Log in

Generating Functions for Local Symplectic Groupoids and Non-perturbative Semiclassical Quantization

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

This paper contains three results about generating functions for Lie-theoretic integration of Poisson brackets and their relation to quantization. In the first, we show how to construct a generating function associated to the germ of any local symplectic groupoid and we provide an explicit (smooth, non-formal) universal formula \(S_\pi \) for integrating any Poisson structure \(\pi \) on a coordinate space. The second result involves the relation to semiclassical quantization. We show that the formal Taylor expansion of \(S_{t\pi }\) around \(t=0\) yields an extract of Kontsevich’s star product formula based on tree-graphs, recovering the formal family introduced by Cattaneo, Dherin and Felder in [6]. The third result involves the relation to semiclassical aspects of the Poisson Sigma model. We show that \(S_\pi \) can be obtained by non-perturbative functional methods, evaluating a certain functional on families of solutions of a PDE on a disk, for which we show existence and classification.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Germany)

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. Notice that for the linear function \(H(x,p)= p_{0j} x^j\) (independent of p) where \(p_0\) is fixed, the hamiltonian flow with \(p(0)=0\) is given by \(p(t)=t p_0\) (instead of \(-t p_0\)). This is, ultimately, the reason for our choice of sign convention in \(\omega _c\).

  2. We use the conventions \((\pi ^\sharp \gamma )^j = \pi ^{ij}\gamma _i\) and \((\omega ^\flat (V))_j=\omega _{ij}V^i\), so that \(\sharp \) and \(\flat \) denote contraction with the first index/entry.

  3. For a formal star product \(\star _\hbar \), the right hand side of (1) can be understood as a formal power series on \(\hbar ,\hbar ^{-1},p_1\) and \(p_2\).

  4. The formal generating function depends on the Poisson structure through the symbols of the Kontsevich operators. In [6], the scaling of the Poisson structure is different from the one used here: with the present notation, they work with \(\bar{S}^K_{2\pi }\) instead of \(\bar{S}^K_\pi \), resulting in the fact that their formal family integrates the Poisson structure \(2\epsilon \pi \) instead of our \(\epsilon \pi \); see also [2, §4.1].

  5. This functional could be called the ’PSM action with sources’.

  6. In principle, these provide only particular semiclassical contributions consisting of critical points of \(A'\) without being restricted to a gauge-fixed subset; as we see in the following subsection, it turns out that these are the contributions that yield \(S_\pi \) irrespective of the gauge fixing condition.

References

  1. Butcher, J.C.: An algebraic theory of integration methods. Math. Comput. 26(117), 79–106 (1972)

    Article  MathSciNet  Google Scholar 

  2. Cabrera, A., Dherin, B.: Formal symplectic realizations. Int. Math. Res. Notices 2016(7), 1925–1950 (2016)

    Article  MathSciNet  Google Scholar 

  3. Cabrera, A., Marcut, I., Salazar, M.A.: On local integration of Lie brackets. J. Reine Angew. Math. (Crelle’s J.) 760, 267–293 (2020)

    Article  MathSciNet  Google Scholar 

  4. Cabrera, A., Marcut, I., Salazar, M. A.: Local formulas for multiplicative forms, to appear in Transformation Groups, ar**v:1809.01546 [math.DG]

  5. Cabrera, A., Drummond, T.: van Est isomorphism for homogeneous cochains. Pac. J. Math. 287–2, 297–336 (2017)

    Article  MathSciNet  Google Scholar 

  6. Cattaneo, A.S., Dherin, B., Felder, G.: Formal symplectic groupoid. Comm. Math. Phys. 253, 645–674 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  7. Cattaneo, A.S., Dherin, B., Weinstein, A.: Symplectic microgeometry II: generating functions. Bull. Brazilian Math. Soc. 42, 507–536 (2011)

    Article  MathSciNet  Google Scholar 

  8. Cattaneo, A.S., Dherin, B., Weinstein, A.: Symplectic Microgeometry III: monoids. J. Sympl. Geom. 11(3), 319–341 (2013)

