Skip to main content
Log in

Different types of spatial correlation functions for non-ergodic stochastic processes of macroscopic systems

  • Regular Article - Flowing Matter
  • Published:
The European Physical Journal E Aims and scope Submit manuscript

Abstract

Focusing on non-ergodic macroscopic systems, we reconsider the variances \(\delta \mathcal{O}^2\) of time averages \(\mathcal{O}[\mathbf {x}]\) of time-series \(\mathbf {x}\). The total variance \(\delta \mathcal{O}^2_{\mathrm {tot}}= \delta \mathcal{O}^2_{\mathrm {int}}+ \delta \mathcal{O}^2_{\mathrm {ext}}\) (direct average over all time series) is known to be the sum of an internal variance \(\delta \mathcal{O}^2_{\mathrm {int}}\) (fluctuations within the meta-basins) and an external variance \(\delta \mathcal{O}^2_{\mathrm {ext}}\) (fluctuations between meta-basins). It is shown that whenever \(\mathcal{O}[\mathbf {x}]\) can be expressed as a volume average of a local field \(\mathcal{O}_{\mathbf{r}}\) the three variances can be written as volume averages of correlation functions \(C_{\mathrm {tot}}(\mathbf{r})\), \(C_{\mathrm {int}}(\mathbf{r})\) and \(C_{\mathrm {ext}}(\mathbf{r})\) with \(C_{\mathrm {tot}}(\mathbf{r}) = C_{\mathrm {int}}(\mathbf{r}) + C_{\mathrm {ext}}(\mathbf{r})\). The dependences of the \(\delta \mathcal{O}^2\) on the sampling time \(\varDelta \tau \) and the system volume V can thus be traced back to \(C_{\mathrm {int}}(\mathbf{r})\) and \(C_{\mathrm {ext}}(\mathbf{r})\). Various relations are illustrated using lattice spring models with spatially correlated spring constants.

Graphical abstract

.

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
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

Data Availability Statement

It was not possible to store all the \(N_\mathrm {c}\times N_\mathrm {k}\times N_\mathrm {t}\times L^2\) primary fields \(x_{ckt\mathbf{q}}\) which were immediately deleted after having been analyzed. Tables of the global averages have been kept, however, for a broad range of \(\varDelta \tau \) and V and different LSM variants. These data sets are available from the corresponding author on reasonable request. My manuscript has associated data in a data repository

Notes

  1. Other definitions of non-ergodicity parameters may be found in the literature [5]. Our definition does not rely on the properties of a specific model or a theoretical assumption. It can be made system-size independent by multiplying it with the appropriate non-universal volume dependence \(V^{\gamma }\).

  2. The inverse Fourier transform is \(f_{\mathbf{r}}= \sum _{\mathbf{q}} f_{\mathbf{q}} \exp (-i \mathbf{q}\cdot \mathbf{r})\).

  3. We note f for the functional argument of the averaged correlation function C and \(f_{\mathbf{r}}\) for the functional argument of the non-averaged correlation function \(K\).

  4. Using Parseval’s theorem, it is seen that \(\langle g_{\mathbf{q}} g_{-\mathbf{q}} \rangle =1/N_{\mathbf{r}}\).

  5. In this case, \(v[\mathbf {x}]\) has the dimension of a (free) energy density just as the stress (pressure) of the system.

  6. The covariance \(v_{\mathbf{r}}\) must be distinguished from the purely local variance \({\tilde{v}}_{\mathbf{r}} = \beta V \mathbf {E}^t (x_{t\mathbf{r}}-x_{\mathbf{r}})^2\).

  7. To show the second relation, it is used that the \(x_{\mathbf{r}}-a_{\mathbf{r}}\) are decorrelated for \(J \rightarrow 0\) albeit their first and second moments may be correlated. \(v_{\mathbf{r}}\rightarrow b_{\mathbf{r}}\) holds for all V and \(\beta \) due to prefactor \(\beta V\) in the definition of \(v_{\mathbf{r}}\).

  8. Following Ref. [3] one simple possibility to characterize \(\tau _{\mathrm {b}}\) is to set \(\mathcal{O}(\varDelta \tau =\tau _{\mathrm {b}},V) = f \mathcal{O}(V)\) using a fixed fraction f close to unity. We use \(f=0.95\).

  9. A spacer time interval \(\tau _{\mathrm {spacer}}\approx \varDelta \tau \) is used between each measured time series of length \(\varDelta \tau \). It may have been more efficient to use instead \(\tau _{\mathrm {spacer}}\approx \max (\varDelta \tau ,\tau _{\mathrm {b}}(J))\) to make the only asymptotically exact Eq. (38) applicable for smaller \(N_\mathrm {k}\).

  10. We often suppress in this subsection the possible additional V-dependences, i.e., we write, e.g., \(v(\varDelta \tau )\), instead of \(v(\varDelta \tau ,V)\).

  11. If \(\delta v_{\mathrm {tot}}(\varDelta \tau )\) is known for a broad range of \(\varDelta \tau \), one may plot \(\delta v_{\mathrm {tot}}(\varDelta \tau )\) as a function of \(1/\sqrt{\varDelta \tau }\) in linear coordinates. \(\varDelta _{\mathrm {ne}}\) may then be obtained for \(N_\mathrm {k}=1\) from the intercept of the vertical axis of a linear data fit. This procedure allows to avoid the determination of \(\delta v_{\mathrm {ext}}(\varDelta \tau ,N_\mathrm {k})\).

