/ericcarlen.pdf (2009)

  12. Couillet, R., Debbah, M.: Random Matrix Methods for Wireless Communications. Cambridge University Press, New York (2011)

    Book  MATH  Google Scholar 

  13. Chaudhury, K.N., Khoo, Y., Singer, A.: Global registration of multiple point clouds using semidefinite programming. SIAM J. Optim. 25(1), 126–185 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  14. Charikar, M., Wirth, A.: Maximizing quadratic programs: extending Grothendieck’s inequality. In: Proceedings of the 45th Annual IEEE Symposium on Foundations of Computer Science, FOCS ’04, pp. 54–60. IEEE Computer Society, Washington (2004)

  15. Fan, K., Hoffman, A.J.: Some metric inequalities in the space of matrices. Proc. Am. Math. Soc. 6(1), 111–116 (1955)

    Article  MathSciNet  MATH  Google Scholar 

  16. Gradshteyn, I.S., Ryzhik, I.M.: Table of Integrals, Series, and Products, Fifth Edition, 5th edn. Academic Press, Cambridge (1994)

    MATH  Google Scholar 

  17. Grothendieck, A.: Resume de la theorie metrique des produits tensoriels topologiques. Bol. Soc. Mat. Sao Paulo, p. 179 (1996). (French)

  18. Gotze, F., Tikhomirov, A.: On the rate of convergence to the Marchenko–Pastur distribution. arXiv:1110.1284 [math.PR] (2011)

  19. Goemans, M.X., Williamson, D.P.: Improved apprximation algorithms for maximum cut and satisfiability problems using semidefine programming. J. Assoc. Comput. Mach. 42, 1115–1145 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  20. Higham, N.J.: Computing the polar decomposition-with applications. SIAM J. Sci. Stat. Comput. 7, 1160–1174 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  21. Keller, J.B.: Closest unitary, orthogonal and hermitian operators to a given operator. Math. Mag. 48(4), 192–197 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  22. Khot, S.: On the unique games conjecture (invited survey). In: Proceedings of the 2010 IEEE 25th Annual Conference on Computational Complexity, CCC ’10, pp. 99–121. IEEE Computer Society, Washington (2010)

  23. Leveque, O.: Random matrices and communication systems: Wishart random matrices: marginal eigenvalue distribution. http://ipg.epfl.ch/~leveque/Matrix/ (2012)

  24. Livan, G., Vivo, P.: Moments of Wishart–Laguerre and Jacobi ensembles of random matrices: application to the quantum transport problem in chaotic cavities. Acta Phys. Pol. B 42, 1081 (2011)

    Article  MathSciNet  Google Scholar 

  25. Man-Cho So, A., Zhang, J., Ye, Y.: On approximating complex quadratic optimization problems via semidefinite programming relaxations. Math. Program. 110(1), 93–110 (2007)

  26. Nemirovski, A.: Sums of random symmetric matrices and quadratic optimization under orthogonality constraints. Math. Program. 109(2–3), 283–317 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  27. Nesterov, Y.: Semidefinite relaxation and nonconvex quadratic optimization. Optim. Methods Softw. 9(1–3), 141–160 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  28. Nesterov, Y.: Introductory Lectures on Convex Optimization: A Basic Course, vol. 87 of Applied Optimization. Springer, Berlin (2004)

    Book  MATH  Google Scholar 

  29. Nemirovski, A., Roos, C., Terlaky, T.: On maximization of quadratic form over intersection of ellipsoids with common center. Math. Program. 86(3), 463–473 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  30. Naor, A., Regev, O., Vidick, T.: Efficient rounding for the noncommutative Grothendieck inequality. In: Proceedings of the 45th annual ACM Symposium on Symposium on theory of computing, STOC ’13, pp. 71–80. ACM, New York (2013)

  31. Pisier, G.: Grothendieck’s theorem, past and present. Bull. Am. Math. Soc. 49, 237323 (2011)

    MathSciNet  Google Scholar 

  32. Raghavendra, P.: Optimal algorithms and inapproximability results for every CSP. In: Proceedings of 40th ACM STOC, pp. 245–254 (2008)

  33. Schonemann, P.H.: A generalized solution of the orthogonal procrustes problem. Psychometrika 31(1), 1–10 (1966)

    Article  MathSciNet  MATH  Google Scholar 

  34. Shen, J.: On the singular values of Gaussian random matrices. Linear Algebra Appl. 326(13), 1–14 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  35. Singer, A.: Angular synchronization by eigenvectors and semidefinite programming. Appl. Comput. Harmon. Anal. 30(1), 20–36 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  36. So, A.-C.: Moment inequalities for sums of random matrices and their applications in optimization. Math. Program. 130(1), 125–151 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  37. Singer, A., Shkolnisky, Y.: Three-dimensional structure determination from common lines in Cryo-EM by eigenvectors and semidefinite programming. SIAM J. Imaging Sci. 4(2), 543–572 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  38. Tulino, A.M., Verdú, S.: Random matrix theory and wireless communications. Commun. Inf. Theory 1(1), 1–182 (2004)

    MATH  Google Scholar 

  39. Vanderberghe, L., Boyd, S.: Semidefinite programming. SIAM Rev. 38, 49–95 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  40. Vershynin, R.: Introduction to the non-asymptotic analysis of random matrices. In: Eldar, Y., Kutyniok, G. (eds.) Chapter 5 of: Compressed Sensing, Theory and Applications. Cambridge University Press, Cambridge (2012)

    Google Scholar 

  41. Wen, Z., Goldfarb, D., Scheinberg, K.: Block coordinate descent methods for semidefinite programming. In: Handbook on Semidefinite, Conic and Polynomial Optimization, vol. 166 of International Series in Operations Research & Management Science, pp. 533–564. Springer, US (2012)

Download references

Acknowledgments

The authors would like to thank Moses Charikar for valuable guidance in context of this work and Jop Briet, Alexander Iriza, Yuehaw Khoo, Dustin Mixon, Oded Regev, and Zhizhen Zhao for insightful discussions on the topic of this paper. Special thanks to Johannes Trost for a very useful answer to a Mathoverflow question posed by the first author. Finally, we would like to thank the reviewers for numerous suggestions that helped to greatly improve the quality of this paper.