    Article  MathSciNet  Google Scholar 

  9. Cattaneo, A. S., Dherin, B., Weinstein, A.: Symplectic Microgeometry IV: Quantization, ar**v:2007.08167 [math.SG]

  10. Cattaneo, A.S., Felder, G.: A path integral approach to the Kontsevich quantization formula. Commun. Math. Phys. 212(3), 591–612 (2000)

    Article  ADS  MathSciNet  Google Scholar 

  11. Cattaneo, A.S., Felder, G.: Poisson sigma models and symplectic groupoids, Quantization of singular symplectic quotients. Progr. Math. 198, 61–93 (2001)

    MATH  Google Scholar 

  12. Cattaneo, A.S., Felder, G.: Poisson sigma models and deformation quantization. Mod. Phys. Lett. A 16, 179–190 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  13. Cattaneo, A. S., Felder, G.: Effective Batalin–Vilkovisky theories, equivariant configuration spaces and cyclic chains, In: Cattaneo A., Giaquinto A., Xu P. (eds) Higher Structures in Geometry and Physics. Progress in Mathematics, vol 287. Birkhauser, Boston, MA, (2011)

  14. Chartier, P., Hairer, E., Vilmart, G.: Algebraic structures of B-series. Found. Comput. Math. 10(4), 407–427 (2010)

    Article  MathSciNet  Google Scholar 

  15. Coste, A., Dazord, P., Weinstein, A.: Groupoides symplectiques, Publications du Departement de Mathematiques, Nouvelle Serie A, Vol. 2, Univ. Claude-Bernard, Lyon, (1987). http://www.numdam.org/article/PDML_1987___2A_1_0.pdf

  16. Crainic, M., Fernandes, R.L.: Integrability of Lie brackets. Ann. Math. 157(2), 575–620 (2003)

    Article  MathSciNet  Google Scholar 

  17. Crainic, M., Marcut, I.: On the existence of symplectic realizations. J. Symplectic Geom. 9, 435–444 (2011)

    Article  MathSciNet  Google Scholar 

  18. Dherin, B.: The universal generating function of analytical Poisson structures. Lett. Math. Phys. 75, 129–149 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  19. Duistermaat, J., Kolk, J.: Lie Groups. Universitext, Springer-Verlag, New York (2000)

    Book  Google Scholar 

  20. Fernandes, R.L., Michiels, D.: Associativity and integrability. Trans. AMS 373(7), 5057–5110 (2020)

    Article  MathSciNet  Google Scholar 

  21. Guillemin, V., Sternberg, S.: Semiclassical Analysis, International Press of Boston, Incorporated (2013)

  22. Ikeda, N.: Two-dimensional gravity and nonlinear gauge theory. Ann. Phys. 235, 435–464 (1994)

    Article  ADS  MathSciNet  Google Scholar 

  23. Karabegov, A.V.: Formal symplectic groupoid of a deformation quantization. Comm. Math. Phys. 258, 223–256 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  24. Karasev, M.V.: Analogues of objects of the theory of Lie groups for nonlinear Poisson brackets, (Russian). Izv. Akad. Nauk SSSR Ser. Mat. 50, 508–538 (1986)

    MathSciNet  MATH  Google Scholar 

  25. Karasev, M. V., Maslov, V. P.: Nonlinear Poisson Brackets. Geometry and Quantization, Translations of Mathematical Monographs (Book 119), American Mathematical Society (1993)

  26. Kontsevich, M.: Deformation quantization of Poisson manifolds. Lett. Math. Phys. 66, 157–216 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  27. Schaller, P., Strobl, T.: Poisson structure induced (topological) field theories Modern Phys. Lett. A 9(33), 3129–3136 (1994)

    MathSciNet  MATH  Google Scholar 

  28. Zhang, Y.: “Associativity and Local Symplectic Groupoids”, work in progress

Download references

Acknowledgements

The author is indebted with several colleagues for stimulating conversations that helped constructing the contents of this paper. In particular, with Alberto Cattaneo, Marco Gualtieri, Rui Loja-Fernandes and Gonçalo “Robin” Oliveira. I also mention that a key set of ideas regarding the relation to the PSM was kindly pointed out to me by A. Cattaneo. The author was supported by CNPq grants 305850/2018-0 and 429879/2018-0 and by FAPERJ grant JCNE E26/203.262/201.

