hep-ph0609295/ds.tex
1: \documentclass[12pt]{JHEP3}
2: 
3: \usepackage[dvips]{epsfig}
4: 
5:  \newlength{\wth}
6:  \setlength{\wth}{10cm}
7: 
8: \newcommand{\twographst}[2]{%
9:  \unitlength=1.1in
10:  \begin{picture}(5.8,2.3)(0.5,0.25)
11:  \put(-0.04,2.54){\epsfig{file=#1, width=0.698 \wth,angle=270}}
12:  \put(0.85,0.5){\epsfig{file=#12, width=0.68 \wth}}
13:  \put(2.66,2.54){\epsfig{file=#2, width=0.698 \wth, angle=270}}
14:  \put(3.55,0.5){\epsfig{file=#22, width=0.68 \wth}}
15:  \put(0.5,2.1){(a)}
16:  \put(3.2,2.1){(b)}
17:  \end{picture}
18: }
19: 
20: \newcommand{\fourgraphst}[4]{%
21:  \unitlength=1.1in
22:  \begin{picture}(5.8,4)(0.5,0.4)
23: \put(0,2){\put(-0.04,2.54){\epsfig{file=#1, width=0.698 \wth,angle=270}}
24:  \put(0.85,0.5){\epsfig{file=#12, width=0.68 \wth}}
25:  \put(2.66,2.54){\epsfig{file=#2, width=0.698 \wth, angle=270}}
26:  \put(3.55,0.5){\epsfig{file=#22, width=0.68 \wth}}
27:  \put(0.5,2.1){(a)}
28:  \put(3.2,2.1){(b)}
29: }
30: 
31: \put(0,0){\put(-0.04,2.54){\epsfig{file=#3, width=0.698 \wth,angle=270}}
32:  \put(0.85,0.5){\epsfig{file=#32, width=0.68 \wth}}
33:  \put(2.66,2.54){\epsfig{file=#4, width=0.698 \wth, angle=270}}
34:  \put(3.55,0.5){\epsfig{file=#42, width=0.68 \wth}}
35:  \put(0.5,2.1){(c)}
36:  \put(3.2,2.1){(d)}
37: }
38:  \end{picture}
39: }
40: 
41: \newcommand{\fourgraphs}[4]{%
42:  \unitlength=1.1in
43:  \begin{picture}(5.8,4.4)(0.3,0.3)
44: \put(0,2.4){\put(0.5,0){\epsfig{file=#1, width=0.698 \wth}}
45:  \put(2.9,0){\epsfig{file=#2, width=0.698 \wth}}
46:  \put(0.5,2.2){(a)}
47:  \put(3,2.2){(b)}}
48: \put(0,0){\put(0.5,0){\epsfig{file=#3, width=0.698 \wth}}
49:  \put(2.9,0){\epsfig{file=#4, width=0.698 \wth}}
50:  \put(0.5,2.2){(c)}
51:  \put(3,2.2){(d)}}
52:  \end{picture}
53: }
54: 
55: \newcommand{\onegraph}[1]{%
56:  \unitlength=1.1in
57:  \begin{picture}(2.5,2.3)(0.5,0.25)
58:  \put(-0.04,2.54){\epsfig{file=#1, width=0.698 \wth,angle=270}}
59:  \put(0.85,0.5){\epsfig{file=#12, width=0.68 \wth}}
60:  \end{picture}
61: }
62: 
63:  \newcommand{\twographs}[2]{%
64:  \unitlength=1.1in
65:  \begin{picture}(5.8,2.3)(0,0.25)
66:  \put(0,0){\epsfig{file=#1.eps, width=0.698 \wth}}
67:  \put(2.7,0){\epsfig{file=#2.eps, width=0.698 \wth}}
68:  \put(0.0,2.3){(a)}
69:  \put(2.7,2.3){(b)}
70:  \end{picture}
71: }
72: 
73: \title{The Dark Side of mSUGRA}
74: 
75: \author{Benjamin C Allanach$^{1}$, Christopher G Lester$^{2}$ and Arne M
76:   Weber$^{3}$ \\ 
77: $^{1}$ DAMTP, CMS, Wilberforce Road, Cambridge CB3 0WA, UK\\
78: $^{2}$ Cavendish Laboratory, J.J. Thomson Avenue, Cambridge CB3 0HE, UK\\
79: $^3$ Max Planck Inst.\ f\"{u}r Phys., F\"{o}hringer Ring 6, D-80805 Munich,
80:   Germany\\ 
81: }
82: 
83: \keywords{Supersymmetry Effective Theories, Cosmology of Theories beyond the
84:   SM, Dark Matter}
85: \abstract{We study the $\mu<0$ branch of the minimal supergravity
86:   ansatz of the minimal supersymmetric standard model. The extent to
87:   which $\mu<0$ is disfavoured compared to $\mu>0$ in global fits is
88:   calculated 
89:   with Markov Chain Monte Carlo methods and bridge sampling. The fits
90:   include   state-of-the-art two-loop MSSM contributions to the electroweak
91:   observables 
92:   $M_W$ and $\sin^2 \theta_w^l$, as well as the anomalous magnetic moment of the
93:   muon $(g-2)_\mu$, the relic density of dark matter and other
94:   relevant indirect observables. $\mu<0$ is only
95:   marginally disfavoured in global fits and should be considered in mSUGRA
96:   analyses. We estimate that the ratio of %marginalised
97:   probabilities is $P(\mu<0) / P(\mu>0)=0.07-0.16$. 
98: }
99: 
100: \preprint{DAMTP-2006-75\\MPP--2006--122\\ Cavendish-HEP-2006-024 \\
101:   hep-ph/0609295} 
102: 
103: \begin{document}
104: 
105: \section{Introduction}
106: 
107: There has been increasing recent attention on global fits of various indirect
108: data to minimal supergravity (mSUGRA), also sometimes called the constrained
109: minimal supersymmetric standard model
110: (CMSSM)~\cite{Ellis:2003si,Profumo:2004at,Baltz:2004aw,Ellis:2004tc,Stark:2005mp}.
111: mSUGRA makes phenomenological 
112: analysis of the minimal supersymmetric standard model (MSSM)
113: tractable via the low number of free parameters.
114: In fact, the scalar masses $m_0$, 
115: gaugino masses $M_{1/2}$ and trilinear coupling $A_0$ are assumed to be
116: universal at a gauge unification scale  $M_{GUT} \sim 2 \times 10^{16}$ GeV. 
117: If the MSSM is present in nature and if the mSUGRA
118: universality assumptions are approximately correct, chi-squared or probability
119: distributions for potential collider/dark matter observables
120: can be derived. Early
121: fits~\cite{Robros,Ellis:2003si,Profumo:2004at,Baltz:2004aw,Ellis:2004tc}
122: necessarily had fixed input parameters to reduce 
123: the dimensionality of the input parameter space, making scans practicable.
124: It is usually assumed that neutralinos constitute the current cold dark matter
125: content of the universe, since they are weakly interacting, electrically and
126: colour neutral and stable.
127: The predicted value of dark matter relic density $\Omega_{DM} h^2$ is a very
128: strong constraint on viable mSUGRA parameter space, effectively reducing its
129: dimensionality by 1.
130: 
131: The accuracy of the inferred value of $\Omega_{DM} h^2$ from WMAP data makes a
132: global fit to all of the relevant mSUGRA parameters potentially difficult
133: because the system is rather under-constrained, possessing narrow, steep
134: valleys of degenerate $\chi^2$ minima.  
135: If the MSSM is confirmed in colliders, it will hopefully be possible to break
136: such degeneracies with collider observables. This does not help us at present,
137: where we want to provide a sort of `weather forecast' for future colliders.
138: It was indicated in ref.~\cite{Baltz:2004aw} that the powerful Markov Chain
139: Monte Carlo (MCMC) technique might allow us to find the probability
140: distribution of a  fully global fit to indirect data. 
141: Two of us went on~\cite{Allanach:2005kz} to demonstrate that MCMCs do indeed
142: allow such a fit, investigating collider observables. One of us examined the
143: effect of a naturalness prior~\cite{Allanach:2006jc}. Our results were
144: confirmed and expanded in Ref.~\cite{deAustri:2006pe}, also utilising the MCMC
145: method and including a one-loop MSSM calculation of the W-boson mass $M_W$ and 
146: the weak leptonic mixing angle $\sin^2 \theta_w^l$ in the likelihood density.
147: The purpose of MCMC mSUGRA global fits is two-fold: as well as producing
148: interesting and useful physics results in themselves, we may profit from the
149: experience of utilising and developing the MCMC tools, which could prove very
150: useful when analysing future collider data. 
151: 
152: It is our purpose in the present paper to extend the previous $\mu>0$ global
153: fits to $\mu<0$. 
154: Besides the observables studied in~\cite{Allanach:2005kz} we now also include
155: $M_W$ and $\sin \theta_w^l$ in our analysis, as done in
156: e.g. \cite{Ellis:2004tc,deAustri:2006pe}. Being highly sensitive to new physics
157: these very accurately measured quantities play a key role in the electroweak
158: sector and are therefore also of great interest when it comes to further
159: constraining the mSUGRA parameter space. It was shown in the literature and
160: that the one-loop predictions for
161: the two observables alone do not bear enough accuracy to make reliable
162: predictions. In fact, the pure one-loop predictions can lead to results
163: contradictory to the state-of-the-art predictions \cite{Heinemeyer:2006px}
164: used in our analysis. These contain the known higher order contributions from
165: both the Standard Model and the MSSM\@. Extending our analysis to negative
166: values of 
167: $\mu$ it is crucial to further use a very accurate prediction for
168: $(g-2)_\mu$. The dominant two-loop corrections \cite{2loop} to this quantity
169: are therefore also taken into account in the present analysis. 
170: It is well known that the measured anomalous value of the magnetic
171: moment of the muon $(g-2)_\mu$ is roughly 2$\sigma$ above the Standard
172: Model (SM) predicted value. This positive contribution is predicted by
173: some regions of mSUGRA parameter space, provided $\mu>0$. The ``dark
174: side'' of mSUGRA (i.e.\ $\mu<0$) provides a negative contribution,
175: thereby being disfavoured by the $(g-2)_\mu$ measurement. In a global
176: fit, one can trade likelihood penalties between different observables
177: and the conclusion that $\mu<0$ is disfavoured to roughly 2$\sigma$ is
178: not at all obvious. We will calculate the extent to which the dark
179: side is ruled out by using MCMCs with ``bridge sampling''
180: \cite{radford}.  
181: We will encounter
182: problems associated with isolated likelihood density maxima in the
183: dark side, potentially ruining MCMC convergence. Fortunately, bridge
184: sampling provides a solution to the convergence issue and we are able
185: to calculate the degree to which the dark side is disfavoured with
186: respect to $\mu>0$. 
187: As well as extending
188: previous analyses to $\mu<0$, we have made several technical
189: improvements in the calculation of the likelihood compared with previous
190: attempts in the literature.
191: 
192: If the lightest supersymmetric particle decays into Standard Model
193: particles, as is the case in R-parity violation for instance, its
194: relic density will be essentially zero today. In that case, one
195: requires to obtain the WMAP fitted $\Omega_{DM} h^2$ from some other
196: source than neutralinos (gravitinos or hidden sector matter for
197: instance).  In order to investigate this case, we will also perform
198: the fits for the case where all relevant data {\em except}
199: $\Omega_{DM} h^2$ are included in the likelihood density.  Such fits
200: will help us to understand the impact of $\Omega_{DM} h^2$ in
201: constraining the model, as well as being relevant for the R-parity
202: violating mSUGRA~\cite{rpvSUGRA} in the limit of small R-parity
203: violating couplings.
204: 
205: We now go on to detail the various constraints used on the model in
206: section~\ref{sec:const}. The results of the dark side fitting
207: procedure are compared and contrasted against the better known $\mu>0$ ones
208: in section~\ref{sec:dark}, before the effect of dropping the dark-matter
209: constraint is examined in section~\ref{sec:noDM}. 
210: Closing remarks are presented in section~\ref{sec:conc}. A presentation of the
211: fitting procedures is confined to the appendix: 
212: Markov Chain Monte Carlos and bridge sampling are discussed in
213: appendix~\ref{sec:bridge}. Convergence problems
214: and their resolution are discussed in appendix~\ref{sec:conv}. 
215: 
216: \section{Constraints \label{sec:const}}
217: 
218: \TABULAR[r]{|c|c|}{\hline
219: mSUGRA parameter & range \\ \hline
220: $A_0$ & -4 TeV to 4 TeV \\
221: $m_0$ & 60 GeV to 4 TeV \\
222: $M_{1/2}$ & 60 GeV to 2 TeV \\
223: $\tan \beta$ &  2 to 62 \\ \hline
224: SM parameter & constraint \\ \hline
225: $1/\alpha^{\overline{MS}}$ & 127.918$\pm$0.018 \\
226: $\alpha_s^{\overline{MS}}(M_Z)$ & 0.1176$\pm$0.002 \\
227: $m_b(m_b)^{\overline   MS}$ & 4.24$\pm$0.11 GeV \\
228: $m_t$ & 171.4$\pm$2.1 \\ \hline
229: }{Input parameters \label{tab:inp}}
230: We vary 8 input parameters relevant to the model. 