A. S. Bandeira was supported by AFOSR Grant No. FA9550-12-1-0317. A. Singer was partially supported by Award Number FA9550-12-1-0317 and FA9550-13-1-0076 from AFOSR, by Award Number R01GM090200 from the NIGMS, and by Award Number LTR DTD 06-05-2012 from the Simons Foundation. Parts of this work have appeared in C. Kennedy’s senior thesis at Princeton University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Afonso S. Bandeira.

Additional information

Most of this work was done while ASB was at Princeton University.

Appendices

Appendix 1: Technical proofs—analysis of algorithm for the Stiefel Manifold setting

Lemma 17

Let \(r\ge d\). Let G be a \(d\times r\) Gaussian random matrix with real valued i.i.d. \(\mathcal {N}\left( 0,\frac{1}{r}\right) \) entries and let \(\alpha _{\mathbb {R}}(d,r)\) as defined in Definition 11. Then,

$$\begin{aligned} \mathbb {E}\left( \mathcal {P}_{d,r}(G) G^T\right) = \mathbb {E}\left( G\mathcal {P}_{d,r}(G)^T\right) = \alpha _{\mathbb {R}}(d,r)I_{d\times d}. \end{aligned}$$

Furthermore, if G is a \(d\times r\) Gaussian random matrix with complex valued i.i.d. \(\mathcal {N}\left( 0,\frac{1}{r}\right) \) entries and \(\alpha _{\mathbb {C}}(d,r)\) the analogous constant (Definition 11), then

$$\begin{aligned} \mathbb {E}\left( \mathcal {P}_{d,r}(G) G^H\right) = \mathbb {E}\left( G\mathcal {P}_{d,r}(G)^H\right) = \alpha _{\mathbb {C}}(d,r)I_{d\times d}. \end{aligned}$$

The proof of this Lemma is a simple adaptation of the proof of Lemma 6.

Proof

We restrict the presentation to the real case. As before, all the arguments are equivalent to the complex case, replacing all transposes with Hermitian adjoints and \(\alpha _{\mathbb {R}}(d,r)\) with \(\alpha _{\mathbb {C}}(d,r)\).

Let \(G = U [\Sigma \ 0] V^T\) be the singular value decomposition of G. Since \(GG^T = U\Sigma ^2 U^T\) is a Wishart matrix, it is well known that its eigenvalues and eigenvectors are independent and U is distributed according to the Haar measure in \(\mathcal {O}_d\) (see e.g. Lemma 2.6 in [38]). To resolve ambiguities, we consider \(\Sigma \) ordered such that \(\Sigma _{11} \ge \Sigma _{22} \ge \cdots \ge \Sigma _{dd}\).

Let \(Y = \mathcal {P}_{(d,r)}(G) G^T\). Since

$$\begin{aligned} \mathcal {P}_{(d,r)}(G) = \mathcal {P}_{(d,r)}(U [\Sigma \ 0] V^T) = U[I_{d\times d} \ 0]V^T, \end{aligned}$$

we have

$$\begin{aligned} Y = \mathcal {P}_{(d,r)}(U [\Sigma \ 0]V^T) (U [\Sigma \ 0] V^T)^T = U [I_{d\times d} \ 0]V^T V \Sigma U^T = U \Sigma U^T. \end{aligned}$$

Note that \( G\mathcal {P}_{(d,r)}(G)^T = U \Sigma U^T = Y\).

Since \(Y_{ij} = u_i \Sigma u_j^T\), where \(u_1,\ldots ,u_d\) are the rows of U, and U is distributed according to the Haar measure, we have that \(u_j\) and \(-u_j\) have the same distribution conditioned on any \(u_i\), for \(i\ne j\), and \(\Sigma \). This implies that, if \(i\ne j, Y_{ij} = u_i \Sigma u_j^T\) is a symmetric random variable, and so \(\mathbb {E}Y_{ij} = 0\). Also, \(u_i\sim u_j\) implies that \(Y_{ii}\sim Y_{jj}\). This means that \(\mathbb {E}Y = c I_{d\times d}\) for some constant c. To obtain c,

$$\begin{aligned} c = c\frac{1}{d}{\text {tr}}(I_{d\times d}) = \frac{1}{d}\mathbb {E}{\text {tr}}(Y) = \frac{1}{d}\mathbb {E}{\text {tr}}(U \Sigma U^T) = \frac{1}{d}\mathbb {E}\sum _{k=1}^n\sigma _{k}(G) = \alpha _{\mathbb {R}}(d,r), \end{aligned}$$

which shows the lemma. \(\square \)

Lemma 18

Let \(r\ge d\). Let \(M,N\in \mathbb {R}^{d\times nd}\) such that \(MM^T = NN^T = I_{d \times d}\). Let \(R \in \mathbb {R}^{nd \times r}\) be a Gaussian random matrix with real valued i.i.d. entries \(\mathcal {N}\left( 0, \frac{1}{r} \right) \). Then

$$\begin{aligned} \mathbb {E}\left[ \mathcal {P}_{(d,r)}(M R) (N R)^T\right] = \mathbb {E}\left[ (M R) \mathcal {P}_{(d,r)}(N R)^T\right] = \alpha _{\mathbb {R}}(d,r) MN^T, \end{aligned}$$

where \(\alpha _{\mathbb {R}}(d,r)\) is the constant in Definition 11.

Analogously, if \(M,N\in \mathbb {C}^{d\times nd}\) such that \(MM^H = NN^H = I_{d \times d}\), and \(R \in \mathbb {C}^{nd \times r}\) is a Gaussian random matrix with complex valued i.i.d. entries \(\mathcal {N}\left( 0, \frac{1}{r} \right) \), then

$$\begin{aligned} \mathbb {E}\left[ \mathcal {P}_{(d,r)}(M R) (N R)^H\right] = \mathbb {E}\left[ (M R) \mathcal {P}_{(d,r)}(N R)^H\right] = \alpha _{\mathbb {C}}(d,r) MN^H, \end{aligned}$$

where \(\alpha _{\mathbb {C}}(d,r)\) is the constant in Definition 11.

