Skip to main content
Log in

Should we still believe in constrained supersymmetry?

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

Abstract

We calculate partial Bayes factors to quantify how the feasibility of the constrained minimal supersymmetric standard model (CMSSM) has changed in the light of a series of observations. This is done in the Bayesian spirit where probability reflects a degree of belief in a proposition and Bayes’ theorem tells us how to update it after acquiring new information. Our experimental baseline is the approximate knowledge that was available before LEP, and our comparison model is the Standard Model with a simple dark matter candidate. To quantify the amount by which experiments have altered our relative belief in the CMSSM since the baseline data we compute the partial Bayes factors that arise from learning in sequence the LEP Higgs constraints, the XENON100 dark matter constraints, the 2011 LHC supersymmetry search results, and the early 2012 LHC Higgs search results. We find that LEP and the LHC strongly shatter our trust in the CMSSM (with M 0 and M 1/2 below 2 TeV), reducing its posterior odds by approximately two orders of magnitude. This reduction is largely due to substantial Occam factors induced by the LEP and LHC Higgs searches.

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

Similar content being viewed by others

Notes

  1. Here ‘more extreme’ can be defined in numerous ways.

  2. Strictly, some prior dependence remains due to the choice of parameter values considered possible by the prior, most often arising from the choice of scan range; however, this is the same kind of dependence that exists in a frequentist analysis. As well as this there exists the possibility that d 1 strictly forbids certain values of θ, and these too should be excluded from the computation of \(P(d_{2}|H_{i},\hat{\theta})\).

  3. The full volume of parameter space viable at this inference step, V total, is defined by the informative prior. If the likelihood function for the new data was constant in a region V and zero outside of it, then the fraction f=V/V total would be the Occam factor.

  4. I.e. in a generic event counting experiment we assume the expected number of signal events at the best-fit point to be close to zero.

  5. Note that a very small value for the evidence from learning some data implies a very large amount of information was gained about the model. This may sound like a good thing; however, it means that little was known about the model before this data arrived and so the model was not very useful for predicting what that data would be. PBFs penalise this failure; however, if the information gain was sufficiently large then the model may in fact become highly predictive about future data, and may thus fare much better in future PBF tests.

  6. The reader may protest that the SM+DM is not just a fine-tuning “puzzle”, it is a very extreme example of fine-tuning! However, this is only true if one considers it from a pre-‘electroweak data’ perspective. The SM+DM presumably suffers a very large PBF penalty for failing to predict the electroweak scale (and for this scale being observed very far from, say, the Planck scale, where a priori arguments based on the hierarchy problem may place it); however, these considerations enter before the ‘baseline’ data we choose for our inference sequence and so do not directly enter our PBFs. The complete assessment of which model best reflects reality should of course take these matters into account.

  7. Except for an early LHCb lower bound on BR(B s μμ).

  8. This is a small lie; we do compute marginalised posteriors for each update, which indeed correspond to the “informative” priors for the subsequent update. Nevertheless we do not explicitly use them in this fashion.

  9. The rest of the Standard Model parameters of course also enter explicitly, but we may reasonably consider priors over those to be statistically independent of the CMSSM parameters, such that measuring the values of these parameters results in PBFs of 1.

  10. Within the range of m h values compatible with d new, i.e. the dark sector theory is permitted to exclude values of m h which are also well excluded by d new.

  11. For a scale parameter this is the Jeffreys prior.

  12. These Δχ 2 curves are almost quadratic in logm h , implying a close to Gaussian likelihood function; however, we have digitised the most loose boundaries of the displayed curves to be conservative. As a result the likelihood function we reconstruct has a flat maximum from ∼80 GeV to ∼100 GeV.

  13. If, due to tuning arguments, we except m h to adopt a value on the largest allowed scale, rather than all scales being equally likely, then a flat prior cut off at this scale may indeed better represent this belief. The lack of sensitivity of the informative “pre-LEP” prior to this choice reflects the fact that before the “pre-LEP” update the Standard Model prediction for the Higgs mass is already quite well constrained.

  14. Aside from nested sampling and several variants of Markov Chain Monte Carlo methods, the list of techniques used to scan the CMSSM has expanded in recent years to also include genetic algorithms and neural networks [76, 77].

  15. The authors estimate these Bayes factors using both flat and log priors; here we refer to the log prior results only since we do not use flat priors. In addition, the δa μ constraint is shown to strongly drive the preference of μ>0 so if the validity of this constraint is questioned (we consider the effects of removing it in Sect. 7) then the impact of ignoring the μ<0 branch may also merit revisitation.

  16. We do this because P eff requires renormalisation group running to be evaluated, i.e. our spectrum generator needs to be run before we can evaluate P eff.

  17. In practice a larger scan volume will decrease the scan resolution and reduce the accuracy of results, so scan prior volume dependence would still exist in this indirect form.

  18. We ignore the variation of a with the predicted signal rate as it is small for small signal.

  19. Actually the CL s method is used so the limit is drawn where p=p s /(1−p b )=0.1 [106], but this correction weakens the limit so it is conservative to ignore it and in this case makes little difference anyway, given our other approximations.

  20. In Sect. 6.3 we construct the ATLAS Higgs search likelihood function using almost identical techniques, but argue that each fitted slice needs to be normalised relative to the others using the likelihood of the best-fit point on each slice. This occurs because the best-fit point of each slice lies a varying number of standard deviations from the zero signal point (zero cross section), which we know to have the same likelihood for every slice. A similar normalisation is in principle required to recover the true likelihood function computed by Xenon; however, the variance of the limit appears to be approximately Gaussian in the logarithm of the cross section, making extrapolation of the likelihood to the zero cross section point extremely unreliable. In addition, the reconstruction method we use for the ATLAS Higgs search likelihood relies on plots of the signal best fit against m X , whereas here we use a plot of the 90 % confidence limit. Performing the extraction using the limit curve requires more assumptions than a best-fit curve, so combined with the logarithmic difficulty we judge that this technique would produce poor results, and so we prefer to stick with the simpler technique described. The Xenon limit turns out to be of very minor importance to our final inferences anyway so we are not concerned with small errors in our reconstructed likelihood. In hindsight we expect that even simply applying a hard cut at the observed Xenon limit would negligibly affect our inferences.

  21. From Eq. (5), using the suggested σ s =50±8 MeV and σ l =47±9 MeV [112], with m s /m l =m s /(2(m u +m d ))=26±4 [86].

  22. We demonstrate this in Appendix A.

  23. The point was found during a wide ranging scan of the CMSSM parameter space, hence the esoteric choice of parameters.

  24. All other points in the test sample are “not excluded” and labelled as such, or “excluded” and labelled as such.

  25. The apparent ‘thinning’ of the likelihood toward higher M 0 and M 1/2 is a mere sampling artefact.

  26. We note that the new limits cut off much less than half of the posterior remaining in the “ATLAS-Higgs” dataset (shown in the last frame of Fig. 14), so the corresponding “additional” PBF is likewise much less than two.