231: The range of intrinsically mSUGRA parameters considered is shown in
232: Table~\ref{tab:inp}, where $\tan \beta$ is the ratio of the two MSSM Higgs
233: doublet vacuum expectation values.
234: We use Ref.~\cite{PDG} for the QED coupling
235: constant $\alpha^{\overline{MS}}$, the strong coupling constant
236: $\alpha_s^{\overline{MS}}(M_Z)$ and the running mass of the bottom quark 
237: $m_b(m_b)^{\overline   MS}$, all in the $\overline{MS}$ renormalisation scheme
238: The recent Tevatron top mass $m_t$ measurement \cite{mtmeas} is also employed. 
239: These SM inputs are shown in Table~\ref{tab:inp}. 
240: Experimental errors are so small on the mass of the $Z^0$ boson $M_Z$ 
241: and the muon decay constant $G_\mu$ that we 
242: fix them to their central values of 91.1876 GeV and $1.16637 \times 10^{-5}$
243: GeV$^{-2}$ respectively. 
244: The Standard Model (SM) input parameters are allowed to vary within 4$\sigma$
245: of their central values but a $\chi^2$ penalty 
246: \begin{equation}
247: \chi_i^2 = \frac{(c_i - p_i({\bf m}))^2}{\sigma_i^2} \label{chisq}
248: \end{equation}
249: is applied for observable $i$.  $c_i$ denotes the central value of the
250: experimental measurement, $p_i({\bf m})$ represents the value ``{\bf
251: p}redicted'' at any stage of the MCMC sampling given knowledge of the
252: model ${\bf m}$ presumed to be ``true'' at that point.  Finally
253: $\sigma_i$ is the standard error of the measurement.  Equivalently,
254: expressing this in the language of likelihoods, we are assuming that
255: each of these measurements have Gaussian errors,\footnote{The
256: experimental constraints on BR($B_s \rightarrow \mu^+ \mu^-$) and the
257: LEP constraints on the Higgs mass, each described later, are not
258: Gaussian constraints and must therefore be treated differently.
259: Nothing prevents us from continuing to parametrise their likelihood
260: distributions in the same way, however, but it should be realised that
261: a consequence of this is that the ``$\chi$-squared penalty'' (i.e. $-2
262: \log{{\mathcal L}_i}$) will not be parabolic, as ${\mathcal L}_i$ is
263: not a Gaussian distribution in these cases.} and that the likelihood
264: distribution ${\mathcal L}_i \equiv p(c_i | {\bf m})$ for any one
265: measurement may be written in the following way:
266: \begin{equation}
267: {\mathcal L}_i \equiv p(c_i | {\bf m}) =
268: \frac 1 {\sqrt{2 \pi \sigma_i^2}}
269: \exp \left[ {-\chi_i^2 /2} \right].
270: \end{equation}
271: The normalisation constant $\sqrt{2 \pi \sigma_i^2}$ may be ignored
272: in subsequent calculations as the absolute value of ${\mathcal L}_i$ will
273: never be needed.  It will only be necessary to know the {\em ratios}
274: \/of values of ${\mathcal L}_i$ at neighbouring points in the MCMC
275: chain, or between chains in which the neglected constants are identical.
276: 
277: 
278: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
279: 
280: 
281: In order to calculate predictions for observables from the inputs in
282: Table~\ref{tab:inp}, we use {\tt
283:   SOFTSUSY2.0.7}~\cite{softsusy} to first calculate the MSSM spectrum.
284: \TABULAR{|cc|cc|cc|cc|}
285: {\hline
286: $m_{\chi_1^0}$ & 37  & $m_{\chi^\pm_1}$ & 67.7 & 
287: $m_{\tilde g}$ & 195 &
288: $m_{{\tilde \tau}_1}$ & 76 \\ $m_{{\tilde l}_R}$ & 88 & $m_{{\tilde t}_1}$ &
289:   86.4 &
290: $m_{{\tilde b}_1}$ & 91 & $m_{{\tilde q}_R}$ & 250 \\
291: $m_{{\tilde \nu}_{e,\mu}}$ & 43.1 & & & & & & \\
292: \hline}{Lower bounds applied to sparticle mass predictions (in GeV). \label{tab:bds}}
293: We apply the bounds in
294: Table~\ref{tab:bds} in order to take into account 95$\%$ limits coming from
295: negative sparticle searches~\cite{PDG}. Any point transgressing these bounds 
296: is given a zero likelihood density (or, equivalently, an infinite $\chi^2$).
297: Also, we set a zero likelihood for any inconsistent point, e.g.\ one which does
298: not break electroweak symmetry correctly, or a point that contains tachyonic
299: sparticles. 
300: For points that are not ruled out,
301: we then link the MSSM spectrum via the SUSY Les Houches Accord~\cite{slha} to
302:   {\tt 
303:   micrOMEGAs1.3.6}~\cite{micromegas}, which then calculates $\Omega_{DM} h^2$, the
304: branching ratios $BR(b \rightarrow s \gamma)$ and $BR(B_s \rightarrow \mu^+
305: \mu^-)$ and $(g-2)_\mu$.  
306: 
307: The measured value of the anomalous magnetic moment $(g-2)_\mu$ is in conflict
308: with the SM predicted value by~\cite{PDG}
309: \begin{equation}
310: \delta a_\mu \equiv \delta \frac{(g-2)_\mu}{2} = (22 \pm 10) \times 10^{-10}.
311: \end{equation}
312: This excess may be explained by a supersymmetric contribution,
313: the sign of which is identical to the sign of the superpotential $\mu$
314: parameter~\cite{susycont}. After obtaining the one-loop MSSM value of
315: $(g-2)_\mu$ from {\tt micrOMEGAs}, we add the following dominant 2-loop
316: corrections~\cite{2loop,private}: the logarithmic piece of the 2-loop QED
317: contribution, two-loop stop-higgs and chargino-stop/sbottom contributions.
318: 
319: \EPSFIGURE{bsmumupen.eps,height=8cm,angle=270}{$\chi^2$ penalty on
320:   BR($B_s \rightarrow \mu^+ \mu^-$) from
321:   \cite{private2}\label{fig:bsmumu}} The Tevatron has recently been
322:   instrumental in bounding the branching ratio of the rare decay
323:   channel $B_s \rightarrow \mu^+ \mu^-$~\cite{bsmumu}.  Such bounds
324:   help constrain the mSUGRA parameter space~\cite{Ellis:2005sc}.  We
325:   apply a $\chi^2$ penalty on the value predicted by {\tt
326:   micrOMEGAs1.3.6} derived from CDF Tevatron Run II
327:   data~\cite{private2}. The resulting penalty is shown in
328:   Fig.~\ref{fig:bsmumu}.
329: 
330: Recently, it has been claimed that light sparticles are preferred by
331: the two weak observables $\sin^2 \theta_w^l$ and $M_W$
332: \cite{Ellis:2004tc,Ellis:2006ix}.  In Ref.~\cite{deAustri:2006pe},
333: $M_W$ and $\sin^2 \theta_w^l$ were used at one loop order to help
334: constrain mSUGRA in a global MCMC fit. The preference for such light
335: SUSY was not particularly evident in the global fits.  We examine the
336: mSUGRA predictions for $M_W$ and $\sin^2 \theta_w^l(\mbox{eff})$ in
337: Figs.~\ref{fig:mw} and~\ref{fig:st} for $A_0=0$, $\tan \beta=10$,
338: $\mu>0$, equal $m_0$ and $M_{1/2}$ and central experimental values for
339: the other inputs.  For the ``SOFTSUSY'' lines, the default {\tt
340: SOFTSUSY} calculation is used.  This contains the full SOFTSUSY MSSM
341: contributions to the leptonic mixing angle $\sin^2 \theta_w^l$ and
342: $M_W$. It also contains the dominant 2-loop Standard Model
343: contributions to $M_W$.  For the lines marked ``2-loop'', the SUSY Les
344: Houches Accord is used to communicate with a currently private code
345: that calculates the W-boson mass $M_W$~\cite{Heinemeyer:2006px}, and
346: the effective leptonic mixing angle variable $\sin^2\theta^l_w$,
347: calculated to two loops in the dominant MSSM parameters. We use the
348: most general MSSM result for the full one-loop contributions.  Besides
349: all known corrections due to SUSY particles, the full SM contributions
350: are also included in the predictions for $M_W$ and $\sin^2\theta^l_w$,
351: leading to the currently most accurate predictions within the MSSM\@.
352: The ``SM'' lines show the SM limit, where all corrections involving
353: sparticles are dropped, i.e.\ the state-of-the-art SM results
354: \cite{MWSM,Awramik:2004ge} with $M_{H^{\mathrm{SM}}}=M_h$.  They vary
355: slightly with $m_0=M_{1/2}$ because the varying mSUGRA parameters
356: produce different values of the Higgs boson mass $m_h$.  The
357: horizontal lines on the figures show the current $1\sigma$
358: experimental limits~\cite{mw,sinth}
359: \begin{equation}
360: M_W = 80.392\pm0.031\mbox{~GeV}, 
361: \qquad \sin^2 \theta_w^l = 0.23153 \pm 0.00020,
362: \end{equation}
363: where we have added experimental and theoretical errors in
364: quadrature. 
365: 
366: \DOUBLEFIGURE{mw,height=8cm}{sinth,height=8cm}{Various approximations to $M_W$
367:   in mSUGRA \label{fig:mw}}{Various
368:   approximations to $\sin^2 \theta_w$ in mSUGRA. \label{fig:st}}
369: The theoretical errors in the predicted $M_W$ and $\sin^2
370: \theta_w^l$  are estimated to be $10$ MeV and 12$\times 10^{-5}$
371: respectively~\cite{Ellis:2003si}. We use these uncertainties for the
372: purposes of comparison, although they have been slightly reduced recently by
373:   the addition of additional two-loop corrections taken into account in the
374:   present analysis~\cite{Heinemeyer:2006px,drMSSMal2B}. 
375: We see from the SOFTSUSY line in the figure that
376: the  prediction of $M_W$ does not have a strong preference for light SUSY,
377: since the model is within the 1$\sigma$ errors up until $m_0=M_{1/2}=4$
378: TeV. In actual fact, only very light SUSY masses are disfavoured by the
379: ``SOFTSUSY'' line, leading to predictions above the $1\sigma$-range.   
380: %
381: The situation is similar for the $\sin^2\theta^l_w$ ``SOFTSUSY'' line, where
382: again only very light SUSY masses lead to predictions outside the
383: $1\sigma$-range. % 
384: %Also, there is a even a preference for heavy SUSY from the $\sin^2 \theta_w^l$
385: %``SOFTSUSY'' line. 
386: However, using the best available predictions, corresponding to the ``2-loop''
387: lines, a 
388: preference for light $m_0=M_{1/2}$ can be seen in the prediction for
389: $M_W$. 
390: The SM curve which lies just below the $1\sigma$-interval is approached
391: from above in the decoupling limit, furthermore indicating a slight preference
392: of the MSSM over the SM\@. 
393: The preference for light SUSY is not as striking for $\sin^2\theta_w^l$. Here
394: most of the $m_0=M_{1/2}$ values are doing equally well, which is mainly due
395: to the fact 
396: that the SM prediction for $\sin^2\theta_w^l$ is already well within the
397: $1\sigma$-range.
398: With the behaviour of the ``one-loop'' curve and the best available result
399: being qualitatively different, it is
400: desirable to use the more accurate result for $M_W$ and
401: $\sin^2\theta_w^l$  when calculating 
402: the $\chi^2$ contributions of $M_W$ and $\sin^2 \theta_w^l$  using
403: Eq.\ref{chisq}. Although Ref.~\cite{deAustri:2006pe} only used the one-loop
404: predictions, the theoretical errors were correspondingly enlarged in order to
405: take the larger uncertainty from higher order terms into account.
406: 
407:   \EPSFIGURE{mh,height=7cm}{LEP2 higgs $\chi^2$ penalty paid 
408: \label{fig:mhpen}}
409: LEP2 constraints on the lightest CP-even higgs mass are included as a further
410: likelihood penalty following a parameterisation of LEP2 data in the SM
411: limit~\cite{lep2higgs}. For the LEP2 constraints,
412: the SM limit is a good approximation for mSUGRA, since 
413: sparticle mass limits imply that we must be near the decoupling r\'{e}gime of
414: the MSSM~\cite{decouple}. We estimate that the {\tt SOFTSUSY2.0.7}
415: determination of $m_h$ has a 2 GeV theoretical
416: error in mSUGRA, although it may be somewhat larger in the general
417: MSSM~\cite{Allanach:2004rh}.  
418: We therefore smear the parameterised LEP2 Higgs likelihood density ${\mathcal
419:   L}_{LEP2}(m_h)$
420: with a Gaussian distribution of width $\sigma_h=2$ GeV:
421: \begin{equation}
422: {\mathcal L}_h(m_h) = \int_{m_h-4 \sigma_h}^{m_h+4 \sigma_h} dx \frac{1}{\sqrt{2 \pi}
423:   \sigma_h} e^{\frac{-(m_h-x)^2}{2 
424:   \sigma_h^2}} {\mathcal L}_{LEP2}(x).