Similarly to above, the proof of this Lemma is a simple adaptation of the proof of Lemma 5.

Proof

We restrict the presentation of proof to the real case. Nevertheless, all the arguments trivially adapt to the complex case by, essentially, replacing all transposes with Hermitian adjoints and \(\alpha _{\mathbb {R}}(d)\) and \(\alpha _{\mathbb {R}}(d,r)\) with \(\alpha _{\mathbb {C}}(d)\) and \(\alpha _{\mathbb {C}}(d,r)\).

Let \(A = \left[ M^T\text { } N^T\right] \in \mathbb {R}^{dn\times 2d}\) and \(A=QB\) be the QR decomposition of A with \(Q\in \mathbb {R}^{nd\times nd}\) an orthogonal matrix and \(B \in \mathbb {R}^{nd \times 2d}\) upper triangular with non-negative diagonal entries; note that only the first 2d rows of B are nonzero. We can write

$$\begin{aligned} Q^TA = B = \begin{bmatrix} B_{11}&\quad B_{12}\\ 0_d&\quad B_{22}\\ 0_d&\quad 0_d\\\ \vdots&\quad \vdots \\ 0_d&\quad 0_d \end{bmatrix} \in \mathbb {R}^{dn \times 2d}, \end{aligned}$$

where \(B_{11}\in \mathbb {R}^{d\times d}\) and \(B_{22}\in \mathbb {R}^{d\times d}\) are upper triangular matrices with non-negative diagonal entries. Since

$$\begin{aligned} (Q^T M^T)^T_{11}(Q^T M^T)_{11}= & {} (Q^T M^T)^T(Q^T M^T) = M Q Q^T M^T = M I_{nd \times nd} M^T\\= & {} M M^T = I_{d \times d}, \end{aligned}$$

\(B_{11} = (Q^T M^T)_{11}\) is an orthogonal matrix, which together with the non-negativity of the diagonal entries (and the fact that \(B_{11}\) is upper-triangular) forces \(B_{11}\) to be \(B_{11} = I_{d\times d}\).

Since R is a Gaussian matrix and Q is an orthogonal matrix, \(QR \sim R\) which implies

$$\begin{aligned} \mathbb {E}\left[ \mathcal {P}_{(d,r)}(M R) (N R)^T\right] = \mathbb {E}\left[ \mathcal {P}_{(d,r)}(M Q R)(NQ R)^T \right] . \end{aligned}$$

Since \(MQ=[B_{11}^T, 0_{d\times d},\ldots ,0_{d\times d}] = [I_{d\times d}, 0_{d\times d},\ldots ,0_{d\times d}]\) and \(NQ = [B_{12}^T, B_{22}^T, 0_{d\times d},\ldots ,0_{d\times d}]\),

$$\begin{aligned} \mathbb {E}\left[ \mathcal {P}_{(d,r)}(M R) (N R)^T\right] = \mathbb {E}\left[ \mathcal {P}_{(d,r)}(R_1)(B_{12}^T R_1 + B_{22}^T R_2)^T\right] , \end{aligned}$$

where \(R_1\) and \(R_2\) are the first two \(d\times r\) blocks of R. Since these blocks are independent, the second term vanishes and we have

$$\begin{aligned} \mathbb {E}\left[ \mathcal {P}_{(d,r)}(M R) (N R)^T\right] = \mathbb {E}\left[ \mathcal {P}_{(d,r)}(R_1) R_1^T\right] B_{12}. \end{aligned}$$

The Lemma now follows from using Lemma 17 to obtain \(\mathbb {E}\left[ \mathcal {P}_{(d,r)}(R_1) R_1^ T\right] = \alpha _{\mathbb {R}}(d,r)I_{d\times d}\) and noting that \(B_{12} = (Q^TM^T)^T(Q^TN^T) = MN^T\).

The same argument, with \(Q'B'\) the QR decomposition of \(A' = \left[ N^T M^T\right] \in \mathbb {R}^{dn\times 2d}\) instead, shows

$$\begin{aligned} \mathbb {E}\left[ (M R) \mathcal {P}_{(d,r)}(N R)^T\right] = \mathbb {E}\left[ R_1\mathcal {P}_{(d,r)}(R_1)^T\right] MN^T = \alpha _{\mathbb {R}}(d,r)MN^T. \end{aligned}$$

\(\square \)

Appendix 2: Bounds for the average singular value

Lemma 19

Let \(G_{\mathbb {C}} \in \mathbb {C}^{d\times d}\) be a Gaussian random matrix with i.i.d. complex valued \(\mathcal {N}(0,d^{-1})\) entries and define \(\alpha _{\mathbb {C}}(d):= \mathbb {E}\left[ \frac{1}{d} \sum _{j=1}^d \sigma _j(G_{\mathbb {C}})\right] \). We have the following bound

$$\begin{aligned} \alpha _{\mathbb {C}}(d) \ge \frac{8}{3\pi } - \frac{5.05}{d}. \end{aligned}$$

Proof

We express \(\alpha _{\mathbb {C}}(d)\) as sums and products of Gamma functions and then use classical bounds to obtain our result.

