Skip to main content
Log in

Designing optimal experiments: an application to proton Compton scattering

  • Regular Article – Theoretical Physics
  • Published:
The European Physical Journal A Aims and scope Submit manuscript

Abstract

Interpreting measurements requires a physical theory, but the theory’s accuracy may vary across the experimental domain. To optimize experimental design, and so to ensure that the substantial resources necessary for modern experiments are focused on acquiring the most valuable data, both the theory uncertainty and the expected pattern of experimental errors must be considered. We develop a Bayesian approach to this problem, and apply it to the example of proton Compton scattering. Chiral Effective Field Theory (\(\chi \)EFT) predicts the functional form of the scattering amplitude for this reaction, so that the electromagnetic polarizabilities of the nucleon can be inferred from data. With increasing photon energy, both experimental rates and sensitivities to polarizabilities increase, but the accuracy of \(\chi \)EFT decreases. Our physics-based model of \(\chi \)EFT truncation errors is combined with present knowledge of the polarizabilities and reasonable assumptions about experimental capabilities at HI\(\gamma \)S and MAMI to assess the information gain from measuring specific observables at specific kinematics, i.e.  to determine the relative amount by which new data are apt to shrink uncertainties. The strongest gains would likely come from new data on the spin observables \(\Sigma _{2x}\) and \(\Sigma _{2x^\prime }\) at \(\omega \simeq 140\) to 200 MeV and \(40^\circ \) to \(120^\circ \). These would tightly constrain \(\gamma _{E1E1}-\gamma _{E1M2}\). New data on the differential cross section between 100 and 200 MeV and over a wide angle range will substantially improve constraints on \(\alpha _{E1}-\beta _{M1}\), \(\gamma _\pi \) and \(\gamma _{M1M1}-\gamma _{M1E2}\). Good signals also exist around 160 MeV for \(\Sigma _3\) and \(\Sigma _{2z^\prime }\). Such data will be pivotal in the continuing quest to pin down the scalar polarizabilities and refine understanding of the spin polarizabilities.

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

Similar content being viewed by others

Data Availability Statement

This manuscript has no associated data or the data will not be deposited. [Authors’ comment: The data and code is available as Jupyter notebook from github, see Ref. [29]. The Supplemental Material mentioned in the main article is available at https://doi.org/10.1140/epja/s10050-021-00382-2.]

Notes

  1. In the \(\chi \)EFT we are using, the lowest-lying nucleonic resonance, the \(\Delta (1232)\), is retained as an explicit degree of freedom.

  2. In Compton scattering, these include the six nucleon polarizabilities in the combinations defined in Sect. 2: \(\alpha _{E1}\pm \beta _{M1}\) and \(\gamma _0,\,\gamma _\pi ,\,\gamma _{E-},\,\gamma _{M-}\).

  3. Interestingly, the spikes disappear when considering the \(c_n\) for rate-differences \(\,{\mathrm{d}}{\sigma }\times \Sigma _i\), but a Gaußian uncertainty in these and in \(\,{\mathrm{d}}{\sigma }\) does not lead to a simple Gaußian uncertainty in \(\Sigma _i\). Further study of the convergence patterns of \(\Sigma _i\) near \(\omega _\pi \) is needed.

  4. In this regime, the relation between the expansion used here and the notation of Refs. [4,5,6, 28] is \(Q^n=e^2\delta ^n\); the simpler symbol leads to more compact formulae later. The LO defined in this presentation corresponds to performing the \(\chi \)EFT power counting on the structure part of the nucleon Compton amplitude, i.e., what remains after the (relativistic) nucleon and pion Born terms are subtracted, and thus differs in that detail from the power counting described in Refs. [4,5,6].

  5. A single design \({\mathbf {d}}\) in this work is specified by an observable and a set of kinematic points at which to measure it, and possibly the experimental noise levels. The space D is the set of all considered experiments over which the utility is optimized, e.g., all possible 5-angle measurements at a given energy.

  6. The notation \({\mathcal {N}}(\vec {\mu }_0, V_0)\) denotes a Gaußian with mean \(\vec {\mu }_0\) and covariance \(V_0\). See also discussion of notation below Eq. (9).