Funding

The author’s research was supported by the following Brazilian public research-funding agencies: CNPq (Conselho Nacional de Desenvolvimento Científico e Tecnológico) grants 305850/2018-0 and 429879/2018-0, and by FAPERJ (Fundação de Amparo à Pesquisa do Estado do Rio de Janeiro) grant JCNE E26/203.262/201.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alejandro Cabrera.

Ethics declarations

Conflict of Interest

The author has no conflicts of interest to declare and there is no financial interest to report.

Additional information

Publisher's Note

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

Appendix: Kontsevich Graphs, Butcher Series and Networks

Appendix: Kontsevich Graphs, Butcher Series and Networks

In this Appendix, we first recall basic definitions involving Kontsevich graphs and symbols as well as of Butcher series parameterized by rooted trees. In Sect. A.3, we introduce a certain type of graphs built from rooted trees which we call “networks of rooted trees” and define associated symbols. These networks make the bridge between certain Butcher series expressions for generating functions to the Kontsevich-trees generating function (33), as explained in Sect. 4.3.

1.1 Kontsevich trees and their symbols

A Kontsevich graph of type (nm) is a graph (VE) whose vertex set is partitioned in two sets of vertices \(V=V^{a}\sqcup V^{g}\), the aerial vertices \(V^{a}=\{1,\dots ,n\}\) and the terrestrial vertices \(V^{g}=\{\bar{1},\dots ,\bar{m}\}\) such that

  • all edges start from the set \(V^{a}\),

  • loops are not allowed,

  • there are exactly two edges going out of a given vertex \(k\in V^{a}\),

  • the two edges going out of \(k\in V^{a}\) are ordered, the first (“left”) one being denoted by \(e_{k}^{1}\) and the second (“right”) one by \(e_{k}^{2}\).

An aerial edge is an edge whose end vertex is aerial, and a terrestrial edge is an edge whose end vertex is terrestrial. We denote by \(G_{n,m}\) the set of Kontsevich graphs of type (nm).

Given a Poisson structure \(\pi \) and a Kontsevich graph \(\Gamma \in G_{n,m}\), one can associate a m-differential operator \(B_{\Gamma }(\pi )\) on \({\mathbb {R}}^{n}\) in the following way: For \(f_{1},\dots ,f_{m}\in C^{\infty }({\mathbb {R}}^{n})\), we define

$$\begin{aligned} B_{\Gamma }(\pi )(f_{1}\dots ,f_{m}):=\sum _{I:E_{\Gamma }\rightarrow \{1,\dots ,d\}}\big [\prod _{k\in V_{\Gamma }^{a}}(\prod _{\begin{array}{c} e\in E_{\Gamma }\\ e=(*,k) \end{array} }\partial _{I(e)})\pi ^{I(e_{k}^{1})I(e_{k}^{2})}\big ]\prod _{i\in V_{\Gamma }^{g}} \big (\prod _{\begin{array}{c} e\in E_{\Gamma }\\ e=(*,i) \end{array} }\partial _{I(e)}\big )f_{i}. \end{aligned}$$

The symbol \(\hat{B}_{\Gamma }\) of \(B_{\Gamma }\) is defined by

$$\begin{aligned} B_{\Gamma }(e^{p_{1}x},\dots ,e^{p_{m}x})=\hat{B}_{\Gamma }(p_{1},\dots ,p_{m})e^{(p_{1} +\dots +p_{m})x}. \end{aligned}$$

Example A.1

Figure 2b shows an example of a Kontsevich graph \(\Gamma \equiv \Gamma _\rho \) of type (n, 2). The corresponding symbol is given by

$$\begin{aligned} \hat{B}_\Gamma (\pi )(p_1,p_2) = \pi ^{k_1k_2} \ \pi ^{j_1j_2} \ \partial _{k_2}\partial _{j_1}\pi ^{i_1i_2} \ \partial _{k_1}\pi ^{l_1l_2} \ p_{1i_1}p_{1l_1}p_{2i_2}p_{2l_2}p_{2j_2}. \end{aligned}$$

The order of the arrows (left vs. right arrows) is important because flip**, for example, the order of the first aerial vertex order would introduce a sign, since \(\pi ^{ij}=-\pi ^{ji}\).