Recall that from equation (16),

$$\begin{aligned} \alpha _{\mathbb {C}}(d) = d^{-3/2} \sum _{n=0}^{d-1} T_n, \end{aligned}$$
(26)

where

$$\begin{aligned} T_n = \int _0^{\infty } x^{1/2} e^{-x} L_n(x)^2 dx, \end{aligned}$$

and \(L_n(x)\) is the nth Laguerre polynomial,

$$\begin{aligned} L_n(x) = \sum _{k=0}^n \left( {\begin{array}{c}n\\ k\end{array}}\right) \frac{(-1)^k}{k!} x^k. \end{aligned}$$

This integral can be expressed as [see [16] section 7.414 equation 4(1)]

$$\begin{aligned} T_n = \frac{ \Gamma (n+3/2)}{\Gamma (n+1)} \sum _{m=0}^n \frac{\left( \frac{-1}{2}\right) _m (-n)_m}{(m!)^2 (-n-\frac{1}{2})_m}, \end{aligned}$$
(27)

where \((x)_m\) is the Pochhammer symbol

$$\begin{aligned} (x)_m = \frac{\Gamma (x+m)}{\Gamma (x)}. \end{aligned}$$

The next lemma states a couple basic facts about the Gamma function that we will need in the subsequent computations. \(\square \)

Lemma 20

The Gamma function satisfies the following inequalities:

$$\begin{aligned} \frac{1}{\sqrt{n}} \le \frac{\Gamma (n)}{\Gamma (n+1/2)} \le \frac{1}{\sqrt{n-1/2}} \end{aligned}$$
$$\begin{aligned} \sqrt{n} \le \frac{\Gamma (n+1)}{\Gamma (n+1/2)} \le \sqrt{n+1/2}. \end{aligned}$$

Proof

See [4] page 255. \(\square \)

We want to bound the summation in (27), which we rewrite as

$$\begin{aligned} \sum _{m=0}^n \frac{\left( \frac{-1}{2}\right) _m (-n)_m}{(m!)^2 (-n-\frac{1}{2})_m}= & {} \sum _{m=0}^{\infty } \frac{(\frac{-1}{2})_m^2}{(m!)^2} - \sum _{m= n+1}^{\infty }\frac{(\frac{-1}{2})_m^2}{(m!)^2}\\&-\sum _{m=0}^n \frac{(\frac{-1}{2})_m^2}{(m!)^2} \left( 1 - \frac{(-n)_m}{(-n-\frac{1}{2})_m} \right) . \end{aligned}$$

For simplicity define

$$\begin{aligned} (I):= & {} \sum _{m=0}^{\infty } \frac{(\frac{-1}{2})^2_m}{(m!)^2}\\ (II):= & {} \sum _{m= n+1}^{\infty }\frac{(\frac{-1}{2})_m^2}{(m!)^2}\\ (III):= & {} \sum _{m=0}^n \frac{(\frac{-1}{2})_m^2}{(m!)^2} \left( 1 - \frac{(-n)_m}{(-n-\frac{1}{2})_m} \right) , \end{aligned}$$

so that (27) becomes

$$\begin{aligned} T_n = \frac{\Gamma (n+3/2)}{\Gamma (n+1)} ((I) + (II) + (III)). \end{aligned}$$

The first term we can compute explicitly (see [16]) as

$$\begin{aligned} (I) = \frac{4}{\pi }. \end{aligned}$$

For the second term we use the fact that \((\frac{-1}{2})_m = \Gamma (m-1/2)/\Gamma (-1/2)\) to get

$$\begin{aligned} (II)= \sum _{m = n+1}^{\infty } \frac{1}{\Gamma (-1/2)^2} \frac{\Gamma (m-1/2)^2}{\Gamma (m+1)^2} = \frac{1}{4\pi } \sum _{m=n+1}^{\infty } \frac{\Gamma (m-1/2)^2}{\Gamma (m+1)^2}. \end{aligned}$$

Using the first inequality in Lemma 20 and the multiplication formula for the Gamma function,

$$\begin{aligned} \frac{\Gamma (m-1/2)}{\Gamma (m+1)} = \frac{1}{m-1/2} \frac{\Gamma (m+1/2)}{\Gamma (m+1)} \le \frac{1}{(m-1/2)\sqrt{m}} \end{aligned}$$

so we have

$$\begin{aligned} (II) \le \frac{1}{4\pi } \sum _{m=n+1}^{\infty } \frac{1}{(m-1/2)^2 m} \le \frac{1}{4\pi } \int _{n-1/2}^{\infty } \frac{1}{x^3} dx = \frac{1}{2\pi (2n-1)^2}. \end{aligned}$$

For the third term, we use the formula \((x)_m = \frac{\Gamma (x+n)}{\Gamma (x)}\) to deduce

$$\begin{aligned} (III)&= \sum _{m=0}^n \frac{(\frac{-1}{2})_m^2}{(m!)^2} \left( 1 - \frac{(-n)_m}{(-n-\frac{1}{2})_m} \right) \\&= \frac{1}{4\pi } \sum _{m=0}^n \frac{\Gamma (m-1/2)^2}{\Gamma (m+1)^2} \left( 1 - \frac{\Gamma (n+1) \Gamma (n-m+3/2)}{\Gamma (n+3/2) \Gamma (n-m+1)}\right) \nonumber \\&= \frac{\Gamma (n+1)}{\Gamma (n+3/2)} \frac{1}{4\pi } \sum _{m=0}^{n} \frac{\Gamma (m-1/2)^2}{\Gamma (m+1)^2} \left( \frac{\Gamma (n+3/2)}{\Gamma (n+1)} - \frac{\Gamma (n-m+3/2)}{\Gamma (n-m+1)}\right) . \end{aligned}$$

Using the second bound in Lemma 20,

$$\begin{aligned} \frac{\Gamma (n-m+3/2)}{\Gamma (n-m+1)} \ge \sqrt{n-m+1/2}, \end{aligned}$$

and also

$$\begin{aligned} \frac{\Gamma (n+3/2)}{\Gamma (n+1)} \le \sqrt{n+1}, \end{aligned}$$

so that

$$\begin{aligned} (III) \le \frac{1}{4\pi \sqrt{n+1/2}} \sum _{m=0}^n \left( \frac{1}{(m-1/2)\sqrt{m+1/2}}\right) ^2 \left( \sqrt{n+1} - \sqrt{n-m+1/2}\right) . \end{aligned}$$