425: \label{Higgserror}
426: \end{equation}
427: The result of this procedure
428:   leads to the effective $\Delta \chi^2=-2 \ln {\mathcal L}_h$ penalty shown in
429:   Fig.~\ref{fig:mhpen}. The slight excess of candidate Higgs events
430:   over the background prediction at LEP2 can be seen by a negative
431:   $\Delta\chi^2$ penalty in the figure for $m_h \sim 116-121$ GeV.
432: 
433: The rare bottom quark branching ratio is
434: $BR(b \rightarrow s \gamma)$ is constrained to be~\cite{hfg}
435: \begin{equation}
436: BR(b \rightarrow s \gamma)=  (3.55\pm0.38) \times 10^{-4}, \label{bsg}
437: \end{equation}
438: obtained by adding the experimental error with the estimated theory
439: error~\cite{gamb} of $0.3 \times 10^{-4}$ in quadrature.
440: Very recent estimates~\cite{Misiak:2006zs,Andersen:2006hr}
441: of $BR(b \rightarrow s \gamma)$ are compatible with Eq.~\ref{bsg} at the
442: 1$\sigma$ level, although
443: the error has decreased.
444: Our prediction of $BR(b \rightarrow s \gamma)$ is substituted for $p_i$ in
445: Eq.~\ref{chisq} in order to calculate $\chi_{BR(b \rightarrow s \gamma)}^2$.
446: 
447: We use the WMAP3~\cite{wmap} power law $\Lambda$-CDM fitted value of 
448: \begin{equation}
449: \Omega \equiv \Omega_{DM} h^2 = 0.104^{+0.0073}_{-0.0128} \label{omega}
450: \end{equation}
451: for the dark matter relic density of the universe. We initially assume that the
452: neutralinos are stable and that they constitute the whole of the dark matter
453: relic density. 
454: Eq.~\ref{chisq} is used to calculate $\chi^2_{\Omega}$, with
455: $\sigma_{\Omega}=0.0073$ for a prediction lower than the central
456: experimental value and 0.0128 otherwise.
457: 
458: 
459: Having described the calculation of the likelihood associated
460: with each individual measurement, we are now in a position to define
461: the likelihood of the set of all measurements or observables, taken
462: together.  We are required to calculate the joint (total) likelihood
463: $\mathcal L$ of all the measurements given the truth, {\em i.e.}\ in the
464: notation of Eq~\ref{chisq} we want to know:
465: \begin{eqnarray}
466: {\mathcal L} &=& p( \mbox{measurements} | \mbox{true model}) \\
467: &=& p(c_1, c_2, \ldots | {\bf m}) \\
468: &=& p(c_1 | {\bf m}) \cdot p(c_2 | c_1 , {\bf m}) \cdot p(c_3 |
469: c_1, c_2, {\bf m}) \cdot \ldots \\
470: &=& {\mathcal L}_1 \cdot p(c_2 | c_1 , {\bf m}) \cdot p(c_3 |
471: c_1, c_2, {\bf m}) \cdot \ldots
472: \end{eqnarray}
473: which is not in general equal to
474: \begin{equation}
475: {\mathcal L}_1 \cdot
476: {\mathcal L}_2 \cdot
477: {\mathcal L}_3 \cdot
478: \ldots
479: \end{equation}
480: unless we can be confident that for each measurement
481: \begin{equation}
482: p(c_3 | c_1, c_2, {\bf m}) \approx p(c_3 | {\bf m}) \mbox{\qquad{\em
483: etc}}.\label{eq:whatwewantapprox}
484: \end{equation}
485: Fortunately we can be confident that Eq~\ref{eq:whatwewantapprox} {\em
486: does}\/ hold in our situation, as we have deliberately constructed ${\bf
487: m}$ to be broad enough such that all fundamental parameters which
488: might reasonably be expected to correlate any two of the measurements
489: are included within it\footnote{If the design of ${\bf m}$ were not
490: broad enough, then ${\bf m}$ would have to be extended.  For example:
491: were it the case that the up quark mass $m_u$ was expected to 
492: significantly correlate two
493: or more of the observables, then for Eq~\ref{eq:whatwewantapprox} to
494: continue to hold, $m_u$ would have to be added to ${\bf m}$ enlarging
495: its dimension by one.}.  We are therefore at liberty to write:
496: \begin{eqnarray}
497: {\mathcal L} &=& {\mathcal L}_1 \cdot
498: {\mathcal L}_2 \cdot
499: {\mathcal L}_3 \cdot
500: \ldots \\ 
501: &=& e^{-\sum_i \chi_i^2/2}
502: \end{eqnarray}
503: for the total likelihood, and we can be confident that this product
504: takes into account the expected correlations between all the
505: observables contained, due to the nature of the space $\left\{{\bf m}
506: \right\}$ of models considered.
507: 
508: 
509: \section{Dark Side Fits \label{sec:dark}}
510: 
511: \FIGURE{\twographst{m0m12Neg}{m0m12Pos}
512: \caption{mSUGRA Fits for (a) $\mu<0$ (b) $\mu>0$ marginalised to the
513:   $m_0$-$M_{1/2}$ plane. The posterior probability is indicated by the
514:   bar on the right hand side. The inner (outer) contours show the
515:   $68\%$ $(95\%)$ confidence region respectively.\label{fig:m0m12}} }
516:   We now compare and contrast the dark side fits to those with
517:   $\mu>0$.  In Figs.~\ref{fig:m0m12}(a) and~\ref{fig:m0m12}(b), we
518:   show the posterior probabilities $P$ marginalised\footnote{For readers
519:   unfamiliar with the term: marginalisation
520:   means ``integrated over the unseen dimensions of parameter space'' in this
521:   context.} to the 
522:   $m_0-M_{1/2}$ plane for both signs of $\mu$.  As with all
523:   2-dimensional marginalised plots in this paper, we bin the plane into
524:   75$\times$75 2-dimensional bins. The colour bar on the right hand side
525:   of the figures shows the posterior probability $P$ of each bin divided
526:   by the maximum posterior probability of any bin in the plot. In
527:   Fig.~\ref{fig:m0m12}(a), the only 68$\%$ contour\footnote{Note that the
528:     confidence regions in Figs.~\ref{fig:m0m12} (and those in later plots)
529:     should strictly be referred to as ``Bayesian credible intervals''
530:     (each region contains a fixed amount of the posterior probability) to
531:     distinguish them from the related concept from Frequentist Statistics
532:     called a ``confidence interval''. Usage of the term ``credible
533:     interval'' is not common in High Energy Physics, however, and we will
534: stick to ``confidence region''.} is at the lowest
535:   values of $m_0$: all other contours are 95$\%$ confidence region contours
536:   due to   the lowish likelihood densities. The brokenness of the contours is a
537:   result of statistical fluctuations in the results. Although
538:   these are visible, there is clearly reliable information 
539:   their trajectories in the plot.  The $\mu<0$ plot displays
540:   two isolated local maxima,
541:   whereas the 95$\%$ confidence region of $\mu>0$ is continuously
542:   connected. We discuss the relative normalisation of the two $\mu<0$ maxima
543:   in appendix~\ref{sec:conv}.
544:   In the 95$\%$ confidence region of Fig.~\ref{fig:m0m12}a closest to the
545:   origin,   the  relic 
546:   density is dominantly depleted 
547:   by either stop co-annihilation~\cite{Boehm:9911,arnie,Ellis:2001nx} ${\tilde
548:   t} \chi_1^0   \rightarrow t g$ or  
549:   stau co-annihilation~\cite{Griest:1990kh} ${\tilde \tau}_1 \chi_1^0 \rightarrow \tau \gamma$.
550: On the other hand, the 95$\%$ region at higher $m_0-M_{1/2}$ 
551: consists of resonant Higgs annihilation
552: regions~\cite{Drees:1992am,Arnowitt:1993mg,Djouadi:2005dz}, where
553: $\chi_1^0\chi_1^0 \rightarrow h,A^0 \rightarrow b {\bar b}/\tau^+
554: \tau^-$ and the focus point region where the LSP has a significant
555: higgsino component and $\chi_1^0\chi_1^0 \rightarrow ZZ, WW, t
556: \bar{t}$~\cite{Feng:1999mn,Feng:1999zg,Feng:2000gh}.
557: There was no significant
558: stop co-annihilation~\cite{Boehm:9911,arnie,Ellis:2001nx} region for
559: $\mu>0$.  
560: 
561: \FIGURE{\fourgraphst{m12a0}{a0tb}{m12tb}{m0tb}
562: \caption{Constraints from global fits with $\mu<0$ mSUGRA marginalised to
563:   2-dimensional parameter planes.
564: We have assumed a flat prior. The posterior probability is indicated by the bar
565:   on the right hand side. The inner (outer) contours show the
566:   boundary of a $68\%$
567:   $(95\%)$ confidence region.
568: \label{fig:mulz}}
569: } We now include some other marginalisations on the other
570: 2-dimensional parameter planes for $\mu<0$ mSUGRA with a flat
571: prior. They are displayed in
572: Figs.~\ref{fig:mulz}(a)-(d). Figs.~\ref{fig:mulz}(a) and~\ref{fig:mulz}(b)
573: show that the probability density for the 2-chain 
574: co-annihilation sample is not separated from the other sample in either
575: the $M_{1/2}-A_0$ plane or the $\tan \beta$-$A_0$ plane (the almost
576: disconnected region at the bottom of Fig.~\ref{fig:mulz}(a) consists
577: of the light $h^0$-pole region). There is only a modest separation in
578: the $M_{1/2}$-$\tan \beta$ plane: the 68$\%$ contours do not connect
579: the two regions, whereas the 95$\%$ contours
580: do. Fig.~\ref{fig:mulz}(d) shows that the connection between the two
581: samples in the $m_0-\tan \beta$ plane is marginal. The $m_0-A_0$
582: marginalisation was useful for investigating the physics behind the two
583: isolated probability maxima, and is displayed in appendix~\ref{sec:conv}.
584: 
585: \FIGURE{\fourgraphs{higgs}{gluino}{squark}{neut}
586: \caption{Probability distributions in mass of (a) the lightest CP even higgs,
587:  (b) gluino, (c) the left-handed squark and (d) the neutralino. Flat priors
588:  have been assumed.
589: \label{fig:mass}}
590: }
591: The probability distributions of the masses of selected MSSM particles are
592: shown in Fig.~\ref{fig:mass} for
593: both signs of $\mu$ in mSUGRA\@. 
594: The $\mu<0$ sample has not been
595: normalised to the correct relative normalisation compared to $\mu>0$ in the
596: figure. The lightest CP-even higgs, the gluino and the lightest neutralino all
597: have mass distributions that are remarkably similar for either sign of $\mu$. 
598: However, in Fig.~\ref{fig:mass}(c), we see that the $\mu<0$ sample has a flat
599: plateau for higher squark masses, whereas the $\mu>0$ sample tails off
600: somewhat. High squark masses
601: will result in smaller total SUSY cross-sections at the LHC but the lighter
602: gluino should still provide enough events for discovery if $m_{\tilde g}<2$
603: TeV~\cite{Armstrong:1994it,CMS}.
604: Sharp peaks at low gluino and neutralino masses are due to the
605: $h^0$-resonance annihilation region~\cite{Allanach:2005kz}. The broader peak
606: of the $\mu<0$ curve in Fig.~\ref{fig:mass}(c) is due mostly to the
607: co-annihilation sample. The significant probability densities for large gluino
608: and squark masses are rather alarming, as then SUSY detection at the LHC would
609: require a longer running time.
610: Gauginos are not so sensitive to the range of the
611: prior in Table~\ref{tab:inp} but the sfermions are~\cite{deAustri:2006pe} and
612: a reduced range makes lighter sfermions more likely.
613: Also, naturalness priors~\cite{Allanach:2006jc} have a large impact, reducing
614: the likely sparticle masses. 
615: 
616: \subsection{Further investigations of the fits}
617: 
618: We examine the best-fit points from each sampling in Table~\ref{tab:bestfit}.
619: \TABULAR{|c|cc|c|cc|}{\hline
620:       & $\mu<0$ & $\mu>0$ & & $\mu<0$ & $\mu>0$  \\ \hline
621: $m_0$/GeV &  3610   & 156   & $\delta a_\mu/10^{-10}$ &-0.4 & 14.2 \\
622: $M_{1/2}$/GeV & 93 & 569  & $BR(b \rightarrow s \gamma)/10^{-4}$ &3.65 &3.41 \\
623: $A_0$/GeV &   -56   & 270  & $BR(B_s \rightarrow \mu^+ \mu^-)/10^{-9}$ &3.1 &3.7 \\
624: $\tan \beta$ &6.0 &24.1 & $\sin^2 \theta_w^l(\mbox{eff})$ & 0.23153&0.23152\\
625: $\Omega_{DM} h^2$ & 0.102 & 0.101 & $M_W$/GeV & 80.382& 80.368\\
626: $m_h$/GeV & 117.5 & 115.8 & $\chi^2$ & 4.5& 1.5\\
627: \hline
628: }{Best-fit points from the MCMC samplings for each sign of $\mu$. 