  12. In in statistical mechanics, Eq. (44) is closely related to the equivalence of the Green–Kubo and the Einstein relations for transport coefficients [2, 16, 23].

  13. No general relation such as Eq. (45) for \(\delta v_{\mathrm {int}}(\varDelta \tau )\) is known at present for \(\delta v_{\mathrm {ext}}(\varDelta \tau )\).

  14. It is well known that \(v_c\) depends on whether the average intensive variable of the basin is imposed or its conjugated extensive variable [22].

  15. Without invoking here thermostatistics this argument demonstrates that v must be V-independent whenever the spatial correlations are short-ranged.

  16. While “short-range” is often reserved for ultimately exponentially decaying correlation functions, it is used here also for correlations decaying sufficiently fast such that the volume average does not depend on the upper integration boundary L.

  17. We use here \(N_\mathrm {k}=1000\) for \(\varDelta \tau \le 10^4\) to obtain for \(\delta v_{\mathrm {ext}}(\varDelta \tau ,N_\mathrm {k})\) a sufficiently accurate \(N_\mathrm {k}\)-extrapolation \(\delta v_{\mathrm {ext}}(\varDelta \tau )\) for small sampling times.

  18. According to Eq. (57) and assuming c(r) to be continuous, the crossover length \(\xi _{\star }\) may be defined by \(c(r=\xi _{\star }) \approx 1/V\).

  19. Data collapse for different V and \(J>0\) can be achieved (not shown) by obtaining first \(c(r) = (2y-1)/V\) and by plotting then c(r)/c(0) as a function of \(x = r/\xi _{\mathrm {ind}}(J,V)\).

  20. As seen using Eq. (83), this is, however, the case for \(m_{\mathbf{r}} = \mathbf {E}^t x_{t\mathbf{r}}\) for decorrelated primary instantaneous fields \(x_{t\mathbf{r}}\).

References

  1. L. Klochko, J. Baschnagel, J.P. Wittmer, A.N. Semenov, J. Chem. Phys. 151, 054504 (2019)

    Article  ADS  Google Scholar 

  2. G. George, L. Klochko, A. Semenov, J. Baschnagel, J.P. Wittmer, EPJE 44, 13 (2021)

    Google Scholar 

  3. G. George, L. Klochko, A.N. Semenov, J. Baschnagel, J.P. Wittmer, EPJE 44, 54 (2021)

    Google Scholar 

  4. G. George, L. Klochko, A.N. Semenov, J. Baschnagel, J.P. Wittmer, EPJE 44, 125 (2021)

    Google Scholar 

  5. W. Götze, Complex Dynamics of Glass-Forming Liquids: A Mode-Coupling Theory (Oxford University Press, Oxford, 2009)

    MATH  Google Scholar 

  6. A. Heuer, J. Phys.: Condens. Matter 20, 373101 (2008)

  7. I. Procaccia, C. Rainone, C.A.B.Z. Shor, M. Singh, Phys. Rev. E 93, 063003 (2016)

    Article  ADS  Google Scholar 

  8. J.F. Lutsko, J. Appl. Phys. 64, 1152 (1988)

    Article  ADS  Google Scholar 

  9. J.F. Lutsko, J. Appl. Phys 65, 2991 (1989)

  10. J.L. Barrat, Microscopic Elasticity of Complex Systems, in Computer Simulations in Condensed Matter Systems: From Materials to Chemical Biology -, vol. 704, ed. by M. Ferrario, G. Ciccotti, K. Binder (Springer, Berlin and Heidelberg, 2006), pp.287–307

  11. J.P. Wittmer, H. Xu, P. Polińska, F. Weysser, J. Baschnagel, J. Chem. Phys. 138, 12A533 (2013)

    Article  Google Scholar 

  12. A. Lemaître, Phys. Rev. Lett. 113, 245702 (2014)

    Article  ADS  Google Scholar 

  13. A. Lemaître, J. Chem. Phys. 143, 164515 (2015)

    Article  ADS  Google Scholar 

  14. D.P. Landau, K. Binder, A Guide to Monte Carlo Simulations in Statistical Physics (Cambridge University Press, Cambridge, 2000)

    MATH  Google Scholar 

  15. W. Press, S. Teukolsky, W. Vetterling, B. Flannery, Numerical Recipes in FORTRAN: The Art of Scientific Computing (Cambridge University Press, Cambridge, 1992)