If we multiply top and bottom by \(\sqrt{n+1} + \sqrt{n-m+1/2}\) and use the fact that

$$\begin{aligned} \frac{m+1/2}{\sqrt{n+1} + \sqrt{n-m+1/2}} \le \frac{m+1/2}{\sqrt{n+1}}, \end{aligned}$$

then

$$\begin{aligned} (III)&\le \frac{1}{4\pi \sqrt{n+1/2}} \sum _{m=0}^n \frac{1}{(m-1/2)^2 (m+1/2)} \frac{m+1/2}{\sqrt{n+1}}\\&\le \frac{1}{2\pi (n+1)} \sum _{m=0}^n \frac{1}{(m-1/2)^2}\\&\le \frac{1}{n+1} \frac{8+\pi ^2}{2\pi }\\&\le \frac{3}{n+1}. \end{aligned}$$

Combining our bounds for (I), (II) and (III),

$$\begin{aligned} T_n&= \frac{\Gamma (n+3/2)}{\Gamma (n+1)} \left[ (I) - (II)-(III)\right] \\&\ge \frac{\Gamma (n+3/2)}{\Gamma (n+1)}\left( \frac{4}{\pi } - \frac{1}{2\pi (2n-1)^2} - \frac{3}{n+1}\right) \\&\ge \sqrt{n+1/2} \left( \frac{4}{\pi } - \frac{1}{2\pi (2n-1)^2} - \frac{3}{n+1}\right) , \end{aligned}$$

and by (26),

$$\begin{aligned} \alpha _{\mathbb {C}}(d) \ge \frac{1}{d^{3/2}} \sum _{n=1}^{d-1} \sqrt{n+1/2} \left( \frac{4}{\pi } - \frac{1}{2\pi (2n-1)^2} - \frac{3}{n+1}\right) . \end{aligned}$$

The term \(\frac{1}{d^{3/2}} \sum _{n=1}^{d-1} 4 \sqrt{n+1/2}/ \pi \) is the main term and can be bounded below by

$$\begin{aligned} \frac{1}{d^{3/2}} \sum _{n=1}^{d-1} \frac{4 \sqrt{n+1/2}}{ \pi }&\ge \frac{1}{d^{3/2}} \frac{8}{3\pi } \left( (d-1/2)^{3/2} - (1/2)^{3/2}\right) \\ {}&\ge (1 - (2d)^{-1}) \frac{8}{3\pi } - (2d)^{-3/2}\\&\ge \frac{8}{3\pi } - \left( \frac{8}{3\pi } + \frac{1}{2}\right) d^{-1} . \end{aligned}$$

The other error terms are at most

$$\begin{aligned} d^{-3/2} \sum _{n=1}^{d-1} \sqrt{n+1/2}\left( \frac{1}{2\pi (2n-1)^2} + \frac{3}{n+1}\right)&\le \frac{1}{d^{3/2}} \sum _{n=1}^{d-1}\frac{4}{\pi (n+1)} \sqrt{n+1/2}\\&\le \frac{1}{d^{3/2}} \sum _{n=1}^{d-1} \frac{4}{\pi (n+1)^{1/2}}\\&\le \frac{4}{\pi d^{3/2}}2\sqrt{d+1}. \end{aligned}$$

Combining the main and error term bounds, the lemma follows. \(\square \)

Lemma 21

For \(G_{\mathbb {K}} \in \mathbb {K}^{d\times d}\) a Gaussian random matrix with i.i.d. \(\mathbb {K}\) valued \(\mathcal {N}(0,d^{-1})\) entries, define \(\alpha _{\mathbb {K}}(d) := \mathbb {E}\left[ \frac{1}{d} \sum _{j=1}^d \sigma _j(G_{\mathbb {K}})\right] \). The following holds

$$\begin{aligned} \alpha _{\mathbb {C}}(d) - \alpha _{\mathbb {R}}(d) \le 4.02 d^{-1}. \end{aligned}$$

Proof

To find an explicit formula for \(\alpha _{\mathbb {R}}(d)\), we need an expression for the spectral distribution of the wishart matrix \(d G_{\mathbb {R}} G_{\mathbb {R}}^T\), which we call \(p_d^{\mathbb {R}}(x)\), given by equation (16) in [24]:

$$\begin{aligned} p_d^{\mathbb {R}}(x) = \frac{1}{2d}\left( 2R_d(x) - \frac{\Gamma \left( \frac{d}{2}+\frac{1}{2} \right) }{\Gamma \left( \frac{d}{2} \right) }L_{d-1}(x) \left\{ \psi _1\left( x\right) - \psi _2\left( x\right) \right\} \right) , \end{aligned}$$

where

$$\begin{aligned} \psi _1(x)= & {} e^{-x} \sum _{k=0}^{(\kappa +d-2)/2} \delta _k L_{2k+1-\kappa }(x),\\ \psi _2(x)= & {} \left( \frac{x}{2}\right) ^{-1/2}e^{-\frac{x}{2}} \left[ (1-\kappa )\frac{2\Gamma \left( \frac{1}{2},\frac{x}{2} \right) }{\Gamma \left( \frac{1}{2}\right) } + 2\kappa -1 \right] ,\\ R_d(x)= & {} e^{-x}\sum _{m=0}^{d-1}\left( L_{m}(x) \right) ^2,\\ \delta _k= & {} \frac{\Gamma \left( k + 1 - \frac{\kappa }{2} \right) }{\Gamma \left( k + \frac{3}{2} - \frac{\kappa }{2}\right) }, \end{aligned}$$

\(\kappa = d\mod 2\) and \(\Gamma (a,y) = \int _y^\infty t^{a-1}e^{-t}dt\) is the incomplete Gamma function.