References

  1. H .W. Griesshammer, J .A. McGovern, D .R. Phillips, G. Feldman, Prog. Part. Nucl. Phys. 67, 841 (2012). https://doi.org/10.1016/j.ppnp.2012.04.003. arXiv:1203.6834

    Article  ADS  Google Scholar 

  2. B .R. Holstein, S. Scherer, Ann. Rev. Nucl. Part. Sci. 64, 51 (2014). https://doi.org/10.1146/annurev-nucl-102313-025555. arXiv:1401.0140 [hep-ph]

    Article  ADS  Google Scholar 

  3. F. Hagelstein, R. Miskimen, V. Pascalutsa, Prog. Part. Nucl. Phys. 88, 29 (2016). https://doi.org/10.1016/j.ppnp.2015.12.001. arXiv:1512.03765 [nucl-th]

    Article  ADS  Google Scholar 

  4. H .W. Griesshammer, J .A. McGovern, D .R. Phillips, Eur. Phys. J. A 52, 139 (2016). https://doi.org/10.1140/epja/i2016-16139-5. arXiv:1511.01952 [nucl-th]

    Article  ADS  Google Scholar 

  5. J .A. McGovern, D .R. Phillips, H .W. Griesshammer, Eur. Phys. J. A 49, 12 (2013). https://doi.org/10.1140/epja/i2013-13012-1. arXiv:1210.4104

    Article  ADS  Google Scholar 

  6. H .W. Griesshammer, J .A. McGovern, D .R. Phillips, Eur. Phys. J. A 54, 37 (2018). https://doi.org/10.1140/epja/i2018-12467-8. arXiv:1711.11546

    Article  ADS  Google Scholar 

  7. A. Margaryan, B. Strandberg, H .W. Grießhammer, J .A. Mcgovern, D .R. Phillips, D. Shukla, Eur. Phys. J. A 54, 125 (2018). https://doi.org/10.1140/epja/i2018-12554-x. arXiv:1804.00956 [nucl-th]

    Article  ADS  Google Scholar 

  8. H. W. Grießhammer, Eur. Phys. J. A 49, 100 (2013), https://doi.org/10.1140/epja/i2018-12502-x, https://doi.org/10.1140/epja/i2013-13100-2, https://doi.org/10.1140/epja/i2017-12311-9 [Erratum: Eur. Phys. J.A53,no.5,113(2017); Erratum: Eur. Phys. J.A54,no.4,57(2018)], arXiv:1304.6594 [nucl-th]

  9. P. Martel (A2), Proceedings, 9th International Workshop on Chiral Dynamics (CD18): Durham (NC), United States, PoS CD2018, 038 (2019) https://doi.org/10.22323/1.317.0038

  10. P. Martel, M. Biroth, C. Collicott, D. Paudyal, A. Rajabi (A2), Proceedings, International Symposium Advances in Dark Matter and Particle Physics (ADMPP16): Messina, Italy, October 24-27, 2016, EPJ Web Conf. 142, 01021 (2017) https://doi.org/10.1051/epjconf/201714201021

  11. G. M. Huber, C. Collicott, in Proceedings, 12th Conference on the Intersections of Particle and Nuclear Physics (CIPANP 2015): Vail, Colorado, USA, May 19-24, 2015 (2015) arXiv:1508.07919 [nucl-ex]

  12. M. W. Ahmed, Proceedings, 9th International Workshop on Chiral Dynamics (CD18): Durham (NC), United States, PoS CD2018, 001 (2020) https://doi.org/10.22323/1.317.0001

  13. H .R. Weller, M .W. Ahmed, H. Gao, W. Tornow, Y .K. Wu, M. Gai, R. Miskimen, Prog. Part. Nucl. Phys. 62, 257 (2009). https://doi.org/10.1016/j.ppnp.2008.07.001

    Article  ADS  Google Scholar 

  14. M. Catacora-Rios, G. King, A. Lovell, F. Nunes, Phys. Rev. C 100, 064615 (2019). https://doi.org/10.1103/PhysRevC.100.064615

    Article  ADS  Google Scholar 

  15. K. Chaloner, I. Verdinelli, Stat. Sci. 10, 273 (1995)

    Article  Google Scholar 

  16. J. Liepe, S. Filippi, M. Komorowski, M .P .H. Stumpf, PLoS Comput. Biol. 9, e1002888 (2013). https://doi.org/10.1371/journal.pcbi.1002888

    Article  ADS  Google Scholar 

  17. E .G. Ryan, C .C. Drovandi, J .M. McGree, A .N. Pettitt, Int. Stat. Rev. 84, 128 (2016). https://doi.org/10.1111/insr.12107. https://onlinelibrary.wiley.com/doi/pdf/10.1111/insr.12107

  18. J. I. Myung, M. Pitt, Y. Tang, D. R. Cavagnaro (2009). http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.703.392&rep=rep1&type=pdf

  19. I. Vernon, M. Goldstein, R .G. Bower, Bayesian Anal. 5, 619 (2010). https://doi.org/10.1214/10-BA524

    Article  MathSciNet  Google Scholar 

  20. L .M. Berliner, Y. Kim, J. Clim. 21, 1891 (2008). https://doi.org/10.1175/2007JCLI1619.1

    Article  ADS  Google Scholar 

  21. J .A. Cumming, M. Goldstein, Technometrics 51, 377 (2009). https://doi.org/10.1198/TECH.2009.08015

    Article  MathSciNet  Google Scholar 

  22. M. Farrow, M. Goldstein, J. Stat. Plan. Inference 136, 498 (2006)

    Article  Google Scholar 

  23. T.J. Santner, B.J. Williams, W.I. Notz, The design and analysis of computer experiments, vol. 1 (Springer, Berlin, 2003)

    Book  Google Scholar 

  24. C. Currin, T. Mitchell, M. Morris, D. Ylvisaker, A Bayesian approach to the design and analysis of computer experiments, Tech. Rep. (Oak Ridge National Lab., TN (USA), 1988)

  25. M. Jones, M. Goldstein, P. Jonathan, D. Randell, J. Stat. Plan. Inference 171, 115 (2016)

    Article  Google Scholar 

  26. G. E. P. Box, in Robustness in Statistics, edited by R. L. Launer and G. N. Wilkinson (Academic Press, 1979) p. 201. https://doi.org/10.1016/B978-0-12-438150-6.50018-2

  27. J .A. Melendez, R .J. Furnstahl, D .R. Phillips, M .T. Pratola, S. Wesolowski, Phys. Rev. C 100, 044001 (2019). https://doi.org/10.1103/PhysRevC.100.044001. arXiv:1904.10581

    Article  ADS  Google Scholar 

  28. V. Pascalutsa, D .R. Phillips, Phys. Rev. C 67, 055202 (2003). https://doi.org/10.1103/PhysRevC.67.055202. arXiv:nucl-th/0212024

    Article  ADS  Google Scholar 

  29. BUQEYE collaboration (2020) https://buqeye.github.io/software/

  30. A. Walker-Loud, C .E. Carlson, G .A. Miller, Phys. Rev. Lett. 108, 232301 (2012). https://doi.org/10.1103/PhysRevLett.108.232301. arXiv:1203.0254 [nucl-th]

    Article  ADS  Google Scholar 

  31. A. Walker-Loud, Proceedings, 9th International Workshop on Chiral Dynamics (CD18): Durham (NC), United States, PoS CD2018, 045 (2019), https://doi.org/10.22323/1.317.0045. arXiv:1907.05459 [nucl-th]

  32. J. Gasser, M. Hoferichter, H. Leutwyler, A. Rusetsky, Eur. Phys. J. C 75, 375 (2015). https://doi.org/10.1140/epjc/s10052-015-3580-9. arXiv:1506.06747 [hep-ph]

    Article  ADS  Google Scholar 

  33. J. Gasser, H. Leutwyler, A. Rusetsky, Phys. Lett. B 814, 136087 (2021). https://doi.org/10.1016/j.physletb.2021.136087

  34. G. Blanpied et al., Phys. Rev. C 64, 025203 (2001). https://doi.org/10.1103/PhysRevC.64.025203

    Article  ADS  Google Scholar 

  35. V. Sokhoyan et al. (A2), Eur. Phys. J. A53, 14 ( 2017), https://doi.org/10.1140/epja/i2017-12203-0. arXiv:1611.03769 [nucl-ex]

  36. C. Collicott, Ph.D. thesis, Dalhousie University (2015), https://wwwa2.kph.uni-mainz.de/images/publications/phd/thesis_Collicott-Cristina-2015.pdf

  37. P. P. Martel et al. (A2), Phys. Rev. Lett. 114, 112501 (2015), https://doi.org/10.1103/PhysRevLett.114.112501. arXiv:1408.1576 [nucl-ex]

  38. P. P. Martel, Ph.D. thesis, University of Massachusetts, Amherst (2012), https://wwwa2.kph.uni-mainz.de/images/publications/phd/pmartel_dis.pdf

  39. D. Paudyal et al. (A2), Phys. Rev. C 102, 035205 (2020), https://doi.org/10.1103/PhysRevC.102.035205. arXiv:1909.02032 [nucl-ex]

  40. S. Wesolowski, R .J. Furnstahl, J .A. Melendez, D .R. Phillips, J. Phys. G 46, 045102 (2019). https://doi.org/10.1088/1361-6471/aaf5fc. arXiv:1808.08211

    Article  ADS  Google Scholar 

  41. R .J. Furnstahl, N. Klco, D .R. Phillips, S. Wesolowski, Phys. Rev. C 92, 024005 (2015). https://doi.org/10.1103/PhysRevC.92.024005. arXiv:1506.01343