References

  1. S. Weinberg, The Quantum Theory of Fields. Vol. III: Supersymmetry (Cambridge University Press, Cambridge, 2000)

    Book  Google Scholar 

  2. G.L. Kane, Supersymmetry: Squarks, Photinos, and the Unveiling of the Ultimate Laws of Nature (Perseus Publishing, Cambridge, 2001)

    Google Scholar 

  3. M. Drees, R. Godbole, P. Roy, Theory and Phenomenology of Sparticles: An Account of Four-Dimensional N=1 Supersymmetry in High Energy Physics (World Scientific, Hackensack, 2004)

    Google Scholar 

  4. H. Baer, X. Tata, Weak scale supersymmetry: From superfields to scattering events (Cambridge University Press, Cambridge, 2006)

    Book  Google Scholar 

  5. P. Binetruy, Supersymmetry: Theory, Experiment and Cosmology (Oxford University Press, Oxford, 2006)

    Google Scholar 

  6. J. Terning, Modern Supersymmetry: Dynamics and Duality (Oxford University Press, Oxford, 2006)

    Google Scholar 

  7. N. Polonsky, Supersymmetry: structure and phenomena. Extensions of the standard model. Lect. Notes Phys. M 68, 1–169 (2001). arXiv:hep-ph/0108236 [hep-ph]

    MathSciNet  Google Scholar 

  8. H. Pagels, J.R. Primack, Supersymmetry, cosmology and new TeV physics. Phys. Rev. Lett. 48, 223 (1982)

    Article  ADS  Google Scholar 

  9. H. Goldberg, Constraint on the photino mass from cosmology. Phys. Rev. Lett. 50, 1419 (1983)

    Article  ADS  Google Scholar 

  10. P. Ramond, Journeys Beyond the Standard Model (Perseus Books, Cambridge, 1999)

    Google Scholar 

  11. H. Baer, C. Balazs, M. Brhlik, P. Mercadante, X. Tata et al., Aspects of supersymmetric models with a radiatively driven inverted mass hierarchy. Phys. Rev. D 64, 015002 (2001). arXiv:hep-ph/0102156 [hep-ph]

    ADS  Google Scholar 

  12. C. Balazs, M.S. Carena, A. Menon, D. Morrissey, C. Wagner, The supersymmetric origin of matter. Phys. Rev. D 71, 075002 (2005). arXiv:hep-ph/0412264 [hep-ph]

    ADS  Google Scholar 

  13. D.V. Nanopoulos, K.A. Olive, M. Srednicki, K. Tamvakis, Primordial inflation in simple supergravity. Phys. Lett. B 123, 41 (1983)

    Article  MathSciNet  ADS  Google Scholar 

  14. R. Holman, P. Ramond, G.G. Ross, Supersymmetric inflationary cosmology. Phys. Lett. B 137, 343–347 (1984)

    Article  ADS  Google Scholar 

  15. S. Dimopoulos, H. Georgi, Softly broken supersymmetry and SU(5). Nucl. Phys. B 193, 150 (1981)

    Article  ADS  Google Scholar 

  16. A.H. Chamseddine, R.L. Arnowitt, P. Nath, Locally supersymmetric grand unification. Phys. Rev. Lett. 49, 970 (1982)

    Article  ADS  Google Scholar 

  17. H. Baer, C. Balazs, Chi**2 analysis of the minimal supergravity model including WMAP, g(mu)-2 and b → s gamma constraints. J. Cosmol. Astropart. Phys. 0305, 006 (2003). arXiv:hep-ph/0303114 [hep-ph]

    Article  ADS  Google Scholar 

  18. J.R. Ellis, K.A. Olive, Y. Santoso, V.C. Spanos, Likelihood analysis of the CMSSM parameter space. Phys. Rev. D 69, 095004 (2004). arXiv:hep-ph/0310356 [hep-ph]

    ADS  Google Scholar 

  19. P. Bechtle, K. Desch, M. Uhlenbrock, P. Wienemann, Constraining SUSY models with Fittino using measurements before, with and beyond the LHC. Eur. Phys. J. C 66, 215–259 (2010). arXiv:0907.2589 [hep-ph]

    Article  ADS  Google Scholar 

  20. P. Bechtle, K. Desch, H. Dreiner, M. Kramer, B. O’Leary et al., Present and possible future implications for mSUGRA of the non-discovery of SUSY at the LHC. arXiv:1105.5398 [hep-ph]

  21. S. Heinemeyer, G. Weiglein, Predicting supersymmetry. Nucl. Phys. Proc. Suppl. 205–206, 283–288 (2010). arXiv:1007.0206 [hep-ph]

    Article  Google Scholar 

  22. O. Buchmueller, R. Cavanaugh, D. Colling, A. De Roeck, M. Dolan et al., Frequentist analysis of the parameter space of minimal supergravity. Eur. Phys. J. C 71, 1583 (2011). arXiv:1011.6118 [hep-ph]

    Article  ADS  Google Scholar 

  23. D.E. Lopez-Fogliani, L. Roszkowski, R. Ruiz de Austri, T.A. Varley, A Bayesian analysis of the constrained NMSSM. Phys. Rev. D 80, 095013 (2009). arXiv:0906.4911 [hep-ph]

    ADS  Google Scholar 

  24. M.E. Cabrera, J.A. Casas, R. Ruiz de Austri, MSSM forecast for the LHC. J. High Energy Phys. 1005, 043 (2010). arXiv:0911.4686 [hep-ph]

    Article  ADS  Google Scholar 

  25. O. Buchmueller, R. Cavanaugh, D. Colling, A. De Roeck, M. Dolan et al., Supersymmetry and dark matter in light of LHC 2010 and Xenon100 data. Eur. Phys. J. C 71, 1722 (2011). arXiv:1106.2529 [hep-ph]

    Article  ADS  Google Scholar 

  26. J. Ellis, K.A. Olive, Revisiting the Higgs mass and dark matter in the CMSSM. arXiv:1202.3262 [hep-ph]

  27. P. Bechtle, T. Bringmann, K. Desch, H. Dreiner, M. Hamer et al., Constrained supersymmetry after two years of LHC data: a global view with Fittino. arXiv:1204.4199 [hep-ph]

  28. B. Allanach, Impact of CMS multi-jets and missing energy search on CMSSM fits. Phys. Rev. D 83, 095019 (2011). arXiv:1102.3149 [hep-ph]

    ADS  Google Scholar 

  29. B. Allanach, T. Khoo, C. Lester, S. Williams, The impact of the ATLAS zero-lepton, jets and missing momentum search on a CMSSM fit. J. High Energy Phys. 1106, 035 (2011). arXiv:1103.0969 [hep-ph]

    Article  ADS  Google Scholar 

  30. G. Bertone, D.G. Cerdeno, M. Fornasa, R. Ruiz de Austri, C. Strege et al., Global fits of the cMSSM including the first LHC and XENON100 data. J. Cosmol. Astropart. Phys. 1201, 015 (2012). arXiv:1107.1715 [hep-ph]

    Article  ADS  Google Scholar 

  31. A. Fowlie, A. Kalinowski, M. Kazana, L. Roszkowski, Y.S. Tsai, Bayesian implications of current LHC and XENON100 search limits for the constrained MSSM. arXiv:1111.6098 [hep-ph]

  32. O. Buchmueller, R. Cavanaugh, A. De Roeck, M. Dolan, J. Ellis et al., Supersymmetry in light of 1/fb of LHC data. arXiv:1110.3568 [hep-ph]

  33. O. Buchmueller, R. Cavanaugh, A. De Roeck, M. Dolan, J. Ellis et al., Higgs and supersymmetry. arXiv:1112.3564 [hep-ph]

  34. G.D. Starkman, R. Trotta, P.M. Vaudrevange, Introducing doubt in Bayesian model comparison. arXiv:0811.2415 [physics.data-an]

  35. M.E. Cabrera, J. Casas, V.A. Mitsou, R. Ruiz de Austri, J. Terron, Histogram comparison as a powerful tool for the search of new physics at LHC. Application to CMSSM. arXiv:1109.3759 [hep-ph]