629: $\chi^2 \equiv
630:   \sum_i \chi_i^2$ and the Standard Model inputs are close to their
631:   experimental central values in each case. \label{tab:bestfit}}
632: We see from the table that, as expected, the $\mu>0$ sample has a better
633: best-fit point and a  correspondingly lower $\chi^2$, mainly due to the much
634: better fit to $\delta a_\mu$. In agreement with
635: Refs.~\cite{Ellis:2004tc,Ellis:2006ix}, the best-fit $\mu>0$ point is at light
636: SUSY masses. 
637: The $m_0$ and $M_{1/2}$ parameters are smaller
638: for the $\mu>0$ case than for $\mu<0$, corresponding to lighter sparticles and
639: the larger contribution to the anomalous magnetic moment of the muon. 
640: If we take the $\mu<0$ best-fit point and {\em flip}\/ the sign of $\mu$, we
641: find that the point does not break electroweak symmetry properly and is
642: excluded. The $\mu>0$ best-fit 
643: point has a total $\chi^2$ of 181 when the sign of $\mu$ is flipped, mostly
644: due to an increase in the predicted value of $\Omega_{DM} h^2$ to 0.195.
645: 
646: The region of smaller $m_0$, $M_{1/2}$ is more
647:   probable for $\mu>0$ than for $\mu<0$: this will lead to a relatively
648:   heavier $\mu<0$ 
649:   spectrum.  Our $\mu>0$ results are generally similar to previous analyses
650:   which did {\em not} \/include $M_W$ and $\sin^2 \theta_w^l$ as
651:   constraints in Ref.~\cite{Allanach:2005kz} and to those which included the
652:   one-loop {\tt SOFTSUSY} prediction for $M_W$, $\sin^2 \theta_w^l$ with
653:   enlarged theoretical errors~\cite{deAustri:2006pe}.  
654:   This seems in contradiction to the conclusions of
655:   Ellis {\em et al}~\cite{Ellis:2004tc,Ellis:2006ix}, where it is claimed that
656:   the electroweak   variables prefer light SUSY\@. Indeed, Fig.~\ref{fig:st}
657:   indicates that $M_W$, $\sin^2 \theta_w^l$ do mildly prefer light SUSY but
658:   our results show that this preference is washed out in the global fits. 
659:   Our results allow for heavier sparticles than Ellis {\em et al}, mainly
660:   because we have chosen to allow more relevant parameters to vary: 8 compared
661:   to 2 in their paper (one dimension is fixed by requiring the relic density
662:   prediction to be the central WMAP-constrained value). Their fits are for
663:   different discrete values of 
664:   fixed $\tan \beta$, but if it were allowed to vary, we believe that the
665:   confidence regions there would be enlarged. In the present paper, we also
666:   obtain additional smearing from allowing $m_t$, $\alpha_s$, $\alpha$
667:   and $m_b$ to vary.
668:   \FIGURE{\twographs{sugraOppmw}{sugraOppstw}
669: \caption{Posterior probability distributions for weak observables in
670:   mSUGRA\@. \label{fig:weakobs}} } 
671: Figs.~\ref{fig:weakobs}(a) and~\ref{fig:weakobs}(b) show the probability
672:   distributions for $M_W$, 
673:   $\sin^2 \theta_w^l$ coming from the mSUGRA fits for both signs of
674:   $\mu$, although the sign does not make much difference.  We see that
675:   the $M_W$ prediction coming from the fits is skewed towards values
676:   lower than the central empirical value, corresponding to a
677:   preference for heavy SUSY from the rest of the
678:   fits. Fig.~\ref{fig:mw} confirms that heavier SUSY tends to have
679:   lighter values of $M_W$.  From Fig.~\ref{fig:st}, we see that heavy
680:   SUSY tends to be on the upper 1$\sigma$ empirical limit of $\sin^2
681:   \theta_w^l$.  Fig.~\ref{fig:weakobs}(b) does show evidence for this
682:   skew, which is rather mild.
683: \DOUBLEFIGURE{sugraOppomega,width=7cm}{sugraOppmt,width=7cm}{Dark
684:   matter relic density probability distribution \label{fig:whsq}}{Top
685:   mass probability distribution
686:   \label{fig:mt}} 
687: 
688: \EPSFIGURE{tb,width=7cm}{Probability distribution for $\tan \beta$ in mSUGRA
689: \label{fig:tb}}
690: The strongest constraint in the fits comes from the dark matter relic
691: density.  In Fig.~\ref{fig:whsq}, we show the probability densities
692: resulting from the fits for the dark matter relic density. Each curve
693: is normalised slightly differently to allow better viewing of the
694: results. We see that both the $\mu>0$ and the $\mu<0$ $\Omega_{DM}
695: h^2$ distributions follow the empirical constraint closely, except for
696: a slight excess just lower than the central value.
697: 
698: Another important aspect influencing the fits is the integrated {\em
699: volume} \/of the probability density, which is automatically taken
700: into account in a Bayesian analysis, as was demonstrated and pointed out in
701: Ref.~\cite{deAustri:2006pe}.  The usual arguments based purely on
702: values of $\chi^2$ above the best-fit value are valid when the probability
703: density function is Gaussian in the interesting parameters
704: (which is certainly not the case here, as even a
705: cursory glance at Fig.~\ref{fig:m0m12} allows). Thus even if, say, the
706: stop co-annihilation region had a much lower $\chi^2$ than other
707: regions of the fits,\footnote{In reality, the stop co-annihilation
708: region fits the data rather marginally.} the fact that its volume in
709: 8-dimensional input parameter space is much smaller than the other
710: regions will automatically disfavour it since its integrated
711: probability will be low.
712: 
713: The LEP2 higgs constraints shown in Fig.~\ref{fig:mhpen}, along with
714: the current empirical value of $m_t$ shown in Table~\ref{tab:inp}
715: favours rather heavy mSUGRA\@.  Indeed, Fig.~\ref{fig:mt} shows that
716: $m_t$ is skewed to somewhat higher masses than the empirical
717: constraint, illustrating this tension (again, we have altered the
718: normalisation of the curves slightly for clarity).  Inspection of the
719: $\alpha(M_Z)^{\overline{MS}}$, $\alpha_s(M_Z)^{\overline{MS}}$,
720: $m_b(m_b)$ inputs show that they follow their empirical probability
721: distributions very closely.  There is a large volume of parameter
722: space for the $A^0$ dark matter annihilation region at high $\tan
723: \beta>10$ particularly for $\mu>0$, as shown in Fig.~\ref{fig:tb}.
724: High values of $\tan \beta$ mean that light SUSY is disfavoured by
725: $BR[b \rightarrow s \gamma]$ and $BR[B_s \rightarrow \mu^+ \mu^-]$
726: since the data disfavours SUSY contributions~\cite{Ellis:2005sc},
727: which are approximately proportional to $\tan^2 \beta / M_{SUSY}^4$
728: and $\tan^6 \beta / M_{SUSY}^4$ respectively.  The distributions of
729: these two observables are shown in Fig.~\ref{fig:bs}.  
730: As can be seen
731: from the figure, the sign of the SUSY contribution to each observable
732: depends upon the sign of $\mu$.  The maxima in each figure correspond
733: to observables being close to their SM limit. 
734: $BR[b   \rightarrow s \gamma]$ prefers $\mu>0$ mildly, as $\mu<0$ tends to
735: predict a $BR[b   \rightarrow s \gamma]$ larger than the central empirical
736: value. 
737: \FIGURE{\twographs{sugraOppbsg}{sugraOppBsmumu}
738: \caption{Probability distributions of (a) $BR(b \rightarrow s
739:     \gamma)$ and (b)$BR(B_s \rightarrow \mu^+ \mu^-)$ in
740:     mSUGRA\@. \label{fig:bs}} 
741: }
742: 
743: \FIGURE{\twographs{sugraOppg-2}{vargm2}
744: \caption{Probability distributions for $\delta a_\mu$ in mSUGRA (a)
745:   for flat priors and both signs of $\mu$ and (b) for $\mu>0$ and
746:   various different priors. The key is explained in the
747:   text.\label{fig:gm2}} } $(g-2)_\mu$ is expected to be the observable
748:   that most strongly discriminates between the two signs of $\mu$. We
749:   plot its distribution in each case in Fig.~\ref{fig:gm2}(a). Since
750:   $\mu<0$ has the wrong sign of $\delta a_\mu$ compared to experiment,
751:   the probability density bunches around zero. Clearly heavy SUSY with
752:   a less negative SUSY contribution is favoured. However, we see that
753:   the $\mu>0$ distribution also prefers smaller SUSY contributions
754:   than the data in the global fits. This was initially unexpected, and
755:   will lead to $\mu<0$ mSUGRA being less ruled out.
756: %
757: In order to understand this behaviour better, we re-weight the $\mu>0$ sample
758: in various ways. $\delta a_\mu$ is approximately proportional to $\tan^2 \beta
759:   / M_{SUSY}^4$ and we need large $\tan \beta$ and rather light SUSY in order to
760:   get a sizable value in line with the central experimental value. As
761:   explained above, this is somewhat in conflict with LEP2 Higgs constraints
762:   and    $BR(b \rightarrow s \gamma)$. We re-weight the $\mu>0$ chains,
763:   dividing by a 
764:   number that removes the likelihood contribution from the LEP2 Higgs
765:   constraint and the  $BR(b \rightarrow s \gamma)$ measurement, i.e.\
766:   ${\mathcal L}_h {\mathcal L}_{BR(b \rightarrow s \gamma)}$. The probability
767:   distribution of $\delta a_\mu$ resulting from this procedure  is marked 
768:   as the ``reduced'' curve in Fig.~\ref{fig:gm2}(b). It extends to somewhat
769:   higher and more central values of $\delta a_\mu$, showing that some of the
770:   skew in the $\delta a_\mu$ distribution does come from the LEP2 Higgs and
771:   $BR(b \rightarrow s \gamma)$ measurements. However, the effect is rather
772:   small and the resulting distribution is rather far from the experimental
773:   distribution, indicating a further effect. There could be a significant
774:   volume effect if regions of parameter space that have central values
775:   of   $\delta a_\mu$ have a small volume. This would render our results
776:   sensitive to the prior, since by changing the measure of the input
777:   parameters, we can change the volume measure~\cite{Allanach:2006jc}.
778: %
779: In order to investigate the effects of this, we re-weight the $\mu>0$ chains
780: by a factor $1/(m_0 M_{1/2})$. Such a re-weighting mimics the effect of using
781:   a logarithmic prior on $m_0$ and $M_{1/2}$ since
782: \begin{eqnarray}
783: \int P(m_0, M_{1/2})\  d \ln m_0\ d \ln M_{1/2} &=& \int P(m_0, M_{1/2}) \frac{d \ln m_0}{d m_0}
784: \frac{d \ln M_{1/2}}{d M_{1/2}} \ dm_0 \ dM_{1/2} \nonumber \\
785: & = & \int \left(\frac{P(m_0,
786:   M_{1/2})}{m_0 M_{1/2}} \right) \ dm_0 \ dM_{1/2}. \label{rewt}
787: \end{eqnarray}
788: One might consider such a prior on technical naturalness grounds. 
789: The ``log prior'' results are displayed in Fig.~\ref{fig:gm2}(b). They have a
790: fatter tail out to higher values of $\delta a_\mu$ than the flat prior sample
791: or the ``reduced'' sample. When we perform the re-weighting in Eq.~\ref{rewt}
792: as well as the one to remove $BR(b \rightarrow s \gamma)$ and LEP2 Higgs
793: constraints, we obtain the ``log prior, reduced'' curve in the figure.
794: This shows a yet fatter tail and is starting to approach the empirical
795: probability distribution imposed on the fits. 
796: Our results are obviously somewhat dependent upon the prior. This is
797: essentially because there is not enough precise data yet to constrain 
798: mSUGRA very strongly. We must bear in mind this dependence upon the prior, and 
799: investigate different priors when we estimate $P_-/P_+$ below. 
800: 
801: \TABULAR{|c|ccc|}{\hline
802: prior &  flat & small & log \\ \hline
803: $P_-/P_+$ & 0.16 & 0.12 & 0.07\\
804: \hline
805: }{Ratios of integrated probability for different signs of $\mu$.
806:   \label{tab:ratios}}
807: The total normalisation of the $\mu>0$, $\mu<0$ samples is 
808: shown in
809: Table~\ref{tab:ratios} for various different priors. 
810: The ``small'' prior is a flat prior with a reduced range compared to the one
811: displayed in Table~\ref{tab:inp}. We filter the chains to only include points
812: with $m_0<2$ TeV and $|A_0|<2$ TeV. The table shows that
813: $\mu<0$ is somewhat disfavoured for the smaller ranges and more disfavoured
814: still for the log prior. The result is somewhat sensitive to the prior,
815: indicating the need for more data. We therefore prefer to quote a range 
816: for $P_-/P_+=0.07-0.16$ depending upon the prior. This range is a focal result
817: of the present paper.