    Article  ADS  Google Scholar 

  42. J .A. Melendez, S. Wesolowski, R .J. Furnstahl, Phys. Rev. C 96, 024003 (2017). https://doi.org/10.1103/PhysRevC.96.024003. arXiv:1704.03308

    Article  ADS  Google Scholar 

  43. J. Sacks, W.J. Welch, T.J. Mitchell, H.P. Wynn, Stat. Sci. 4, 409 (1989)

    Google Scholar 

  44. N. Cressie, Terra Nova 4, 613 (1992). https://doi.org/10.1002/9781119115151

    Article  ADS  Google Scholar 

  45. C. Rasmussen, C. Williams, Gaussian Processes for Machine Learning, Adaptative computation and machine learning series (University Press Group Limited, 2006)

  46. D. J. C. MacKay, in Neural Networks and Machine Learning, NATO ASI Series, Vol. 168, edited by C. M. Bishop ( Springer, 1998) pp. 133–166

  47. D.J.C. MacKay, Information theory, inference, and learning algorithms (Cambridge University Press, Cambridge, New York, 2003)

    MATH  Google Scholar 

  48. A. Gelman, J.B. Carlin, H.S. Stern, D.B. Dunson, A. Vehtari, D.B. Rubin, Bayesian data analysis (CRC Press, Boca Raton, 2013)

    Book  Google Scholar 

  49. O. Gryniuk, F. Hagelstein, V. Pascalutsa, Phys. Rev. D 92, 074031 (2015). https://doi.org/10.1103/PhysRevD.92.074031. arXiv:1508.07952 [nucl-th]

    Article  ADS  Google Scholar 

  50. M .I. Levchuk, Nucl. Phys. A 674, 449 (2000). https://doi.org/10.1016/S0375-9474(00)00145-7. arXiv:nucl-th/9909066 [nucl-th]

    Article  ADS  Google Scholar 

  51. L .S. Myers et al., Phys. Rev. C 92, 025203 (2015). https://doi.org/10.1103/PhysRevC.92.025203. arXiv:1503.08094 [nucl-ex]

    Article  ADS  Google Scholar 

  52. L. S. Myers et al. (COMPTON@MAX-lab), Phys. Rev. Lett. 113, 262506 (2014), https://doi.org/10.1103/PhysRevLett.113.262506. arXiv:1409.3705 [nucl-ex]

  53. B. Pasquini, P. Pedroni, D. Drechsel, Phys. Lett. B 687, 160 (2010). https://doi.org/10.1016/j.physletb.2010.03.007. arXiv:1001.4230 [hep-ph]

    Article  ADS  Google Scholar 

  54. M. Schumacher, Prog. Part. Nucl. Phys. 55, 567 (2005). https://doi.org/10.1016/j.ppnp.2005.01.033. arXiv:hep-ph/0501167 [hep-ph]

    Article  ADS  Google Scholar 

  55. O. Gryniuk, F. Hagelstein, V. Pascalutsa, Phys. Rev. D 94, 034043 (2016). https://doi.org/10.1103/PhysRevD.94.034043. arXiv:1604.00789 [nucl-th]

    Article  ADS  Google Scholar 

  56. V. Lensky, J. McGovern, V. Pascalutsa, Eur. Phys. J. C 75, 604 (2015). https://doi.org/10.1140/epjc/s10052-015-3791-0. arXiv:1510.02794 [hep-ph]

    Article  ADS  Google Scholar 

  57. D. Babusci, G. Giordano, A .I. L’vov, G. Matone, A .M. Nathan, Phys. Rev. C 58, 1013 (1998). https://doi.org/10.1103/PhysRevC.58.1013. arXiv:hep-ph/9803347 [hep-ph]

    Article  ADS  Google Scholar 

  58. R .P. Hildebrandt, H .W. Grießhammer, T .R. Hemmert, B. Pasquini, Eur. Phys. J. A 20, 293 (2004). https://doi.org/10.1140/epja/i2003-10144-9. arXiv:nucl-th/0307070 [nucl-th]

    Article  ADS  Google Scholar 

  59. B .R. Holstein, D. Drechsel, B. Pasquini, M. Vanderhaeghen, Phys. Rev. C 61, 034316 (2000). https://doi.org/10.1103/PhysRevC.61.034316. arXiv:hep-ph/9910427 [hep-ph]

    Article  ADS  Google Scholar 

  60. B. Pasquini, P. Pedroni, S. Sconfietti, Phys. Rev. C 98, 015204 (2018). https://doi.org/10.1103/PhysRevC.98.015204. arXiv:1711.07401 [hep-ph]

    Article  ADS  Google Scholar 

  61. N. Krupina, V. Lensky, V. Pascalutsa, Phys. Lett. B 782, 34 (2018). https://doi.org/10.1016/j.physletb.2018.04.066. arXiv:1712.05349 [nucl-th]

    Article  ADS  Google Scholar 

  62. B. Pasquini, P. Pedroni, S. Sconfietti, J. Phys. G 46, 104001 (2019). https://doi.org/10.1088/1361-6471/ab323a. arXiv:1903.07952 [hep-ph]

    Article  ADS  Google Scholar 

  63. P. P. Martel, E. Downie, M. Ahmed, G. Feldman, Private Communication (2020)

  64. V .O. de León et al., Eur. Phys. J. A 10, 207 (2001). https://doi.org/10.1007/s100500170132

    Article  ADS  Google Scholar 

  65. A. Zieger, R. Van de Vyver, D. Christmann, A. De Graeve, C. Van den Abeele, B. Ziegler, Phys. Lett. B 278, 34 (1992). https://doi.org/10.1016/0370-2693(92)90707-B

    Article  ADS  Google Scholar 

  66. R .P. Hildebrandt, H .W. Grießhammer, T .R. Hemmert, Eur. Phys. J. A 46, 111 (2010). https://doi.org/10.1140/epja/i2010-11024-y. arXiv:nucl-th/0512063 [nucl-th]

    Article  ADS  Google Scholar 

  67. S. Jackson, Design of physical system experiments using bayes linear emulation and history matching methodology with application to arabidopsis Thaliana (Durham University, Doctoral, 2018)

    Google Scholar 

  68. A. Gelman, J. Carlin, H. Stern, D. Dunson, A. Vehtari, D. Rubin, Bayesian data analysis, third edition, chapman & hall/crc texts in statistical science (Taylor & Francis, Milton Park, 2013)

    Google Scholar 

  69. A. O’Hagan, Bayesian Inference, Kendall’s Advanced Theory of Statistics (Edward Arnold, London, 1994)

  70. C. Rasmussen, Bayesian Stat. 7, 651 (2003)

    Google Scholar 

  71. E. Solak, R. Murray-smith, W. E. Leithead, D. J. Leith, C. E. Rasmussen, in Advances in Neural Information Processing Systems 15, edited by S. Becker, S. Thrun, and K. Obermayer (MIT Press, 2003) pp. 1057–1064

  72. D. Eriksson, K. Dong, E. Lee, D. Bindel, A. G. Wilson, in Advances in Neural Information Processing Systems 31, edited by S. Bengio, H. Wallach, H. Larochelle, K. Grauman, N. Cesa-Bianchi, and R. Garnett (Curran Associates, Inc., 2018) pp. 6867–6877, arXiv:1810.12283

  73. M. Chilenski, M. Greenwald, Y. Marzouk, N. Howard, A. White, J. Rice, J. Walk, Nucl. Fus. 55, 023012 (2015). https://doi.org/10.1088/0029-5515/55/2/023012

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We thank Ian Vernon for useful discussions, and M. Ahmed, E. Downie, G. Feldman, P. P. Martel, as well as the MAMI-A2/CB and Compton@HI\(\gamma \)S teams for their patience in discussing experimental constraints. We gratefully acknowledge the stimulating atmosphere created by organizers and participants of the workshops Uncertainty Quantification at the Extremes (ISNET-6) at T.U. Darmstadt (Germany) and Bayesian Inference in Subatomic Physics - A Wallenberg Symposium (ISNET-7) at Chalmers U. (Göteborg, Sweden), which triggered and expanded these investigations. H.W.G. gratefully acknowledges the warm hospitality and financial support of the A2/Crystall Ball Collaboration Meeting 2020 at MAMI (U. Mainz, Germany), of both Ohio University and the Ohio State University, and of the University of Manchester, where part of this work was conducted. The work of R.J.F. and J.A.M. was supported in part by the National Science Foundation under Grant Nos. PHY–1614460 and PHY–1913069 and the NUCLEI SciDAC Collaboration under US Department of Energy MSU subcontract RC107839-OSU. The work of D.R.P. was supported by the US Department of Energy under contract DE-FG02-93ER-40756 and by the ExtreMe Matter Institute EMMI at the GSI Helmholtzzentrum für Schwerionenphysik, Darmstadt, Germany. The work of H.W.G. was supported in part by the US Department of Energy under contract DE-SC0015393, by the High Intensity Gamma-Ray Source HI\(\gamma \)S of the Triangle Universities Nuclear Laboratory TUNL in concert with the Department of Physics of Duke University, and by The George Washington University: by the Dean’s Research Chair programme and an Enhanced Faculty Travel Award of the Columbian College of Arts and Sciences; and by the Office of the Vice President for Research and the Dean of the Columbian College of Arts and Sciences. His work was conducted in part at GW’s Campus in the Closet. The work of J.McG. was supported by the UK Science and Technology Facilities Council Grant ST/P004423/1. The work of M.T.P. was supported in part by the King Abdullah University of Science and Technology (KAUST) Office of Sponsored Research (OSR) under Award No. OSR-2018-CRG7-3800.3.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to H. W. Grießhammer.

Additional information

Communicated by Vittorio Somà

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 35089 KB)

Appendices

Appendix A: Experimental design details

Suppose that our theoretical model \(y_k({x};\vec {a})\) is related to measurements \(y_{exp }({x})\) via additive theoretical and experimental noise, as in Eq. (5). We can linearize \(y_k({x};\vec {a})\) about some point \(\vec {a}_\star \) by keeping only the first order terms in its Taylor expansion, i.e.,

$$\begin{aligned} y_k({x};\vec {a})&\approx y_k({x}; \vec {a}_\star ) + \sum _i b_i({x}) [\vec {a}_i - \vec {a}_\star ] \nonumber \\&= c({x}; \vec {a}_\star ) + \vec {b}({x}) \cdot \vec {a}\,, \end{aligned}$$
(A1)

where \(\vec {b}({x}) \equiv \partial y_k({x};\vec {a})/\partial \vec {a}\) evaluated at \(\vec {a}_\star \) are our basis functions and \(c({x};\vec {a}_\star ) \equiv y_k({x};\vec {a}_\star ) - \vec {b} \cdot \vec {a}_\star \) is constant with respect to the polarizabilities \(\vec {a}\) but depends on the kinematic point \({x}\). Thus, the vector of N measurements \(\mathbf {y}\) is related to the polarizabilities via the likelihood

$$\begin{aligned} \mathbf {y}\,|\,\vec {a}\sim {\mathcal {N}}[B\vec {a}+ {\mathbf {c}}, \Sigma ] \end{aligned}$$
(A2)

where \(B \equiv \vec {b}({\mathbf {x}})\) is an \(N \times 6\) matrix, \({\mathbf {c}} \equiv c({\mathbf {x}}; \vec {a}_\star )\) is a length N vector, and \(\Sigma \) is the \(N \times N\) covariance matrix due to theoretical and experimental error. That is, given some experimental covariance \(\Sigma _{\mathrm{exp}}\) and a theoretical covariance \({\bar{c}}^2 R_{\delta k}\) from Eqs. (13) and (14), then

$$\begin{aligned} \Sigma = {\bar{c}}^2 R_{\delta k} + \Sigma _{\mathrm{exp}} \, . \end{aligned}$$
(A3)

Note that \(R_{\delta k}\) depends on the values of the tuned \(\ell _\omega \) and \(\ell _\theta \), whose estimates from the order-by-order convergence pattern are given in Table 1.

The linear model of Eq. (A2) is well known in the statistics literature [68, 69], so here we will simply state the relevant results. If a Gaußian prior with mean \(\vec {\mu }_0\) and covariance \(V_0\) is placed on \(\vec {a}\) as in Eq. (20), then the resulting posterior is also Gaußian, with mean and covariance given by

$$\begin{aligned} \vec {\mu }&= V\left[ V_0^{-1}\vec {\mu }_0 + B^\intercal \Sigma ^{-1} (\mathbf {y}- {\mathbf {c}})\right] \, , \end{aligned}$$
(A4)
$$\begin{aligned} V&= (V_0^{-1} + B^\intercal \Sigma ^{-1} B)^{-1} \, . \end{aligned}$$
(A5)

Importantly to our study of experimental design, the posterior covariance V depends on the kinematic points \({\mathbf {x}}\) where the experiment is performed, and on the specifics of the observable through \(\Sigma \), but not on the exact results of the experiment \(\mathbf {y}\).

Given that we choose to maximize the expected information gain in the polarizabilities, then the integrals of Eq. (22) must still be performed. The integral over \(\vec {a}\) splits into the difference of two terms: the differential entropy of the prior \({{\,\mathrm{pr}\,}}(\vec {a})\) and of the posterior \({{\,\mathrm{pr}\,}}(\vec {a}\,|\,\mathbf {y}, {\mathbf {d}})\). The differential entropy of a Gaußian \({\mathcal {N}}(\mu , \Sigma )\) is well known to be \(\frac{1}{2}\ln {|2\pi e\Sigma |}\). Therefore

$$\begin{aligned} U_{\text {KL}}({\mathbf {d}})&= - \int \ln [{{\,\mathrm{pr}\,}}(\vec {a})] {{\,\mathrm{pr}\,}}(\vec {a}) \,{\mathrm{d}}{\vec {a}} \nonumber \\&\quad + \int \ln [{{\,\mathrm{pr}\,}}(\vec {a}\,|\,\mathbf {y}, {\mathbf {d}})]{{\,\mathrm{pr}\,}}(\vec {a}\,|\,\mathbf {y}, {\mathbf {d}}) \,{\mathrm{d}}{\vec {a}} {{\,\mathrm{pr}\,}}(\mathbf {y}\,|\,{\mathbf {d}})\,{\mathrm{d}}{\mathbf {y}} \nonumber \\&= \frac{1}{2}\ln {|2\pi e V_0|} - \frac{1}{2}\ln {|2\pi e V|} \int {{\,\mathrm{pr}\,}}(\mathbf {y}\,|\,{\mathbf {d}})\,{\mathrm{d}}{\mathbf {y}} \nonumber \\&= \frac{1}{2} \ln \frac{|V_0|}{|V|} \, , \end{aligned}$$
(A6)

where we used the fact that V does not depend on \(\mathbf {y}\) and then performed the trivial integration over all possible measurements \(\mathbf {y}\).

Appendix B: Observable constraints and EFT truncation model details

Constraints on Compton observables are discussed in detail in Ref. [6]. Some of this is reproduced here, with particular attention paid to nth-order chiral corrections to observables \(\Delta y_n\) rather than the value of the observable y itself. The \(\Delta y_n\) impact the distribution for the \(\chi \)EFT uncertainty \(\delta y_k\), but because we restrict the “experimentally accessible regime” in this study from small \(\omega \), and forwards/backwards angles, these constraints are not as important as they otherwise would be. These constraints are summarized for particular \(\omega _{\text {lab}}\) and \(\theta _{\text {lab}}\) values in Table 5.

Table 5 The constraints on corrections to observables \(\Delta y\) and their derivatives \(\Delta y'\) at particular \(\theta \) and \(\omega \). The LO amplitude as well as all calculated higher orders fulfill them automatically, so these must only be enforced in the GP. The observables marked by a dagger \(\dagger \) are zero below the pion-production threshold, but we impose no constraint on them at \(\omega =\omega _\pi \), as discussed in the text

All observables that are nonzero below \(\omega _\pi \) approach the Thomson limit as \(\omega \rightarrow 0\) [1]. Thus, higher-order corrections must vanish there , and approach \(\omega = 0\) as at least \(\omega ^2\). Therefore, at least the first derivative of all corrections must vanish there as well.

The remaining observables must vanish for \(\omega \le \omega _\pi \), but there is no constraint on the the derivative of corrections at \(\omega = \omega _\pi \). We have found that the corrections approach 0 very quickly, so that imposing the constraint \(\Delta y_n(\omega _\pi ,\theta ) = 0\) for all higher order terms is actually a worse approximation than not imposing the constraint at all; see, e.g., Fig. S.4 in the Supplemental Material. This comes back to the large cusps in the spin-observable \(c_n\) found near \(\omega _\pi \), discussed in Sect. 3.2.1, which remain an unresolved aspect of this model.

Due to the coordinate singularity at \(\theta =0^\circ \) and \(180^\circ \), observables or their derivative with respect to \(\theta \) must vanish there [6]. But this does not preclude both the value and their derivatives from vanishing there. These constraints can be deduced by symmetry arguments, and are summarized in Table 5.

The hyperparameters \({\bar{c}}^2\) and \(\ell _i\), shown in Table 1, are tuned to coefficients \(c_n\) at the best known \(\vec {a}\) (see Table 3) for \(\Lambda _b = 650\, \text {MeV}\). The training data are on a grid with \(\theta _{\text {lab}}= \{30^\circ , 50^\circ , 70^\circ , 90^\circ , 110^\circ , 130^\circ \}\) and \(\omega _{\text {lab}}= \{200, 225, 250\}\, \text {MeV}\) for observables which are zero below \(\omega _\pi \). For observables that are non-zero below \(\omega _\pi \), the additional training points \(\omega _{\text {lab}}= \{50, 75, 100, 125\}\, \text {MeV}\) are included, and common \(\ell _\omega \) and \(\ell _\theta \) are used between the two regions. The training region is well outside the kinematic endpoints where additional constraints arise on observables or their derivatives, and excludes the pion-production threshold region.

Because the first nonzero order often behaves differently than the corrections, we do not use it for induction on the \(c_n\); that is we only train the hyperparameters on corrections. Hence, we train on \(c_2\)\(c_4\) for \(\,{\mathrm{d}}{\sigma }\) and \(\Sigma _3\), but otherwise we train on \(c_3\) and \(c_4\).

The coefficients for various observable slices are shown in Figs. 12 for the cross section, and Figs. S.1–S.6 of the Supplemental Material for the spin observables. These plots also include uncertainty bands for higher order coefficients, with the symmetry constraints given in Table 5 included. These constraints on both the coefficient functions and their derivatives propagate directly to the truncation error \(\delta y_k\) by replacing \(r(x,x';\ell _\omega , \ell _\theta )\) in Eq. (14) by its conditional form \({\tilde{r}}(x,x';\ell _\omega , \ell _\theta )\), see Refs. [27, 45]. For example, if the value of \(c_n\) is known at the set of points \({\mathbf {x}}\), then one can compute its conditional GP, with covariance kernel given by

$$\begin{aligned} {\tilde{r}}(x, x') = r(x, x') - r(x, {\mathbf {x}}) r({\mathbf {x}}, {\mathbf {x}})^{-1} r({\mathbf {x}}, x') \, . \end{aligned}$$
(B1)

See Refs. [70,71,72] for details about adding derivative observations to GPs. Because the RBF kernel [Eq. (11)] is separable in \(\omega \) and \(\theta \), these constraints can simply be applied to each one-dimensional kernel separately, and multiplied to yield the total constrained kernel. We employ the gptools python package for easily implementing derivative constraints [73].

Fig. 11
figure 11

A component of the standard deviation due to \(\chi \)EFT uncertainty at \(\mathrm {N}^{4}\mathrm {LO} ^+\), see Eq. (14). The factor of \({\bar{c}}\) is unique to each observable, and is not included. See Table 1

Fig. 12
figure 12

Coefficients for the differential cross section \(\,{\mathrm{d}}{\sigma }\)

For completeness, we also provide the profile for the truncation error standard deviation (up to factors of \({\bar{c}}\), which vary by observable); see Fig. 11. It assumes the form of Q provided in Eq. (8) along with the first omitted \(\chi \)EFT order given in Eq. (18).

This allows us to return to the discussion of the omitted constraints \(\Sigma _i \in [-1, 1]\) on the spin observables in Sect. 4.1. Over the physically interesting kinematic range, the actual value of most spin observables lies in the much more narrow interval \([-0.7,0.7]\); see Fig. 5 in Ref. [6]. So, then the question becomes: are the mean prediction and its theory uncertainty contained in \([-1,1]\) with a high degree of probability? From Eq. (12), one can see that the \(1\sigma \) interval for the truncation error \(\delta y_k\) is \(y_{\mathrm {ref}}{\bar{c}}\) times another factor \(Q^{\nu _{\delta k}(\omega )}/\sqrt{1-Q^2(\omega )}\). Here \(y_{\mathrm {ref}}=1\) and \({\bar{c}}\lesssim 0.7\) for most spin observables (see Table 1). The third factor is plotted in Fig. 11 and does not exceed \(\approx 0.3\) at \(\omega \lesssim 230\, \text {MeV}\), where our analysis shows the biggest sensitivities. Therefore, even for the spin observables with large magnitudes, the \(1\sigma \) upper range of a GP will only give values about 0.2 larger than the established maximum of 0.7, namely about 0.9 in total. This is close but still below \(|\Sigma _i|=1\). Therefore, a majority of our test functions in the GP will not probe, let alone exceed, the strict bounds on those spin observables. Furthermore, if observables and their truncation errors vanish at \(\theta = 0^\circ \) or \(180^\circ \), this will make the constraint even more trivially satisfied near these regions. We are therefore confident that implementing the constraint \(\Sigma _i\in [-1, 1]\) would not impact our results for \(\omega \lesssim 220\, \text {MeV}\), and cautiously optimistic that the impact would be small even at higher energies.

Though we likewise do not constrain the cross section to be non-negative, we are confident that within our constrained angle range, corrections are highly unlikely to be large enough for this to be a worry. According to Fig. 4 in Ref. [6], the \(\mathrm {N}^{4}\mathrm {LO}\) cross section is small (\(<10\,\mathrm {nb/sr}\)) in a narrow region at forward angles around \(\omega _\pi \). Figure 11 shows that the expansion parameter is small, and Fig. 12 shows that the coefficients \(c_i\) are natural-sized. Therefore, the GP corrections are highly unlikely to exceed the size of the predicted cross section and create negative (unphysical) values.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Melendez, J.A., Furnstahl, R.J., Grießhammer, H.W. et al. Designing optimal experiments: an application to proton Compton scattering. Eur. Phys. J. A 57, 81 (2021). https://doi.org/10.1140/epja/s10050-021-00382-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epja/s10050-021-00382-2

Navigation