  36. S. AbdusSalam, B. Allanach, H. Dreiner, J. Ellis, U. Ellwanger et al., Benchmark models, planes, lines and points for future SUSY searches at the LHC. Eur. Phys. J. C 71, 1835 (2011). arXiv:1109.3859 [hep-ph]

    Article  ADS  Google Scholar 

  37. S. Sekmen, S. Kraml, J. Lykken, F. Moortgat, S. Padhi et al., Interpreting LHC SUSY searches in the phenomenological MSSM. arXiv:1109.5119 [hep-ph]

  38. C. Strege, G. Bertone, D. Cerdeno, M. Fornasa, R. Ruiz de Austri et al., Updated global fits of the cMSSM including the latest LHC SUSY and Higgs searches and XENON100 data. arXiv:1112.4192 [hep-ph]

  39. L. Roszkowski, E.M. Sessolo, Y.-L.S. Tsai, Bayesian implications of current LHC supersymmetry and dark matter detection searches for the constrained MSSM. arXiv:1202.1503 [hep-ph]

  40. A. O’Hagan, Fractional bayes factors for model comparison. J. Royal Stat. Soc. Ser. B, Methodol. 57(1), 99–138 (1995)

    MathSciNet  MATH  Google Scholar 

  41. J. Berger, L. Pericchi, The intrinsic bayes factor for model selection and prediction. J. Am. Stat. Assoc. 91(433), 109–122 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  42. J. Berger, J. Mortera, Default bayes factors for nonnested hypothesis testing. J. Am. Stat. Assoc. 94(446), 542–554 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  43. B. Allanach, Naturalness priors and fits to the constrained minimal supersymmetric standard model. Phys. Lett. B 635, 123–130 (2006). arXiv:hep-ph/0601089 [hep-ph]

    Article  ADS  Google Scholar 

  44. B.C. Allanach, K. Cranmer, C.G. Lester, A.M. Weber, Natural priors, CMSSM fits and LHC weather forecasts. J. High Energy Phys. 0708, 023 (2007). arXiv:0705.0487 [hep-ph]

    Article  ADS  Google Scholar 

  45. M.E. Cabrera, J.A. Casas, R. Ruiz de Austri, Bayesian approach and naturalness in MSSM analyses for the LHC. J. High Energy Phys. 03, 075 (2009). arXiv:0812.0536 [hep-ph]

    Article  ADS  Google Scholar 

  46. M.E. Cabrera, Bayesian study and naturalness in MSSM forecast for the LHC. arXiv:1005.2525 [hep-ph]

  47. L.J. Hall, D. Pinner, J.T. Ruderman, A natural SUSY Higgs near 126 GeV. arXiv:1112.2703 [hep-ph]

  48. P. Athron, D.J. Miller, A new measure of fine tuning. Phys. Rev. D 76, 075010 (2007). arXiv:0705.2241 [hep-ph]

    ADS  Google Scholar 

  49. S. Cassel, D. Ghilencea, G. Ross, Testing SUSY. Phys. Lett. B 687, 214–218 (2010). arXiv:0911.1134 [hep-ph]

    Article  ADS  Google Scholar 

  50. D. Horton, G. Ross, Naturalness and focus points with non-universal Gaugino masses. Nucl. Phys. B 830, 221–247 (2010). arXiv:0908.0857 [hep-ph]

    Article  ADS  Google Scholar 

  51. S. Cassel, D. Ghilencea, G. Ross, Testing SUSY at the LHC: electroweak and dark matter fine tuning at two-loop order. Nucl. Phys. B 835, 110–134 (2010). arXiv:1001.3884 [hep-ph]

    Article  ADS  MATH  Google Scholar 

  52. S. Akula, M. Liu, P. Nath, G. Peim, Naturalness, supersymmetry and implications for LHC and dark matter. arXiv:1111.4589 [hep-ph]

  53. A. Arbey, M. Battaglia, F. Mahmoudi, Implications of LHC searches on SUSY particle spectra: the pMSSM parameter space with neutralino dark matter. Eur. Phys. J. C 72, 1847 (2012). arXiv:1110.3726 [hep-ph]

    Article  ADS  Google Scholar 

  54. S. Cassel, D. Ghilencea, S. Kraml, A. Lessa, G. Ross, Fine-tuning implications for complementary dark matter and LHC SUSY searches. J. High Energy Phys. 1105, 120 (2011). arXiv:1101.4664 [hep-ph]

    Article  ADS  Google Scholar 

  55. M. Papucci, J.T. Ruderman, A. Weiler, Natural SUSY endures. arXiv:1110.6926 [hep-ph]

  56. T. Li, J.A. Maxin, D.V. Nanopoulos, J.W. Walker, Natural predictions for the Higgs boson mass and supersymmetric contributions to rare processes. Phys. Lett. B 708, 93–99 (2012). arXiv:1109.2110 [hep-ph]

    Article  ADS  Google Scholar 

  57. Z. Kang, J. Li, T. Li, On the naturalness of the (N)MSSM. arXiv:1201.5305 [hep-ph]

  58. E. Jaynes, G. Bretthorst, Probability Theory: The Logic of Science (Cambridge University Press, Cambridge, 2003)

    Book  Google Scholar 

  59. F. Feroz, B.C. Allanach, M. Hobson, S.S. AbdusSalam, R. Trotta et al., Bayesian selection of sign(mu) within mSUGRA in global fits including WMAP5 results. J. High Energy Phys. 0810, 064 (2008). arXiv:0807.4512 [hep-ph]

    Article  ADS  Google Scholar 

  60. S.S. AbdusSalam, B.C. Allanach, M.J. Dolan, F. Feroz, M.P. Hobson, Selecting a model of supersymmetry breaking mediation. Phys. Rev. D 80, 035017 (2009). arXiv:0906.0957 [hep-ph]

    ADS  Google Scholar 

  61. F. Feroz, M.P. Hobson, L. Roszkowski, R. Ruiz de Austri, R. Trotta, Are \(BR(\bar{B} \to X_{s} \gamma)\) and (g−2) μ consistent within the constrained MSSM? arXiv:0903.2487 [hep-ph]

  62. M.E. Cabrera, J. Casas, R. Ruiz de Austri, R. Trotta, Quantifying the tension between the Higgs mass and (g−2) μ in the CMSSM. Phys. Rev. D 84, 015006 (2011). arXiv:1011.5935 [hep-ph]

    ADS  Google Scholar 

  63. M. Pierini, H. Prosper, S. Sekmen, M. Spiropulu, Model inference with reference priors. arXiv:1107.2877 [hep-ph]

  64. D. MacKay, Information Theory, Inference, and Learning Algorithms (Cambridge University Press, Cambridge, 2003)

    MATH  Google Scholar 

  65. R. Solomonoff, A formal theory of inductive inference. Part I. Inf. Control 7(1), 1–22 (1964). http://www.sciencedirect.com/science/article/pii/S0019995864902232

    Article  MathSciNet  MATH  Google Scholar 

  66. S. Fichet, Quantified naturalness from Bayesian statistics. arXiv:1204.4940 [hep-ph]

  67. P. Bock, J. Carr, S. De Jong, F. Di Lodovico, E. Gross, P. Igo-Kemenes, P. Janot, W. Murray, M. Pieri, A.L. Read, V. Ruhlmann-Kleider, A. Sopczak (ALEPH, DELPHI, L3, OPAL, LEP Electroweak Working Group Collaboration), Lower bound for the standard model Higgs boson mass from combining the results of the four lep experiments. Tech. rep., CERN, Geneva (1998). http://cdsweb.cern.ch/record/353201