\(T_{n,2}\) is a subset of \(G_{n,2}\) that we now define:

Definition A.2

Let \(\Gamma \in G_{m,n}\) be a Kontsevich graph. The interior of \(\Gamma \) is the graph \(\Gamma _{i}\) obtained from \(\Gamma \) by removing all terrestrial vertices and terrestrial edges. A Kontsevich graph is a Kontsevich tree if its interior is a tree in the usual sense (i.e. it has no cycles). We denote by \(T_{n,m}\) the set of Kontsevich’s trees of type (nm).

We will not need a detailed presentation of Kontsevich weights \(\Gamma \mapsto W_\Gamma \in {\mathbb {R}}\), the reader can find a rough account with the conventions that we follow in this paper in [2, App. A].

1.2 Rooted trees, elementary differentials and Butcher series

A graph is the data (VE) of a finite set of vertices \(V=\{v_{1},\dots ,v_{n}\}\) together with a set of edges E, which is a subset of \(V\times V\). The number of vertices is called the degree of the graph and is denoted by \(|\Gamma |\). We think of \((v_{1},v_{2})\in E\) as an arrow that starts at the vertex \(v_{1}\) and ends at \(v_{2}\). Two graphs are isomorphic if there is a bijection between their vertices that respects theirs edges. The set \({\bar{\Gamma }}\) of all isomorphic graphs to a given graph \(\Gamma \) is called a topological graph. A symmetry of a graph is an automorphism of the graph (i.e. a relabeling of its vertices that leaves the graph unchanged). The group of symmetries of a given graph \(\Gamma \) will be denoted by \(E(\Gamma )\). Note that the number of symmetries of all graphs sharing the same underlying topological graph is equal; we define the symmetry coefficient \(\sigma ({\overline{\Gamma }})\) of a topological graph \({\overline{\Gamma }}\) to be the number of elements in \(E(\Gamma )\), where \(\Gamma \in {\overline{\Gamma }}\).

A rooted tree is a graph that (1) contains no cycle, (2) has a distinguished vertex called the root, (3) whose set of edges is oriented toward the root. We will denote the set of rooted trees by RT and the set of topological rooted trees by [RT]. The set of topological rooted trees can be described recursively as follows: The single vertex graph \(\bullet \) is in [RT] and if \(t_{1},\dots ,t_{n}\in [RT]\) then so is

$$\begin{aligned} t=[t_{1},\dots ,t_{n}]_{\bullet }, \end{aligned}$$

where the bracket is to be thought as grafting the roots of \(t_{1},\dots ,t_{n}\) to a new root, which is symbolized by the subscript \(\bullet \) in the bracket \([\,,\dots ,\,]\). We can represent graphically topological rooted trees as the \(\gamma \equiv \gamma (\rho )=[\bullet ,\bullet ]_\bullet \) depicted in Fig. 2a (without the extra orientation of edges present in that figure).

Remark A.3

Since we are dealing with topological rooted trees, the ordering in \([t_{1},\dots , t_{m}]_{\bullet }\) is not important (for instance, we do not distinguish between \([\bullet ,[\bullet ]_{\bullet }]_{\bullet }\) and \([[\bullet ]_{\bullet },\bullet ]_{\bullet }\)).