818: %\TABULAR{|c|ccc|}{\hline
819: %prior &  $P_-^{\tilde t}/P_+$ & $P_-^{\slash\!\!\!{\tilde t}}/P_+$ &
820: %$\mathbf{P_-/P_+}$ \\ \hline
821: %flat & .063 & .097 & {\bf 0.16} \\
822: %small  & .052 & 076 & {\bf 0.12} \\
823: %log  & .031 & .035 & {\bf 0.07} \\
824: %\hline
825: %}{Ratios of integrated probability for different signs of $\mu$
826: %  \label{tab:ratios}}
827: 
828: 
829: 
830: \section{Fits Without WMAP3 \label{sec:noDM}}
831: 
832: \FIGURE{\twographst{m0m12NDMp}{m0m12NDMm}
833: \caption{Constraints from global fits {\em without the dark matter constraint}
834:   \/with (a) $\mu>0$ and (b) $\mu<0$ mSUGRA 
835:   marginalised to the $m_0$-$M_{1/2}$ plane.
836: We have assumed a flat prior. The posterior probability is indicated by the bar
837:   on the right hand side. The inner (outer)  contours show the
838:   boundary of a $68\%$
839:   $(95\%)$ confidence region. \label{fig:noDM}
840: }
841: }
842: We now briefly examine the effect of removing the dark matter $\chi^2$ penalty
843: from the fits. We could in principle re-weight the chains in order to do this,
844: but we find that that leads to statistical fluctuations in the results that
845: are too large. Initial investigations revealed that the efficiency becomes
846: much higher when we remove the dark matter relic density contribution to the
847: total $\chi^2$. We are able to increase the widths of the proposal
848: distribution to 197 GeV for $m_0$, 100 GeV for $M_{1/2}$, 400 GeV for $A_0$,
849: 8.5 for $\tan 
850: \beta$ and 1$\sigma$ for the Standard Model inputs and {\em still} \/achieve an
851: efficiency of around 35$\%$. This has the consequence that the chains explore
852: the parameter space much quicker than in the previous section, and so less
853: MCMC steps are needed.
854: We run a further 9$\times$200~000 MCMC steps for each sign of $\mu$. This
855: time, both signs of $\mu$ have excellent convergence statistics, $\hat R$
856: being different to 1 at the per-mille level only in each case.
857: \FIGURE{\fourgraphs{higgsNDM}{gluinoNDM}{squarkNDM}{neutNDM}
858: \caption{Probability distributions in mass of (a) the lightest CP even higgs,
859:  (b) gluino, (c) the left-handed squark and (d) the neutralino. Flat priors
860:  have been assumed.
861: \label{fig:massNDM}}
862: }
863: 
864: \EPSFIGURE{omega,width=6.4cm}{$\Omega h^2$ distributions without imposing the
865: dark matter constraint. \label{fig:om}}
866: We see from Fig.~\ref{fig:noDM} that removing the dark matter constraint
867:   yields a very different picture for the mSUGRA probability distributions in
868:   the $m_0$-$M_{1/2}$ plane. This confirms our statement that many of the
869:   features seen in the previous section 
870:   are due precisely to that constraint. The probability distribution is now
871:   much flatter in the input parameter space.  
872: The disallowed region at low $m_0$ and high $M_{1/2}$ is due to the no-charged
873:   LSP constraint. The disallowed region at low $m_0$ and low $M_{1/2}$ is due
874:   mostly to the combined effect of the $(g-2)_\mu$, $BR(b\rightarrow s \gamma)$
875:   and $m_h$ constraints. The disallowed region is significantly larger for
876:   $\mu<0$ than for $\mu>0$, due to those three observables. Marginalisations
877:   in other mSUGRA input parameter planes tell a similar
878:   story, more featureless than the fits including the dark matter constraint. 
879: As mentioned in the introduction, Fig.~\ref{fig:noDM} covers the case of
880:   R-parity violating mSUGRA when the R-parity violating couplings are smaller
881:   than about
882:   0.1. For larger R-parity violating couplings, one would have to include
883:   them in the renormalisation group equations to obtain accurate results.
884: 
885: The more featureless fits have predictable effects on the mass spectrum of the
886: MSSM, as shown in Fig.~\ref{fig:massNDM}. We have fixed the ranges of the
887: abscissas to be identical to those in Fig.~\ref{fig:mass} in order to
888: facilitate comparison. The lightest CP-even Higgs probability distribution is
889: broader and shifted to heavier values, due to the bigger volume of parameter
890: space allowed at higher $m_0$ and $M_{1/2}$ values. Gluino mass distributions
891: no longer tail off at higher masses: upper bounds would be mostly determined
892: by the cut-off placed $m_0$ and $M_{1/2}$. The gluino mass and neutralino
893: distributions are 
894: also much less peaked than the ones in Figs.~\ref{fig:mass}b,d particularly for
895: $\mu<0$. The flatter distributions are, of course, an indication that the data
896: aren't strongly constraining. Large volumes at large $m_0,M_{1/2}$ effectively
897: move up the squark masses, as Fig.~\ref{fig:massNDM}c illustrates.
898: 
899: The probability distribution of $\Omega_{DM} h^2$ is shown in
900: Fig.~\ref{fig:om} for each sign of $\mu$, when we drop the relic dark matter
901: $\chi^2$ contribution. In the figure, we see that huge
902: values can result: in fact the mean value for the $\mu>0$ sample
903: tails off at $\Omega_{DM} h^2 \sim 128$, much
904: larger than the WMAP3 value of 0.1. 
905: \TABULAR{|c|ccc|}{\hline
906: prior& flat & small  & log  \\ \hline
907:  ${P_-/P_+}$ &
908:   {0.45} &
909:  {0.43} &
910:  {0.19} \\
911: \hline
912: }{Ratios of integrated probability for different signs of $\mu$ with no dark
913:  matter constraint.
914:   \label{tab:ratios2}}
915: This illustrates the fact that regions of parameter space which fit the dark
916: matter data in mSUGRA need some special annihilation mechanism that is not
917: typical of the whole space. This and similar arguments have led several
918: authors to consider non-universal models~\cite{gian,sfk}, where the relic
919: density might be less fine-tuned. The distributions are highly skewed, having
920: tiny tails up to $\Omega h^2 \sim 1000$. 
921: The relative normalisation of $\mu>0$ to $\mu<0$ is calculated in the same
922: manner as in the previous section for different priors, and displayed in
923: Table~\ref{tab:ratios2}. 
924: $\mu<0$ is hardly disfavoured in R-parity violating mSUGRA (where we can
925: neglect the dark matter constraint), where
926: $P_-/P_+=0.19-0.45$.
927: 
928: 
929: \section{Conclusions \label{sec:conc}}
930: 
931: We have performed global fits to mSUGRA using indirect data and
932: state-of-the-art predictions of the observables. The MCMC technique was
933: successfully employed despite initial non-convergence of the $\mu<0$ chains.
934: Bridge sampling was used to normalise two isolated maxima that had not been
935: traversed by any chain for $\mu<0$. 
936: We found that $\mu<0$ is somewhat disfavoured in comparison to $\mu>0$ but not
937: by huge margins. The rest of the fit prefers rather heavy SUSY and so the SUSY
938: contribution to $(g-2)_\mu$ is small whichever the sign of $\mu$. 
939: $\mu<0$ is only disfavoured marginally, 
940: the ratio of integrated probability densities being
941: $P_-/P_+=0.07-0.16$ depending upon the prior.
942: We see from Fig.~\ref{fig:om} that without
943: the dark matter constraint, $\Omega_{DM} h^2$ is predicted up to values of
944: around 128. This corresponds to a $\chi^2_{\Omega}$ of around 3$\times 10^8$,
945: much larger than is likely from the other observables such as $M_W$ or 
946: $\sin^2 \theta_w^l(\mbox{eff})$.  
947: The fits are therefore completely dominated by the
948: dark matter relic density constraint and volume effects. 
949: Expectations that the SUSY scale will be light because of a preference from
950: weak observables turn out not to be true in the global fits. 
951: If the dark matter constraint is dropped, as would be the
952: case for R-parity violation, $\mu<0$ is hardly disfavoured at all, 
953: $P_-/P_+=0.19-0.45$. $\mu<0$ is much less disfavoured than many seem to
954: assume.  
955: Many recent analyses only consider $\mu>0$ on the grounds that $\mu<0$ is
956: strongly 
957: disfavoured by $(g-2)_\mu$. We have therefore demonstrated that this is {\em
958:   not} \/the case when one considers the entirety of the data and that $\mu<0$ 
959: should still  be considered in mSUGRA analyses.
960: 
961: It could be argued that the flat measure used here in $m_0$, $M_{1/2}$, $A_0$
962: and $\tan \beta$ can be improved upon. For instance, $\tan \beta$ is really
963: a derived quantity and is related to more fundamental Higgs potential
964: parameters, which could be considered more natural to have a flat measure
965: upon. There is also the issue of fine-tuning, recently illuminated in
966: Ref.~\cite{Giudice:2006sn}: we could disfavour regions of parameter space that
967: are highly fine-tuned, for instance~\cite{Allanach:2006jc}. Changes such as
968: these in the prior could potentially change the results of the fits and we
969: intend to investigate them in a future publication. 
970: 
971: Clearly, more data is required to 
972: decrease the dependence of results on the prior.
973: The most helpful data is likely to be that from colliders. 
974: The MCMC fitting technique has been used in an ATLAS study examining how
975: cross-section and kinematic endpoint information constrains
976: mSUGRA and non-universal models~\cite{Lester:2005je}.
977: At this moment, without data from colliders, we are forced to use indirect
978: constraints for the observables. However, in the future it will be desirable
979: to predict $\Omega_{DM} h^2$ given some SUSY collider
980: observables~\cite{Allanach:2004xn,Baltz:2006fm}. If this is in contradiction
981: with the observed value from cosmology, it will point to a wrong cosmological
982: assumption, which could then be changed. In order to really confirm that 
983: dark matter particles have been produced at colliders, one requires
984: compatibility with direct dark matter detection data. 
985: Of course one would like to drop the mSUGRA assumption and perform a general
986: SUSY analysis, but for this it is likely that additional data from a future
987: international linear collider would be
988: required~\cite{Allanach:2004xn,Baltz:2006fm}.  
989: In any case, the techniques investigated in this paper should prove useful for
990: the fits.
991: 
992: \acknowledgments
993: This work has been partially supported by PPARC\@. This paper was produced
994: using the University of Cambridge EScience CAMGRID computing facility.
995: We would like to thank W Hollik for help with electroweak observables, 
996: B Heinemann and C S Lin for the likelihood density penalty of $B_s \rightarrow
997: \mu^+ \mu^-$, R Rattazzi for a discussion about priors,
998: D St\"ockinger for the two-loop contributions to the anomalous magnetic
999: moment of the muon, A Pukhov for help with {\tt micrOMEGAs},
1000: J Ellis, K Olive, T Plehn, L Roszkowski, J Smillie, G Weiglein and the
1001: Cambridge SUSY 
1002: working group for helpful comments and M Calleja for invaluable help with using
1003: CAMGRID. 
1004: 
1005: \appendix
1006: 
1007: \section{Markov Chain Monte Carlos\label{sec:bridge}}
1008: 
1009: Our Markov chain consists of a list of parameter points (${\mathbf
1010: x}^{(t)}$) and associated likelihood densities (${\mathcal L}({\mathbf x}^{(t)})$). Here, $t$ labels the link
1011: number in the chain.  Given some point at the end of the Markov chain
1012: (${\mathbf x}^{(t)}$), the Metropolis-Hastings
1013: algorithm~\cite{Metropolis,Hastings,MacKay} requires one to randomly
1014: pick another potential point (${\mathbf x}$) (typically in the
1015: vicinity of ${\mathbf x}^{(t)}$) using a proposal distribution
1016: $Q({\mathbf x};{\mathbf x}^{(t)})$. There is a large amount of freedom
1017: in the choice of the proposal function $Q$, and this freedom is
1018: usually exploited to improve the efficiency of the sampling process.
1019: In order to ensure that the choice of $Q$ does not bias the final set
1020: of samples in some way, the form of $Q$ is taken into account when
1021: deciding whether to accept or reject the new point.  
1022: If the ratio $\rho$
1023: defined by
1024: \begin{equation}
1025: \rho = 
1026: \frac{{\mathcal L}({\bf x})}{{\mathcal L}{({\bf x}^{(t)} )}}
1027: \frac{Q({\mathbf x}^{(t)};{\mathbf x})}{Q({\mathbf x};{\mathbf x}^{(t)})}
1028: \end{equation} is greater
1029: than one, the new point ${\bf x}$ is appended to the chain.  If $\rho$
1030: is instead less than one, a decision must be made to determine whether
1031: to accept or reject the proposed point ${\bf x}$.  The rule is that
1032: acceptance of ${\bf x}$ must occur with probability $\rho$.  If
1033: accepted, ${\bf x}$ is added to the end of the chain. If not accepted,
1034: the point ${\mathbf x}^{(t)}$ is copied once more onto the end of the
1035: chain.  Whichever point makes it on to the end of the chain is
1036: thereafter known as ${\mathbf x}^{(t+1)}$.