  68. ALEPH, CDF, D0, DELPHI, L3, OPAL, SLD, LEP Electroweak Working Group, Tevatron Electroweak Working Group, SLD Electroweak and Heavy Flavour Groups Collaboration, Precision Electroweak Measurements and Constraints on the Standard Model. arXiv:1012.2367 [hep-ex]

  69. G. Aad et al. (ATLAS Collaboration), Search for the standard model Higgs boson in the diphoton decay channel with 4.9 fb−1 of pp collisions at \(\sqrt{s}=7\) TeV with ATLAS. arXiv:1202.1414 [hep-ex]

  70. G. Aad et al. (ATLAS Collaboration), Search for the Higgs boson in the HWW lνlν decay channel in pp collisions at \(\sqrt{s} = 7\) TeV with the ATLAS detector. arXiv:1112.2577 [hep-ex]

  71. G. Aad et al. (ATLAS Collaboration), Search for the standard model Higgs boson in the decay channel HZZ →4l with 4.8 fb−1 of pp collisions at \(\sqrt {s}=7\) TeV with ATLAS. arXiv:1202.1415 [hep-ex]

  72. G. Aad et al. (ATLAS Collaboration), Combined search for the standard model Higgs boson using up to 4.9 fb−1 of pp collision data at \(\sqrt{s} = 7\) TeV with the ATLAS detector at the LHC. Phys. Lett. B 710, 49–66 (2012). arXiv:1202.1408 [hep-ex]

    Article  ADS  Google Scholar 

  73. F. Feroz, M.P. Hobson, M. Bridges, MultiNest: an efficient and robust Bayesian inference tool for cosmology and particle physics. Mon. Not. R. Astron. Soc. 398, 1601–1614 (2009). arXiv:0809.3437 [astro-ph]

    Article  ADS  Google Scholar 

  74. F. Feroz, M.P. Hobson, Multimodal nested sampling: an efficient and robust alternative to MCMC methods for astronomical data analysis. arXiv:0704.3704 [astro-ph]

  75. J. Skilling, Nested sampling. AIP Conf. Proc. 735(1), 395–405 (2004). http://link.aip.org/link/?APC/735/395/1

    Article  MathSciNet  ADS  Google Scholar 

  76. Y. Akrami, P. Scott, J. Edsjo, J. Conrad, L. Bergstrom, A profile likelihood analysis of the constrained MSSM with genetic algorithms. J. High Energy Phys. 1004, 057 (2010). arXiv:0910.3950 [hep-ph]

    Article  ADS  Google Scholar 

  77. M. Bridges, K. Cranmer, F. Feroz, M. Hobson, R. Ruiz de Austri et al., A coverage study of the CMSSM based on ATLAS sensitivity using fast neural networks techniques. J. High Energy Phys. 1103, 012 (2011). arXiv:1011.4306 [hep-ph]

    Article  ADS  Google Scholar 

  78. F.E. Paige, S.D. Protopopescu, H. Baer, X. Tata, ISAJET 7.69: A Monte Carlo event generator for p p, anti-p p, and e+ e- reactions. arXiv:hep-ph/0312045

  79. G. Belanger, F. Boudjema, A. Pukhov, A. Semenov, micrOMEGAs: a tool for dark matter studies. arXiv:1005.4133 [hep-ph]

  80. G. Belanger, F. Boudjema, A. Pukhov, A. Semenov, Dark matter direct detection rate in a generic model with micrOMEGAs2.1. Comput. Phys. Commun. 180, 747–767 (2009). arXiv:0803.2360 [hep-ph]

    Article  ADS  Google Scholar 

  81. G. Belanger, F. Boudjema, A. Pukhov, A. Semenov, MicrOMEGAs2.0: a program to calculate the relic density of dark matter in a generic model. Comput. Phys. Commun. 176, 367–382 (2007). arXiv:hep-ph/0607059

    Article  ADS  MATH  Google Scholar 

  82. F. Mahmoudi, SuperIso: a program for calculating the isospin asymmetry of B -> K* gamma in the MSSM. Comput. Phys. Commun. 178, 745–754 (2008). arXiv:0710.2067 [hep-ph]

    Article  ADS  MATH  Google Scholar 

  83. F. Mahmoudi, SuperIso v2.3: a program for calculating flavor physics observables in supersymmetry. Comput. Phys. Commun. 180, 1579–1613 (2009). arXiv:0808.3144 [hep-ph]

    Article  ADS  Google Scholar 

  84. A. Djouadi, J. Kalinowski, M. Spira, HDECAY: a program for Higgs boson decays in the standard model and its supersymmetric extension. Comput. Phys. Commun. 108, 56–74 (1998). arXiv:hep-ph/9704448 [hep-ph]

    Article  ADS  MATH  Google Scholar 

  85. F. Feroz, K. Cranmer, M. Hobson, R. Ruiz de Austri, R. Trotta, Challenges of profile likelihood evaluation in multi-dimensional SUSY scans. J. High Energy Phys. 06, 042 (2011). arXiv:1101.3296 [hep-ph]

    Article  ADS  Google Scholar 

  86. K. Nakamura, et al. (Particle Data Group), Review of particle physics. J. Phys. G, Nucl. Part. Phys. 37(7A), 075021 (2010), and 2011 partial update for the 2012 edition. http://stacks.iop.org/0954-3899/37/i=7A/a=075021

    Article  ADS  Google Scholar 

  87. H. Jeffreys, Theory of Probability (1961)

    MATH  Google Scholar 

  88. R. Barbieri, G.F. Giudice, Upper bounds on supersymmetric particle masses. Nucl. Phys. B 306(1), 63–76 (1988)

    Article  ADS  Google Scholar 

  89. L. Roszkowski, R. Ruiz de Austri, R. Trotta, Efficient reconstruction of CMSSM parameters from LHC data—a case study. Phys. Rev. D 82, 055003 (2010). arXiv:0907.0594 [hep-ph]

    ADS  Google Scholar 

  90. D.M. Ghilencea, H.M. Lee, M. Park, Tuning supersymmetric models at the LHC: a comparative analysis at two-loop level, 23 pp., 46 figs. arXiv:1203.0569 [hep-ph]

  91. D. Ghilencea, G. Ross, The fine-tuning cost of the likelihood in SUSY models. arXiv:1208.0837 [hep-ph]

  92. E. Komatsu et al. (WMAP Collaboration), Seven-year Wilkinson microwave anisotropy probe (WMAP) observations: cosmological interpretation. Astrophys. J. Suppl. 192, 18 (2011). arXiv:1001.4538 [astro-ph.CO]

    Article  ADS  Google Scholar 

  93. M. Benayoun, P. David, L. DelBuono, F. Jegerlehner, Upgraded breaking of the HLS model: a full solution to the τ e + e and ϕ decay issues and its consequences on g-2 VMD estimates. Eur. Phys. J. C 72, 1848 (2012). arXiv:1106.1315 [hep-ph]

    Article  ADS  Google Scholar 

  94. D. Asner et al. (Heavy Flavor Averaging Group Collaboration), Averages of b-hadron, c-hadron, and tau-lepton properties. arXiv:1010.1589 [hep-ex]

  95. B. Aubert et al. (BABAR Collaboration), Measurement of branching fractions and CP and isospin asymmetries in BK γ. arXiv:0808.1915 [hep-ex]