This means that

$$\begin{aligned} \alpha _{\mathbb {R}}(d)&= d^{-1/2} \int _0^{\infty } x^{1/2} p_d^{\mathbb {R}}(x) dx\\&= \frac{1}{d^{3/2}}\int _0^\infty x^{1/2}R_d(x)dx\\&\quad - \frac{1}{2d^{3/2}}\int _0^\infty x^{1/2} \frac{\Gamma \left( \frac{d}{2}+\frac{1}{2} \right) }{\Gamma \left( \frac{d}{2} \right) }L_{d-1}(x) \left\{ \psi _1\left( x\right) - \psi _2\left( x\right) \right\} dx. \end{aligned}$$

Recall that (see Sect. 5)

$$\begin{aligned} \alpha _{\mathbb {C}}(d) = d^{-3/2}\sum _{n=0}^{d-1}\int _{0}^\infty x^{1/2}e^{-x}L_n(x)^2 dx \end{aligned}$$

which implies

$$\begin{aligned} \alpha _{\mathbb {R}}(d) = \alpha _{\mathbb {C}}(d) - \frac{1}{2d^{3/2}}\int _0^\infty x^{1/2}\frac{\Gamma \left( \frac{d}{2}+\frac{1}{2} \right) }{\Gamma \left( \frac{d}{2} \right) }L_{d-1}(x) \left\{ \psi _1\left( x\right) - \psi _2\left( x\right) \right\} dx. \end{aligned}$$
(28)

We are especially interested in the following terms which appear in the full expression for \(\alpha _{\mathbb {R}}(d)\):

$$\begin{aligned} Q(m,k)= \int _0^{\infty } x^{1/2} e^{-x} L_m(x) L_k(x) dx. \end{aligned}$$
(29)

From [16] section 7.414 equation 4(1), we have

$$\begin{aligned} Q(m,k) = \frac{1}{4\pi } \sum _{i=0}^{\min \{m,k\}} \frac{\Gamma (i+3/2)}{\Gamma (i+1)}\frac{\Gamma (m-i-1/2)}{\Gamma (m-i+1)}\frac{\Gamma (k-i-1/2)}{\Gamma (k-i+1)}. \end{aligned}$$

The following lemma deals with bounds on sums involving Q(mk) terms. \(\square \)

Lemma 22

For Q(mk) as defined in (29) we have the following bounds

$$\begin{aligned}&\sum _{k=0}^m \frac{\Gamma (k+1/2)}{\Gamma (k+1)} Q(2m,2k) \le 2.8\end{aligned}$$
(30)
$$\begin{aligned}&\sum _{k=1}^m \frac{\Gamma (k+3/2)}{\Gamma (k+1)} Q(2m-1,2k-1) \le 5.6 \end{aligned}$$
(31)

Proof

Note that in (30),

$$\begin{aligned} Q(2m,2k) = \frac{1}{4\pi } \sum _{i=0}^{2k} \frac{\Gamma (i+3/2)}{\Gamma (i+1)} \frac{ \Gamma (2m - i - 1/2)}{\Gamma (2m-i+1)}\frac{ \Gamma (2k - i - 1/2)}{ \Gamma (2k-i+1)} \end{aligned}$$

since \(m\ge k\).

For \(0<i<2k-1\), the ith term in the summation of Q(2m, 2k) can be bounded above by

$$\begin{aligned}&\frac{\Gamma (i+3/2)}{\Gamma (i+1)} \frac{ \Gamma (2m - i - 1/2)}{\Gamma (2m-i+1)}\frac{ \Gamma (2k - i - 1/2)}{ \Gamma (2k-i+1)}\\&\quad \le \sqrt{i+1} \frac{1}{(2k-i)\sqrt{2k-i-1}}\frac{1}{(2m-i)\sqrt{2m-i-1}}\\&\quad \le \sqrt{i+1} \frac{1}{(2k-i-1)^{3/2} (2m-i-1)^{3/2}}. \end{aligned}$$

This means that

$$\begin{aligned} Q(2m,2k)&\le \frac{1}{8 \sqrt{\pi }} \frac{\Gamma (2m-1/2) \Gamma (2k-1/2)}{\Gamma (2m+1) \Gamma (2k+1)}\\&\quad + \frac{1}{4\pi } \sum _{i=1}^{2k-1} \sqrt{i+1} \frac{1}{(2k-i-1)^{3/2} (2m-i+1)^{3/2}}\\&\quad + \frac{1}{4\pi } \sqrt{\pi } \frac{\Gamma (2k+1/2)}{\Gamma (2k)} \frac{\Gamma (2m-2k+1/2)}{\Gamma (2m-2k+2)}\\&\quad + \frac{1}{4\pi } \max \left( \frac{\Gamma (2k+3/2)}{\Gamma (2k+1)} \frac{\Gamma (2m-2k-1/2)}{\Gamma (2m-2k+1)}(-2\sqrt{\pi }) , 0\right) . \end{aligned}$$

We bound the sum from \(i=1\) to \(2k-3\) by

$$\begin{aligned}&\frac{1}{4\pi } \sum _{i=1}^{2k-3} \sqrt{i+1} \frac{1}{(2k-i-1)^{3/2} (2m-i+1)^{3/2}}\\&\quad \le \frac{1}{4\pi } \sum _{i=0}^{2k-3} \sqrt{i+1} \frac{1}{(2k-i-1)^{3/2}} \frac{1}{(2m-2k+1)^{3/2}}\\&\quad \le \frac{1}{4\pi (2m-2k+1)^{3/2}} \int _0^{2k-2} \frac{\sqrt{x+1}}{(2k-x-1)^{3/2}} dx\\&\quad \le \frac{1}{4\pi (2m-2k+1)^{3/2}} \left( \sqrt{8k} + \frac{\sqrt{4k}}{2k-1}\right) , \end{aligned}$$

so that for \(k \ge 1\),

$$\begin{aligned} Q(2m,2k)&\le \frac{1}{8\sqrt{\pi }} \frac{\Gamma (2m-1/2) \Gamma (2k-1/2)}{\Gamma (2m+1) \Gamma (2k+1)}\\&\quad + \frac{1}{4\pi (2m-2k+1)^{3/2}} \left( \sqrt{8k} + \frac{\sqrt{4k}}{2k-1}\right) \\&\quad + \frac{1}{4\pi } \frac{\sqrt{2k-1}}{(2m-2k+3)^{3/2}}\\&\quad + \frac{1}{4\pi } \sqrt{\pi } \frac{\Gamma (2k+1/2)}{\Gamma (2k)} \frac{\Gamma (2m-2k+1/2)}{\Gamma (2m-2k+2)}\\&\quad + \frac{1}{4\pi } \max \left( \frac{\Gamma (2k+3/2)}{\Gamma (2k+1)} \frac{\Gamma (2m-2k-1/2)}{\Gamma (2m-2k+1)}(-2\sqrt{\pi }) , 0\right) . \end{aligned}$$

For \(k=0, Q(2m,0) < 0\) except for the term \(Q(0,0) = \sqrt{\pi }/2\) which also becomes negative in the full sum, so we ignore these terms.

We now turn our attention to the full sum \(\sum _{k=0}^m \frac{\Gamma (k+1/2)}{\Gamma (k+1)} Q(2m,2k)\). As before, we define for clarity

$$\begin{aligned} (I)&:=\frac{1}{8\sqrt{\pi }} \sum _{k=1}^m \frac{\Gamma (k+1/2)}{\Gamma (k+1)} \frac{\Gamma (2m-1/2) \Gamma (2k-1/2)}{\Gamma (2m+1) \Gamma (2k+1)}\\ (II)&:=\frac{1}{4\pi } \sum _{k=1}^m \frac{\Gamma (k+1/2)}{\Gamma (k+1)} \frac{1}{(2m-2k+1)^{3/2}} \left( \sqrt{8k} + \frac{\sqrt{4k}}{2k-1}\right) \\ (III)&:=\frac{1}{4\pi } \sum _{k=1}^m \frac{\Gamma (k+1/2)}{\Gamma (k+1)} \frac{\sqrt{2k-1}}{(2m-2k+3)^{3/2}}\\ (IV)&:= \frac{1}{4\sqrt{\pi } } \sum _{k=1}^m \frac{\Gamma (k+1/2)}{\Gamma (k+1)} \frac{\Gamma (2k+1/2)}{\Gamma (2k)} \frac{\Gamma (2m-2k+1/2)}{\Gamma (2m-2k+2)}\\ (V)&:= \frac{1}{4\pi } \sum _{k=1}^m \frac{\Gamma (k+1/2)}{\Gamma (k+1)} \max \left( \frac{\Gamma (2k+3/2)}{\Gamma (2k+1)} \frac{\Gamma (2m-2k-1/2)}{\Gamma (2m-2k+1)}(-2\sqrt{\pi }) , 0\right) . \end{aligned}$$

Using the bounds in lemma 20,

$$\begin{aligned} (I)&\le \sum _{k=1}^m \frac{1}{32 \sqrt{\pi } k^{1/2}} \frac{1}{mk} \frac{1}{\sqrt{2m-1}\sqrt{2k-1}} \le \frac{1}{32 \sqrt{\pi }},\\ (II)&\le \frac{1}{4\pi } \sum _{k=1}^m \frac{1}{k^{1/2} (2m-2k+1)^{3/2}} (4 \sqrt{k})\\&\le \frac{1}{\pi } \left( 1 - \frac{1}{\sqrt{2m-1}}\right) \le \frac{1}{\pi },\\ (III)&\le \frac{1}{4\pi } \sum _{k=1}^m \frac{1}{k^{1/2}} \frac{\sqrt{2k-1}}{(2m -2k + 3)^{3/2}}\\&\le \frac{1}{\sqrt{24}\pi },\\ (IV)&\le \frac{1}{4\sqrt{\pi }} \left( \sum _{k=1}^m \frac{(2k)^{1/2}}{k^{1/2}} \frac{1}{(2m-2k+1)\sqrt{2m-2k}} + \sqrt{\pi }\right) \\&\le \frac{1}{2\pi } + \frac{1}{2},\\ (V)&= \frac{1}{4\pi } 4 \pi \frac{\Gamma (2m+3/2)}{\Gamma (2m+1)}\frac{\Gamma (m+1/2)}{\Gamma (m+1)}\\&\le \frac{\sqrt{2m+1}}{\sqrt{m}} \le \sqrt{3}. \end{aligned}$$

Finally,

$$\begin{aligned} \sum _{k=0}^m \frac{\Gamma (k+1/2)}{\Gamma (k)} Q(2m,2k)&\le (I) + (II) + (III) + (IV) + (V)\\&\le 2.8. \end{aligned}$$

To deduce the inequality (31), we use the previously derived bounds to show that

$$\begin{aligned} Q(2m-1,2k-1)&\le \frac{1}{4\pi } \sum _{i=1}^{2k-3} \sqrt{i+1} \frac{1}{(2k-i-2)^{3/2} (2m-i)^{3/2}}\\&\quad + \frac{1}{4\pi } \sqrt{\pi } \frac{\Gamma (2k-1/2)}{\Gamma (2k-1)} \frac{\Gamma (2m-2k+1/2)}{\Gamma (2m-2k+2)}\\&\quad + \frac{1}{4\pi } \max \left( \frac{\Gamma (2k+1/2)}{\Gamma (2k)} \frac{\Gamma (2m-2k+1/2)}{\Gamma (2m-2k+2)}(-2\sqrt{\pi }) , 0\right) , \end{aligned}$$

so that \(Q(2m-1,2k-1) \le Q(2m,2k)\). Now it suffices to note that in the full sum, \(\sum _{k=1}^m \frac{\Gamma (k+3/2)}{\Gamma (k+1)} Q(2m-1,2k-1) \le 2 \sum _{k=1}^m \frac{\Gamma (k+1/2)}{\Gamma (k)} Q(2m-1,2k-1)\) and we get

$$\begin{aligned} \sum _{k=1}^m \frac{\Gamma (k+3/2)}{\Gamma (k+1)} Q(2m-1,2k-1)\le 2 \sum _{k=1}^m\frac{\Gamma (k+1/2)}{\Gamma (k)} Q(2m,2k) \le 5.6. \end{aligned}$$

\(\square \)

We now return our focus to finding a bound on the expression for \(\alpha _{\mathbb {R}}(d)\) given in (28). Since \(\psi _1,\psi _2\) depend on the parity of d, we split in to two cases.

Odd \(d=2m+1\)

From (see [16] section 7.414 equation 6),

$$\begin{aligned} \int _0^{\infty } e^{-x/2} L_{2m}(x) dx = 2, \end{aligned}$$

thus Eq. (28) becomes

$$\begin{aligned}&\alpha _{\mathbb {C}}(2m+1) - \alpha _{\mathbb {R}}(2m+1)\\&\quad = \frac{1}{(2m+1)^{3/2}} \frac{\Gamma (m+1)}{\Gamma (m+1/2)} \left( \sum _{k=0}^{m} \frac{\Gamma (k+1/2)}{\Gamma (k)} Q(2m,2k) - 2^{1/2}\right) , \end{aligned}$$

and using the first bound in Lemma 22,

$$\begin{aligned} \alpha _{\mathbb {C}}(2m+1) - \alpha _{\mathbb {R}}(2m+1) \le 2.8 \sqrt{m+1/2} \frac{1}{(2m+1)^{3/2}} \le m^{-1}. \end{aligned}$$

Even \(d=2m\)

For \(d=2m\), we have

$$\begin{aligned} \alpha _{\mathbb {R}}(2m) = \alpha _{\mathbb {C}}(2m) - \frac{1}{2(2m)^{3/2}} \int _0^{\infty } x^{1/2} \frac{\Gamma (m+1/2)}{\Gamma (m)} L_{2m-1}(x) \{ \psi _1(x) - \psi _2(x)\} dx. \end{aligned}$$

We split the integral into two parts,

$$\begin{aligned} (I):= & {} \frac{1}{2(2m)^{3/2}} \int _0^{\infty } x^{1/2} \frac{\Gamma (m+1/2)}{\Gamma (m)} L_{2m-1}(x) \psi _1(x) dx\\ (II):= & {} \frac{-1}{2(2m)^{3/2}} \int _0^{\infty } x^{1/2} \frac{\Gamma (m+1/2)}{\Gamma (m)} L_{2m-1}(x) \psi _2(x) dx. \end{aligned}$$

Expanding from the definition of \(\psi _1\) above, we have

$$\begin{aligned} (I)&= \frac{1}{2(2m)^{3/2}} \int _0^{\infty } \frac{\Gamma (m+1/2)}{\Gamma (m)} x^{1/2} L_{2m-1}(x)e^{-x} \sum _{k=0}^{m-1} \frac{\Gamma (k)}{\Gamma (k+1/2)} L_{2k-1}(x) dx\\&= \frac{1}{2(2m)^{3/2}} \frac{\Gamma (m+1/2)}{\Gamma (m)} \sum _{k=1}^{m} \frac{\Gamma (k+3/2)}{\Gamma (k+1)} Q(2m-1,2k-1), \end{aligned}$$

so by Lemma 22,

$$\begin{aligned} (I)&\le \frac{1}{2(2m)^{3/2}} \frac{\Gamma (m+1/2)}{\Gamma (m)} 5.6 \le \frac{1}{ m^{1/2}}. \end{aligned}$$

The other part of the integral is

$$\begin{aligned} (II)&= \frac{-1}{2(2m)^{3/2}} \int _0^{\infty } x^{1/2} \frac{\Gamma (m+1/2)}{\Gamma (m)} L_{2m-1}(x) (x/2)^{-1/2}\\&\quad \times e^{-x/2} \left[ \frac{2 \Gamma (1/2,x/2)}{\Gamma (1/2)} - 1\right] dx\\&= \frac{-1}{4m^{1/2}} \int _0^{\infty } \frac{\Gamma (m+1/2)}{\Gamma (m)} L_{2m-1}(x) e^{-x/2} \frac{2 \Gamma (1/2,x/2)}{\Gamma (1/2)} dx \\&\quad + \frac{1}{2m^{3/2}} \frac{\Gamma (m+1/2)}{\Gamma (m)}, \end{aligned}$$

where we use the fact that for odd \(2m-1\) (see [16] section 7.414 equation 6),

$$\begin{aligned} \int _0^{\infty } L_{2m-1}(x) e^{-x/2} dx = -2. \end{aligned}$$

We can bound the first integral in the expression of (II) by

$$\begin{aligned}&\left| \int _0^{\infty } L_{2m-1}(x) e^{-x/2} \Gamma (1/2,x/2) dx\right| \\&\quad \le \left( \int _0^{\infty } e^{-x} L_{2m-1}(x)^2 dx\right) ^{1/2} \left( \int _0^{\infty } \Gamma (1/2,x/2)^2 dx\right) ^{1/2}\\&\quad = \left[ \int _0^{\infty } \left( \int _x^{\infty } t^{-1/2} e^{-t} dt\right) ^2 dx\right] ^{1/2}\\&\quad = \left[ \int _0^1 \left( \int _x^{\infty } t^{-1/2} e^{-t} dt \right) ^2 dx + \int _1^{\infty } \left( \int _x^{\infty } t^{-1/2} e^{-t} dt\right) ^2 dx\right] ^{1/2}\\&\quad \le \left( \Gamma (1/2)^2 + \int _1^{\infty } (e^{-x})^2 dx\right) ^{1/2} \le (\pi + e^{-2}/2)^{1/2}, \end{aligned}$$

so finally

$$\begin{aligned} (II)&\le \frac{ (\pi + 1/2 e^{-2})^{1/2}}{2\sqrt{\pi } m^{3/2}} \frac{\Gamma (m+1/2)}{\Gamma (m)} + \frac{m^{1/2}}{2m^{3/2}}\\&\le 1.01 m^{-1}. \end{aligned}$$

Combining the above bounds we see that in the case of even \(d=2m\),

$$\begin{aligned} \alpha _{\mathbb {C}}(2m) - \alpha _{\mathbb {R}}(2m) =(I) + (II) \le 2.01 m^{-1}. \end{aligned}$$

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bandeira, A.S., Kennedy, C. & Singer, A. Approximating the little Grothendieck problem over the orthogonal and unitary groups. Math. Program. 160, 433–475 (2016). https://doi.org/10.1007/s10107-016-0993-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-016-0993-7

Keywords

Mathematics Subject Classification

Navigation