    MATH  Google Scholar 

  16. M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids, 2nd edn. (Oxford University Press, Oxford, 2017)

    Book  Google Scholar 

  17. A.L. Barbabási, H. Stanley, Fractal Concepts in Surface Growth (Cambridge University Press, Cambridge, 1995)

    Book  Google Scholar 

  18. M. Maier, A. Zippelius, M. Fuchs, Phys. Rev. Lett. 119, 265701 (2017)

    Article  ADS  Google Scholar 

  19. M. Maier, A. Zippelius, M. Fuchs, J. Chem. Phys. 149, 084502 (2018)

    Article  ADS  Google Scholar 

  20. F. Vogel, A. Zippelius, M. Fuchs, Europhys. Lett. 125, 68003 (2019)

    Article  ADS  Google Scholar 

  21. L. Klochko, J. Baschnagel, J.P. Wittmer, A.N. Semenov, Soft Matter 14, 6835 (2018)

    Article  ADS  Google Scholar 

  22. J.L. Lebowitz, J.K. Percus, L. Verlet, Phys. Rev. 153, 250 (1967)

    Article  ADS  Google Scholar 

  23. J.P. Hansen, I.R. McDonald, Theory of Simple Liquids, 3rd edn. (Academic Press, New York, 2006)

    MATH  Google Scholar 

Download references

Acknowledgements

We acknowledge computational resources from the HPC cluster of the University of Strasbourg.

Author information

Authors and Affiliations

Authors

Contributions

JPW designed and wrote the project benefiting from contributions of all authors.

Corresponding author

Correspondence to J. P. Wittmer.

Appendices

A Spatial correlations of periodic microcells

1.1 A.1 Some useful general relations

Let us begin by stating several useful general relations for spatial correlation functions of d-dimensional, real, discrete and periodic fields. As defined by Eq. (13) or Eq. (14) we consider the instantaneous correlation function \(K[y_{\mathbf{r}}](\mathbf{r})\) of a field \(y_{\mathbf{r}}\) of volume average \(y \equiv \mathbf {E}^{\mathbf{r}} y_{\mathbf{r}}\). Obviously,

$$\begin{aligned} K[y_{\mathbf{r}}-y](\mathbf{r}) = K[y_{\mathbf{r}}](\mathbf{r}) - y^2. \end{aligned}$$
(58)

Let us assume that the field \(y_{l\mathbf{r}}\) additionally depends on an index l. We use below the averages \(y_l=\mathbf {E}^{\mathbf{r}} y_{l\mathbf{r}}\), \(y_{\mathbf{r}}=\mathbf {E}^l y_{l\mathbf{r}}\) and \(y=\mathbf {E}^l y_l = \mathbf {E}^{\mathbf{r}} y_{\mathbf{r}}\). Rewriting Eq. (58) and summing over l gives

$$\begin{aligned} \mathbf {E}^lK[y_{l\mathbf{r}}-y_l](\mathbf{r}) = \mathbf {E}^l K[y_{l\mathbf{r}}](\mathbf{r}) - \mathbf {E}^ly_l^2. \end{aligned}$$
(59)

Also it is seen by expansion using Eq. (13) that

$$\begin{aligned} \mathbf {E}^lK[y_{l\mathbf{r}}-\lambda y_{\mathbf{r}}](\mathbf{r}) = \mathbf {E}^lK[y_{l\mathbf{r}}](\mathbf{r}) - \lambda (2-\lambda ) K[y_{\mathbf{r}}](\mathbf{r})\qquad \end{aligned}$$
(60)

for any real constant \(\lambda \).

We remind that \(\mathbf {V}^l y_l = \mathbf {E}^l y_l^2 - y^2 = \mathbf {E}^l (y_l-y)^2\). Using again Eq. (58) and the periodicity of the grid the variance \(\mathbf {V}^l y_l\) may be written as the volume average

$$\begin{aligned} \mathbf {V}^l y_l= & {} \mathbf {E}^{\mathbf{r}} C(\mathbf{r}) \text{ with } \end{aligned}$$
(61)
$$\begin{aligned} C(\mathbf{r})\equiv & {} \mathbf {E}^l K[y_{l\mathbf{r}}-y](\mathbf{r}) \equiv \mathbf {E}^l K[y_{l\mathbf{r}}](\mathbf{r}) - y^2 \end{aligned}$$
(62)

being the l-averaged correlation function in real space. By comparing Eq. (59) and Eq. (61), we may also write

$$\begin{aligned} C(\mathbf{r}) = \mathbf {E}^l K[y_{l\mathbf{r}}-y_l](\mathbf{r}) + \mathbf {V}^l y_l \end{aligned}$$
(63)

which using Eq. (61) implies \(\mathbf {E}^{\mathbf{r}}\mathbf {E}^l K[y_{l\mathbf{r}}-y_l](\mathbf{r}) =0\). It is useful to state two important limits for \(C(\mathbf{r})\): (i) At the origin, we have

$$\begin{aligned} C(\mathbf{r}=\mathbf{0}) = \mathbf {V}^{l\mathbf{r}'} y_{l\mathbf{r}'} = \mathbf {E}^l \mathbf {V}^{\mathbf{r}'} y_{l\mathbf{r}'} + \mathbf {V}^l \mathbf {E}^{\mathbf{r}'} y_{l\mathbf{r}'} \end{aligned}$$
(64)

and (ii) \(C(\mathbf{r})\) exactly vanishes if and only if

$$\begin{aligned} \mathbf {E}^l \mathbf {E}^{\mathbf{r}'} y_{\mathbf{r}'+\mathbf{r}} y_{\mathbf{r}'} = \mathbf {E}^l \mathbf {E}^{\mathbf{r}'} y_{\mathbf{r}'+\mathbf{r}} \times \mathbf {E}^l \mathbf {E}^{\mathbf{r}'} y_{\mathbf{r}'} = y^2 \end{aligned}$$
(65)

as it happens for most (albeit not all) fields for sufficiently large \(r=|\mathbf{r}|\). See Appendix B.1 for an exception relevant for the present study.