Let \(X=X^{i}(x)\partial _{x^i}\) be a vector field on \({\mathbb {R}}^{d}\). We define the elementary differential of X recursively as follows: For the single vertex tree, we define \(D_{\bullet }^{u}X=X^{u}(x)\), and for \(t=[t_{1},\dots ,t_{m}]\) in [RT], we define

$$\begin{aligned} D_{t}^{u}X(x)=D_{t_{1}}^{i_{1}}X(x)\cdots D_{t_{m}}^{i_{m}}X(x) \ \partial _{i_{1}}\dots \partial _{i_{m}}X^{u}(x), \end{aligned}$$
(57)

where we used the Einstein summation convention. Similarly, given a smooth function \(H:{\mathbb {R}}^d \rightarrow {\mathbb {R}}\) we define recursively for \(t=[t_{1},\dots ,t_{m}]\),

$$\begin{aligned} F^{H,X}_{t}(x) = D^{i_1}_{t_1}X(x)\cdots D^{i_m}_{t_m}X(x) \ \partial _{i_1}\dots \partial _{i_m}H (x), \ F^{H,X}_{\bullet }(x)=H(x). \end{aligned}$$

Following [1], one has that the n-th iterated Lie derivative is given by

$$\begin{aligned} L_X^n x^i = \sum _{t \in [ RT], \ |t|=n} \frac{|t|!}{t! \sigma (t)} D^i_t X, \end{aligned}$$
(58)

where the tree factorial is recursively defined as \(t!= |t| t_1! \cdots t_k!\) for \(t=[t_1,..,t_k]\), \(\bullet ! =1\). The symmetry factor \(\sigma (t)\) also admits a recursive formula, namely, \(\sigma ([t_1^{n_1},..,t_1^{n_k}]) = n_1!\sigma (t_1)^{n_1} ... n_k!\sigma (t_k)^{n_k}\) where the trees \(t_1,..,t_k\) are assumed different and the exponent \(n_i\) denotes it is repeated \(n_i\)-times.

Given a map \(a:[RT] \rightarrow {\mathbb {R}}\) and a vector field \(X\equiv X^i(x)\partial _{x^i}\) on \({\mathbb {R}}^n\) the associated Butcher series is defined by

$$\begin{aligned} B(a, \epsilon X, x) = x + \sum _{t \in [RT]} \frac{\epsilon ^{|t|}}{\sigma (t)} \ a_t \ D_t X |_x. \end{aligned}$$

One can extend this assignment to \(B(a, \epsilon X, H)\equiv H|_{y= B(a, \epsilon X, x)}\) for any function \(H\equiv H(x)\) by expanding formally around \(\epsilon =0\) and using the symbols \(F^{H,X}_t\) defined above (see “S-series” in [14]). The reader is referred to the foundational work of Butcher [1] on elementary differentials and the use of trees and Butcher series in ordinary differential equations for more information (see also [14]).

Example A.4

Following [2, Thm. 21], the formal Taylor expansion around \(t=0\) of the realization map \(\alpha _{t\pi }(x,p)\) defined by Eq. (27) is given by the Butcher series of Eq. (38). The coefficients \(t\mapsto c^B_t\) generalize Bernoulli numbers and can be computed by iterated integrals (see [2, Thm. 24]).

1.3 Networks of rooted trees and their symbols

The motivation for introducing networks of rooted trees comes from considering elementary differentials of the form \(D_\gamma ^iX\), with \(\gamma \in [RT]\) and \(X=X^{p_1\alpha }\) a hamiltonian vector field in \(M\times M^*\) as in (37) (M is a coordinate space), and substituting the Butcher series (39) in place of \(\alpha :M\times M^*\cdot \cdot _{M\times 0}\dashrightarrow M\). The idea is that each derivative in the elementary differential acting on each term of \(\alpha \) produces, by the Leibniz rule, a series of terms which can be arranged into these “network” graphs. To account for the two terms in the hamiltonian vector field \(X^{p_1\alpha }\) we consider an extra orientation on the (“skeleton”) edges of \(\gamma \), which tells us which term is acting.

With these considerations, a network of rooted trees \(\rho \in NRT\) consists of the following data:

  • a rooted tree \(\gamma \equiv \gamma (\rho )\in RT\), called the skeleton of \(\rho \), endowed with an additional orientation on its edges (besides the natural one towards the root);

  • for each vertex v of \(\gamma \), a rooted tree \(\rho (v) \in [RT]\);

  • for each additionally oriented edge e of \(\gamma \) going from v to w, a pair of vertices \(\rho (e)_1 \in \rho (v) \ ,\rho (e)_2 \in \rho (w)\).

We also consider the case in which there is an extra marked vertex \(\rho ^* \in \rho (r(\gamma ))\), where \(r(\gamma )\in \gamma \) is the root, and in this case we denote \(\rho \in NRT^*\). The sum of all the vertices on the various \(\rho (v), \ v\in \gamma ,\) yields the total number of vertices in the network, denoted \(|\rho |\), and we say \(\rho \in NRT_{|\rho |}\). An example of a network of rooted trees is illustrated in Fig. 2a.

Fig. 2
figure 2

In a we have a network of rooted trees \(\rho \in NRT_{4}\) and we depicted its skeleton \(\gamma (\rho )\in [RT]_3\) together with the additional orientation on its edges. In b we have the associated Kontsevich graph \(\Gamma _\rho \) of type (n, 2), following the assignment \(\rho \mapsto \Gamma _\rho \) of Sect. 4.3. The terrestrial vertices are on a horizontal line and the aerial vertices are placed above them. The first (“left”) arrow stemming out of an aerial vertex has a solid black head, while the second (“right”) one has a hollow white one

Consider a vector field \(V=V^j(x,p)\partial _{x^j}\) as in Eq. (39) and \(p_2 \in M^*\). We now associate a symbol map

$$\begin{aligned} \rho \mapsto E_\rho \equiv E_\rho ^{p_2,V} \in C^\infty (T^*M), \end{aligned}$$

to each network, by giving symbolic rules. For each internal vertex in some \(\rho (v), \ v\in \gamma \), we write \(p_{2k}V^k\) if it is the root or \(V^k\) otherwise. For each internal edge in each tree \(\rho (v)\), we write \(\partial _{x^k}\) on the left if the edge is arriving or on the right if it is departing towards the root. For each skeleton-oriented edge \(e \in \gamma \) connecting \(\rho (e)_1\) to \(\rho (e)_2\), we write \(\pm \partial _{p_j}\) on the left at \(\rho (e)_1\) and \(\partial _{x^j}\) also on the left at \(\rho (e)_2\), where we take \(-1\) if the orientation of e is towards the root of \(\gamma \) or \(+1\) otherwise. When \(\rho \in NRT^*\) has an extra marked vertex \(\rho ^*\) as above, the associated symbol requires an additional choice of ’decoration’ by \(x^j\) or \(p_j\) and is denoted

$$\begin{aligned} E_\rho ^{x^j} \ or \ E_\rho ^{p_j}. \end{aligned}$$

This symbol is computed with the same rules as before and by adding at \(\rho ^*\) the extra symbol \(-\partial _{p_j}\) or \(\partial _{x^j}\) on the left, respectively for decorations \(x^j\) or \(p_j\).

Example A.5

Consider the network \(\rho \in NRT_4\) of Fig. 2a. Following the above rules we obtain that the corresponding symbol is given by

$$\begin{aligned} E_\rho ^{p_2,V}(x,p)&= - p_{2k_1} \partial _{p_{j_1}}V^{k_1}|_{(x,p)} \ \partial _{p_{j_2}}V^{l_2}|_{(x,p)} \ \partial _{x^{j_1}}\partial _{x^{l_2}}V^{k_2}|_{(x,p)}\\&\qquad p_{2k_2} \ \partial _{x^{j_2}}V^{k_3}|_{(x,p)} p_{2k_3}. \end{aligned}$$

When \(V=V_\pi = \pi ^{ij} p_j \partial _{x^i}\) as in Eq. (38), then

$$\begin{aligned} E_\rho ^{p_2,V}(x,p_1) = - \hat{B}_{\Gamma _\rho }(\pi )(p_1,p_2), \end{aligned}$$

where \(\Gamma _\rho \) is the associated Kontsevich graph, computed following Sect. 4.3 and depicted in Fig. 2b, and its Kontsevich symbol was computed in Example A.1.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cabrera, A. Generating Functions for Local Symplectic Groupoids and Non-perturbative Semiclassical Quantization. Commun. Math. Phys. 395, 1243–1296 (2022). https://doi.org/10.1007/s00220-022-04453-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-022-04453-3

Navigation