  96. B. Aubert et al. (BABAR Collaboration), Observation of the semileptonic decays \(B \to D^{*} \tau^{-} \bar{\nu}_{\tau}\) and evidence for \(B \to D \tau^{-} \bar{\nu}_{\tau}\). Phys. Rev. Lett. 100, 021801 (2008). arXiv:0709.1698 [hep-ex]

    Article  ADS  Google Scholar 

  97. M. Antonelli et al. (FlaviaNet Working Group on Kaon Decays Collaboration), Precision tests of the standard model with leptonic and semileptonic kaon decays. arXiv:0801.1817 [hep-ph]

  98. W.-M. Yao et al. (Particle Data Group), Review of particle physics. J. Phys. G 33 (2006). http://pdg.lbl.gov

  99. V. Barger, P. Langacker, H.-S. Lee, G. Shaughnessy, Higgs sector in extensions of the MSSM. Phys. Rev. D 73, 115010 (2006). arXiv:hep-ph/0603247

    ADS  Google Scholar 

  100. E. Aprile et al. (XENON100 Collaboration), Dark matter results from 100 live days of XENON100 data. Phys. Rev. Lett. 107, 131302 (2011). arXiv:1104.2549 [astro-ph.CO]

    Article  ADS  Google Scholar 

  101. M.-O. Bettler, Search for B s,d μμ at LHCb with 300 pb−1. arXiv:1110.2411 [hep-ex]

  102. G. Aad et al. (ATLAS Collaboration), Search for squarks and gluinos using final states with jets and missing transverse momentum with the ATLAS detector in \(\sqrt{s} = 7~\mbox {TeV}\) proton–proton collisions. arXiv:1109.6572 [hep-ex]

  103. O. Buchmueller, R. Cavanaugh, D. Colling, A. de Roeck, M. Dolan et al., Implications of initial LHC searches for supersymmetry. Eur. Phys. J. C 71, 1634 (2011). arXiv:1102.4585 [hep-ph]

    Article  ADS  Google Scholar 

  104. E. Aprile et al. (XENON100 Collaboration), Likelihood approach to the first dark matter results from XENON100. Phys. Rev. D 84, 052003 (2011). arXiv:1103.0303 [hep-ex]

    ADS  Google Scholar 

  105. G. Cowan, K. Cranmer, E. Gross, O. Vitells, Asymptotic formulae for likelihood-based tests of new physics. Eur. Phys. J. C, Part. Fields 71(2), 1–19 (2011). arXiv:1007.1727 [data-an]

    Article  Google Scholar 

  106. A.L. Read, Presentation of search results: the CL s technique. J. Phys. G, Nucl. Part. Phys. 28(10), 2693 (2002). http://stacks.iop.org/0954-3899/28/i=10/a=313

    Article  ADS  Google Scholar 

  107. J.M. Alarcon, J.M. Camalich, J.A. Oller, The chiral representation of the πN scattering amplitude and the pion-nucleon sigma term. arXiv:1110.3797 [hep-ph]

  108. M.M. Pavan, I.I. Strakovsky, R.L. Workman, R.A. Arndt, The pion nucleon Sigma term is definitely large: results from a GWU analysis of pi N scattering data. PiN Newslett. 16, 110–115 (2002). arXiv:hep-ph/0111066

    Google Scholar 

  109. J. Gasser, H. Leutwyler, M. Sainio, Sigma-term update. Phys. Lett. B 253(1–2), 252–259 (1991). http://www.sciencedirect.com/science/article/pii/037026939191393A

    Article  ADS  Google Scholar 

  110. R. Koch, A new determination of the pi N sigma term using hyperbolic dispersion relations in the (nu**2, t) plane. Z. Phys. C 15, 161–168 (1982)

    ADS  Google Scholar 

  111. J. Giedt, A.W. Thomas, R.D. Young, Dark matter, the CMSSM and lattice QCD. Phys. Rev. Lett. 103, 201802 (2009). arXiv:0907.4177 [hep-ph]

    Article  ADS  Google Scholar 

  112. R.D. Young, A.W. Thomas, Octet baryon masses and sigma terms from an SU(3) chiral extrapolation. Phys. Rev. D 81, 014503 (2010). arXiv:0901.3310 [hep-lat]

    Article  ADS  Google Scholar 

  113. J. Gasser, H. Leutwyler, Quark masses. Phys. Rep. 87, 77–169 (1982)

    Article  ADS  Google Scholar 

  114. B. Borasoy, U.-G. Meissner, Chiral expansion of baryon masses and sigma-terms. Ann. Phys. 254, 192–232 (1997). arXiv:hep-ph/9607432

    Article  ADS  Google Scholar 

  115. M.E. Sainio, Pion nucleon sigma-term: a review. PiN Newslett. 16, 138–143 (2002). arXiv:hep-ph/0110413

    Google Scholar 

  116. M. Knecht, Working group summary: pi N sigma term. PiN Newslett. 15, 108–113 (1999). arXiv:hep-ph/9912443

    Google Scholar 

  117. J.R. Ellis, K.A. Olive, C. Savage, Hadronic uncertainties in the elastic scattering of supersymmetric dark matter. Phys. Rev. D 77, 065026 (2008). arXiv:0801.3656 [hep-ph]

    ADS  Google Scholar 

  118. G. Aad et al. (ATLAS Collaboration), The ATLAS experiment at the CERN Large Hadron Collider. J. Instrum. 3, S08003 (2008)

    Article  Google Scholar 

  119. R. Adolphi et al. (CMS Collaboration), The CMS experiment at the CERN LHC. J. Instrum. 3, S08004 (2008)

    Article  Google Scholar 

  120. G. Aad et al. (ATLAS Collaboration), Search for diphoton events with large missing transverse momentum in 1 fb−1 of 7 TeV proton–proton collision data with the ATLAS detector. arXiv:1111.4116 [hep-ex]

  121. G. Aad et al. (ATLAS Collaboration), Searches for supersymmetry with the ATLAS detector using final states with two leptons and missing transverse momentum in \(\sqrt{s} = 7\) TeV proton–proton collisions. arXiv:1110.6189 [hep-ex]

  122. G. Aad et al. (ATLAS Collaboration), Search for new phenomena in final states with large jet multiplicities and missing transverse momentum using \(\sqrt{s}=7\) TeV pp collisions with the ATLAS detector. J. High Energy Phys. 1111, 099 (2011). arXiv:1110.2299 [hep-ex]

    Article  ADS  Google Scholar 

  123. G. Aad et al. (ATLAS Collaboration), Search for supersymmetry in final states with jets, missing transverse momentum and one isolated lepton in \(\sqrt{s} = 7\) TeV pp collisions using 1 fb−1 of ATLAS data. arXiv:1109.6606 [hep-ex]

  124. S. Chatrchyan et al. (CMS Collaboration), Search for supersymmetry at the LHC in events with jets and missing transverse energy. http://cdsweb.cern.ch/record/1381201

  125. S. Chatrchyan et al. (CMS Collaboration), Search for supersymmetry in all-hadronic events with MT2. http://cdsweb.cern.ch/record/1377032

  126. S. Chatrchyan et al. (CMS Collaboration), Search for supersymmetry in all-hadronic events with missing energy. http://cdsweb.cern.ch/record/1378478

  127. N. Desai, B. Mukhopadhyaya, Constraints on supersymmetry with light third family from LHC data. arXiv:1111.2830 [hep-ph]

  128. C. Beskidt, W. de Boer, D. Kazakov, F. Ratnikov, E. Ziebarth et al., Constraints from the decay \(B_{s}^{0} \to\mu^{+} \mu^{-}\) and LHC limits on supersymmetry. Phys. Lett. B 705, 493–497 (2011). arXiv:1109.6775 [hep-ex]

    Article  ADS  Google Scholar 

  129. B. Allanach, T. Khoo, K. Sakurai, Interpreting a 1 fb−1 ATLAS search in the minimal anomaly mediated supersymmetry breaking model. arXiv:1110.1119 [hep-ph]

  130. S. Gieseke, D. Grellscheid, K. Hamilton, A. Papaefstathiou, S. Platzer et al., Herwig++ 2.5 release note. arXiv:1102.1672 [hep-ph]

  131. S. Ovyn, X. Rouby, V. Lemaitre, DELPHES, a framework for fast simulation of a generic collider experiment. arXiv:0903.2225 [hep-ph]

  132. W. Beenakker, R. Hopker, M. Spira, PROSPINO: a program for the production of supersymmetric particles in next-to-leading order QCD. arXiv:hep-ph/9611232 [hep-ph]

  133. S. Agostinelli et al. (GEANT4 Collaboration), GEANT4: a simulation toolkit. Nucl. Instrum. Methods A 506, 250–303 (2003)

    Article  ADS  Google Scholar 

  134. A. Buckley, A. Shilton, M.J. White, Fast supersymmetry phenomenology at the Large Hadron Collider using machine learning techniques. arXiv:1106.4613 [hep-ph]

  135. A. Hoecker, P. Speckmayer, J. Stelzer, J. Therhaag, E. von Toerne, H. Voss, TMVA: toolkit for multivariate data analysis. PoS ACAT, 040 (2007). arXiv:physics/0703039

    Google Scholar 

  136. J.-H. Zhong, R.-S. Huang, S.-C. Lee, R.-S. Huang, S.-C. Lee, A program for the Bayesian neural network in the ROOT framework. Comput. Phys. Commun. 182, 2655–2660 (2011). arXiv:1103.2854 [physics.data-an]

    Article  ADS  MATH  Google Scholar 

  137. G. Aad et al. (ATLAS Collaboration), Search for squarks and gluinos using final states with jets and missing transverse momentum with the atlas detector in \(\sqrt{s}\) = 7 tev proton–proton collisions. Tech. rep., CERN, Geneva, Mar (2012). http://cdsweb.cern.ch/record/1432199

  138. S. Chatrchyan et al. (CMS Collaboration), Search for supersymmetry with the razor variables at cms. http://cdsweb.cern.ch/record/1430715

  139. G. Aad et al. (ATLAS Collaboration), Combination of Higgs boson searches with up to 4.9 fb−1 of pp collisions data taken at a center-of-mass energy of 7 TeV with the ATLAS experiment at the LHC. Tech. Rep. ATLAS-CONF-2011-163, CERN, Geneva, Dec (2011). http://cdsweb.cern.ch/record/1406358

  140. S. Chatrchyan et al. (CMS Collaboration), Combination of sm Higgs searches. http://cdsweb.cern.ch/record/1406347

  141. S. Chatrchyan et al. (CMS Collaboration), Combined results of searches for the standard model Higgs boson in pp collisions at \(\sqrt{s} = 7\) TeV. arXiv:1202.1488 [hep-ex]

  142. S. Akula, B. Altunkaynak, D. Feldman, P. Nath, G. Peim, Higgs boson mass predictions in SUGRA unification, recent LHC-7 results, and dark matter. Phys. Rev. D 85, 075001 (2012). arXiv:1112.3645 [hep-ph]

    ADS  Google Scholar 

  143. M. Kadastik, K. Kannike, A. Racioppi, M. Raidal, Implications of the 125 GeV Higgs boson for scalar dark matter and for the CMSSM phenomenology. arXiv:1112.3647 [hep-ph]

  144. H. Baer, V. Barger, A. Mustafayev, Neutralino dark matter in mSUGRA/CMSSM with a 125 GeV light Higgs scalar. arXiv:1202.4038 [hep-ph]

  145. A. Azatov, R. Contino, J. Galloway, Model-independent bounds on a light Higgs. arXiv:1202.3415 [hep-ph]

  146. A. Hoecker, The hadronic contribution to the muon anomalous magnetic moment and to the running electromagnetic fine structure constant at MZ—overview and latest results. Nucl. Phys. Proc. Suppl. 218, 189–200 (2011). arXiv:1012.0055 [hep-ph]

    Article  ADS  Google Scholar 

  147. T. Goecke, C.S. Fischer, R. Williams, Hadronic light-by-light scattering in the muon g-2: a Dyson–Schwinger equation approach. Phys. Rev. D 83, 094006 (2011). arXiv:1012.3886 [hep-ph]

    ADS  Google Scholar 

  148. K. Hagiwara, R. Liao, A.D. Martin, D. Nomura, T. Teubner, (g−2) μ and alpha(\(M_{Z}^{2}\)) re-evaluated using new precise data. J. Phys. G 38, 085003 (2011). arXiv:1105.3149 [hep-ph]

    Article  ADS  Google Scholar 

  149. S. Bodenstein, C. Dominguez, K. Schilcher, Hadronic contribution to the muon g-2: a theoretical determination. Phys. Rev. D 85, 014029 (2012). arXiv:1106.0427 [hep-ph]

    ADS  Google Scholar 

  150. T. Goecke, C.S. Fischer, R. Williams, Hadronic contribution to the muon g-2: a Dyson–Schwinger perspective. arXiv:1111.0990 [hep-ph]

  151. G. Aad et al. (ATLAS Collaboration), Observation of a new particle in the search for the standard model Higgs boson with the ATLAS detector at the LHC. Phys. Lett. B (2012). arXiv:1207.7214 [hep-ex]

  152. S. Chatrchyan et al. (CMS Collaboration), Observation of a new boson at a mass of 125 GeV with the CMS experiment at the LHC. Phys. Lett. B (2012). arXiv:1207.7235 [hep-ex]

Download references

Acknowledgements

The authors are indebted to Sudhir Gupta and Doyoon Kim for their assistance with the calculation of Higgs boson production cross sections and decays. B.F. is thankful to Farhan Feroz for assistance with MultiNest. M.J.W. thanks Teng Jian Khoo and Ben Allanach for conversations regarding the calculation of ATLAS-based likelihoods for candidate SUSY models. This research was funded in part by the ARC Centre of Excellence for Particle Physics at the Tera-scale, and in part by the Project of Knowledge Innovation Program (PKIP) of Chinese Academy of Sciences Grant No. KJCX2.YW.W10. A.B. acknowledges the support of the Scottish Universities Physics Alliance. The use of Monash Sun Grid (MSG) and Edinburgh ECDF high-performance computing facilities is also gratefully acknowledged. Most numerical calculations were performed on the Australian National Computing Infrastructure (NCI) National Facility SGI XE cluster and Multi-modal Australian ScienceS Imaging and Visualisation Environment (MASSIVE) cluster.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Benjamin Farmer.

Appendices

Appendix A: Fast approximation to combined CLs limits for correlated likelihoods

In this appendix we offer a brief justification of the simplified method used to combine the ATLAS CL s limits on sparticle production in our analysis. In this method an approximate combined confidence limit is obtained for a specified model point by simply taking the most powerful observed (lowest CL s value) limit from one of several signal regions, or search channels. Our aim is to demonstrate a set of minimal conditions under which this procedure is conservative. This will be done by demonstrating conditions under which the following inequality holds:

$$ \min(CL_{s_1},CL_{s_2} ) \ge CL_{s_{1,2}} $$
(35)

where \(CL_{s_{1}}\) is the value of the CL s statistic for some signal model, derived from a dataset which we may call ‘channel 1’; \(CL_{s_{2}}\) is the value of CL s under the same signal model but derived from a correlated dataset ‘channel 2’; and \(CL_{s_{1,2}}\) is the value of CL s for this signal model derived from the full combination of the two datasets, accounting rigorously for correlations between datasets. This inequality does not hold in general, but if the experimental situation is such that it does hold, it means that the combined dataset results in a more powerful limit than either of the individual datasets alone, or conversely that considering only the most constraining of the two individual dataset limits is conservative. In the course of this exercise we will make use of the asymptotic results obtained in Ref. [105].

We remind the reader that the CL s statistic is defined as

$$ CL_{s} = \frac{p_{s+b}}{1-p_b} $$
(36)

where p s+b and p b are p-values derived using the null hypotheses ‘s+b’ and ‘b’, respectively. s+b is the hypothesis that the data is generated from the nominal signal plus background model, while b supposes that the data contains background events only. In the CL s method these p-values are computed using the likelihood ratio statistic

$$ q = -2\ln\frac{L_{s+b}}{L_b}=-2\ln\frac{L(\mu =1,\hat{\theta}(1))}{L(\mu=0,\hat{\theta}(0))}, $$
(37)

where L s+b and L b are the likelihoods of the ‘s+b’ and ‘b’ models, respectively. The second equality defines the background model as one which can be obtained by scaling the signal model by an appropriate ‘signal strength’ parameter μ, which is set to zero. \(\hat{\theta}(1)\) and \(\hat{\theta}(0)\) are the profiled values of any nuisance parameters. In the asymptotic limit (which requires sufficiently many candidate events) this statistic is given by the Wald approximation, with μ as the parameter of interest, as

$$ q = \frac{(\hat{\mu}-1)^2}{\sigma^2} - \frac{\hat {\mu}^2}{\sigma^2} = \frac{1-2\hat{\mu}}{\sigma^2}, $$
(38)

where \(\hat{\mu}\) is the best fit value of μ given some dataset, and σ 2 is the variance of \(\hat{\mu}\) (which is normally distributed) under either the ‘s+b’ or the ‘b’ models, that is, σ takes the values σ s+b and σ b when the μ=1 and μ=0 models are assumed to be generating the data, respectively. Using so-called ‘Asimov’ datasets, which when observed cause \(\hat{\mu}\) to adopt its true value (either 1 or 0; see Ref. [105]) we can obtain σ 2 as

$$\begin{aligned} &\sigma^2 = \frac{1-2\mu'}{q_A}, \quad \mbox{so that}\quad \sigma^2_{s+b} = \frac{1}{\vert q_{A_{s+b}}\vert} \quad \mathrm{and} \end{aligned}$$
(39)
$$\begin{aligned} & \sigma^2_{b} = \frac{1}{\vert q_{A_b}\vert} \end{aligned}$$
(40)

where μ′ is the assumed true value of μ and q A is the value of q obtained using the relevant Asimov dataset.

The asymptotic distribution f of the statistic q is normal with mean (1−2μ′)/σ 2 and variance 4/σ 2 so the p- in Eq. (36) can be computed by values

$$\begin{aligned} p_{s+b} = \int_{q_{\mathrm{obs}}}^{\infty} f(q|s+b) \,dq &= \ 1 - \varPhi\biggl(\frac{q_{\mathrm{obs}} + 1/\sigma^2_{s+b}}{2/\sigma_{s+b}} \biggr) \\ &= 1 - \varPhi\biggl(\frac{q_{\mathrm{obs}} - q_{A_{s+b}}}{2\sqrt {|q_{A_{s+b}}|}} \biggr) \end{aligned}$$
(41)

and

$$ \begin{aligned}[b] p_{b} = \int_{-\infty}^{q_{\mathrm{obs}}} f(q|b) \,dq &= \ \varPhi\biggl(\frac{q_{\mathrm{obs}} -1/\sigma^2_{b}}{2/\sigma _{b}} \biggr) \\ &= \varPhi\biggl(\frac{q_{\mathrm{obs}} - q_{A_b}}{2\sqrt {|q_{A_b}|}} \biggr). \end{aligned} $$
(42)

Let us now go to the case where the observed events in all channels are in accordance with the background hypothesis, such that \(\hat{\mu}\sim0\). Then \(q_{\mathrm{obs}}\sim q_{A_{b}}\). Furthermore, in this case the 95 % CL s limit lies near model points which predict low signals, so we may further take σσ s+b σ b (since the distribution f under both s+b and b hypotheses will be very similar). Also note that in this limit \(q_{A_{s+b}}=-q_{A_{b}}\). Our p-values can thus be simplified to

$$ p_{s+b} = 1 - \varPhi\bigl(\sqrt{|q_A|} \bigr) \quad \mathrm{and} \quad p_{b} = \varPhi(0 ) = \frac{1}{2} $$
(43)

(where we have also used the knowledge that \(\mathrm {sign}(q_{A_{s+b}}) = -1\)). We can thus write the inequality of Eq. (35) as:

$$\begin{aligned} &\frac{p_{s+b;1}}{1-p_{b;1}} \ge\frac{p_{s+b;1,2}}{1-p_{b;1,2}} \\ &\quad \rightarrow\quad 1 - \varPhi\bigl(\sqrt{|q_{1_{A}}|} \bigr) \ge 1 - \varPhi\bigl(\sqrt{|q_{1,2_{A}}|} \bigr) \end{aligned}$$
(44)

where we have assumed WLOG that \(CL_{s_{1}}\le CL_{s_{2}}\). The function Φ(x) is monotonically increasing with x, so our inequality will hold if

$$ \vert q_{1_{A}}\vert\le\vert q_{1,2_{A}} \vert. $$
(45)

To determine when this is the case, we need to express \(q_{1,2_{A}}\) in terms of the parameters describing \(q_{1_{A}}\) and \(q_{2_{A}}\). We can do this by obtaining the two parameter Wald expansion for the combined test statistic q 1,2 (i.e. taking a Taylor expansion of q about the best fit values of μ 1 and μ 2, up to second order):

$$\begin{aligned} q_{1,2}&= \frac{1}{1-\rho} \biggl(\frac{1-2\hat {\mu}_1}{\sigma_1^2} + \frac{1-2\hat{\mu}_2}{\sigma_2^2} \\ &\quad {}-2\rho\frac{1-\hat{\mu}_1-\hat{\mu}_2}{\sigma_1\sigma_2} \biggr), \end{aligned}$$
(46)

where ρ characterises linear correlations between the two channels, taking values in the domain (−1,1), and \(\hat{\mu}_{1},\hat {\mu}_{2}\) and \(\sigma_{1}^{2},\sigma_{2}^{2}\) are the best fit μ values and their variances, as obtained above for each individual channel. Again we use the Asimov dataset for the background hypothesis, which sets \(\hat{\mu}_{1}=\hat{\mu}_{2}=0\), to find \(q_{1,2_{A,b}}\):

$$ q_{1,2_{A,b}} = \frac{1}{1-\rho} \biggl( \frac{1}{\sigma_1^2} + \frac{1}{\sigma_2^2} -2\rho\frac{1}{\sigma _1\sigma_2} \biggr), $$
(47)

which, like \(q_{1_{A,b}}\), is strictly positive. Using this expression together with Eq. (39) we can rewrite the inequality of Eq. (45) as

$$ \frac{1}{\sigma^2_1} \le\frac{1}{1-\rho} \biggl( \frac{1}{\sigma_1^2} + \frac{1}{\sigma_2^2} -2\rho\frac{1}{\sigma _1\sigma_2} \biggr). $$
(48)

One can readily see that Eq. (48) holds in the case ρ=0, i.e. when no correlations exist between channels. Knowing this, we may vary ρ from this point and see where the equality is achieved in order to check if the inequality may be violated. Setting the equality we solve for σ 1, finding the two general solutions

$$ \sigma_1=\sigma_2 \bigl(\rho\pm \sqrt{(\rho-1)\rho} \bigr), $$
(49)

from which it is apparent that no real solutions exist for 0<ρ<1, while such solutions do exist for −1<ρ<0. We could convert this to a bound on the allowed values of σ 1/σ 2, since only the positive root solution can give a positive σ 1, but negative correlations are not relevant for our signal regions, which are correlated due to shared events, so we are done.

We can thus conclude that if channel correlations are linear and positive, the observed event counts are not far from the expected background, the nominal signal hypothesis at the limit is small, and enough events are observed for asymptotic formulae to hold, then we can safely take the most powerful limit from among several channels as an estimate of the full combination, without overestimating the combined limit. Violations of these conditions may result in the target inequality of Eq. (35) being violated, with a particular concern being that this can occur as the observed events differ from the background expectation; however, it is difficult to determine the general conditions under which this happens. Certainly if one channel sees an excess above the background while another does not then in general the combined limit will be weaker than one obtained using only the more constraining (background-like) channel.

Nevertheless, in our special case we may be confident that our method remains approximately valid thanks to the procedure used by ATLAS to produce their official limits (in Ref. [102]), to which our approximate limits are fitted. ATLAS also do not attempt to rigorously account for the correlations between channels; they follow a similar procedure to us and, for each point in the CMSSM parameter space, take the limit from the channel with the best expected limit. We, on the other hand, take the channel with the best observed limit, which, following the discussion of this appendix, can be expected to less reliably approximate the rigorous combination.

We follow our more approximate procedure because ATLAS do not provide the expected limits on the mean signal for each channel; however, it is possible to estimate these using the asymptotic formulae discussed in this appendix, and so we use these estimates to gauge the seriousness of the difference between our method and the one used by ATLAS.

To do this a model of the likelihood in each signal region is needed. Taking the random variable to be the best fit signal strength \(\hat {\mu}\), the simplest option is the normal limit of a Poisson likelihood, with standard deviation σ modified by convolution with normal signal and background systematics σ s and σ b . The mean and variance are then simply

$$\begin{aligned} &\mu= \frac{n-b}{s} \end{aligned}$$
(50)
$$\begin{aligned} &\sigma^2 = \bigl(\mu'\bigl(s + \sigma_s^2 \bigr) + \sigma_b^2\bigr)/s^2 \end{aligned}$$
(51)

where n=μs+b is the expected total number of events (and μ′=1 or 0 as before). ATLAS provide estimates of σ b so we use these, however σ s is not provided since it varies point to point. This variation would require a large effort to model so we simply fit a single value for σ s for each channel, ensuring that the observed 95 % CL s limits obtained from our simplified likelihood agree with ATLAS (we have also checked that varying this value has little effect on our results).

We then use this model likelihood to estimate the expected limits on the signal yield in each channel for each point in our training dataset, and obtain an estimate of the ATLAS combined observed limit by taking the observed CL s value of each training data point to be the one obtained from the channel with the lowest expected CL s value for that point (i.e. following ATLAS’s method). We find the difference between this estimate of the ATLAS limit and the one used in our analysis to be very small: of the 26491 training points there are 100 which are classified (into excluded/not excluded) differently by the two limits. We show these points in Fig. 10; they predominantly occur in a group clustered at low m 0, and for most of them the observed strongest limit comes from R1, while we estimate that the expected strongest limit comes from R2.

Fig. 10
figure 10

Classification of training data for the ATLAS 1 fb−1 jets+MET search used in the main analysis. Two methods for combining the ATLAS limits for each search channel are used: the method used in this analysis uses the most constraining observed CL s value from the set of channels at each training data point to determine its classification, while ATLAS use the observed CL s value from the signal region with the most powerful expected exclusion. We have estimated the limit that would be obtained from the ATLAS method using asymptotic approximations for the signal likelihood. Training data model points which are excluded at 95 % CL s by both limits are coloured red, while model points not excluded by either are coloured green. Points where conflict exists are coloured black. The official ATLAS limit is overlaid for comparison. Points are sampled from the full CMSSM parameter space as described in the text, but are projected onto the (m 0,m 1/2) plane for visualisation (Color figure online)

Appendix B: Plots of CMSSM profile likelihoods and marginalised posteriors

This appendix contains the figures referred to in Sect. 7. We refer the reader to that section for further information.

Fig. 11
figure 11

The evolution of the profile of the (log-)likelihood function from the “pre-LEP” situation (first row), to including the LEP Higgs search and XENON100 data (second row), to adding the 1 fb−1 LHC sparticle searches (third row), to folding in the 2012 February Higgs search results. Contours containing 68 % and 95 % confidence regions are shown. The above results were obtained using the log prior. Results obtained using the CCR prior (not shown) show variations consistent with the different sampling density but are qualitatively similar

Fig. 12
figure 12

The evolution of the profile of the (log-)likelihood function from the “pre-LEP” situation (first row), to including the LEP Higgs search and XENON100 data (second row), to adding the LHC sparticle searches (third row), to folding in the 2012 February Higgs search results. Contours containing 68 % and 95 % confidence regions are shown. The above results were obtained using the log prior and have been reweighted to estimate the effect of removing the δa μ constraint. Significant deterioration of the sampling is seen due to the shift in the preferred regions away from the originally sampled regions, however the general impact of removing the δa μ constraint can be seen in the motion of the preferred regions upwards in the mass parameters. Of particular note is the very strong shift to high A 0 when the ATLAS Higgs search results are imposed, which is much less pronounced in Fig. 11, indicating very strong tension between the ATLAS Higgs search results and the δa μ constraint. Results obtained using the CCR prior (not shown) show variations consistent with the different sampling density but are qualitatively similar

Fig. 13
figure 13

The evolution of the CMSSM marginalised posterior probability distributions from the “pre-LEP” situation (first row), to including the LEP Higgs search and XENON100 data (second row), to adding the LHC sparticle searches (third row), to folding in the 2012 February Higgs search results. Log priors are used and 68 % and 95 % credible regions are shown

Fig. 14
figure 14

The evolution of the CMSSM marginalised posterior probability distributions from the “pre-LEP” situation (first row), to including the LEP Higgs search and XENON100 data (second row), to adding the LHC sparticle searches (third row), to folding in the 2012 February Higgs search results. Natural (“CCR”) priors are used and 68 % and 95 % credible regions are shown. The natural prior can be seen to favour lower M 0 and tanβ than the log prior

Rights and permissions

Reprints and permissions

About this article

Cite this article

Balázs, C., Buckley, A., Carter, D. et al. Should we still believe in constrained supersymmetry?. Eur. Phys. J. C 73, 2563 (2013). https://doi.org/10.1140/epjc/s10052-013-2563-y

Download citation

  • Received:

  • Revised:

  • Published:

  • DOI: https://doi.org/10.1140/epjc/s10052-013-2563-y

Keywords

Navigation