Moreover, with \(z=\mathbf {E}^{\mathbf{r}} z_{\mathbf{r}}\) being an l-independent quantity it follows from Eq. (61) that

$$\begin{aligned} \mathbf {V}^l (y_l-z) = \mathbf {E}^{\mathbf{r}} \ \mathbf {E}^l K[(y_{l\mathbf{r}}-z_{\mathbf{r}})-(y-z)](\mathbf{r}). \end{aligned}$$
(66)

Since \(\mathbf {V}^l y_l = \mathbf {V}^l (y_l - z)\), this implies quite generally that

$$\begin{aligned} \mathbf {E}^{\mathbf{r}} \mathbf {E}^l K[\delta y_{l\mathbf{r}}](\mathbf{r}) = \mathbf {E}^{\mathbf{r}} \mathbf {E}^l K[\delta y_{l\mathbf{r}}-\delta z_{\mathbf{r}}](\mathbf{r}) \end{aligned}$$
(67)

with \(\delta y_{l\mathbf{r}}=y_{l\mathbf{r}}-y\) and \(\delta z_{\mathbf{r}}=z_{\mathbf{r}}-z\), i.e., the correlation function of a field \(y_{l\mathbf{r}}\) can be shifted by an l-independent field \(z_{\mathbf{r}}\) without changing the l-averaged volume average.

1.2 A.2 Derivation of correlation functions

Using these general relations, it is readily seen that the correlation functions defined as

$$\begin{aligned} C_{\mathrm {tot}}(\mathbf{r})\equiv & {} \mathbf {E}^l K[\mathcal{O}_{l\mathbf{r}}-\mathcal{O}](\mathbf{r}) = \mathbf {E}^l K[\mathcal{O}_{l\mathbf{r}}](\mathbf{r}) - \mathcal{O}^2 \end{aligned}$$
(68)
$$\begin{aligned} C_{\mathrm {ext}}(\mathbf{r})\equiv & {} \mathbf {E}^c K[\mathcal{O}_{c\mathbf{r}}-\mathcal{O}](\mathbf{r}) = \mathbf {E}^c K[\mathcal{O}_{c\mathbf{r}}](\mathbf{r}) - \mathcal{O}^2 \end{aligned}$$
(69)
$$\begin{aligned} C_{\mathrm {int}}(\mathbf{r})\equiv & {} \mathbf {E}^c \mathbf {E}^k K[\mathcal{O}_{ck\mathbf{r}}-\mathcal{O}_{c\mathbf{r}}](\mathbf{r}) \end{aligned}$$
(70)

are consistent with Eqs. (36). The index l runs again over all independent configurations c and all time-series k for each c and the expectation value \(\mathcal{O}\) is defined in Eq. (32). The corresponding equations in reciprocal space are given in Sec. 5.1, Eqs. (4749). That Eq. (68) is consistent with \(\delta \mathcal{O}^2_{\mathrm {tot}}= \mathbf {E}^{\mathbf{r}} C_{\mathrm {tot}}(\mathbf{r})\) and Eq. (69) with \(\delta \mathcal{O}^2_{\mathrm {ext}}= \mathbf {E}^{\mathbf{r}} C_{\mathrm {ext}}(\mathbf{r})\) is directly implied by Eqs. (61) and (61). To show that Eq. (70) is consistent with \(\delta \mathcal{O}^2_{\mathrm {int}}= \mathbf {E}^{\mathbf{r}} C_{\mathrm {int}}(\mathbf{r})\) and that all three correlation functions sum up according to Eq. (6), let us first note that due to Eq. (60) for \(\lambda =1\) the internal correlation function may be rewritten as

$$\begin{aligned} C_{\mathrm {int}}(\mathbf{r}) = \mathbf {E}^c \left\{ \mathbf {E}^k K[\mathcal{O}_{ck\mathbf{r}}](\mathbf{r}) - K[\mathcal{O}_{c\mathbf{r}}](\mathbf{r})\right\} . \end{aligned}$$
(71)

Using Eqs. (68) and (69), this implies \(C_{\mathrm {int}}(\mathbf{r}) = C_{\mathrm {tot}}(\mathbf{r}) - C_{\mathrm {ext}}(\mathbf{r})\) in agreement with the key relation Eq. (6) stated in the Introduction. In turn we thus have

$$\begin{aligned} \mathbf {E}^{\mathbf{r}} C_{\mathrm {int}}(\mathbf{r})= & {} \mathbf {E}^{\mathbf{r}} \left( C_{\mathrm {tot}}(\mathbf{r}) - C_{\mathrm {ext}}(\mathbf{r})\right) \nonumber \\= & {} \delta \mathcal{O}^2_{\mathrm {tot}}- \delta \mathcal{O}^2_{\mathrm {ext}}= \delta \mathcal{O}^2_{\mathrm {int}}\end{aligned}$$
(72)

where we have used Eq. (1) in the last step.

Please note that due to Eq. (67) \(\delta \mathcal{O}^2_{\mathrm {int}}= \mathbf {E}^{\mathbf{r}} C_{\mathrm {int}}(\mathbf{r})\) would also be solved by the more general internal correlation function

$$\begin{aligned} C_{\mathrm {int}}(\mathbf{r},\lambda ) \equiv \mathbf {E}^c \mathbf {E}^k K[(\mathcal{O}_{ck\mathbf{r}}-\mathcal{O}_c)-\lambda (\mathcal{O}_{c\mathbf{r}}-\mathcal{O}_c)](\mathbf{r})\qquad \end{aligned}$$
(73)

which reduces to Eq. (70) for \(\lambda =1\), since it is possible to shift \(\mathcal{O}_{ck\mathbf{r}}-\mathcal{O}_c\) with the k-independent field \(\lambda (\mathcal{O}_{c\mathbf{r}}-\mathcal{O}_c)\) without changing the k-averaged volume average. The trouble with such alternative definitions is that Eq. (6) does not hold anymore in general, e.g., it can be shown that Eq. (73) leads to

$$\begin{aligned} C_{\mathrm {tot}}(\mathbf{r}) = C_{\mathrm {int}}(\mathbf{r},\lambda ) + \lambda (2-\lambda ) C_{\mathrm {ext}}(\mathbf{r}) + (\lambda -1)^2 \delta \mathcal{O}^2_{\mathrm {ext}}.\nonumber \\ \end{aligned}$$
(74)

Due to the last term and since \(\delta \mathcal{O}^2_{\mathrm {ext}}> 0\) for non-ergodic systems all three correlation functions may in principle only vanish for the same \(\mathbf{r}\) for \(\lambda =1\) and for exactly this limit Eq. (74) reduces to Eq. (6). We therefore set \(\lambda =1\).

1.3 A.3 Important limits

We have omitted for clarity in the preceding subsection all possible dependences on \(N_\mathrm {c}\), \(N_\mathrm {k}\), \(\varDelta \tau \) and V. However, it is assumed below that \(N_\mathrm {c}\) and \(N_\mathrm {k}\) are arbitrarily large, i.e., all properties are \(N_\mathrm {c}\)- and \(N_\mathrm {k}\)-independent. Moreover, we focus on the limit \(\varDelta \tau \gg \tau _{\mathrm {b}}\), i.e., both \(\mathcal{O}_{c\mathbf{r}} = \mathbf {E}^k \mathcal{O}_{ck\mathbf{r}}\) and its average \(\mathcal{O}= \mathbf {E}^c \mathbf {E}^{\mathbf{r}} \mathcal{O}_{c\mathbf{r}}\) are \(\varDelta \tau \)-independent to leading order. Due to Eq. (69) the same holds for the external correlation function, i.e.,

$$\begin{aligned} C_{\mathrm {ext}}(\mathbf{r},\varDelta \tau ,V) \simeq C_{\mathrm {ext}}(\mathbf{r},V) \text{ for } \varDelta \tau \gg \tau _{\mathrm {b}}. \end{aligned}$$
(75)

The indicated V-dependence drops out if \(\mathcal{O}_{c\mathbf{r}}\) is V-independent as in all the models of this work.

The internal correlation function, Eq. (70), characterizes the correlations of the difference \(\mathcal{O}_{ck\mathbf{r}}(\varDelta \tau ) - \mathcal{O}_{c\mathbf{r}}\). While \(\mathcal{O}_{ck\mathbf{r}}(\varDelta \tau )\) depends in general not only on k but also on \(\varDelta \tau \), both dependences drop out for \(\varDelta \tau \rightarrow \infty \). Hence, \(\mathcal{O}_{ck\mathbf{r}}(\varDelta \tau ) \rightarrow \mathcal{O}_{c\mathbf{r}}\) and in turn

$$\begin{aligned} \lim _{\varDelta \tau \rightarrow \infty } C_{\mathrm {int}}(\mathbf{r},\varDelta \tau ,V) = 0. \end{aligned}$$
(76)

To obtain the internal correlation function for finite \(\varDelta \tau \gg \tau _{\mathrm {b}}\), it should be remembered that \(\mathcal{O}_{ck\mathbf{r}}(\varDelta \tau )\) is a time-averaged moment over \(N_\mathrm {t}=\varDelta \tau /\delta \tau \) data entries from one stored time-series. The internal correlation function can thus be written as an average

$$\begin{aligned} C_{\mathrm {int}}(\mathbf{r}=\mathbf{r}_2-\mathbf{r}_1,\varDelta \tau ,V) = \mathbf {E}^c \mathbf {E}^{\mathbf{r}_1} \left( \mathbf {E}^{t_1} \mathbf {E}^{t_2} \mathbf {E}^k \ldots \right) \end{aligned}$$
(77)

over entries measured at discrete times \(t_1\) and \(t_2\). A specific example is worked out in Appendix B.1. If one assumes for simplicity that \(\delta \tau \gg \tau _{\mathrm {b}}\) only contributions with \(t_1=t_2\) can contribute. Using also that the time-average \(\mathbf {E}^t\) is normalized by \(N_\mathrm {t}\propto \varDelta \tau \), this shows that quite generally the internal correlation function must decay to leading order for all \(\mathbf{r}\) as

$$\begin{aligned} C_{\mathrm {int}}(\mathbf{r},\varDelta \tau ,V) \propto \frac{1}{\varDelta \tau } \text{ for } \varDelta \tau \gg \tau _{\mathrm {b}}\end{aligned}$$
(78)

as expected from \(\delta \mathcal{O}^2_{\mathrm {int}}(\varDelta \tau ) = \mathbf {E}^{\mathbf{r}} C_{\mathrm {int}}(\mathbf{r}) \propto 1/\varDelta \tau \).

B Scaling of \(C_{\mathrm {int}}[v](\mathbf{r},\varDelta \tau ,V)\)

1.1 B.1 Predictions for LSM-A

Table 2 k-averages \(\mathbf {E}^k A_1A_2\), \(\mathbf {E}^k A_1B_2\), \(\mathbf {E}^k A_2B_1\) and \(\mathbf {E}^k B_1B_2\) for LSM-A. The different relevant cases for \(\mathbf{r}_1\), \(\mathbf{r}_2\), \(\mathbf{r}_3\), \(\mathbf{r}_4\) are indicated in the first column. Note that \(r = |\mathbf{r}|\) with \(\mathbf{r}=\mathbf{r}_2-\mathbf{r}_1\). The first two cases indicate contributions for \(r=0\), the last two cases contributions for \(r>0\). \(\mathbf {E}^k A_1B_2\) and \(\mathbf {E}^k A_2B_1\) are identical by symmetry. The leading contributions of order \(\mathcal{O}(\varDelta \tau ^{-1})\) are due to \(\mathbf {E}^k A_1A_2\) (second column). The second case (\(\mathbf{r}_1=\mathbf{r}_2 \ne \mathbf{r}_3=\mathbf{r}_4\)) yields contributions proportional to the system size

As noted in Appendix A.1, all correlation functions \(C[f](\mathbf{r})\) discussed in the present work must vanish if Eq. (65) holds, i.e., if two typical points of the field f at a respective distance \(\mathbf{r}\) are uncorrelated. Here we draw attention to the fact that although the primary instantaneous field \(x_{t\mathbf{r}}\) may be uncorrelated this may not be the case for the field \(\mathcal{O}_{\mathbf{r}}\) associated with the time-averaged functional \(\mathcal{O}[\mathbf {x}]\) of the time series \(\mathbf {x}\).Footnote 20 As we shall see, this matters specifically for the covariance field, Eq. (30),

$$\begin{aligned} v_{ck\mathbf{r}}= & {} V \left\{ \mathbf {E}^t x_{ckt\mathbf{r}}x_{ckt} - x_{ck\mathbf{r}} x_{ck} \right\} \nonumber \\= & {} \mathbf {S}^{\mathbf{r}'} \left\{ \mathbf {E}^t x_{ckt\mathbf{r}} x_{ckt\mathbf{r}'} - \mathbf {E}^t \mathbf {E}^{t'} x_{ckt\mathbf{r}} x_{ckt'\mathbf{r}'}\right\} \end{aligned}$$
(79)

restated for convenience omitting the irrelevant prefactor \(\beta \) and introducing the sum \(\mathbf {S}^{\mathbf{r}'} \equiv V\mathbf {E}^{\mathbf{r}'}\). In the second step we have made explicit the crucial contributions with \(\mathbf{r}'\ne \mathbf{r}\) to the averages \(x_{ckt}\) and \(x_{ck}\).

Without restricting the generality of the argument, let us focus on the model LSM-A with \(J=0\), i.e., the \(x_{ckt\mathbf{r}}\) of different microcells \(\mathbf{r}\) are uncorrelated. Ultimately, we want to expand \(v_{ck\mathbf{r}}\) for large \(\varDelta \tau \). Reminding Eq. (31), it is thus useful that the \(x_{ckt\mathbf{r}}\) in Eq. (79) can be replaced by \(\delta x_{ckt\mathbf{r}} \equiv x_{ckt\mathbf{r}}-a_{c\mathbf{r}}\) without changing \(v_{ck\mathbf{r}}\). Moreover, let us assume that the data is sampled with large time increments \(\delta \tau \gg \tau _{\mathrm {b}}\), i.e., different times \(t'\) and \(t''\) can be considered to be uncorrelated. This implies that

$$\begin{aligned} \mathbf {E}^k \delta x_{ckt'\mathbf{r}'}^{p'} \delta x_{ckt''\mathbf{r}''}^{p''} = \mathbf {E}^k \delta x_{ckt'\mathbf{r}'}^{p'} \mathbf {E}^k \delta x_{ckt''\mathbf{r}''}^{p''} \end{aligned}$$
(80)

whenever \(t'\ne t''\) or \(\mathbf{r}'\ne \mathbf{r}''\) (\(p'\) and \(p''\) being integers) and, moreover, the following useful relations hold:

$$\begin{aligned} \mathbf {E}^{t}\mathbf {E}^{t'} \delta _{tt'}= & {} 1/N_\mathrm {t}, \end{aligned}$$
(81)
$$\begin{aligned} \mathbf {E}^k \delta x_{ckt\mathbf{r}}= & {} 0, \end{aligned}$$
(82)
$$\begin{aligned} \mathbf {E}^k \delta x_{ckt\mathbf{r}} \delta x_{ckt'\mathbf{r}'}= & {} b_{c\mathbf{r}} \ \delta _{\mathbf{r}\mathbf{r}'} \delta _{tt'}, \end{aligned}$$
(83)
$$\begin{aligned} \mathbf {E}^k \delta x_{ckt\mathbf{r}}^2 \delta x_{ckt\mathbf{r}'}^2= & {} 3 b_{c\mathbf{r}}^2 \ \delta _{\mathbf{r}\mathbf{r}'} + b_{c\mathbf{r}}b_{c\mathbf{r}'}(1-\delta _{\mathbf{r}\mathbf{r}'}). \end{aligned}$$
(84)

We have used above that the lattice springs are Gaussian variables, Eq. (12), i.e., for each independent spring of length \(x_{c\mathbf{r}}\) (dropping the indices k and t) we have \(\left<\delta x_{c\mathbf{r}} \right>= 0\), \(\left<\delta x_{c\mathbf{r}}^2 \right>= b_{c\mathbf{r}}\) and \(\left<\delta x_{c\mathbf{r}}^4 \right>= 3 b_{c\mathbf{r}}^2\) with \(\left<\ldots \right>\) denoting the thermal average within each basin c. Using Eqs. (8183), we obtain, e.g., the k-average

$$\begin{aligned} v_{c\mathbf{r}} = \mathbf {E}^{k} v_{ck\mathbf{r}} = b_{c\mathbf{r}} \left( 1-1/N_\mathrm {t}\right) \end{aligned}$$
(85)

and, hence, the total average \(v=\mathbf {E}^c\mathbf {E}^{\mathbf{r}} v_{c\mathbf{r}} = b (1-1/N_\mathrm {t})\) with \(b= \mathbf {E}^c\mathbf {E}^{\mathbf{r}} b_{c\mathbf{r}}\).

To compute the internal correlation function, we write

$$\begin{aligned} C_{\mathrm {int}}[v](\mathbf{r})= & {} \mathbf {E}^c C_c(\mathbf{r}) \text{ with } \end{aligned}$$
(86)
$$\begin{aligned} C_c(\mathbf{r})= & {} \mathbf {E}^k K[(v_{ck\mathbf{r}}-b_{c\mathbf{r}})-(v_{c\mathbf{r}}-b_{c\mathbf{r}})](\mathbf{r}) \end{aligned}$$
(87)
$$\begin{aligned}= & {} \mathbf {E}^k K[v_{ck\mathbf{r}}-b_{c\mathbf{r}}](\mathbf{r}) - K[v_{c\mathbf{r}}-b_{c\mathbf{r}}](\mathbf{r}) \end{aligned}$$
(88)

using Eq. (71) in the last step. Due to Eq. (85), it is clear that the second term in Eq. (88) is of order \(\mathcal{O}(\varDelta \tau ^{-2})\). We can focus on the first term which we rewrite as volume average \(\mathbf {E}^{\mathbf{r}_1} {\tilde{C}}_c(\mathbf{r}_1,\mathbf{r}_2=\mathbf{r}_1+\mathbf{r})\) with

$$\begin{aligned} {\tilde{C}}_{c\mathbf{r}_1\mathbf{r}_2}(\mathbf{r}_1,\mathbf{r}_2)\equiv & {} \mathbf {E}^k (v_{ck\mathbf{r}_1}-b_{c\mathbf{r}_1}) (v_{ck\mathbf{r}_2}-b_{c\mathbf{r}_2}) \end{aligned}$$
(89)
$$\begin{aligned}= & {} \mathbf {E}^k (A_1 - B_1) (A_2 -B_2) \end{aligned}$$
(90)

where Eq. (79) is used for the definition of the terms

$$\begin{aligned} A_1= & {} \mathbf {S}^{\mathbf{r}_3} \mathbf {E}^{t_1} \left( \delta x_{ckt_1\mathbf{r}_1} \delta x_{ckt_1\mathbf{r}_3} - b_{c\mathbf{r}_1} \delta _{\mathbf{r}_1\mathbf{r}_3}\right) , \end{aligned}$$
(91)
$$\begin{aligned} A_2= & {} \mathbf {S}^{\mathbf{r}_4} \mathbf {E}^{t_2} \left( \delta x_{ckt_2\mathbf{r}_2} \delta x_{ckt_2\mathbf{r}_4} - b_{c\mathbf{r}_2} \delta _{\mathbf{r}_2\mathbf{r}_4}\right) , \end{aligned}$$
(92)
$$\begin{aligned} B_1= & {} \mathbf {S}^{\mathbf{r}_3} \mathbf {E}^{t_1} \mathbf {E}^{t_3} \delta x_{ckt_1\mathbf{r}_1} \delta x_{ckt_3\mathbf{r}_3} \ \text{ and } \end{aligned}$$
(93)
$$\begin{aligned} B_2= & {} \mathbf {S}^{\mathbf{r}_4} \mathbf {E}^{t_2} \mathbf {E}^{t_4} \delta x_{ckt_2\mathbf{r}_2} \delta x_{ckt_4\mathbf{r}_4}. \end{aligned}$$
(94)

We expand the different contributions and k-average using the identities Eqs. (8084). As summarized in Table 2 it is helpful to distinguish four cases for \(\mathbf{r}_1\), \(\mathbf{r}_2\), \(\mathbf{r}_3\) and \(\mathbf{r}_4\). Most contributions are of order \(\mathcal{O}(\varDelta \tau ^{-2})\) and only three contributions of \(\mathbf {E}^kA_1A_2\) (second column) do matter. A central result is that due to the last case (\(\mathbf{r}_1=\mathbf{r}_4 \ne \mathbf{r}_2=\mathbf{r}_3\)), the internal correlation function must remain finite for \(r > 0\). Note also that all terms for the second case (\(\mathbf{r}_1=\mathbf{r}_2 \ne \mathbf{r}_3=\mathbf{r}_4\)) increase linearly with V due to the sum \(\mathbf {S}^{\mathbf{r}_3}\mathbf {S}^{\mathbf{r}_4} \delta _{\mathbf{r}_3\mathbf{r}_4} (1-\delta _{\mathbf{r}_1\mathbf{r}_3})\). In fact, using \(b_c=\mathbf {E}^{\mathbf{r}_3}b_{c\mathbf{r}_3}\), the indicated term for \(\mathbf {E}^kA_1A_2\) can be rewritten as

$$\begin{aligned} b_{c\mathbf{r}_1} (\mathbf {S}^{\mathbf{r}_3\ne \mathbf{r}_1}b_{c\mathbf{r}_3})/N_\mathrm {t}= b_{c\mathbf{r}_1} b_c V/N_\mathrm {t}- b_{c\mathbf{r}_1}^2/N_\mathrm {t}. \end{aligned}$$
(95)

We finally average over \(\mathbf{r}_1\) and c using that the \(b_{c\mathbf{r}'}\) are decorrelated for different \(\mathbf{r}'\). Summarizing all terms, we obtain to leading order

$$\begin{aligned} C_{\mathrm {int}}[v](\mathbf{r}) \simeq \frac{b^2}{N_\mathrm {t}} \times \left\{ \begin{array}{ll} V+1 &{} \text{ for } r = 0,\\ 1 &{} \text{ for } r > 0. \end{array} \right. \end{aligned}$$
(96)

As a consequence \(\delta v^2_{\mathrm {int}}(\varDelta \tau ,V) = \mathbf {E}^{\mathbf{r}} C_{\mathrm {int}}[v](\mathbf{r}) = 2b^2/N_\mathrm {t}\propto V^0/\varDelta \tau \), as expected.

1.2 B.2 Scaling for general Gaussian fields

It is clear that the above result can be generalized to other models with short-range correlations and general \(\delta \tau \) including \(\delta \tau \ll \tau _{\mathrm {b}}\). This merely requires a renormalization of space and time. Especially this suggests to replace \(1/N_\mathrm {t}\) by \(\delta v^2_{\mathrm {int}}(\varDelta \tau )\). It is in this context of relevance that the above result Eq. (96) can be recast as

$$\begin{aligned} C_{\mathrm {int}}[v](\mathbf{r}) \simeq \left\{ V (1-\alpha ) c(\mathbf{r})+\alpha \right\} \delta v^2_{\mathrm {int}}(\varDelta \tau ) \end{aligned}$$
(97)

with \(\mathbf {S}^{\mathbf{r}} c(\mathbf{r}) = 1\), \(c(\mathbf{r}) \rightarrow 0\) for large r (with \(r \ne 0\) for LSM-A) and \(\alpha =1/2\). Note that \(\mathbf {E}^{\mathbf{r}} C_{\mathrm {int}}[v](\mathbf{r}) = \delta v^2_{\mathrm {int}}(\varDelta \tau )\) holds for all coefficients \(\alpha \). As discussed in Sect. 5.2, the numerical results of all our LSM variants are consistent with this generalization of the direct calculation for the simple LSM-A model. (As far as we can tell this even holds reasonably for systems with long-range correlations.) There is in fact a general reason for expecting Eq. (97) to hold for many models: For a given c the \(x_{t\mathbf{r}}\)-fluctuations are often nearly Gaussian. [For the LSM variants the joint distributions of the \(x_{t\mathbf{r}}\) are in fact exactly Gaussian since the total energy is quadratic in \(x_{t\mathbf{r}}\), Eq. (12).] This allows for a theoretical treatment of \(C_{\mathrm {int}}[v](\mathbf{r})\) based on the cumulant formalism (“Wick’s theorem”) similar to the calculation which leads to Eq. (45) for the global variance \(\delta v^2_{\mathrm {int}}(\varDelta \tau )\) [2, 21]. It is thus possible to show that Eq. (97) must hold for general fluctuating Gaussian fields with \(\alpha =1/2\). This calculation is beyond of the scope of the present work.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wittmer, J.P., Semenov, A.N. & Baschnagel, J. Different types of spatial correlation functions for non-ergodic stochastic processes of macroscopic systems. Eur. Phys. J. E 45, 65 (2022). https://doi.org/10.1140/epje/s10189-022-00222-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epje/s10189-022-00222-1

Navigation