1037: 
1038: As a result of following the above steps, the sampling density of
1039: points in the chain becomes proportional to the density of the target
1040: distribution (such as the posterior probability density, or the
1041: likelihood when the prior is uniform) as the number of links goes to
1042: infinity, under the circumstances described in Ref.~\cite{MacKay}.  The
1043: Metropolis-Hastings MCMC algorithm is typically much more efficient
1044: than a straightforward scan for the dimensionality of input parameter
1045: space $D>3$; the number of required steps scales roughly linearly with
1046: $D$ rather than as a power law.  We take the proposal function $Q$ to
1047: be a product of Gaussian distributions along each dimension
1048: $k=1,2,\ldots,D$ centred on the location of the current point along
1049: that dimension, i.e.\ $x^{(t)}_k$:
1050: \begin{equation}
1051: Q({\mathbf x}; {\mathbf x}^{(t)}) = 
1052:   \prod_{k=1}^D \frac{1}{\sqrt{2 \pi} l_k} e^{-(x_k -x_k^{(t)} )^2
1053:   / 2 l_k^2},
1054: \end{equation}
1055: where $l_k$ denotes the width of the distribution along direction $k$. 
1056: For the case where we include the dark matter relic density in the
1057: calculation, 
1058: we choose $l_{m_0}=100$ GeV, $l_{M_{1/2}}=50$ GeV, $l_{A_0}=400$ GeV and
1059:   $l_{\tan \beta}=3$. 
1060: For the Standard Model inputs, we choose $l_k = 8 \sigma_k / 20$.  We
1061: discuss why these particular values were chosen in the next section.
1062: 
1063: In order to start the chain we follow the following procedure, which finds a
1064: point at random in parameter space that is not a terrible fit to the data.
1065: We pick some ${\mathbf y}^{(0)}$ at random in the mSUGRA parameter
1066: space using a flat distribution for its probability density function (pdf). 
1067: The Markov chain for ${\mathbf y}$ is evolved through 2000 steps.
1068: We then set ${\mathbf x}^{(0)}={\mathbf y}^{(2000)}$,
1069: continuing the Markov chain in ${\mathbf x}$ and discarding the ``burn-in''
1070: chain ${\mathbf y}$. A reasonable-fit point is typically found long before
1071: 2000 iterations of the Markov chain.
1072: We must make sure that we perform enough iterations after this point 
1073: that the chain traverses the remaining viable parameter space. 
1074: We will provide a convergence test to this effect.
1075: 
1076: \subsection{Efficiency}
1077: 
1078: The efficiency of a chain can be defined as ``the
1079: number of links whose coordinates differ from those of their
1080: predecessor in the chain'' divided by ``the total number of points in
1081: the chain''.  There will always be a tension between efficiency and
1082: convergence\footnote{For a discussion of convergence, see
1083: appendix~\ref{sec:conv}.} in chains:
1084: if the $l_k$ are set to be too
1085: small, efficiency will increase but the chain will take too long to
1086: achieve convergence whereas if they are too large, the efficiency will
1087: be so small that the sampling will contain large statistical
1088: fluctuations. 
1089: In practice, the bulk of our posterior
1090: probability density is contained in a very thin hyper-surface in the
1091: 8-dimensional input parameter space~\cite{Allanach:2005kz}. It is thin
1092: because $\Omega_{DM} h^2$ varies very rapidly over mSUGRA space
1093: compared to the high accuracy of the empirical constraint. With the
1094: $l_k$ listed above, we found that efficiencies were at the per-mille
1095: level, too small to achieve a statistically stable result in a
1096: reasonable amount of CPU time.  To achieve a significantly larger
1097: efficiency, we had to reduce $l_k$ to such a level that we lost
1098: convergence because the chains had not traversed the viable parameter
1099: space.  In order to counter this, we expanded the errors on
1100: $\Omega_{DM} h^2$ to $\pm 0.02$ while calculating the likelihood
1101: density in the chain. This artificially thickens the ``surface''
1102: containing the bulk of the posterior probability density, increasing
1103: the efficiency to much more reasonable values of around 5--7$\%$.  In
1104: order to correct for this artificial thickening and re-impose the
1105: required constraint of Eq.~\ref{omega}, it was therefore necessary to
1106: re-weight each link of each chain at the end of the sampling.  Each
1107: link is re-weighted by the ratio of the proper likelihood density
1108: ${\mathcal L}_P$ to the likelihood density with inflated errors
1109: ${\mathcal L}_I$:
1110: \begin{equation}
1111: \frac{{\mathcal L}_P}{{\mathcal L}_I} =
1112:         \exp \left(
1113: 	  {-\frac{(c_{\Omega}-p_{\Omega})^2}{2\sigma^2_{\Omega
1114:         }}} \right) \div
1115: 	\exp \left(
1116: -\frac{(c_\Omega - p_\Omega)^2}{2 \times 0.02^2}
1117: \right) 
1118: \label{reweight}
1119: \end{equation} 
1120: in order to
1121: impose the correct penalty on the links. We ignore additional
1122: constants that are independent of $\Omega_{DM} h^2$ in this expression
1123: since the overall normalisation of the likelihood density is here
1124: undetermined.  The re-weighting procedure necessarily degrades the
1125: statistical spread of the results, however we find that the increase
1126: in efficiency more than compensates for this effect. Below, we
1127: re-weight different variables in order to investigate various features
1128: in the results, but the method remains analogous to the one described
1129: here.
1130: 
1131: \subsection{Bridge Sampling}
1132: 
1133: One of the numbers we will require from our MCMC samples is the ratio
1134: of integrated posterior probabilities of $\mu>0$ ($P_+$) and $\mu<0$
1135: ($P_-$). This ratio will tell us the extent to which $\mu<0$ is
1136: disfavoured over $\mu>0$.  Assuming a flat prior in the variables of
1137: the model, the posterior probability is equal to the integrated
1138: likelihood divided by a factor which does not depend upon model
1139: hypotheses or parameters. Thus $P_-/P_+=\int d{\bf x} {\mathcal
1140: L}_-({\bf x}) / \int d{\bf x}{\mathcal L}_+({\bf x})$, where
1141: ${\mathcal L}_{+,-}({\bf x})$ is the likelihood density of $\mu>0$
1142: $(<0)$ mSUGRA respectively at parameter point $({\bf x})$.  One way to
1143: estimate this ratio would be to include the sign of $\mu$ as a free
1144: parameter in the Metropolis-Hastings procedure, to be chosen randomly
1145: in a proposal point. This algorithm leads to large inefficiencies
1146: because the $\mu>0$ and $\mu<0$ likelihood surfaces have a limited
1147: overlap, meaning that too many proposals for an opposite sign of $\mu$
1148: will be rejected. Also, the procedure would likely provide large
1149: statistical fluctuations for the disfavoured sample, which we expect
1150: to be the $\mu<0$ one. Even though it is disfavoured, we should like
1151: to investigate its properties.
1152: 
1153: A simple way one might hope to evaluate the ratio ${P_-}/{P_+}$ is
1154: \begin{equation}
1155: \frac{P_-}{P_+} = \frac{1}{E_- \left[ \frac{{\mathcal L}_+}{{\mathcal L}_-}
1156:     \right]} 
1157: \approx \frac{1}{N} \sum_{t=1}^N \frac{{\mathcal L}_-
1158: ({\mathbf x_i^{(t)}})}{{\mathcal L}_+ ({\mathbf x_i^{(t)}})},
1159: \end{equation}
1160: where $E_-$ denotes the expectation with respect to the $\mu<0$
1161: likelihood distribution \cite{radford}. $N$ denotes the number of MCMC
1162: steps. 
1163: Unfortunately, a simple importance sampling estimate of this kind does
1164: not work if there are any valid points (${\mathcal L} \ne 0$) for one
1165: sign of $\mu$ that are invalid (${\mathcal L} = 0$) for the opposite
1166: sign of $\mu$.  In our case there are plenty of these dangerous
1167: pairings, as sparticle mass or tachyonic bounds move around in
1168: parameter space depending upon the sign of $\mu$.  To get around this
1169: problem, we use a solution known as bridge sampling~\cite{radford}
1170: with a ``geometric bridge''.  This allows us to generate a (biased)
1171: estimator $r$ for the ratio ${P_-}/{P_+}$ as long as there is some viable
1172: region of $\mu>0$ parameter space that is also viable for $\mu<0$.
1173: The estimator for the ratio is constructed as follows:
1174: \begin{equation}
1175:  \frac{P_-}{P_+}  =
1176: \frac{E_+ \left[ \sqrt{\frac{{\mathcal L}_-}{{\mathcal L}_+}}
1177:   \right]}
1178: {E_- \left[ \sqrt{\frac{{\mathcal L}_+}{{\mathcal L}_-}}
1179:   \right]} \approx r \equiv
1180: \frac{\sum_{t=1}^N \sqrt{\frac{{\mathcal L}_- ({\mathbf x}_+^{(t)})
1181: }{{\mathcal L}_+ ({\mathbf x}_+^{(t)})}}}{\sum_{t=1}^N \sqrt{\frac{{\mathcal
1182: 	L}_+ ({\mathbf x}_-^{(t)}) 
1183: }{{\mathcal L}_- ({\mathbf x}_-^{(t)})}}} , \label{bridge}
1184: \end{equation}
1185: where $(\mathbf x)^{(t)}_{+,-}$ are the parameter points of the links
1186: in the $\mu>0$ or $\mu<0$ chains respectively.  Here, we have assumed
1187: an equal number of links in each chain.  In summary, to calculate $r$,
1188: we must run two chains, one for positive $\mu$ and one for negative
1189: $\mu$, and for every link, record the likelihood one would have
1190: obtained for identical input parameters except for the opposite sign
1191: of $\mu$. 
1192: 
1193: \section{Convergence and Normalisation\label{sec:conv}}
1194: 
1195: \EPSFIGURE{conv,width=7cm}{Convergence statistics for the
1196:   MCMCs. 
1197: \label{fig:conv}}In order to evaluate the convergence of the MCMC chains, we always run 9
1198: independent chains with different random starting points. By comparing the 
1199: similarity of the resulting sampling densities of input parameters in
1200: the chains, one can
1201: construct~\cite{gelman} a measure of convergence $\hat R$. $\hat R$ is an upper
1202: bound on the reduction in variance of parameters that would result from
1203: running the chains for an infinite number of steps. The precise implementation
1204: is  listed in Ref.~\cite{Allanach:2005kz}. Values close to 1 indicate
1205: convergence of the chains.
1206: 
1207: \FIGURE{\twographst{stopm0A0}{noStopm0A0}
1208: \caption{The two types of negative $\mu$ samples: (a) the ``$\mu<0$, 2
1209:   chains'' samples (later shown to be co-annihilation samples), and (b)
1210:   the ``$\mu<0$, 7 chains'' samples (later shown to be
1211:   non co-annihilation samples, dominated by resonant Higgs annihilation
1212:   regions).  The posterior probability is indicated by the bar on the
1213:   right hand side. The inner (outer) contours show the boundary of a
1214:   $68\%$ $(95\%)$ confidence region.\label{fig:stop}}}
1215: We run 9 chains of 500~000 points for $\mu>0$ mSUGRA and for the
1216: $\mu<0$ dark side of mSUGRA\@.  The $\mu>0$ curve in
1217: Fig.~\ref{fig:conv} shows good convergence is achieved by 500 000 MCMC
1218: steps. However, the $\mu<0$ curve shows a problem: convergence is
1219: never achieved. This is a serious difficulty as one could not draw any
1220: quantitative statistical inferences from the non-converged chains.
1221: Further inspection of the $\mu<0$ results shows that two of the
1222: $\mu<0$ chains are in a completely different part of parameter space
1223: than the other seven.  This indicates isolated maxima of likelihood
1224: density which the MCMC has not been able to jump between in the finite
1225: number of MCMC steps attempted\footnote{A proposal distribution with
1226: longer tails, such as an $n-$dimensional Cauchy distribution, would
1227: have more chance of making such a jump.}.  There is no balance between
1228: the two isolated maxima in the sample.  Isolating the two anomalous
1229: $\mu<0$ chains and calculating $\hat R$ between just them, we obtain
1230: the ``$\mu<0$, 2 chains'' curve, which closely approaches 1 by 500~000
1231: MCMC steps. Thus within this isolated maximum, convergence is
1232: achieved.  The same can be said of the other ``$\mu<0$, 7 chains''
1233: samples: they also converge amongst themselves. Thus, the shapes of
1234: each isolated maximum are well determined, but the relative
1235: normalisation of the two different types of negative $\mu$ samples is
1236: not. 
1237: 
1238: \FIGURE{\onegraph{sugraOppmchimstop}
1239: \caption{Probability density in the lightest stop-lightest neutralino mass
1240:   plane for the 2-chain sample. The posterior probability is indicated
1241:   by the bar on the right hand side. The inner (outer) contours show
1242:   the boundary of a $68\%$ $(95\%)$ confidence
1243:   region. \label{fig:stop2}} }
1244: 
1245: In order to illustrate the two maxima, we marginalise the two types of
1246: negative $\mu$ samples onto the $m_0-A_0$ plane in
1247: Figs.~\ref{fig:stop}(a) and~\ref{fig:stop}(b).
1248:   The maxima are isolated in this
1249: plane (as well as in some other 2-parameter planes).  
