607Accesses
Abstract
We study a Hermitian matrix model with a kinetic term given by\( Tr (H \Phi ^2 )\), whereH is a positive definite Hermitian matrix, similar as in the Kontsevich Matrix model, but with its potential\(\Phi ^3\) replaced by\(\Phi ^4\). We show that its partition function solves an integrable Schrödinger-type equation for a non-interactingN-body Harmonic oscillator system.
Similar content being viewed by others
Avoid common mistakes on your manuscript.
1Introduction
Let\(\Phi \) be a Hermitian\(N\times N\) matrix,E be a positive diagonal\(N\times N\) matrix\(E := diag (E_1, E_2 , \ldots ,E_N )\) without degenerate eigenvalues, and\(\eta \) be a positive real number as a coupling constant. We deal in this paper with the following one Hermitian matrix model defined using thisE:
This matrix model is obtained by changing the potential of the Kontsevich model [15] from\(\Phi ^3\) to\(\Phi ^4\). It was introduced while studying a scalar field defined on a deformed four-dimensional space-time and studied over years [7,8]. An additional oscillator term was added in order to resolve the IR-UV mixing problem. This term leads to an external matrixE with equally spaced eigenvalues. Recent developments are summarized in Branahl et al. [2].
The main theorem of this paper is expressed as follows.
Theorem 1.1
Let\(Z(E, \eta )\) be the partition function defined by
Let\(\Delta (E)\) be the Vandermonde determinant\(\Delta (E) := \prod _{k<l} (E_l -E_k)\). Then the function
is a zero-energy solution of a Schrödinger-type differential equation being 2-nd order in each of its variables,
where\(\mathcal {H}_{HO}\) is the Hamiltonian of theN-body harmonic oscillator without interaction:
In this sense, this matrix model is a solvable system.
2Schwinger–Dyson equation
Let\(\Phi \) be a Hermitian\(N\times N\) matrix. LetH be a positive Hermitian\(N\times N\) matrix with nondegenerate eigenvalues\(\{E_1, E_2 , \ldots ,E_N ~ | ~ E_i \ne E_j ~\text{ for }~ i \ne j \}\).\(\eta \) is a real positive number. We consider the following action
The partition function is defined by
and we denote the expectation value with this actionS by\(\displaystyle \langle O \rangle := \int \nolimits _{h_N} d \Phi ~ O e^{-S} \). Note that we do not normalize it here, i.e.,\(\langle 1 \rangle = Z(E, \eta ) \ne 1\). Here, the integral measure is the ordinary Haar measure. Using the real variables defined by\(\Phi _{ij}= \Phi _{ij}^{Re} + i \Phi _{ij}^{Im}\), the measure is given as\(\displaystyle \int \nolimits _{h_N} d \Phi := \prod \nolimits _{i}^N \int _{-\infty }^{\infty } d\Phi _{ii} \prod \nolimits _{k<l} \int _{-\infty }^{\infty } d\Phi _{kl}^{Re} \int _{-\infty }^{\infty } d\Phi _{kl}^{Im} \). Note that the partition function\(Z(E, \eta ) \) depends only on the eigenvalues ofH because the integral measure isU(N) invariant. Indeed\(Z(E, \eta ) \) is equal to the partition function obtained from the action defined by\(S'\) in (1.1).
In the following, we use the notation:
For the diagonal elements\(\Phi _{ii} (i= 1,2, \ldots ,N)\), the corresponding partial derivatives are the usual ones. The Schwinger–Dyson equation is derived from
which is expressed as
Taking sum over the indicesi, j and using
a partial differential equation is obtained:
Here,\(\mathcal {L}_{SD}^H \) is a second-order differential operator defined by
Next we rewrite this Schwinger–Dyson equation by using eigenvalues ofH, i.e.,\(E_n (n= 1,2, \ldots , N)\). References [13,14] are helpful in the following calculations. LetP(x) be the characteristic polynomial:
Using thisP(x),
is obtained. Here,\( | M |_{kj} \) denote the minors of the matrixM defined by the determinant of the smaller matrix obtained by removing thek-th row andj-th column fromM. Using the formula (2.9),
The first term in the last line is equal to\(\displaystyle -\sum _{k,i} \frac{P(E_k)}{P'(E_k)} \frac{\partial Z(E, \eta )}{\partial E_{k}} = 0\), becauseP(x) is the characteristic polynomial and\(E_k\) is one of eigenvalues ofH.\(\sum _{i,j} \delta _{ij} (-1)^{i+j} | E_k~ Id_N - H |_{ij} \) in the second term is\(P'(E_k)\). Then we find
Next step, we rewrite the Laplacian\(\displaystyle \sum \nolimits _{i,k} \left( \frac{\partial }{\partial H_{ki}}\frac{\partial }{\partial H_{ik}} \right) Z(E, \eta )\) by\(E_k\). It is a well-known fact that by using the Vandermonde determinant\(\Delta (E) := \prod _{k<l} (E_l -E_k)\) the Jacobian for the change of variables is obtained as follows:
Then, the Laplacian is rewritten as
Here,\(\displaystyle \sum \nolimits _{i \ne j}\) means\(\displaystyle \sum \nolimits _{i,j=1 , i \ne j}^N\). From (2.7), (2.11), and (2.12), we obtain the following.
Theorem 2.1
The partition function defined by (2.2) satisfies
where
In AppendixA, this Schwinger–Dyson equation is checked by using perturbative calculations, as a cross-check.
3Diagonalization of\(\mathcal {L}_{SD}\)
In this section, we prove the main theorem (Theorem1.1).
As the first step, we prove the following proposition.
Proposition 3.1
The differential operator\(\mathcal {L}_{SD} \) defined in (2.14) is transformed into the Hamiltonian of theN-body harmonic oscillator as
We denote this Hamiltonian by\(\mathcal {H}_{HO}\):
Proof
The proof is done by direct calculations.
We calculate\(\Delta ^{-1}(E) e^{\frac{N}{2\eta } \sum _i E_i^2} \frac{\eta }{N} \sum _{i=1}^N \left( \frac{\partial }{\partial E_i} \right) ^2 e^{-\frac{N}{2\eta } \sum _i E_i^2} \Delta (E)\) at first.
(3.3) is equal to 0 since
(3.4) is written as follows.
Then, we obtain
\(\square \)
We introduce a transformed partition function\(\Psi (E, \eta )\) by
Note that this transformation is invertible. Then, the following theorem follows from Proposition3.1 immediately.
Theorem 3.2
The transformed partition function\(\Psi (E, \eta ) \) is a zero-energy solution of the Schrödinger-type differential equation:
Here,\(\mathcal {H}_{HO}\) is theN-body harmonic oscillator Hamiltonian (3.2).
ThisN-body harmonic oscillator system has no interaction terms between the oscillators, so it is a trivial quantum integrable system.
Theorem1.1 is proved as above. In the next section, we calculate the solution\(\Psi (E, \eta ) \) more concretely and give another proof using it.
4From partition function to zero-energy solution
A new expression of the zero-energy solution of theN-body harmonic oscillator system is constructed by a direct calculation of the partition function.
Let us carry out the integration of the off-diagonal components of\(\Phi \) in the definition of the partition function after the change of variables to\(U(N)\times {\mathbb R}^N\). We denote the eigenvalues of\(\Phi \) by\(x_1 , x_2 , \ldots , x_N\). By using a unitary matrixU,\(\Phi \) is diagonalized as\(X=U \Phi U^\dagger \), where\(X= diag(x_1, x_2 , \ldots , x_N )\). Then,
where\(V(x):= \frac{\eta }{4}x^4\).
Let us use the Harish–Chandra–Itzykson–Zuber integral [12,19] for the unitary groupU(N) :
Here,A andB are Hermitian matrices whose eigenvalues are denoted by\(\lambda _{i}(A)\) and\(\lambda _{i}(B)\)\((i=1,\ldots ,N)\), respectively.t is a nonzero complex parameter,\(\displaystyle \Delta (\lambda (A)):=\prod \nolimits _{1\le i<j\le N}(\lambda _{j}(A)-\lambda _{i}(A))\) is the Vandermonde determinant, and\(\displaystyle \tilde{c}_{N}:=\left( \prod \nolimits _{i=1}^{N-1}i!\right) \times \pi ^{\frac{N(N-1)}{2}}\).\(\left( \exp \left( t\lambda _{i}(A)\lambda _{j}(B)\right) \right) \) is the\(N\times N\) matrix with thei-th row and thej-th column being\(\exp \left( t\lambda _{i}(A)\lambda _{j}(B)\right) \). After adapting this formula, the partition function is described by
where\(c_N = \tilde{c}_N (-1/ N)^{\frac{N^2-N}{2}}\) and\(S_N\) denotes the symmetric group. This integral representation (4.3) should be regarded as a Cauchy principal value. Consider the change of variables\(x_i \mapsto x_{\sigma ^{-1}(i)}\). Note that the sign of\(\prod _{l < k}(x_k -x_l)\) changes as\((-1)^\sigma \prod _{l < k}(x_k -x_l)\) , and the following formula is obtained by this change of variables.
Then, the zero-energy solution of (3.9) is obtained by (1.2).
Theorem 4.1
The function
satisfies the Schrödinger-type differential equation (3.9).
Since this fact follows from Theorem3.2, there is no need to prove it, but it would be worthwhile to show the differential equation (3.9) directly from expression (4.5) as a confirmation.
At first, we prove the following Lemma:
Lemma 4.2
Proof
For simplicity,\(\displaystyle \sum \nolimits _i^N N \left( \frac{\eta }{4} x_i^4 + E_i x_i^2 + \frac{1}{2\eta } E_i^2 \right) \) will be abbreviated asf(x, E). From the following identity:
we obtain the formal expression:
Then, we get
From the identity
(4.6) is obtained.\(\square \)
Using Lemma4.2, let us prove Theorem4.1 by direct calculations.
Proof
From (4.5),
For the last equality, we used Lemma4.2.\(\square \)
\(\Psi (E, \eta )\) can also be expressed using Pfaffian. It is described in AppendixB.
As described above, we have also directly proved that the function\(\Psi (E, \eta )\) obtained from the partition function of the matrix model satisfies the Schrödinger-type differential equation for theN-body harmonic oscillator system without interactions.
5Discussions and remarks for\(N=1\)
The matrix model studied in this paper is related to a renormalizable scalar\(\Phi ^4\) theory on Moyal space [8] in the largeN limit. There are mainly two approaches to study the question of integrability of this matrix model: One relies on the model, where one replaces the\(\Phi ^4\) interaction by a constant times\(\Phi ^3\). This gives the Kontsevich model, for which it is known, that the logarithm of the partition function is the\(\tau \) function for the KdV hierarchy and fulfills a Hirota bilinear equation [10,13,15,21]. Another approach follows topological recursion. While the Kontsevich model follows topological recursion, it turned out that the\(\Phi ^4\) model follows the more sophisticated blobbed topological recursion, (proven for genus one and two) [2,3,11]. Due to these complications, it was unexpected, to obtain such a simple answer. ThisN-body harmonic oscillator system is known as an integrable system and this system has been studied for a long time. See, for example, [17,18,20] and references therein. Note that the solution required by the Schrödinger-type equation (3.9) is a zero-energy solution, which is different from the well-known harmonic oscillator solutions by using Hermite polynomials for nonzero-energy solutions. In particular, the case\(N=1\) corresponds to what is called the Weber equation. In the following, we will consider the case\(N=1\) as a particularly simplest case and see how\(\Psi (E, \eta )\) corresponds to a solution to the Weber equation.
Introducing new variables\(\displaystyle u_i := \sqrt{\frac{N}{\eta }}E_i\), the Schrödinger-type equation (3.9) is deformed into
So the\(N=1\) case, this is a kind of the Weber equation [6]:
The series solution of this Weber equation is given as follows. For\(y(u)= \sum _{n=0}^{\infty } a_n u^n\), (5.2) requires
So the boundary conditions are given by\(y(0)=a_0\) and\(y'(0) =a_1\). For\(N=1\) case, the partition function is
and using this\(Z(u,\eta ),\) (3.8) implies that the solution of (5.2) is given as
Indeed, we can prove that\(\Psi (u)\) satisfies (5.2) as follows. As similar to the proof for Lemma4.2,
where\(-f(u)= -\sqrt{\eta } u x^2 - \frac{\eta }{4} x^4 -\frac{u^2}{2}\). Using this formula, (5.2) is derived:
Furthermore, for\(u>0\) , by using modified Bessel function of the second kind\(K_{\frac{1}{4}}(u)\),\(\Psi (u) \) can also be written as
We find that the boundary condition for this solution is required as
Thus, in the case of\(N=1\), the results are derived using known special functions.
Additional comments
During the peer-review process of this paper, a follow-up paper [9] was submitted by the authors, Kanomata and Wulkenhaar, which was published first. Paper [9] corresponds to the application of the technology of this paper. For the benefit of the readers of this journal, we comment on it below.
In this paper, we have seen that the Schwinger–Dyson equation satisfied by the partition function of the Hermitian matrix model (1.1) derives the Schrödinger equation for the Hamiltonian ofN-body harmonic oscillator system. TheN-body harmonic oscillator system can be extended to the integrable Calogero–Moser model [5,16]. It is thus natural to think that there should be matrix models whose partition functions satisfy the Schrödinger equation for the Calogero–Moser model. The calculations with Hermitian matrices in this paper are performed by replacing them with real symmetric matrices and showing that the Schwinger–Dyson equation satisfied by the partition function corresponds to the Schrödinger equation of the Calogero–Moser model in Grosse et al. [9]. Furthermore, since the Calogero–Moser model admits a Virasoro algebra representation, it gives rise to a family of differential equations satisfied by the partition function\({Z}(E,\eta )\) in Grosse et al. [9]. It is also possible to construct Virasoro algebra representations forN-body harmonic oscillator systems. Therefore, a similar family of differential equations satisfied by the partition function\({Z}(E,\eta )\) can also be constructed for the Hermitian matrix model of this paper.
Next, we also comment on the generalization of this model. As mentioned above, the matrix model of the Hermitian matrix and the matrix model of the real symmetric matrix correspond to the harmonic oscillator system and the Calogero–Moser model, respectively. As a similar extension, it is natural to extend the matrix\(\Phi \) from a Hermitian to a quaternion self-dual matrix without changing the form of the action (1.1). In this case, it is still expected to correspond to the Schrödinger equation in the Calogero–Moser model.
A more non-trivial generalization is to consider higher-order potentials like\(\Phi ^6 , \Phi ^8 ,\) etc. In the case of this paper, the\(\Phi ^4\) interaction term leads to the Laplacian in the Schrödinger equation, as seen from its derivation process. Equations in (2.6) and the subsequent calculations show this. By similar considerations, it can be expected that higher-order differential terms will appear, as the order of interactions increases. In such cases, it remains a challenging problem for the future what kind of theory can be developed. We intend to return to such models and to treat also different observables beyond partition functions.
Data availability
No datasets were generated or analyzed during the current study.
References
Borot, G., Wulkenhaar, R.: A short note on BKP for the Kontsevich matrix model with arbitrary potential. [arXiv:2306.01501 [math-ph]]
Branahl, J., Grosse, H., Hock, A., Wulkenhaar, R.: From scalar fields on quantum spaces to blobbed topological recursion. J. Phys. A55(42), 423001 (2022)
Branahl, J., Hock, A., Wulkenhaar, R.: Blobbed topological recursion of the quartic Kontsevich model I: loop equations and conjectures. Commun. Math. Phys. (2022).https://doi.org/10.1007/s00220-022-04392-z
de Bruijn, N.G.: On some multiple integrals involving determinants. J. Indian Math. Soc. New Ser.19, 133–151 (1955)
Calogero, F.: Solution of the one-dimensional N body problems with quadratic and/or inversely quadratic pair potentials. J. Math. Phys.12, 419–436 (1971).https://doi.org/10.1063/1.1665604
Darwin, C.G.: On Weber’s Function. Quart. J. Mech. Appl. Math.2(3), 311–320 (1949).https://doi.org/10.1093/qjmam/2.3.311
Grosse, H., Steinacker, H.: Exact renormalization of a noncommutative\(\phi ^{3}\) model in 6 dimensions. Adv. Theor. Math. Phys.12(3), 605–639 (2008)
Grosse, H., Wulkenhaar, R.: Self-dual noncommutative\(\phi ^4\) -theory in four dimensions is a non-perturbatively solvable and non-trivial quantum field theory. Commun. Math. Phys.329, 1069–1130 (2014)
Grosse, H., Kanomata, N., Sako, A., Wulkenhaar, R.: Real symmetric\(\Phi ^4\) -matrix model as Calogero-Moser model. Lett Math Phys114, 25 (2024).https://doi.org/10.1007/s11005-024-01772-5arXiv:2311.10974 [hep-th]
Harnad, J., Balogh, F.: Tau Functions and their Applications (Cambridge Monographs on Mathematical Physics). Cambridge University Press, Cambridge (2021).https://doi.org/10.1017/9781108610902
Hock, A., Wulkenhaar, R.: Blobbed topological recursion of the quartic Kontsevich model II: Genus=0, [arXiv:2103.13271 [math-ph]]
Itzykson, C., Zuber, J.B.: The planar approximation. 2. J. Math. Phys.21, 411 (1980).https://doi.org/10.1063/1.524438
Itzykson, C., Zuber, J.B.: Combinatorics of the modular group. 2. The Kontsevich integrals. Int. J. Mod. Phys. A7, 5661–5705 (1992).https://doi.org/10.1142/S0217751X92002581. [arXiv:hep-th/9201001 [hep-th]]
Kimura, T.: Mathematical Physics of Random Matrices. Morikita Publishing, Tokyo (2021)
Kontsevich, M.: Intersection theory on the moduli space of curves and the matrix Airy function. Commun. Math. Phys.147, 1–23 (1992).https://doi.org/10.1007/BF02099526
Moser, J.: Three integrable Hamiltonian systems connected with isospectral deformations. Adv. Math.16, 197–220 (1975).https://doi.org/10.1016/0001-8708(75)90151-6
Olshanetsky, A., Perelomov, A.M.: Quantum integrable systems related to lie algebras. Phys. Rep.94, 313–404 (1983).https://doi.org/10.1016/0370-1573(83)90018-2
Semay, C., Buisseret, F., Silvestre-Brac, B.: The quantum N-body problem and the auxiliary field method. Few Body Syst.50, 211–213 (2011).https://doi.org/10.1007/s00601-010-0174-9
Tao, T.:http://terrytao.wordpress.com/2013/02/08/the-harish-chandra-itzykson-zuber-integral-formula/
Willemyns, C.T., Semay, C.: Some specific solutions to the translation-invariant N-body harmonic oscillator Hamiltonian. J. Phys. Comm.5(11), 115002 (2021).https://doi.org/10.1088/2399-6528/ac314e.arXiv:2108.05171 [quant-ph]
Witten, E.: Two-dimensional gravity and intersection theory on moduli space. Surveys Diff. Geom.1, 243–310 (1991).https://doi.org/10.4310/SDG.1990.v1.n1.a5
Acknowledgements
Authors are grateful to Raimar Wulkenhaar and Naoyuki Kanomata for lots of meaningful discussions at ESI. We also thank Kenji Iohara, Taro Kimura, and Ryuichi Nakayama for sharing various techniques and information with us. A.S. was supported by JSPS KAKENHI Grant Number 21K03258. This study was supported by Erwin Schrödinger International Institute for Mathematics and Physics (ESI) through the project “Research in Teams Project Integrability.”
Funding
Open Access funding provided by Tokyo University of Science.
Author information
Authors and Affiliations
Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo, 162-8601, Japan
Akifumi Sako
Erwin Schrödinger International Institute for Mathematics and Physics, University of Vienna, Boltzmanngasse 9, 1090, Vienna, Austria
Harald Grosse & Akifumi Sako
Faculty of Physics, University of Vienna, Boltzmanngasse 5, 1090, Vienna, Austria
Harald Grosse
- Harald Grosse
You can also search for this author inPubMed Google Scholar
- Akifumi Sako
You can also search for this author inPubMed Google Scholar
Corresponding author
Correspondence toAkifumi Sako.
Ethics declarations
Conflict of interest
On behalf of all authors, the corresponding author states that there is no conflict of interest.
Additional information
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Appendices
Appendix: Perturbative check for Schwinger–Dyson equation
In AppendixA, we check that the partition function given in Sect. 4 satisfies the Schwinger–Dyson equation (2.13) by using perturbation theory.
For simplicity, we use the action (1.1) in the following perturbative calculations. We use the fact that the theory itself is equivalent even if the action is changed from (2.1) to (1.1). Then, the partition function is written as
Note that
The Schwinger–Dyson equation (2.13) is expressed as
So, we verify this equation (A.4) by perturbation expansion up to first order in\(\eta \). We calculate the three expectation values (A.1), (A.2), and (A.3) by using Feynman rules.
We shall now summarize the Feynman rules. We denote the free model action by
Introducing an\(N\times N\) Hermitian matrixJ as an external field, the partition function for the free theory is defined as
After Gaussian integral, this is expressed as
where
We denote the vacuum expectation value of the free action\(S_f\) by\(\langle O \rangle _f := \int _{h_N} d\Phi ~ O e^{-S_f}\). Note that we do not normalize this vacuum expectation value, i.e.,\(\langle 1 \rangle _f = Z_f \). From (A.7), the propagator is given by

the interaction is given as

and each loop corresponds to a sum\(\sum _n\). The Feynman diagrams contributing to the 2-point function\(\langle \Phi _{ij}\Phi _{kl} \rangle \) to zeroth and first order in\(\eta \) are shown in Fig.1. If the terms are written in the same order as each graph in Fig.1, they contribute as follows.
whereB represents the terms from bubble graphs in Fig.1:
(See also (A.19).) This equation can be rearranged as follows.
Now that we are ready to calculate contributuions by using Feynman rules, we shall immediately carry out the calculations of the three expectation values (A.1), (A.2), and (A.3). At first, we estimate (A.1). By using Wick contraction, we get
Substituting (A.13) for (A.14), finally we obtain
Next, after substituting (A.13), (A.2) is given as
Similarly, (A.3) is given as
For the right hand side of (A.4), next we calculate
\(\eta ^0\)-term is\(N^2 \langle 1 \rangle _f = N^2 Z_f\).\(\eta ^1\)-term is given by\( -\frac{ N^3 \eta }{4} \langle Tr \Phi ^4 \rangle _f \). This is the same calculation forB:
Now is the time to check the Schwinger–Dyson equation (A.4). (A.1), (A.2), and (A.3) are calculated as (A.15), (A.16), and (A.17), respectively. Note thatB is the term proportional to\(\eta \). The\(\eta ^0\)-term in (A.1)+ (A.2)+ (A.3) exists only in (A.17):
On the other hand,\(\eta ^0\)-term from\(N^2 Z(E, \eta )\), that is the right hand side of (A.4), is\(N^2 Z_f\). Thus, it is confirmed that Eq. (A.4) holds in zeroth order of\(\eta \).
Next order terms in (A.15) + (A.16) + (A.17) are summarized as
Here, we use the same way in (A.20) to obtain the last term. From the following observation,
we find (A.21) is given by
On the other hand,\(\eta \)-linear terms from\(N^2 Z(E, \eta )\) are given by (A.19), then we checked that Eq. (A.4) holds in first order of\(\eta \).
Appendix: Pfaffian expression
We can rewrite\(Z(E, \eta )\) and\(\Psi (E, \eta )\) by using Pfaffian. For simplicity, we consider only 2N case of (4.4).
As noted in Sect. 4, the integral shall be treated as a principal value integral. We use\(C= ((2N)!)^2 c_{2N}\) below. After applying Bruijn’s formula [4], this is obtained by
where
The justification for this process is discussed in Borot and Wulkenhaar [1]. Therefore, the zero-energy solution is given by
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visithttp://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Grosse, H., Sako, A. Integrability of\( \Phi ^4\) matrix model asN-body harmonic oscillator system.Lett Math Phys114, 48 (2024). https://doi.org/10.1007/s11005-024-01783-2
Received:
Revised:
Accepted:
Published:
Share this article
Anyone you share the following link with will be able to read this content:
Sorry, a shareable link is not currently available for this article.
Provided by the Springer Nature SharedIt content-sharing initiative