1250: %As with all
1251: %2-dimensional marginalised plots in this paper, we bin the plane into
1252: %75$\times$75 2-dimensional bins. The colour bar on the right hand side
1253: %of the figures shows the posterior probability $P$ of each bin divided
1254: %by the maximum posterior probability of any bin in the plot. 
1255: The two
1256: regions are completely separated. Their shape is primarily determined
1257: by regions which efficiently deplete the relic density of neutralinos
1258: which, in mSUGRA, is often higher than the WMAP3 constraint.  We
1259: investigate the 2-chain sample in Fig.~\ref{fig:stop2}. 
1260: In the figure,
1261: there are two good-fit regions: where the stau
1262: co-annihilates~\cite{Griest:1990kh} (at moderate values of $A_0$, higher
1263: values of $m_{\tilde t}$ in the figure) with
1264: the LSP ${\tilde \tau}_1 \chi_1^0 \rightarrow \tau \gamma$ and where
1265: the lightest stop co-annihilates (at $A_0<-3$ TeV) with the LSP
1266: ${\tilde t} \chi_1^0 \rightarrow t g$ in the early universe, where
1267: $m_{\chi_1^0} \approx m_{{\tilde t}_1}$, the lower
1268: strip in the figure. There was no significant
1269: stop co-annihilation~\cite{Boehm:9911,arnie,Ellis:2001nx} region for
1270: $\mu>0$.  On the other hand, Fig.~\ref{fig:stop}(b) dominantly
1271: consists of resonant Higgs annihilation
1272: regions~\cite{Drees:1992am,Arnowitt:1993mg,Djouadi:2005dz}, where
1273: $\chi_1^0\chi_1^0 \rightarrow h,A^0 \rightarrow b {\bar b}/\tau^+
1274: \tau^-$ and the focus point region where the LSP has a significant
1275: higgsino component and $\chi_1^0\chi_1^0 \rightarrow ZZ, WW, t
1276: \bar{t}$~\cite{Feng:1999mn,Feng:1999zg,Feng:2000gh}.
1277: 
1278: We need a method to determine the relative normalisation of the 
1279: 2-chain co-annihilation sample and the 7-chain resonant higgs
1280: annihilation sample.  Equivalently we need a method to determine the
1281: ratio of the posterior probability ${\tilde P}_-^{\tilde t}$ of the 2-chain
1282: co-annihilation sample and the posterior probability
1283: ${\tilde P}_-^{\slash\!\!\!{\tilde t}}$ of the 7-chain resonant higgs
1284: annihilation sample. We use ${\tilde P}$ to denote the fact that the
1285: posterior probabilities are un-normalised.
1286: Fortunately, Eq.~\ref{bridge} provides us with a
1287: solution: we first determine the normalisation of the $\mu>0$ sample
1288: with respect to each separate $\mu<0$ sample, i.e.\ ${\tilde P}_-^{\tilde
1289: t}/P_+$ and ${\tilde P}_-^{\slash\!\!\!{\tilde t}}/P_+$.  Since these
1290: quantities individually have good convergence properties, their ratio is also
1291: well determined:
1292: \begin{equation}
1293: \frac{{\tilde P}_-^{\slash\!\!\!{\tilde t}}}{{\tilde P}_-^{\tilde t}} = 
1294:  \frac{{\tilde P}_-^{\slash\!\!\!{\tilde t}}}{P_+} \div \frac{{\tilde
1295:  P}_-^{\tilde  t}}{P_+} 
1296: =0.097 \div 0.063 = 1.53. \label{RofR}
1297: \end{equation}
1298: Normalising the probabilities as $a{\tilde P}_-^{{\tilde t}} \equiv
1299: P_-^{{\tilde t}}$, $b{\tilde P}_-^{\slash\!\!\!{\tilde t}} \equiv
1300: P_-^{\slash\!\!\!{\tilde t}}$, we fix $a$ and $b$ by imposing ${
1301:   P}_-^{{\tilde t}} + {P}_-^{\slash\!\!\!{\tilde t}} = 1$ and
1302: Eq.~\ref{RofR}. 
1303: The relative posterior probabilities ratios are re-calculated
1304: whenever alternative priors are investigated.
1305: In section~\ref{sec:dark} where we present the total $\mu<0$
1306: sample results, we present posterior probability densities with the correct
1307: normalisation, according to this prescription.
1308: The ratio of the probability of $\mu<0$ to $\mu>0$ is then determined simply
1309: by: 
1310: \begin{equation}
1311:   \frac{P_-}{P_+} = \frac{P_-^{\tilde
1312:  t}}{P_+} + \frac{P_-^{\slash\!\!\!{\tilde t}}}{P_+}. 
1313: \end{equation} 
1314: 
1315: It is worth noting that, had we been unlucky, we might have obtained
1316: only chains like those in the 7-chain sample. In that case we would
1317: have carried on with the analysis without realising about the
1318: different 2-chain sample, therefore any results achieved would have
1319: been incomplete. An obvious question is: were there any other local
1320: maxima that we have missed by not running enough chains?
1321: Unfortunately, any fitting procedure is susceptible to this caveat and
1322: there is no satisfactory answer. Finding a high but very narrow global
1323: maximum is an unsolved problem in any number of dimensions.
1324:  
1325: \begin{thebibliography}{99}
1326: 
1327: \bibitem{Ellis:2003si}
1328:   J.~R.~Ellis, K.~A.~Olive, Y.~Santoso and V.~C.~Spanos,
1329:   {\em Likelihood analysis of the CMSSM parameter space},
1330:   Phys.\ Rev.\ D {\bf 69} (2004) 095004
1331:   [arXiv:hep-ph/0310356].
1332:   %%CITATION = HEP-PH 0310356;%%
1333: \bibitem{Profumo:2004at}
1334:   S.~Profumo and C.~E.~Yaguna,
1335:    {\em A statistical analysis of supersymmetric dark matter in the MSSM after
1336:   WMAP},
1337:   Phys.\ Rev.\ D {\bf 70} (2004) 095004
1338:   [arXiv:hep-ph/0407036].
1339:   %%CITATION = HEP-PH 0407036;%%
1340: \bibitem{Baltz:2004aw}
1341:   E.~A.~Baltz and P.~Gondolo,
1342:   {\em Markov chain Monte Carlo exploration of minimal supergravity with
1343:   implications for dark matter},
1344:   JHEP {\bf 0410} (2004) 052
1345:   [arXiv:hep-ph/0407039].
1346:   %%CITATION = HEP-PH 0407039;%
1347: \bibitem{Ellis:2004tc}
1348:   J.~R.~Ellis, S.~Heinemeyer, K.~A.~Olive and G.~Weiglein,
1349:   {\em Indirect sensitivities to the scale of supersymmetry},
1350:   JHEP {\bf 0502} (2005) 013
1351:   [arXiv:hep-ph/0411216].
1352:   %%CITATION = HEP-PH 0411216;%%
1353: \bibitem{Stark:2005mp}
1354:   L.~S.~Stark, P.~Hafliger, A.~Biland and F.~Pauss,
1355:  {\em New allowed mSUGRA parameter space from variations of the trilinear
1356:    scalar
1357:   coupling A0},
1358:   JHEP {\bf 0508} (2005) 059
1359:   [arXiv:hep-ph/0502197].
1360:   %%CITATION = HEP-PH 0502197;%%
1361: \bibitem{Robros}
1362:  R.~G.~Roberts and L.~Roszkowski,
1363:   {\em Implications for minimal supersymmetry from grand unification and the
1364:   neutrino relic abundance},
1365:   Phys.\ Lett.\ B {\bf 309} (1993) 329
1366:   [arXiv:hep-ph/9301267].
1367:   %%CITATION = HEP-PH 9301267;%%
1368: 
1369: %\cite{Allanach:2005kz}
1370: \bibitem{Allanach:2005kz}
1371:   B.~C.~Allanach and C.~G.~Lester,
1372:   {\em Multi-dimensional mSUGRA likelihood maps},
1373:   Phys.\ Rev.\ D {\bf 73} (2006) 015013
1374:   [arXiv:hep-ph/0507283].
1375:   %%CITATION = HEP-PH 0507283;%%
1376: 
1377: %\cite{Allanach:2006jc}
1378: \bibitem{Allanach:2006jc}
1379:   B.~C.~Allanach,
1380:    {\em Naturalness priors and fits to the constrained minimal supersymmetric
1381:   standard model},
1382:   Phys.\ Lett.\ B {\bf 635} (2006) 123
1383:   [arXiv:hep-ph/0601089].
1384:   %%CITATION = HEP-PH 0601089;%%
1385: \bibitem{deAustri:2006pe}
1386:   R.~R.~de Austri, R.~Trotta and L.~Roszkowski,
1387:   {\em A Markov chain Monte Carlo analysis of the CMSSM},
1388:   JHEP {\bf 0605} (2006) 002
1389:   [arXiv:hep-ph/0602028].
1390:   %%CITATION = HEP-PH 0602028;%%
1391: %cite{Heinemeyer:2006px}
1392: \bibitem{Heinemeyer:2006px}
1393:   The code is forthcoming in a publication by A.~M.~Weber et al.;
1394:   S.~Heinemeyer, W.~Hollik, D.~St\"ockinger, A.~M.~Weber and G.~Weiglein,
1395:   {\em Precise prediction for M(W) in the MSSM},
1396:   arXiv:hep-ph/0604147.
1397: \bibitem{2loop}
1398: S. Heinemeyer, D.~St\"ockinger and G.~Weiglein, 
1399: {\em  Electroweak and supersymmetric two-loop corrections to (g-2)(mu)},
1400: Nucl. Phys. {\bf B690} (2004) 103, [arXiv:hep-ph/0405255];
1401: S.~Heinemeyer, D.~St\"ockinger and G.~Weiglein,
1402:   {\em Two-loop SUSY corrections to the anomalous magnetic moment of the muon},
1403:   Nucl.\ Phys.\ B {\bf 690} (2004) 62
1404:   [arXiv:hep-ph/0312264].
1405:   %%CITATION = HEP-PH 0312264;%%
1406: \bibitem{radford}
1407: C.H.~Bennett, 
1408: {\em Efficient estimation of free energy differences from Monte Carlo data}, 
1409: Jnl.\ of Comp. Phys. {\bf 22} (1976) 245;
1410: A.~Gelman and X.-Li~Meng, 
1411: {\em Simulating normalizing constants: from importance sampling to bridge
1412:   sampling to path sampling}, Stat. Sci. {\bf 13} (1998) 163;
1413: R.~M.~Neal {\em Estimating ratios of normalizing constants using Linked
1414:   Importance Sampling}, Technical Report No. 0511 (2005), Dept.\ of Statistics,
1415: University of Toronto 
1416: \bibitem{rpvSUGRA}
1417: B..~C.~Allanach, A.~Dedes and H.~K.~Dreiner,
1418:   {\em The R parity violating minimal supergravity model},
1419:   Phys.\ Rev.\ D {\bf 69}, 115002 (2004)
1420:   [Erratum-ibid.\ D {\bf 72}, 079902 (2005)]
1421:   [arXiv:hep-ph/0309196];
1422:   %%CITATION = HEP-PH 0309196;%%
1423:   B.~C.~Allanach, M.~A.~Bernhardt, H.~K.~Dreiner, C.~H.~Kom and P.~Richardson,
1424:   {\em Mass Spectrum in R-Parity Violating mSUGRA and Benchmark Points},
1425:   arXiv:hep-ph/0609263.
1426:   %%CITATION = HEP-PH 0609263;%%
1427: \bibitem{PDG}
1428: W.-M.~Yao {\em et al}, {\em The review of particle physics}, Jnl.\ Phys. {\bf
1429:   G33} (2006) 1, http://pdg.lbl.gov/
1430: \bibitem{mtmeas}
1431: The Tevatron Electroweak Working Group, 
1432: {\em Combination of CDF and D0 results on the mass of the top quark},
1433: [arXiv:hep-ex/0608032].
1434: \bibitem{softsusy}
1435: B.C. Allanach, {\em SOFTSUSY: A program for calculating supersymmetric
1436:   spectra}, Comput. Phys. Commun. {\bf 143} (2002) 305, [arXiv:hep-ph/0104145].
1437: \bibitem{slha}
1438: P. Skands {\em et al}, {\em SUSY Les Houches accord: Interfacing SUSY spectrum
1439:   calculators, decay packages, and event generators}, JHEP {\bf 0407} (2004)
1440: 036, [arXiv:hep-ph/0311123].
1441: \bibitem{micromegas}
1442:  G. B\'{e}langer, F. Boudjema, A. Pukhov and A. Semenov,
1443: {\em micrOMEGAs: Version 1.3},
1444:  Comput. Phys. Commun. {\bf 174} (2006) 577 [arXiv:hep-ph/0405253];
1445: G. B\'{e}langer, F. Boudjema, A. Pukhov and A. Semenov,
1446: {\em micrOMEGAs: A program for calculating the relic density in the MSSM},
1447:  Comput. Phys. Commun. {\bf 149} (2002) 103 [arXiv:hep-ph/0112278].
1448: \bibitem{susycont}
1449:  U.~Chattopadhyay and P.~Nath,
1450:   {\em Probing supergravity grand unification in the Brookhaven g-2 experiment},
1451:   Phys.\ Rev.\ D {\bf 53} (1996) 1648,
1452:   [arXiv:hep-ph/9507386].
1453:   %%CITATION = HEP-PH 9507386;%%
1454: \bibitem{private}
1455: D.~St\"{o}ckinger,
1456: {\em  The muon magnetic moment and supersymmetry},
1457: arXiv:hep-ph/0609168.
1458: \bibitem{private2}
1459: We thank C S Lin for providing us with the likelihoods. 
1460: \bibitem{bsmumu}
1461: CDF Collaboration, CDF Public Note 8176,
1462: {\em Search for the Rare Decays $B_{s(d)}$}, CDF Public Note 8176,
1463: http://www-cdf.fnal.gov/physics/new/bottom/060316.blessed-bsmumu3/
1464: \bibitem{Ellis:2005sc}
1465:  A.~Dedes and B.~T.~Huffman,
1466:    {\em Bounding the MSSM Higgs sector from above with the Tevatron's $B/s
1467:  \rightarrow 
1468:  \mu^+ \mu^-$},
1469:   Phys.\ Lett.\ B {\bf 600}, 261 (2004)
1470:   [arXiv:hep-ph/0407285];
1471:   %%CITATION = HEP-PH 0407285;%%
1472:   J.~R.~Ellis, K.~A.~Olive and V.~C.~Spanos,
1473:   {\em On the interpretation of $B/s \rightarrow \mu^+ \mu^-$ in the CMSSM},
1474:   Phys.\ Lett.\ B {\bf 624} (2005) 47
1475:   [arXiv:hep-ph/0504196].
1476:   %%CITATION = HEP-PH 0504196;%%
1477: \bibitem{Ellis:2006ix}
1478:   J.~R.~Ellis, S.~Heinemeyer, K.~A.~Olive and G.~Weiglein,
1479:   {\em Phenomenological indications of the scale of supersymmetry},
1480:   JHEP {\bf 0605} (2006) 005
1481:   [arXiv:hep-ph/0602220].
1482:   %%CITATION = HEP-PH 0602220;%%
1483:  \bibitem{MWSM} M.~Awramik, M.~Czakon, A.~Freitas and G.~Weiglein,  
1484:                 {\em Phys.\ Rev.} {\bf D 69} (2004) 053006,   
1485:                 hep-ph/0311148.
1486:                 %%CITATION = HEP-PH 0311148;%%
1487:  \bibitem{Awramik:2004ge} M.~Awramik, M.~Czakon, A.~Freitas and
1488:                 G.~Weiglein,
1489:                 {\em Phys.\ Rev.\ Lett.} {\bf 93} (2004) 201805,
1490:                 hep-ph/0407317.
1491:                 %%CITATION = HEP-PH 0407317;%%
1492: 
1493: \bibitem{mw}
1494: See the July 2006 numbers from http://lepewwg.web.cern.ch/LEPEWWG/
1495: \bibitem{sinth}
1496: S. Heinemeyer and G. Weiglein, {\em The MSSM in the Light of Precision Data},
1497: arXiv:hep-ph/0307177
1498: \bibitem{drMSSMal2B} J.~Haestier, S.~Heinemeyer, D.~St\"ockinger and
1499:                      G.~Weiglein,
1500:                      {\em JHEP} {\bf 0512} (2005) 027,
1501:                      hep-ph/0508139;
1502:                      %%CITATION = HEP-PH 0508139;%%
1503:                      hep-ph/0506259.
1504:                      %%CITATION = HEP-PH 0506259;%%
1505: \bibitem{lep2higgs}
1506:   R.~Barate {\it et al.}  [LEP Working Group for Higgs boson searches],
1507:   {\em Search for the standard model Higgs boson at LEP},
1508:   Phys.\ Lett.\ B {\bf 565} (2003) 61
1509:   [arXiv:hep-ex/0306033].
1510:   %%CITATION = HEP-EX 0306033;%%
1511: \bibitem{decouple}
1512:   A.~Dobado, M.~J.~Herrero and S.~Penaranda,
1513:   {\em The Higgs sector of the MSSM in the decoupling limit},
1514:   Eur.\ Phys.\ J.\ C {\bf 17} (2000) 487
1515:   [arXiv:hep-ph/0002134].
1516:   %%CITATION = HEP-PH 0002134;%%
1517: \bibitem{Allanach:2004rh}
1518:   B.~C.~Allanach, A.~Djouadi, J.~L.~Kneur, W.~Porod and P.~Slavich,
1519:   {\em Precise determination of the neutral Higgs boson masses in the MSSM},
1520:   JHEP {\bf 0409} (2004) 044
1521:   [arXiv:hep-ph/0406166].
1522:   %%CITATION = HEP-PH 0406166;%%
1523: \bibitem{hfg}
1524: E. Barberio {\em et al} [Heavy Flavour Averaging Group],
1525: {\em Averages of b-hadron Properties at the End of 2005}, arXiv:hep-ex/0603003.
1526: \bibitem{gamb}
1527:   P.~Gambino, U.~Haisch and M.~Misiak,
1528:   {\em Determining the sign of the $b \rightarrow s \gamma$ amplitude},
1529:   Phys.\ Rev.\ Lett.\  {\bf 94} (2005) 061803
1530:   [arXiv:hep-ph/0410155].
1531:   %%CITATION = HEP-PH 0410155;%%
1532: \bibitem{Misiak:2006zs}
1533:   M.~Misiak {\it et al.},
1534:   %``The first estimate of BR(B -> X_s gamma) at O(alpha_s^2),''
1535:   arXiv:hep-ph/0609232.
1536:   %%CITATION = HEP-PH 0609232;
1537: \bibitem{Andersen:2006hr}
1538:   J.~R.~Andersen and E.~Gardi,
1539:   %``Radiative B decay spectrum: DGE at NNLO,''
1540:   arXiv:hep-ph/0609250.
1541:   %%CITATION = HEP-PH 0609250;%%
1542: \bibitem{wmap}
1543:  D.~N.~Spergel {\it et al.},
1544:    {\em Wilkinson Microwave Anisotropy Probe (WMAP) three year results
1545:   Implications for cosmology},
1546:   arXiv:astro-ph/0603449.
1547:   %%CITATION = ASTRO-PH 0603449;%%
1548: 
1549: \bibitem{Boehm:9911}
1550: C.~Boehm, A.~Djouadi and M.~Drees,  {\em Light Scalar Top Quarks and Supersymmetric Dark Matter}, 
1551: {Phys. Rev.} {\bf D62} (2000) 035012
1552:   {\tt hep-ph/9911496}.
1553: 
1554: \bibitem{arnie}
1555:  R.~Arnowitt, B.~Dutta and Y.~Santoso, {\em  Coannihilation Effects in
1556:  Supergravity and D-Brane Models}, {Nucl. Phys.} {\bf B606} (2001) 59
1557:   {\tt hep-ph/0102181}.
1558: 
1559: \bibitem{Ellis:2001nx}
1560: J.~R. Ellis, K.~A. Olive and Y.~Santoso, {\em Calculations of Neutralino-Stop Coannihilation in the CMSSM}, {Astropart. Phys.} {\bf 18} (2003) 395,
1561:   {\tt hep-ph/0112113}.
1562: 
1563: \bibitem{Griest:1990kh}
1564: K.~Griest and D.~Seckel, {\em Three exceptions in the calculation of relic
1565:   abundances}, {Phys. Rev.} {\bf D43} (1991) 3191--3203.
1566: 
1567: \bibitem{Drees:1992am}
1568: M.~Drees and M.~M. Nojiri, {\em The Neutralino Relic Density in Minimal N=1 Supergravity},  {Phys. Rev.} {\bf D47} (1993) 376--408,
1569:   {\tt hep-ph/9207234}.
1570: 
1571: \bibitem{Arnowitt:1993mg}
1572: R.~Arnowitt and P.~Nath, {\em  Cosmological Constraints and SU(5) Supergravity Grand Unification}, {Phys. Lett.} {\bf B299} (1993) 58--63,
1573:   {\tt hep-ph/9302317}.
1574: 
1575: \bibitem{Djouadi:2005dz}
1576:   A.~Djouadi, M.~Drees and J.~L.~Kneur,
1577:   {\em Neutralino dark matter in mSUGRA: Reopening the light Higgs pole  window,},
1578:   Phys.\ Lett.\ B {\bf 624} (2005) 60
1579:   [arXiv:hep-ph/0504090].
1580:   %%CITATION = HEP-PH 0504090;
1581: 
1582: \bibitem{Feng:1999mn}
1583: J.~L. Feng, K.~T. Matchev, and T.~Moroi, {\em Multi-TeV scalars are natural in minimal supergravity},  {Phys. Rev. Lett.} {\bf 84} (2000) 2322--2325,
1584:   {\tt hep-ph/9908309}.
1585: 
1586: \bibitem{Feng:1999zg}
1587: J.~L. Feng, K.~T. Matchev, and T.~Moroi, {\em Focus points and naturalness in supersymmetry},  {Phys. Rev.} {\bf D61} (2000) 075005,
1588:   {\tt hep-ph/9909334}.
1589: 
1590: \bibitem{Feng:2000gh}
1591: J.~L. Feng, K.~T. Matchev, and F.~Wilczek, {\em  Neutralino dark matter in focus point supersymmetry},  {Phys. Lett.} {\bf B482} (2000) 388--399,
1592:   {\tt hep-ph/0004043}.
1593: 
1594: \bibitem{Armstrong:1994it}
1595: {\bf ATLAS} Collaboration, W.~W. Armstrong {\em et.~al.},  
1596: {\em ATLAS detector and physics performance},
1597: CERN-LHCC-94-43.
1598: 
1599: \bibitem{CMS}
1600: {\bf CMS} Collaboration, A. de Roeck {\em et al}, 
1601: {\em CMS physics technical design report vol. II: physics performance},
1602: CERN/LHCC 2006-021.
1603: 
1604: \bibitem{gian}
1605:   N.~Arkani-Hamed, A.~Delgado and G.~F.~Giudice,
1606:   {\em The well-tempered neutralino},
1607:   Nucl.\ Phys.\ B {\bf 741} (2006) 108
1608:   [arXiv:hep-ph/0601041].
1609:   %%CITATION = HEP-PH 0601041;%%
1610: \bibitem{sfk}
1611:  S.~F.~King and J.~P.~Roberts,
1612:  {\em Natural implementation of neutralino dark matter},
1613:   arXiv:hep-ph/0603095.
1614:   %%CITATION = HEP-PH 0603095;%%
1615: \bibitem{Giudice:2006sn}
1616:   G.~F.~Giudice and R.~Rattazzi,
1617:   {\em Living dangerously with low-energy supersymmetry},
1618:   Nucl.\ Phys.\ B {\bf 757} (2006) 19
1619:   [arXiv:hep-ph/0606105].
1620:   %%CITATION = HEP-PH 0606105;%%
1621: \bibitem{Lester:2005je}
1622:   C.~G.~Lester, M.~A.~Parker and M.~J.~White,
1623:    {\em Determining SUSY model parameters and masses at the LHC using
1624:   cross-sections, kinematic edges and other observables},
1625:   JHEP {\bf 0601} (2006) 080
1626:   [arXiv:hep-ph/0508143].
1627:   %%CITATION = HEP-PH 0508143;%%
1628: 
1629: 
1630: 
1631: \bibitem{Allanach:2004xn}
1632:   B.~C.~Allanach, G.~Belanger, F.~Boudjema and A.~Pukhov,
1633:    {\em Requirements on collider data to match the precision of WMAP on
1634:   supersymmetric dark matter},
1635:   JHEP {\bf 0412} (2004) 020
1636:   [arXiv:hep-ph/0410091].
1637:   %%CITATION = HEP-PH 0410091;%%
1638: 
1639: 
1640: 
1641: \bibitem{Baltz:2006fm}
1642:   E.~A.~Baltz, M.~Battaglia, M.~E.~Peskin and T.~Wizansky,
1643:   {\em Determination of dark matter properties at high-energy colliders},
1644:   arXiv:hep-ph/0602187.
1645:   %%CITATION = HEP-PH 0602187;%%
1646: 
1647: 
1648: 
1649: \bibitem{Metropolis}
1650: N.~Metropolis, A.W.~Rosenbluth, M.N.~Teller and E.~Teller,
1651: {\em Equations of State Calculations by Fast Computing Machines},
1652: Journal of Chemical Physics, {\bf 21} (1953) 1087-1091
1653: \bibitem{Hastings}
1654: W.K.~Hastings,
1655: {\em {Monte Carlo} Sampling Methods Using Markov Chains and Their
1656:   Applications},
1657: Biometrika {\bf 57} (1970) 97-109.
1658: \bibitem{MacKay}
1659: D.~MacKay, {\em Information Theory, Inference, and Learning Algorithms}.
1660: \newblock Cambridge University Press, 2003.
1661: 
1662: 
1663: \bibitem{gelman}
1664: A. Gelman and D. Rubin, 
1665: {\em Inference from Iterative Simulation Using Multiple Sequences},
1666: Stat. Sci. {\bf 7} (1992) 457.
1667: 
1668: 
1669: \end{thebibliography}
1670: 
1671: \end{document}
1672: