hep-ph0507283/map.tex
1: \documentclass[aps,prd,twocolumn]{revtex4}
2: \bibliographystyle{apsrev}
3: %\usepackage{multicol}
4: \usepackage[dvips]{epsfig,graphics}
5: \usepackage{graphicx}
6: 
7: %%%%%%%%%%% Chris' macros %%%%%%%%%%%%%%%%
8: \usepackage{subfigure}
9: \usepackage{amsmath}
10: \newcommand\etc{{\em etc}}
11: \newcommand\ie{{\bf i.e.}}
12: 
13: \newcommand\ringfinder{{\tt ring-finder}}
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
15: 
16: %% %%%%%%%%% Graphics macros %%%%%%%%%%%%%%
17:  \newlength{\wth}
18:  \setlength{\wth}{10cm}
19:  \newcommand{\twographs}[2]{%
20:  \unitlength=1.1in
21:  \begin{picture}(5.8,2.6)(-0.3,0)
22:  \put(0,0){\epsfig{file=#1.eps, width=9cm}}
23:  \put(2.7,0){\epsfig{file=#2.eps, width=9cm}}
24:  \put(0.2,2.1){(a)}
25:  \put(2.7,2.1){(b)}
26:  \end{picture}
27: }
28: 
29: \newcommand{\twographsg}[2]{%
30:  \unitlength=1.1in
31:  \begin{picture}(6.,2.6)
32:  \put(-0.2,0){\epsfig{file=#1.eps, width=\wth}}
33:  \put(3.,0){\epsfig{file=#2.eps, width=\wth}}
34:  \put(0,2.1){(a)}
35:  \put(3.0,2.1){(b)}
36:  \end{picture}
37: }
38: 
39: \newcommand{\twographschris}[2]{%
40:  \unitlength=1.1in
41:  \begin{picture}(6.,2.3)
42:  \put(0.2,0){\epsfig{file=#1, width=0.7\wth}}
43:  \put(3.2,0){\epsfig{file=#2, width=0.7\wth}}
44:  \put(0,2.0){(a)}
45:  \put(3.0,2.0){(b)}
46:  \end{picture}
47: }
48: 
49: \newcommand{\twographst}[2]{%
50:  \unitlength=1.1in
51:  \begin{picture}(5.8,2.6)(0.5,0)
52:  \put(0,0){\epsfig{file=#1, width=\wth}}
53:  \put(0.9,0.46){\epsfig{file=#12, width=0.678 \wth}}
54:  \put(2.7,0){\epsfig{file=#2, width=\wth}}
55:  \put(3.6,0.46){\epsfig{file=#22, width=0.678 \wth}}
56:  \put(0.5,2.1){(a)}
57:  \put(3.2,2.1){(b)}
58:  \end{picture}
59: }
60: 
61: \newcommand{\fourgraphs}[4]{%
62: \unitlength=1in
63: \begin{picture}(8.7,4.7)(-0.4,0.3)
64: \put(2.7,2.4){\epsfig{file=#2.eps, width=\wth}}
65: \put(3.69,2.905){\epsfig{file=#22.eps, width=0.678 \wth}}
66: \put(-0.5,0.1){\epsfig{file=#3.eps, width=\wth}}
67: \put(0.49,0.605){\epsfig{file=#32.eps, width=0.678 \wth}}
68: \put(0.1,4.7){(a)}
69: \put(-0.5,2.4){\epsfig{file=#1.eps, width=\wth}}
70: \put(0.49,2.905){\epsfig{file=#12.eps, width=0.678 \wth}}
71: \put(3.3,4.7){(b)}
72: \put(0.1,2.3){(c)}
73: \put(3.3,2.3){(d)}
74: \put(2.7,0.1){\epsfig{file=#4.eps, width=\wth}}
75: \put(3.69,0.605){\epsfig{file=#42.eps, width=0.678 \wth}}
76: %% \put(0,0){\epsfig{file=#3.eps, width=\wth}}
77: %% \put(2.7,0){\epsfig{file=#4.eps, width=\wth}}
78: %% \put(0,2.6){\epsfig{file=#1.eps, width=\wth}}
79: %% \put(2.7,2.6){\epsfig{file=#2.eps, width=\wth}}
80: %% \put(0.1,2.3){(c)}
81: %% \put(0.1,4.9){(a)}
82: %% \put(3.0,2.3){(d)}
83: %% \put(3.0,4.9){(b)}
84: \end{picture}}
85: 
86: \newcommand{\threegraphs}[3]{%
87: \unitlength=1in
88: \begin{picture}(5.8,4.6)(0.15,0)
89: \put(1.35,0){\epsfig{file=#3.eps, width=\wth}}
90: \put(-0.5,2.3){\epsfig{file=#1.eps, width=\wth}}
91: \put(2.7,2.3){\epsfig{file=#2.eps, width=\wth}}
92: \put(1.55,2.2){(c)}
93: \put(0.1,4.5){(a)}
94: \put(3.0,4.5){(b)}
95: \end{picture}}
96: 
97: \newcommand{\sixgraphs}[6]{%
98: \unitlength=1in
99: %\begin{picture}(8.7,6.9)(0.15,0.3)
100: \begin{picture}(8.7,6.9)(-0.4,0.3)
101: \put(0.1,6.9){(a)}
102: \put(-0.5,4.8){\epsfig{file=#1.eps, width=10cm}}
103: \put(3.3,6.9){(b)}
104: \put(2.7,4.8){\epsfig{file=#2.eps, width=10cm}}
105: \put(0.1,4.5){(c)}
106: \put(-0.5,2.3){\epsfig{file=#3.eps, width=10cm}}
107: \put(3.3,4.5){(d)}
108: \put(2.7,2.3){\epsfig{file=#4.eps, width=10cm}}
109: \put(0.1,2.2){(e)}
110: \put(-0.5,0.){\epsfig{file=#5.eps, width=10cm}}
111: \put(3.3,2.2){(f)}
112: \put(2.7,0.){\epsfig{file=#6.eps, width=10cm}}
113: \end{picture}}
114: 
115: \newcommand{\sixgraphsmod}[6]{%
116: \unitlength=1in
117: %\begin{picture}(8.7,6.9)(0.15,0.3)
118: \begin{picture}(8.7,6.9)(-0.4,0.3)
119: \put(-0.5,0.){\epsfig{file=#5.eps, width=10cm}}
120: \put(0.49,0.5){\epsfig{file=#52.eps, width=6.8cm}}
121: \put(-0.5,4.78){\epsfig{file=#1.eps, width=10cm}}
122: \put(0.49,5.3){\epsfig{file=#12.eps, width=6.8cm}}
123: \put(0.1,6.9){(a)}
124: \put(2.7,4.8){\epsfig{file=#2.eps, width=10cm}}
125: \put(3.69,5.3){\epsfig{file=#22.eps, width=6.8cm}}
126: \put(0.1,4.5){(c)}
127: \put(-0.5,2.3){\epsfig{file=#3.eps, width=10cm}}
128: \put(0.49,2.8){\epsfig{file=#32.eps, width=6.8cm}}
129: \put(3.3,4.5){(d)}
130: \put(2.7,2.3){\epsfig{file=#4.eps, width=10cm}}
131: \put(3.69,2.8){\epsfig{file=#42.eps, width=6.8cm}}
132: \put(0.1,2.2){(e)}
133: \put(3.3,2.2){(f)}
134: \put(2.7,0.){\epsfig{file=#6.eps, width=10cm}}
135: \put(3.69,0.5){\epsfig{file=#62.eps, width=6.8cm}}
136: \put(3.3,6.9){(b)}
137: \end{picture}}
138: 
139: %\keywords{Dark Matter, Beyond Standard Model, Cosmology, MSSM, CMSSM, universality}
140: 
141: %\preprint{DAMTP-2005-64\\ hep-ph/0507283}
142: 
143: \begin{document}
144: 
145: \title{Multi-Dimensional mSUGRA Likelihood Maps\footnote{Preprint number: DAMTP-2005-64}}
146: 
147: \author{B.C. Allanach}
148: \affiliation{DAMTP, CMS, Wilberforce Road,
149: Cambridge, CB3 0WA, United Kingdom}
150: 
151: \author{C.G. Lester}
152: \affiliation{Cavendish Laboratory. Madingley Road.
153: Cambridge CB3 0HE, United Kingdom}
154: 
155: \begin{abstract}
156:   We calculate the likelihood map in the full 7 dimensional parameter
157:   space  of the minimal supersymmetric standard model (MSSM) assuming universal
158:   boundary conditions on the supersymmetry breaking terms. Simultaneous
159:   variations of $m_0$, $A_0$, $M_{1/2}$, $\tan \beta$, $m_t$, $m_b$ and
160:   $\alpha_s(M_Z)$ are applied using a Markov chain Monte Carlo algorithm.
161:   We use measurements of $b \rightarrow s \gamma$, $(g-2)_{\mu}$ and
162:   $\Omega_{DM} h^2$ in order to constrain the model. We present likelihood
163:   distributions for some of the sparticle 
164:   masses, for the branching ratio of
165:   $B_s^0 \rightarrow \mu^+ \mu^-$ and for $m_{\tilde \tau}-m_{\chi_1^0}$. 
166:   An upper limit of 2$\times 10^{-8}$ on this branching ratio might be achieved
167:   at the Tevatron, and would rule out 29$\%$ of the currently allowed
168:   likelihood. If one allows for non thermal-neutralino components of dark
169:   matter, this fraction becomes 35$\%$. 
170:   The mass ordering allows the important cascade decay
171:   ${\tilde q}_L \rightarrow {\chi_2^0} \rightarrow {\tilde l}_R
172:   \rightarrow {\chi_1^0}$ with a likelihood of 24$\pm 4\%$. The stop
173:   coannihilation region is highly disfavoured, whereas the light Higgs region 
174:   is marginally disfavoured.
175: \end{abstract}
176: 
177: \maketitle
178: 
179: \section{Introduction}
180: Weak-scale supersymmetry provides a well-documented solution to the technical
181: hierarchy problem~\cite{Martin:1997ns}, which is particularly difficult to
182: solve in a perturbatively calculable model. 
183: Specialising to a minimal extension of the Standard Model, the MSSM, one can
184: provide a weakly interacting massive particle dark matter candidate,
185: provided $R-$parity is respected by the model. 
186: Examples of dark matter candidates are the
187: gravitino~\cite{Ellis:2003dn,Roszkowski:2004jd}, the axino~\cite{Covi:2004rb}
188: and the lightest neutralino~\cite{Ellis:1983ew}, the subject of much
189: recent investigation~\cite{Baer:2002gm,Ellis:2003cw,Battaglia:2003ab,Ellis:2003si,Gomez:2004ek,Ellis:2004tc}. 
190: The general MSSM is rather complicated due to the large number of free
191: parameters in the supersymmetric (SUSY) breaking sector. 
192: However, the observed rareness of flavour changing neutral currents (FCNCs)
193: suggests  
194: that the vast majority of parameter space for general SUSY breaking terms is
195: ruled out. 
196: Particular patterns of SUSY breaking parameters can postdict small enough
197: FCNCs: for instance flavour universality. 
198: One highly studied subset of such terms is that of mSUGRA, often called
199: the Constrained Minimal Supersymmetric Standard Model (CMSSM).
200: In mSUGRA, at some high energy scale (typically taken to be the scale of
201: unification of electroweak gauge couplings), all of the SUSY breaking scalar
202: mass terms are assumed to be equal to $m_0$, the scalar trilinear terms 
203: are set to $A_0$ and the gaugino masses are set equal ($M_{1/2}$). 
204: These are indeed strong assumptions, but they have several advantages for
205: phenomenological analysis as the number of independent SUSY breaking
206: parameters is much 
207: reduced. Indeed, assuming that the MSSM is the correct model, the initial
208: data from the Tevatron or Large Hadron Collider are likely to contain only a
209: few relevant 
210: observables and so one may be able to fit against a simple SUSY breaking model 
211: as an example~\cite{Armstrong:1994it}. As the data become more accurate and
212: additional 
213: relevant observables are measured, the lack of a good fit would propel 
214: extensions 
215: of the simple model. One may then start to consider
216: patterns of non-universality, for instance. 
217: For the rest of this paper though, given the lack of data to the contrary,
218: we will assume mSUGRA\@.
219: Aside from the universal soft terms $m_0, A_0, M_{1/2}$,
220: other non-standard model mSUGRA input parameters are taken to be $\tan
221: \beta$, the ratio of the two Higgs vacuum expectation values, and
222: the sign of $\mu$ (a parameter that appears in the Higgs potential of the
223: MSSM).  
224: 
225: When combined with large scale structure data,
226: the Wilkinson microwave ani\-sot\-ropy probe (WMAP)~\cite{Spergel:2003cb,Bennett:2003bz} has placed 
227: stringent constraints upon the dark matter relic density $\Omega_{DM} h^2$.
228: A common assumption, which we will adhere to here, is
229: that the neutralino makes up the entire cold dark matter relic density.
230: The prediction of the relic density of dark matter in the MSSM depends
231: crucially upon annihilation cross-sections, since in the early universe 
232: SUSY particles will annihilate in the thermal bath.\ 
233: Regions of 
234: mSUGRA that are compatible with the WMAP constraint often predict some of the
235: following annihilation channels ~\cite{Allanach:2004xn}:
236: \begin{itemize}
237: \item
238: Stau (${\tilde \tau}$) co-annihilation~\cite{Griest:1990kh} at small $m_0$
239: where the lightest stau is quasi-degener\-ate with the lightest neutralino
240: ($\chi_1^0$).  
241: \item
242: Pseudoscalar Higgs ($A^0$) funnel region at large $\tan \beta > 45$ where
243: two neutralinos annihilate through an $s$-channel $A^0$
244: resonance~\cite{Drees:1992am,Arnowitt:1993mg}. 
245: \item
246: Light CP-even Higgs ($h^0$) region at low $M_{1/2}$ where
247: two neutralinos annihilate through an $s$-channel $h^0$
248: resonance~\cite{Drees:1992am,Djouadi:2005dz}. 
249: \item
250: Focus point~\cite{Feng:1999mn,Feng:1999zg,Feng:2000gh} at large $m_0$ where a
251: significant 
252: Higgsino component leads to efficient neutralino annihilation into gauge boson
253: pairs.
254: \item
255: Stop co-annihilation~\cite{Boehm:9911,arnie,Ellis:2001nx} at large $(-A_0)$,
256: where the lightest 
257: stop is close in mass to the lightest neutralino.
258: \end{itemize}
259: Many pre-WMAP analyses focused on the so-called bulk
260: region~\cite{Drees:1992am}. The 
261: bulk region is continuously connected to the
262: stau co-annihilation region at low $m_0$ and $M_{1/2}$.
263: There are two reasons why the bulk region has shrunk in size when
264: one takes the current constraints into account: 
265: the WMAP constraint upon the relic density has ruled much of the region out,
266: and the new low value of the top mass mean that the MSSM Higgs mass
267: predictions are sometimes too low for low $m_0$ and $M_{1/2}$
268: and are ruled out by LEP2 constraints. 
269: The (now reduced) bulk region will make an implicit appearance in our
270: results, and we will comment upon this fact later.
271: 
272: Several
273: authors~\cite{Brhlik:2000dm,Drees:2000he,Polesello:2004qy,Bambade:2004tq,Battaglia:2003ab,Allanach:2004jh,Moroi:2005nc}
274: have asked how the annihilation cross-section can be 
275: constrained 
276: by collider measurements in order to provide a more solid prediction of the
277: relic density. This would then be fed into a cosmological model in
278: order to predict  $\Omega_{DM} h^2$ for comparison
279: with the value derived from cosmological observation, allowing a test of
280: cosmological assumptions (and the assumption that there is only one component
281: of cold dark matter). Of course, colliders could not unambiguously identify
282: the lightest 
283: observed SUSY particle as the dark matter since it could always decay
284: unobserved outside the detector. 
285: It would therefore be interesting to combine collider
286: information with  that derived from a possible future direct
287: detection~\cite{Bourjaily:2005ax} of dark matter, providing corroboration and additional
288: empirical information. 
289: Before such observations are made, however, we may  ask how well current data
290: constrain models of new physics. 
291: 
292: This question has been addressed many times for mSUGRA by using the dark
293: matter constraint. Most of the analyses (see, for
294: example~\cite{Roszkowski:2001sb,Baer:2002gm,Baer:2002fv,Baer:2003yh,Ellis:2003cw,Chattopadhyay:2003xi,Lahanas:2003yz,Belanger:2004hk})   
295: fix all but two parameters and examine constraints upon the remaining 2
296: dimensional (2d) 
297: slice of parameter space. The dark matter relic density constraint is the most
298: limiting, but the branching ratio of the decay $b \rightarrow s \gamma$ 
299: and the anomalous magnetic moment of the muon $(g-2)_\mu$ also rule part of
300: the parameter space out. Recent upper bounds from the Tevatron
301:   experiments on the branching ratio $B_s 
302: \rightarrow \mu^+ \mu^-$~\cite{Herndon:2004tk} have the potential to 
303: restrict mSUGRA in the future, but the analysis of ref.~\cite{Ellis:2005sc}
304: shows that the resulting constraints currently subsumed within other
305: constraints. In the above analyses,
306: limits are typically imposed separately, each to some 
307: prescribed confidence level. Such analyses have the advantage of being
308: quite transparent: it is fairly easy to see which constraint rules out which
309: part of parameter space. However, they have the disadvantages of not properly
310: describing the combination of likelihoods coming from different experimental
311: constraints and of having to assume {\em ad hoc}\/ values for several input
312: parameters. In particular, as well as the soft SUSY breaking input parameters,
313: the bottom mass $m_b$, the strong structure
314: constant $\alpha_S(M_Z)$ and the top mass $m_t$ can all have a strong effect on 
315: mSUGRA predictions.
316: A large random scan of flavour diagonal MSSM space involving $10^5$ points
317: that pass various prescribed constraints was presented in
318: ref.\cite{Profumo:2004at}, 
319: however the sampling of the 20d parameter space was necessarily
320: sparse. The analysis is also subject to the limitation that likelihoods have
321: not been combined; instead the measurements have been used as cuts to discard
322: points. 
323: In ref.~\cite{Ellis:2003si}, the likelihood from the observables is
324: calculated, properly combining different constraints, but again 1d and 2d
325: slices through parameter space were taken.
326: Of course the time taken to efficiently sample from a likelihood distribution
327: using the naive method (a scan) scales like a power law with respect to
328: the number of parameters, meaning that in practice even a high
329: resolution 3d scan is difficult. By parameterising lines in 2d that are
330: consistent with 
331: the WMAP dark matter constraint and scanning in two other parameters, the
332: analysis of 
333: ref.~\cite{Ellis:2004tc} 
334: calculates the $\chi^2$ statistic
335: for the 2d part of a 3d parameter space which is
336: consistent with the WMAP constraint on the dark matter relic density.
337: The predicted value of $\Omega_{DM} h^2$ is not combined in the
338:   $\chi^2$ with the other observables
339: for this analysis, and the parameter $\tan \beta$ must be fixed. 
340: As the authors note~\cite{Ellis:2004tc}, parts of the scan were
341: sparse. In Ref.~\cite{Stark:2005mp}, a scan was performed which included
342: variations of $A_0$ and $\tan \beta$ as well as other mSUGRA parameters. It is
343: clear from this paper that the WMAP allowed region 
344: (expressed in the $M_{1/2}$-$m_0$ plane) becomes much larger from the $A_0$
345: variations. No likelihood distribution was given. 
346: 
347: Baltz and Gondolo~\cite{Baltz:2004aw} demonstrated that a Markov chain Monte
348: Carlo (MCMC) algorithm efficiently samples from the mSUGRA parameter space,
349: rendering 4d scans in $m_0, A_0, M_{1/2}, \tan \beta$
350: feasible. However, they were interested in  
351: which parts of parameter space are compatible with the WMAP measurement of
352: $\Omega_{DM} h^2$ and what the prospects are for direct detection there, not
353: in the likelihood distribution. In order
354: to increase the efficiency of their parameter sampling, they changed
355: the simple ``Metropolis-Hastings'' MCMC algorithm in order to achieve a better
356: efficiency. As the authors state in their
357: conclusions, this has the consequence that
358: caution must be exercised when trying to interpret their
359: results as a likelihood distribution.
360: Indeed, we will show in a toy model that changes to the MCMC algorithm like
361: the ones that Baltz and Gondolo made can alter the sampling
362: from a distribution. 
363: 
364: It is our purpose here to utilise the MCMC algorithm in such a way as to
365: reliably calculate the combined likelihood of mSUGRA in the full
366: dimensionality of its parameter space, thereby extending the previous
367: studies. We will then be 
368: able to infer what is known about the multi-dimensional
369: parameter space, including important variations of the SM quantities.
370: These results will have implications for collider searches and  rare decays.
371: 
372: In section~\ref{sec:mcmc}, we briefly review the MCMC algorithm. 
373: We present the implementation used in the present paper to 
374: calculate the likelihood maps of mSUGRA parameter space and then demonstrate
375: that the results are convergent using a particular statistical test.
376: In section~\ref{sec:Lmaps}, we present the likelihood distributions and
377: derived quantities of the 7d
378: mSUGRA parameter space. In section~\ref{sec:uncertainties}, we illustrate the
379: effects of theoretical uncertainties in the sparticle spectrum calculation and
380: in section~\ref{sec:other}, possible effects from allowing an additional non
381: thermal-neutralino component to the relic density are explored.
382: A summary and conclusions are presented in  section~\ref{sec:conc}.
383: In appendix~\ref{sec:badsampling}, we 
384: demonstrate with two different toy models that the algorithm
385: used by Baltz and Gondolo may not provide a sampling proportional to the
386: likelihood of the parameter space.
387: 
388: \section{Implementation of the MCMC Algorithm\label{sec:mcmc}}
389: 
390: \subsection{Likelihoods}
391: Some readers might be unfamiliar with the use of statistics in this paper, and
392: so we include some comments on how to interpret them. The 
393: likelihood is {\em not}
394: dependent upon any priors. 
395: The likelihood ${\mathcal L} \equiv p(d|m)$ is the probability density
396: function (pdf) of reproducing
397: data $d$ 
398: assuming some mSUGRA model $m$. In $p(d|m)$, the model $m$ is specified by the
399: mSUGRA input parameters and so $p(d|m)$ has a dependence upon them.
400: $p(d|m)$ is related to the pdf of the model being the one chosen by nature,
401: given the data, by an application of Bayes' theorem:
402: \begin{equation}
403: p(m|d) = p(d|m) \frac{ p(m) }{ p(d) }, \label{bayes}
404: \end{equation}
405: where $p(m)$ is the probability of the model being correct (the {\em prior})
406: and $p(d)$ is the total probability of the data being reproduced, integrating
407: over all possible models. $p(d)$ is
408: practically impossible to estimate, so we cannot get the quantity that one
409: really wishes to estimate, $p(m|d)$. 
410: However, we may compare the relative probabilities of two different models
411: $m_1$ and $m_2$ (corresponding here to different points of mSUGRA space) by
412: applying 
413: Eq.~\ref{bayes} for each model, implying that
414: \begin{equation}
415: \frac{p(m_1|d)}{p(m_2|d)} = \frac{ p(d|m_1) p(m_1) }{p(d|m_2) p(m_2)}.
416: \end{equation}
417: We note here the appearance of the infamous prior distributions $p(m_1),
418: p(m_2)$. If one assumes that the ratio of these two priors is one (that no
419: region of parameter space is more likely than any other), one may interpret
420: the likelihood ratio of two different points in mSUGRA space as the ratio of
421: probabilities of the models, given the data. 
422: In this paper however, we provide likelihood distributions. If the reader
423: prefers a different ratio of priors to one, they must convolute the likelihood
424: density we give with their preferred ratio of pdfs. 
425: 
426: \subsection{The MCMC Algorithm}
427: 
428: We now briefly review the Metropolis MCMC algorithm, but for a more
429: thorough explanation, see refs.~\cite{Baltz:2004aw,MacKay}.
430: Other adaptive scanning algorithms have recently been
431: suggested in the context of high energy
432: physics~\cite{Allanach:2004my,Brein:2004kh} but (although they can be very
433: useful for 
434: other purposes) they do not yield a likelihood distribution. 
435: A Markov chain consists of a list of parameter points (${\mathbf x}^{(t)}$)
436: and associated 
437: likelihoods (${\mathcal L}^{(t)} \equiv {\mathcal L}({\mathbf x}^{(t)})$). Here
438: $t$ labels 
439: the link number in the chain.
440: Given some point at the end of the Markov chain (${\mathbf x}^{(t)}$),
441: the Metropolis-Hastings algorithm involves randomly picking another potential
442: point (${\mathbf x}^{(t+1)}$) (typically in the vicinity of ${\mathbf
443:   x}^{(t)}$)
444:  using
445: some proposal pdf $Q({\mathbf x};{\mathbf x}^{(t)})$. In this paper we will
446: specialise to the case of symmetric proposal functions, i.e.\ $Q({\mathbf
447:   x_b}; {\mathbf x_a})=Q({\mathbf x_a}; {\mathbf x_b})$. 
448: If ${\mathcal L}^{(t+1)}>{\mathcal L}^{(t)}$, the new point is appended onto
449: the 
450: chain. Otherwise, the proposed point is accepted with probability
451: ${\mathcal L}^{(t+1)} / {\mathcal L}^{(t)}$ and, if accepted, added to the end
452: of the   chain. If 
453: the point ${\mathbf x}^{(t+1)}$ is not accepted, the point ${\mathbf
454:   x}^{(t)}$ is copied on to the end of the chain instead.
455: 
456: Providing ``detailed balance'' is satisfied, it can be shown~\cite{MacKay}
457: that the sampling density of points in the chain is proportional to 
458: the target distribution (in this case, the likelihood) 
459: as the number of links goes to infinity. 
460: In the context of this analysis, detailed balance states that for any two
461: points  ${\mathbf x_a}, {\mathbf x_b}$
462: \begin{equation}
463: T({\mathbf x_a} ; {\mathbf x_b}) {\mathcal L}({\mathbf x_b}) = T({\mathbf x_b}
464: ; {\mathbf x_a}) {\mathcal L}({\mathbf x_a}),
465: \end{equation}
466: where $T({\mathbf x_b}; {\mathbf x_a}) \equiv Q({\mathbf x_b}; {\mathbf x_a})
467: \times \mbox{min}(1, {\mathcal L}({\mathbf x_b})/{\mathcal L}({\mathbf x_a}))$ is the probability of
468:  making a transition from ${\mathbf x_a}$ to ${\mathbf x_b}$ in the case where
469:  the proposal function 
470:  is symmetric. 
471: Thus, the probability of sampling a point ${\mathbf x_a}$ from the likelihood
472: distribution and then making a 
473: transition to ${\mathbf x_b}$ 
474: be equal to the probability of sampling ${\mathbf x_b}$ and making a
475: transition to ${\mathbf x_a}$.
476: 
477: The Metropolis-Hastings MCMC algorithm is typically much more efficient than a 
478: straightforward scan for $D>3$; the number of required steps scales roughly
479: linearly 
480: with $D$ rather than as a power law. The sampling is in principle independent
481: of the form of $Q$ as $t \rightarrow \infty$ as long as it is bigger than zero
482: everywhere. However,
483: $Q$ must be chosen with some care: since in practice we can only sample a
484: finite number of points, the choice of the form of $Q$ can determine whether 
485: the entire parameter space is sampled and how quickly convergence is reached. 
486: 
487: Baltz and Gondolo used a geometrical model for $Q$: choosing a random distance
488: from the point ${\mathbf x^{(t)}}$ and using a direction that was calculated
489:   from the positions of previous 
490: points in the chain. The width of the random radius pdf was calculated
491: depending upon previous points in the chain in order to increase the
492: efficiency of the 
493: calculation, aiming to accept roughly 25$\%$ of potential points. Either of
494: these changes upset detailed balance and may spoil the sampling. We
495: demonstrate in Appendix~\ref{sec:badsampling} with toy models
496: that the width changing modification gives a sampling that is not proportional
497: to the target density. 
498: 
499: \subsection{Parameter Ranges \label{sec:imp}}
500: 
501: \begin{table}
502: \caption{Parameter ranges considered. \label{tab:prior}}
503: \begin{tabular}{|c|c|}
504: \hline
505: parameter & range \\ \hline
506: {sign}$(\mu)$ & +1\ \\
507: $A_0$ & -2 TeV$-$2 TeV \\
508: $m_0$ & 60 GeV$-$2 TeV \\
509: $M_{1/2}$ & 60 GeV$-$2 TeV \\
510: $\tan \beta$ & 2$-$60 \\ \hline
511: \end{tabular}
512: \end{table}
513: On general naturalness grounds\footnote{That
514:   is, to avoid a large cancellation between weak scale SUSY breaking terms in
515:   order to get a small value of $\mu$.}, we expect $m_0$ and $M_{1/2}$ to not
516: be too large: less than say, 2 TeV. 
517: mSUGRA is ruled out by negative results in sparticle searches for
518: $m_0<60$ GeV or $M_{1/2}<60$ GeV. $\mu>0$ is favoured by the measurement of
519: the anomalous magnetic moment of the muon.
520: $\tan \beta$ is bounded from below by
521: negative searches at LEP2 for $h^0$ (and perturbativity of the top Yukawa
522: coupling) and from above by perturbativity of the Yukawa couplings up to the
523: unification scale. 
524: 
525: Upper bounds upon $m_0$ can 
526: exclude the focus point region, which, in our calculation,
527: is at much higher values of $m_0\sim O(8)$ TeV.
528: It has been
529: argued that a quantitative measure of fine-tuning in the focus point region is
530: not too large~\cite{Feng:1999mn,Feng:1999zg}, however the
531: fine-tuning of the top quark Yukawa coupling is
532: enormous~\cite{Allanach:2000ii}. 
533: This makes the focus-point regime practically impossible to reliably
534: calculate starting from mSUGRA inputs. Tiny higher order effects in
535: the top Yukawa coupling strongly
536: change the position of the focus point regime in mSUGRA parameter space.
537: In ref.~\cite{Allanach:2000ii}, it is demonstrated that the focus point
538: regime 
539: moves in the $m_0$ direction by several TeV depending on how exactly the
540: highest order top-quark Yukawa radiative corrections are calculated.
541: Because the calculation cannot be controlled with the current state-of-the-art
542: calculations of the top Yukawa coupling, we will
543: exclude the focus point regime from this analysis by placing an appropriate
544: upper bound upon $m_0$.
545: Here, we restrict the parameter space to that shown in Table~\ref{tab:prior}.
546: 
547: \subsection{Observables and constraints}
548: %% The add\-ed volume of parameter space necessary to include the focus point
549: %% region means that convergence is difficult to achieve with the CPU power
550: %% available to us. Also, focus point region points take much more time to
551: %% accurately calculate the spectra for than other viable mSUGRA 
552: %% points. 
553: %% For now, we will simply neglect the focus point region for convenience. 
554: 
555: \begin{table}
556: \caption{Lower bounds applied to sparticle mass predictions in GeV.\label{tab:constraints}}
557: \begin{tabular}{|cc|cc|cc|cc|}
558: \hline
559: $m_{\chi_1^0}$ & 37  & $m_{\chi^\pm_1}$ & 67.7 & 
560: $m_{\tilde g}$ & 195 &
561: $m_{{\tilde \tau}_1}$ & 76 \\ $m_{{\tilde l}_R}$ & 88 & $m_{{\tilde t}_1}$ &
562:   86.4 &
563: $m_{{\tilde b}_1}$ & 91 & $m_{{\tilde q}_R}$ & 250 \\
564: $m_{{\tilde \nu}_{e,\mu}}$ & 43.1 & & & & & & \\
565: \hline
566: \end{tabular}
567: \end{table}
568: We calculate the MSSM spectrum from mSUGRA
569: parameters, by using the program {\tt SOFTSUSY1.9.2}~\cite{Allanach:2001kg}.
570: Ideally we would like to include data from negative search results from
571: collider data within a combined likelihood. 
572: Unfortunately it is difficult to obtain the data in such a form and so
573: instead, we assign a 
574: zero likelihood to any point for which at least one of the
575: constraints~\cite{PDBook} in Table~\ref{tab:constraints} is not satisfied.
576: We also implement a parameterisation\footnote{Developed by P Slavich for
577:   Ref.~\protect\cite{Allanach:2004rh}.} of the 95$\%$
578: confidence level 
579: limits~\cite{Barate:2003sz} on $m_h(g_{hZZ}/g_{hZZ}^{SM})$, where
580: $g_{hZZ}/g_{hZZ}^{SM}$ is the ratio of the MSSM higgs coupling to two $Z^0$
581: bosons to the equivalent Standard Model coupling. In order to take a 3 GeV
582: uncertainty on  
583: the mSUGRA prediction of $m_h$ into account, we
584: add 3 GeV~\cite{Degrassi:2002fi,Allanach:2004rh}
585: to the $m_{h^0}$ value that is used in the parameterisation. In the MSSM,
586: $g_{hZZ}/g_{hZZ}^{SM}=\sin(\beta - \alpha)$ and
587: in practice, it
588: is easier to apply limits in terms of the {\em inverse} \/parameterisation
589: $\sin^2 (\beta - \alpha)(m_{h^0})$ as shown in Table~\ref{tab:par}.
590: 
591: \begin{table*}
592: \caption{Parameterisation of 95$\%$ condidence level LEP2 Higgs limits on the
593:   $m_{h^0}$-$\sin^2(\beta-\alpha)$ plane. All points with $m_{h^0}<90$ GeV are
594:   ruled out. \label{tab:par}}
595: \begin{tabular}{|c|c|}
596: \hline $m_{h^0}/$GeV range & upper bound on $\sin^2 (\beta -
597:   \alpha)$ \\ \hline
598: 90-99 & -6.1979 + 0.12313 $m_{h^0}/$GeV - 0.00058411 $(m_{h^0}/$GeV$)^2$\\
599: 99-104 & 35.73 - 0.69747 $m_{h^0}/$GeV + 0.0034266 $(m_{h^0}/$GeV$)^2$\\
600: 104-109.5 & 21.379 - 0.403 $m_{h^0}/$GeV + 0.0019211 $(m_{h^0}/GeV)^2$ \\
601: 109.5-114.4 & 1/(60.081 - 0.51624 $m_{h^0}/$GeV) \\ \hline
602: \end{tabular}
603: \end{table*} 
604: The spectrum is transferred via the SUSY Les 
605: Houches Accord~\cite{Skands:2003cj} to the computer program {\tt
606:   micrOMEGAs1.3.5}~\cite{Belanger:2001fz,Belanger:2004yn} in order to
607: calculate several quantities used to calculate the likelihood of a
608: parameter point. 
609: We will use six measurements in order to construct the final likelihood of any
610: given point of parameter space. As mentioned in the introduction, we make the 
611: assumption that the neutralino makes up the entire cold dark matter relic
612: density as constrained by WMAP:
613: \begin{equation}
614: \Omega_{DM} h^2 = 0.1126^{+0.0081}_{-0.0091}. \label{omConst}
615: \end{equation}
616: The anomalous magnetic moment of the muon has been
617:  measured~\cite{Bennett:2004pv} to be higher than the Standard Model
618:  prediction~\cite{Passera:2004bj,deTroconiz:2004tr}. The experimental
619:  measurement is so precise that the comparison is limited by theoretical
620:  uncertainties in the Standard Model prediction. Following
621:  Ref.~\cite{Allanach:2005yq}, we constrain any new physics contribution to 
622:  be
623: \begin{equation}
624: \delta \frac{(g-2)_\mu}{2} = 19.0 \pm 8.4 \times 10^{-10}. 
625: \end{equation}
626: Adding theoretical errors~\cite{Gambino:2004mv} to measurement
627: errors~\cite{hfag} in quadrature 
628: for the branching ratio for the decay $b \rightarrow s \gamma$,
629: one obtains the empirically derived constraint
630: \begin{equation}
631: BR(b \rightarrow s \gamma) = 3.52 \pm 0.42 \times 10^{-4}.
632: \end{equation}
633: 
634: The Standard Model inputs' measurements also contribute to the
635: likelihood. We take these to be~\cite{PDBook}, for the running bottom quark
636: mass in the modified minimal subtraction scheme,
637: \begin{equation}
638: m_b (m_b)^{\overline{MS}} = 4.2 \pm 0.2 \mbox{~GeV},
639: \end{equation}
640: for the pole mass of the top quark\footnote{For an analysis with 
641: $m_t=174.3 \pm 3.4$ GeV, see the original version of this paper on the
642:   hep-ph/ electronic archive.}~\cite{Group:2005cc},  
643: \begin{equation}
644: m_t = 172.7 \pm 2.9 \mbox{~GeV},
645: \end{equation}
646: and for the strong coupling constant in the modified minimal subtraction
647: scheme at $M_Z$
648: \begin{equation}
649: \alpha_s (M_Z)^{\overline{MS}} = 0.1187 \pm 0.002.
650: \end{equation}
651: A prediction $p_i$ of one of these quantities, where
652: \begin{eqnarray}
653: i&\equiv& \{
654: \alpha_s(M_Z)^{\overline{MS}}, m_t, m_b(m_b)^{\overline{MS}},
655: (g-2)_\mu/2, \nonumber \\
656: && \mbox{BR}(b \rightarrow s \gamma), \Omega_{DM} h^2\}
657: \end{eqnarray}
658:  with
659: measurement $m_i\pm s_i$  yields a log likelihood 
660: \begin{equation}
661: \ln {\mathcal L}_{i} = - \frac{(m_i-p_i)^2}{2s^2_i} - \frac{1}{2} \ln (2 \pi) -
662: \ln s_i, 
663: \label{lnL}
664: \end{equation}
665: assuming the usual Gaussian errors. 
666: Note that since Eq.~\ref{omConst}
667: has asymmetric errors, $s_{\Omega_{DM} h^2}=0.0081(0.0091)$ if our prediction
668: is higher (lower) than the observed central value. 
669: To form the combined likelihood, one 
670: takes $\ln {\mathcal L}^{tot} = \sum_{i=1} \ln {\mathcal L}_i$,
671: corresponding to the combination of independent Gaussian likelihoods.
672: In practice, we will ignore the normalisation
673: constants $-\frac{1}{2}\ln(2 \pi) - \ln s_i$, since the likelihood
674: distribution has an arbitrary normalisation anyway. 
675: 
676: We take 
677: the proposal function to be the product of Gaussian distributions along
678: each dimension $k=1,2,\ldots,D$ centred on the location of the current point 
679: along that dimension, i.e.\ $x^{(t)}_k$: 
680: \begin{equation}
681: Q({\mathbf x}^{(t+1)}; {\mathbf x}^{(t)}) = 
682:   \prod_{k=1}^D \frac{1}{\sqrt{2 \pi} l_k} e^{-(x_k^{(t+1)} -x_k^{(t)} )^2
683:   / 2 l_k^2},
684: \end{equation}
685: where $l_k$ denotes the width of the distribution along direction $k$. 
686: By trial and error we find that using values of $l_k$ that are equal to the
687: parameter range of dimension $k$ given in Table~\ref{tab:prior}
688: divided by 25 works well. 
689: For the Standard Model inputs, we choose $l_k = 8 \sigma_k / 25$. 
690: 
691: In order to start the chain we follow the following procedure, which finds a
692: point at random in parameter space that is not a terrible fit to the data.
693: We pick some ${\mathbf y}^{(0)}$ at random in the mSUGRA parameter
694: space using a flat distribution for its pdf. 
695: The Markov chain for ${\mathbf y}$ is evolved a sufficient number of steps
696: ($t$) such that $\ln{\mathcal L }({\mathbf y}^{(t)}) > -5$, i.e.\ the initial
697: chain 
698: has found a reasonable fit. We then set ${\mathbf x}^{(0)}={\mathbf y}^{(t)}$,
699: continuing the Markov chain in ${\mathbf x}$ and discarding the ``burn-in''
700: chain ${\mathbf y}$. The reasonable-fit point is typically found long before
701: 2000 iterations of the Markov chain.
702: 
703: 
704: \subsection{Convergence}
705: 
706: In order that likelihood distributions calculated in this paper be considered
707: reliable, it is important to check convergence of the MCMC\@. This is done by
708: running 9 independent Markov chains, each with random starting positions as
709: described above.  
710: The starting positions are chosen in the ranges presented in
711: Table~\ref{tab:prior} 
712: with a flat pdf, since it is important for the convergence measure that the
713: initial values be over distributed compared to the likelihood function one
714: samples from. By examining the variance and means of input parameters within the chains
715: and between the 9 different chains, we will construct a
716: quantity~\cite{stat}
717: $\hat R$. $\hat R$ will provide an upper bound on the factor of expected
718: decrease of 
719: variance of 1d likelihood distributions
720: if  the chain were iterated to an infinite number of steps. 
721: An ${\hat R}$ value can be constructed for scalar quantities that are
722: associated with a point ${\mathbf x}^{(t)}$. 
723: 
724: The analysis of the $\hat R$ convergence statistic follows Ref.~\cite{stat}
725: closely. 
726: We consider $c=1,\ldots,M$ chains ($M=9$ here), each with $N=10^6$ steps. Then
727: we may define the  average input parameter along direction $k$ for the chain
728: $c$ and the average amongst the ensemble of chains 
729: \begin{equation}
730: [\bar{x}_k]_c = \frac{1}{N} \sum_{t=1}^{N} [{x_k^{(t)}}]_c, \qquad
731: \bar{x}_k = \frac{1}{M} \sum_{c=1}^{M} [\bar{x}_k]_c,
732: \end{equation}
733: respectively.
734: The variance of chain $c$ along direction $k$ is
735: \begin{equation}
736: [V_k]_c = \frac{1}{N-1} \sum_{t=1}^N ([x^{(t)}_k]_c - [\bar{x}_k]_c)^2,
737: \end{equation}
738: so that we have the average of the variances within a chain
739: \begin{equation}
740: w_k = \frac{1}{M} \sum_{c=1}^M [V_k]_c
741: \end{equation}
742: and the variance between chains' averages
743: \begin{equation}
744: B_k/N = \frac{1}{M-1} \sum_{c=1}^M ([\bar{x}_k]_c - \bar{x}_k)^2.
745: \end{equation}
746: The basic ratio constructed corresponds to
747: \begin{equation}
748: R_k = \frac{\frac{N-1}{N} w_k + B_k/N(1 + \frac{1}{M})}{w_k}.
749: \end{equation}
750: As long as the initial seed parameters of the Markov chain are
751: over-distributed, i.e.\ they have larger variance than the likelihood, 
752: this ratio will be larger than one~\cite{stat} if the chains have not
753: converged or if they have not had time to explore the entirety of the
754: parameter space. It tends to one only if both of these conditions are met.
755: In order to construct $\hat R$, we must take into account the sampling
756: variability of $[\bar{x}_k]_c$ and $[V_k]_c$.
757: The variance of chain variances along direction $k$ is estimated to be
758: \begin{equation}
759: v_k = \frac{1}{M-1} \sum_{c=1}^M ([V_k]_c - w_k)^2,
760: \end{equation}
761: and we must take into account the following estimates of
762: co-variances between the values of
763: $[\bar{x}_k]_c$ and $[V_k]_c$:
764: \begin{eqnarray}
765: (\sigma_k)_1 &=& - w_k \bar{x}_k^2 + \frac{1}{M}\sum_{c=1}^M [V_k]_c
766:   [\bar{x}_k]_c^2 , \nonumber \\
767: (\sigma_k)_2 &=& - w_k
768:   \bar{x}_k + \frac{1}{M} \sum_{c=1}^M [V_k]_c [\bar{x}_k]_c.
769: \end{eqnarray}
770: Defining the total estimated variance of the target distribution along
771: direction $k$ 
772: \begin{eqnarray}
773: {\mathcal V_k} &=& \frac{1}{M}(1 - \frac{1}{N})^2 v_k +
774: \frac{2(M+1)^2}{M^2(M-1)} 
775: (B_k/N)^2 +\nonumber \\
776: && 
777: 2 \frac{(N-1)(M+1)}{M^2 N} \left( (\sigma_k)_1 - 2 \bar{x}_k 
778: (\sigma_k)_2
779: \right),
780: \end{eqnarray}
781: where we have degrees of freedom
782: \begin{equation}
783: df_k = 2 \frac{\left(R_k w_k + B_k / (NM)\right)^2}{{\mathcal V}_k},
784: \end{equation}
785: leading us to the final equation for the estimated reduction in the sampled
786: variance as $t \rightarrow \infty$:
787: \begin{equation}
788: \hat{R}_k = R_k \frac{df_k}{df_k - 2}.
789: \end{equation}
790: %% saved correct version
791: \begin{figure}
792:   \epsfig{figure=scan7conv.eps,width=10cm}
793:   \caption{Estimate of potential scale reduction shown as a function of the
794:     number of Markov
795:     chain Monte Carlo steps. The upper bound we require for convergence is shown
796:     as a horizontal line. \label{fig:conv}}
797: \end{figure}
798:  
799: %% \begin{figure}
800: %%  \unitlength=1cm
801: %% \begin{picture}(10,6)
802: %%  \put(0,0){\epsfig{figure=scan7m0A0.eps,width=10cm}}
803: %%    \put(2.51,1.28){\epsfig{figure=scan7m0A02.eps,width=6.8cm}}
804: %%  \end{picture}
805: %%    \caption{Estimate of potential scale reduction shown as a function of the
806: %%      number of Markov
807: %%      chain Monte Carlo steps. The upper bound we require for convergence is
808: %%      shown as a horizontal line. \label{fig:conv}}
809: %%  \end{figure}
810: 
811: Here,
812: we define $r \equiv$max$_k \{ \sqrt{\hat R_k}\}$.
813: Values of
814: $r<1.05$ are considered to signify convergence and compatibility of the
815: chains, since we could only hope to decrease the scale of any of the input
816: parameter distributions by at most $5\%$ by performing further Markov Chain
817: steps.  
818: In Fig.~\ref{fig:conv}, we show the quantity $r$ as a function of step
819: number. $r<1.05$ is met already for 600 000 steps indicating adequate
820:   convergence, although for the 
821:   results we present below we we always use the full 9$\times$1 000 000
822:   sample.
823: 
824: \section{Likelihood Maps\label{sec:Lmaps}}
825: 
826: The number of input parameters exceeds the number of data and
827: the likelihood shows a rough degeneracy along
828: directions 
829: which give iso-lines of $\Omega_{DM} h^2$. The parameters of the best-fit
830: point of the MCMC do not 
831: therefore supply us with much information, but the value of the likelihood at
832: that point is interesting: a very small value would indicate a high $\chi^2$
833: and therefore a bad fit. 
834: The best-fit point sampled by the MCMC with 7d input parameter space was 
835: \begin{eqnarray}
836: &&m_0=964 \mbox{~GeV},\ M_{1/2}=341 \mbox{~GeV},\ A_0=1394 \mbox{~GeV},
837: \nonumber \\
838: &&
839: m_b(m_b)=4.18 \mbox{~GeV},\ m_t=173.0 \mbox{~GeV},\ \nonumber \\
840: && \alpha_s(M_Z)=0.1185,\
841: \tan \beta = 57.9, 
842: \end{eqnarray}
843: leading to predictions of $\delta (g-2)_\mu/2=1.8 \times 10^{-9}$, $BR(b
844: \rightarrow s 
845: \gamma)=3.63 \times 10^{-4}$ and $\Omega_{DM} h^2=0.1124$ and corresponding to
846: a 
847: combined likelihood (ignoring normalisation constants, as stated above) 
848: of ${\mathcal L}=0.95$. 
849: The point is within the $A^0-$pole region, and the centrality of predicted
850: observables 
851: gives us confidence that mSUGRA can fit well to current data.
852: The efficiency of the MCMC algorithm is quite small: only 4.1$\%$. This
853: reflects the fact that the thickness of WMAP-allowed volume is small, making
854: it difficult to sample efficiently. 
855: 
856: We display binned sampled likelihood distributions in
857: Figs.~\ref{fig:like7}a-\ref{fig:like7}f for the full $m_t$, $m_b(m_b)$,
858: $\alpha_s(M_Z)^{\overline{MS}}$, $m_0$, $A_0$, $\tan \beta$ and $M_{1/2}$ parameter space. 
859: The unseen dimensions in each figure have been marginalised with flat priors. 
860: All 2-d or 1-d marginalisations in this paper assume a flat prior (within the
861: ranges of parameters considered in Table~\ref{tab:prior})
862: in all unseen dimensions (or in other words, the likelihood is integrated over
863: them with equal weight). One can view the marginalisation probabilistically,
864: or just as a means of viewing the higher dimensional parameter space. 
865: \begin{figure*}
866: \begin{center}
867: \sixgraphsmod{scan7m0m12}{scan7m0tb}{scan7m0A0}{scan7m12tb}{scan7A0tb}{scan7m12A0}
868: \end{center}
869: \caption{Likelihood maps of mSUGRA parameter space.
870: The graphs show the
871: likelihood distributions sampled from
872:   7d parameter space and marginalised down to two.
873: The likelihood (relative to the likelihood in the highest bin) 
874: is displayed by reference to the bar on the right hand side of each plot.
875: The contours show the 68$\%$ and $95\%$ confidence level limits.
876: \label{fig:like7}}
877: \end{figure*}
878: We have used 75$\times$75 bins, normalising the likelihood in each bin to the
879: maximum likelihood in any bin in each 2d plane.
880: 
881: In each plot, the $h^0$ pole $s$-channel resonant annihilation region
882: is present close to the lowest values of $M_{1/2}$. It can be seen as a
883: vertical sliver in the 
884: top-left hand corner of the $m_0-M_{1/2}$ as in Fig.~\ref{fig:like7}a
885: and the 
886: slim band ranging across the bottom of
887: Figs.~\ref{fig:like7}d,\ref{fig:like7}f.
888: The bright region in Fig.~\ref{fig:like7}a at low values of $m_0$ is
889: primarily a co-annihilation 
890: region where slepton-neutralino annihilation contributes significantly to the
891: depletion of the neutralino relic density in the early universe. However, at
892: the lowest values of $m_0$ and $M_{1/2}$ values visible on the graph, 
893: the bulk region resides, being continuously connected to the co-annihilation
894: tail, as shown in Ref.~\cite{Ellis:2003cw}.
895: The pseudoscalar Higgs ($A^0$) $s$-channel annihilation channel occurs at high
896: $\tan \beta=50-60$ and in the intermediate areas of $m_0=500-1600$ GeV,
897: $M_{1/2}=250-1400$ GeV. 
898: In the literature, the most common way to display mSUGRA results is to 
899: present them in 2d in the $m_0$-$M_{1/2}$ plane, where thin strips are
900: observed (see for example Ref.~\cite{Ellis:2004tc}) that are consistent with
901: the WMAP constraint upon $\Omega_{DM} h^2$. 
902: Fig.~\ref{fig:like7}a demonstrates (in corroboration with Refs.~\cite{Baltz:2004aw,Stark:2005mp})
903: that the strips are truly a result of
904: picking a 2d hyper-surface in parameter space: if one performs a full
905: multi-dimensional scan, there is a large region in the $m_0$-$M_{1/2}$ plane
906: that is 
907: consistent with the data. The bottom right hand side corner of
908: Fig.~\ref{fig:like7}a is ruled out primarily by the fact that the lightest
909: supersymmetric particle (LSP) is
910: charged. Large $M_{1/2}$ is disfavoured by the $(g-2)_\mu$ result. The bottom
911: left-hand corner of Fig.~\ref{fig:like7}a is ruled out by a combination of
912: dark matter and direct search constraints. We see an interesting correlation
913: between $m_0$ and $\tan \beta$ in Fig.~\ref{fig:like7}b: the region extending
914: to low $\tan \beta$ and low $m_0$ is essentially the stau co-annihilation/bulk
915: region.  
916: 
917: \begin{figure*}[t]
918: \begin{center}
919: \fourgraphs{scan7mAmchi}{scan7mhmchi}{scan7mstaumchi}{scan7mchimslep}
920: \end{center}
921: \caption{Likelihood distributions of masses in mSUGRA\@. The graphs show the
922: likelihood distributions marginalised down to 2d.
923: The likelihood (relative to the likelihood in the highest bin) 
924: is displayed by reference to the bar on the right hand side of each plot.
925: The contours show the 68$\%$ and $95\%$ confidence level limits.
926: \label{fig:likeMass7}}
927: \end{figure*}
928: We display some binned likelihood distributions of MSSM particle masses
929: in
930: Figs.~\ref{fig:likeMass7}a-\ref{fig:likeMass7}c that are relevant for dark
931: matter annihilation. In 
932: Fig.~\ref{fig:likeMass7}a, the $A^0$-pole resonance region is clearly
933: discernable: at just above a line $2m_{\chi_1^0}=M_{A^0}$ and just below it,
934: there is 
935: just enough annihilation to produce the observed relic density. Throughout much
936: of the parameter space ($M_A<1$ TeV), the exact resonance condition depletes
937: the relic density
938: $\Omega_{DM} h^2$ to be too small. At around $M_A \sim 1$ TeV, exact resonance is
939: required in order to sufficiently deplete the relic dark matter density. 
940: The rest of the likelihood density is spread over the dark part of
941: the plot and is mostly too diffuse to be visible. 
942: In Fig.~\ref{fig:likeMass7}b, there are two detectable regions: the small one
943: with the maximum 2d binned likelihood at the lowest possible values
944: of $m_{\chi_1^0}$ corresponds to the $h^0$-pole region. The larger upper high
945: likelihood region
946: is an amalgam of the co-annihilation and $A^0$-pole regions.
947: As a by-product we see that values of $m_h>126$ GeV are disfavoured in the
948: mSUGRA model. 
949: In
950: Fig.~\ref{fig:likeMass7}c, the stau co-annihilation region is visible as the
951: diagonal dark line and the $h^0$-pole region as the lower likelihood horizontal
952: dark line. The rest of the likelihood density is diffusely distributed in
953: between these two extremes. 
954: We have shown the likelihood distribution for lightest chargino and slepton
955: masses in Fig.~\ref{fig:likeMass7}d. Unfortunately, most of the likelihood 
956: is in a region where the well-known tri-lepton search
957: channel~\cite{Abazov:2005ku} at the Tevatron 
958: is rather difficult. With 8 fb$^{-1}$ of integrated luminosity,
959: this search channel requires
960: $m_{\chi_1^\pm} < m_{\tilde l}$, $m_{\chi_1^\pm}<250$ 
961: for a discovery~\cite{future}. 
962: 
963: \begin{figure*}
964: \unitlength=1.1in
965: \begin{center}
966: \begin{picture}(5.8,6.0)(0,0.2)%4.8 high normally with 2x2
967: \put(0,4.4){\epsfig{file=scan7ms, width=3.4in}}
968: \put(2.9,4.4){\epsfig{file=scan7dmBig, width=3.4in}}
969: \put(3.69,5.1){\epsfig{file=scan7dm, width=0.59\wth}}
970: \put(0,2.1){\epsfig{file=scan7Bsmumu, width=3.4in}}
971: \put(2.4,2.0){\epsfig{file=scan7m12bsmm, width=3.9in}}
972: \put(3.28,2.45){\epsfig{file=scan7m12bsmm2, width=2.66in}}
973: \put(-0.5,-0.2){\epsfig{file=scan7gmuBsmumu, width=3.9in}}
974: \put(0.38,0.25){\epsfig{file=scan7gmuBsmumu2, width=2.66in}}
975: \put(2.4,-0.2){\epsfig{file=scan7mAtanb, width=3.9in}}
976: \put(3.28,0.25){\epsfig{file=scan7mAtanb2, width=2.66in}}
977: \put(0,6.4){(a)}
978: \put(2.9,6.4){(b)}
979: \put(0,4.1){(c)}
980: \put(2.9,4.1){(d)}
981: \put(0,1.8){(e)}
982: \put(2.9,1.8){(f)}
983: \end{picture}
984: \end{center}
985: \caption{(a) Selected sparticle mass likelihood distributions in mSUGRA, 
986: (b) stau-neutralino mass difference likelihood distribution 
987: where the insert
988:   shows a blow-up of the quasi mass-degenerate region.
989: (c) branching ratio for the decay $B_s \rightarrow \mu^+ \mu^-$. 
990: The Tevatron upper bound is displayed by a vertical line.
991: (d) Likelihood density marginalised to the 2d plane
992: $BR(B_s \rightarrow \mu^+ \mu^-)$ versus $M_{1/2}$. (e)
993: Correlation between $BR(B_s \rightarrow \mu^+ \mu^-)$ and $(g-2)_\mu$.
994: (f) Likelihood marginalised to the $\tan \beta$-$M_A$ plane.
995: The contours show the 68$\%$ and $95\%$ confidence level limits.
996: \label{fig:hists}}
997: \end{figure*}
998: 
999: We show the sampled likelihood distribution for $m_{{\tilde l}_R}$,
1000: $m_{\tilde g}$, $m_{{\tilde t}_1}$ and $m_{{\tilde q}_L}$ in
1001: Fig.~\ref{fig:hists}a. The 
1002: likelihood distributions have been placed into 75 bins of 
1003: widths that are equal for the  
1004: different types of sparticles. They are each normalised to an integrated
1005: likelihood of 1. The spike at low values of the gluino mass corresponds to the 
1006: $h^0$-pole region of mSUGRA\@. This spike has a good chance of being covered
1007: at the Tevatron experiments before the LHC starts running~\cite{future}.
1008: It should be noted that upper limits 
1009: upon the scalar sparticle masses inferred from Fig.~\ref{fig:hists}a
1010: are due largely to the definition of the range of
1011: the initial parameters ($m_0$ being less than 2
1012: TeV). Nevertheless, it is clear that there is already some preference from the
1013: combined data for various ranges of sparticle masses and upper bounds upon the
1014: gaugino masses. For example, values of $m_{\tilde g} > 3.5$ TeV and
1015: $m_{\tilde g}=400-800$ GeV are disfavoured, as well as $m_{{\tilde
1016:     q}_L}<800$ GeV. In Fig.~\ref{fig:hists}b, the mass splitting between the
1017: lightest stau and lightest neutralino is shown. The insert in the figure
1018: displays the quasi-degenerate co-annihilation region at low mass splittings. 
1019: The peaked region at $m_{{\tilde \tau}_1} - m_{\chi_1^0}<$10 GeV is likely to
1020: be difficult to discern at the Large Hadron Collider (LHC). 
1021: One would wish to
1022: measure decays of the 
1023: lightest staus in order to check the co-annihilation region, but reconstructing
1024: a relevant soft $\tau$ resulting from such a decay is likely to prove
1025: problematic. On the other hand, a linear collider with
1026: sufficient energy to produce sparticles could provide the necessary
1027: information~\cite{Nojiri:1996fp,Bambade:2004tq}. 
1028: The predicted likelihood distribution of the $B_s \rightarrow \mu^+ \mu^-$ branching
1029: ratio is shown in Fig.~\ref{fig:hists}c. Possible values for this 
1030: observable were 
1031: found with a random scan of unconstrained MSSM parameter space in
1032: Ref.~\cite{Dedes:2004yc} (no likelihood distribution was given).
1033: The region to the right hand side of
1034: the vertical line is ruled out from the combined
1035: CDF/D0 95$\%$C.L. exclusion~\cite{Acosta:2004xj,Abazov:2004dj}\footnote{There
1036:   are newer CDF/D0 bounds~\cite{Dagu}, for example CDF(D0) have non-combined
1037:   95$\%$ C.L. limits of 2.0(3.0)$\times 10^{-7}$ respectively.}
1038: limit $BR(B_s^0 \rightarrow \mu^+
1039: \mu^-) < 3.4 \times 10^{-7}$. 
1040: We have not cut points violating this
1041: constraint, but this has negligible effect
1042: since there are
1043: only a small number of them. The 2d marginalisation of
1044: the branching ratio versus $M_{1/2}$ shows a peak at very low $M_{1/2}$
1045: values, as Fig.~\ref{fig:hists}d displays.
1046: This indicates that the spike in Fig.~\ref{fig:hists}c at branching ratios
1047: of about $10^{-8.6}$ is due to the $h^0$-pole region. 
1048: Lowering the empirical
1049: upper bound on the $B_s \rightarrow \mu^+ \mu^-$ branching ratio will
1050: significantly cut into the allowed mSUGRA parameter space. 
1051: A significant lowering of the bound upon this branching ratio is expected 
1052: in the coming years from the Tevatron experiments and from LHCb.
1053: For example, it is estimated~\cite{future} that the Tevatron could exclude 
1054: a branching ratio of more than $2 \times 10^{-8}$ with 8 fb$^{-1}$ of
1055: integrated luminosity. This corresponds to ruling out 29$\%$ of the
1056: currently allowed likelihood density. 
1057: Predictions for $BR(B_s \rightarrow \mu^+ \mu^-)$ were correlated
1058: with those for $(g-2)_\mu$ in Ref.~\cite{Dedes:2001fv} in mSUGRA\@. For a given
1059: mSUGRA 
1060: parameter point, a correlation was shown when $\tan \beta$ was varied. 
1061: The authors conclude that for high $\tan \beta$, 
1062: an enhancement of 
1063: $BR(B_s \rightarrow \mu^+ \mu^-)$
1064: is implied by the $(g-2)_\mu$ measurements. We re-examine this statement in view
1065: of the full dimensionality of the mSUGRA parameter space in
1066: Fig.~\ref{fig:hists}e. The correlation is seen to be far from perfect, the
1067: likelihood distribution being highly smeared in terms of the two measurements.
1068: Nevertheless, there is a mild correlation between $BR(B_s \rightarrow \mu^+
1069: \mu^-)$ and the SUSY contribution to $(g-2)_\mu$ 
1070: at the highest likelihoods, as evidenced by the bright
1071: oblique stripe in Fig.~\ref{fig:hists}e.
1072: In Fig.~\ref{fig:hists}f, we show the likelihood distribution in the $\tan
1073: \beta$-$m_A$ plane. There is a significant amount of likelihood density
1074: towards the top left-hand side of the plot, where the Tevatron is expected to
1075: have sensitivity~\cite{future} (covering $\tan \beta>40$ for $m_A<240$ GeV for
1076: 8fb$^{-1}$ of integrated luminosity). LHC experiments should be able to 
1077: observe the $A^0$ for
1078: the entire high likelihood range~\cite{Armstrong:1994it}.
1079: 
1080: We now ask what are the likelihoods of the 
1081: different regions of relic density depletion in mSUGRA parameter
1082: space. In order to sharply delineate the
1083: regions, 
1084: we define them as follows: 
1085: the co-annihilation region is defined such that $m_{\chi_1^0}$ is within $10\%$
1086: of $m_{{\tilde \tau}_1}$.
1087: The $h^0/A^0$-pole regions have $2m_{\chi_1^0}$ within $10\%$ of $m_{h^0}$,
1088: $m_{A^0}$ respectively.
1089: Stop co-annihilation requires a broader definition: it is defined such that 
1090: $m_{\chi_1^0}$ is within 30$\%$ of $m_{{\tilde t}_1}$, since the
1091: annihilation is so much
1092: more efficient~\cite{Boehm:9911,arnie,Ellis:2001nx} than in the other regions.
1093: A better defined procedure might perhaps be to determine regions on the basis
1094: of the {\em dominant}\/ annihilation mechanism, but since we are only looking
1095: for a rough indication of the region involved, the procedure adopted here
1096: will suffice. 
1097: Points that fall in between any of the sharp definitions are either 
1098: from the bulk region or in the smaller tails of the likelihood distribution.
1099: 
1100: \begin{table}
1101: \caption{Likelihood of being in a certain region of mSUGRA parameter
1102:   space. \label{tab:regions}}
1103: \begin{tabular}{|c|c|}
1104: \hline Region & likelihood \\ \hline
1105: $h^0$ pole & 0.02$\pm$0.01 \\
1106: $A^0$ pole & 0.41$\pm$0.03 \\
1107: $\tilde \tau$
1108: co-annihilation & 0.27$\pm$0.04 \\ 
1109: $\tilde t$ co-annihilation & $(2.1\pm4.8)\times 10^{-4}$\\ \hline
1110: \end{tabular}
1111: \end{table}
1112: The likelihoods of these regions are shown in Table~\ref{tab:regions}.
1113: We estimate the uncertainty by calculating the standard deviation on the 9
1114: independent Markov chain samples. The quoted error thus reflects an
1115: uncertainty due not to experimental errors, but to a to finite
1116: simulation time of the Markov chain.
1117: We see that the $h^0$-pole region has a relatively low likelihood whereas for
1118: the $A^0$-pole and $\tau$ co-annihilation regions the likelihood is larger. 
1119: %These likelihoods do not have add to a total of one for several
1120: %reasons. Firstly, some of the points 
1121: %sampled will be a few sigma away from the $\Omega_{DM} h^2$ constraint and may
1122: %fall 
1123: %outside of our definitions of the individual regions. Also, since 
1124: %the full 7d parameter space does not provide a sharp distinction
1125: %between the various regions, there will be some points for which several
1126: %mechanisms significantly contribute to the relic density annihilation
1127: %mechanism, but for which none of the criteria that we have set in order to
1128: %pigeon-hole the point are quite satisfied. 
1129: From the table, we see that the $\tilde t$-co-annihilation region, although
1130: uncertain due to the low statistics, is
1131: negligible, and we now investigate why this is the case.
1132: 
1133: The suppression of the stop-co-annihilation region comes from essentially two
1134: effects: firstly, as already apparent from Ref.~\cite{Ellis:2001nx}, finding a
1135: suitable stop co-annihilation region which is compatible with {\em both} the
1136: $(g-2)_\mu$ and $BR[b   \rightarrow s \gamma]$ measurements is problematic. 
1137: Secondly, the central value of $m_t$ has come down since
1138: ref.~\cite{Ellis:2001nx}. The dominant radiative corrections to $m_{h^0}$ are 
1139: highly correlated with $m_t$~\cite{Allanach:2004rh}, 
1140: with the consequence that the lower predicted Higgs mass 
1141: now rules out more of the stop co-annihilation region. 
1142: We illustrate these points in Fig.~\ref{fig:stopCoan} along the $m_0$
1143: direction for given values of the other mSUGRA parameters (stated in the
1144: caption). 
1145: \begin{figure*}
1146: \unitlength=1.1in
1147:  \begin{picture}(5.8,2.6)(0.5,0)
1148:  \put(0.2,0){\epsfig{file=mt172a.eps, width=0.9\wth}}
1149:  \put(3.2,0){\epsfig{file=mt172b.eps, width=0.9\wth}}
1150:  \put(0.2,2.1){(a)}
1151:  \put(3.1,2.1){(b)}
1152:  \end{picture}
1153: \caption{A line through the stop co-annihilation region in mSUGRA: $\tan
1154:   \beta=10$,  $M_{1/2}=350$ GeV, $A_0=-2000$ GeV and central Standard Model
1155:   inputs. (a) dark matter relic density, fractional stop-neutralino mass
1156:   splitting $\Delta \equiv (m_{{\tilde t}_1} - m_{\chi_1^0}) / m_{\chi_1^0}$ 
1157:   and light Higgs mass. The horizontal lines show the 1$\sigma$ WMAP-derived
1158:   limits upon $\Omega_{DM} h^2$ derived from Eq.~\protect\ref{omConst}, (b)
1159:   muon magnetic moment and $BR[b \rightarrow s \gamma]$ predictions. 
1160: \label{fig:stopCoan}}
1161: \end{figure*}
1162: In Fig.~\ref{fig:stopCoan}a, we plot the fractional stop-neutralino
1163: mass splitting $\Delta$ 
1164: alongside the neutralino relic density $\Omega_{DM} h^2$. 
1165: We see that the fractional mass splitting takes values between 0.1 and 0.23 in
1166: the range of $m_0$ shown. This is the stop-co-annihilation regime, and we see
1167: that around $m_0 \sim 786$ GeV, $\Delta \sim 0.2$ corresponding to
1168: $\Omega_{DM} h^2$ roughly compatible with the WMAP constraint. Unfortunately,
1169: we also see that the lightest CP-even Higgs mass is predicted to be 110.3
1170: GeV here and is ruled out by the LEP2 Higgs limits shown in Table~\ref{tab:par}
1171: ($\sin^2 (\beta-\alpha) =1.0$) for this range of parameters). This problem is
1172: remedied by going to higher values of $m_t$, since $m_{h^0}$ then goes up, but
1173: of course this comes with an associated penalty in the likelihood from being
1174: away from the empirically central values of $m_t$. 
1175: In Fig.~\ref{fig:stopCoan}b, we display predictions for $BR[b\rightarrow s
1176:   \gamma]$ and $\delta (g-2)_\mu/2$ along the chosen range for $m_0$. 
1177: These predictions are both lower than the empirically derived constraints in
1178: the region where the dark matter relic density is in accordance with the WMAP
1179: constraint: including errors in the theoretical prediction as described in the
1180: previous section, 
1181: $BR[b \rightarrow s \gamma]$ is 1.95$\sigma$ lower than the
1182: central value and $\delta (g-2)_\mu/2$ is 1.6$\sigma$ lower. It turns out that
1183: these predictions
1184: are {\em not} very sensitive to changes in $m_t$ and so their likelihood
1185: penalty tends to apply for the higher values.
1186: However, lower values of $(-A_0)$
1187: require lower $m_0$ in order to fit the dark matter constraint and pick up 
1188: a bigger likelihood constraint from the egregious prediction of
1189: $BR[b\rightarrow s \gamma]$.  
1190: These findings are in rough agreement with those of Ref.~\cite{Ellis:2001nx}
1191: except for the more restrictive Higgs bounds, which are a consequence of the
1192: lower experimental value of $m_t$. Ref.~\cite{Ellis:2001nx} only applies
1193: 2$\sigma$ bounds on both $(g-2)_\mu$ and $BR[b \rightarrow s \gamma]$, whereas
1194: our results take into account the likelihood penalty paid by the fact that
1195: neither prediction is close to its central value near the stop co-annihilation
1196: region, which then becomes disfavoured compared to the other regions (where
1197: an almost perfect fit is possible). 
1198: There is also a volume effect: in Table~\ref{tab:regions}, the likelihoods we 
1199: calculate are {\em integrated} over the relevant region. Thus regions that are
1200: very small, such as the stop co-annihilation region, will tend to have a smaller
1201: likelihood than other, larger regions. 
1202: In analyses in following sections, stop 
1203: co-annihilation also turns out to have negligible likelihood and so we will 
1204: neglect it from the results. 
1205: 
1206: As an aside, we note that the decay chain ${{\tilde q}_L} \rightarrow
1207: {\chi_2^0} \rightarrow {\tilde l}_R \rightarrow \chi_1^0$ exists with a
1208: likelihood  
1209: of $0.24\pm0.04$ (this number is just based upon the mass ordering and does
1210: not take into account the branching ratio for the chain). The existence of
1211: such a chain 
1212: allows the extraction of several functions of sparticle masses from kinematic
1213: end-points and they have been used in many LHC analyses,
1214: for example Refs.~\cite{Armstrong:1994it,Allanach:2000kt,Weiglein:2004hn}. 
1215: 
1216: 
1217: \section{Theoretical Uncertainty \label{sec:uncertainties}}
1218: 
1219: \begin{table}
1220: \caption{Likelihood of being in a certain region of mSUGRA parameter
1221:   space including theoretical uncertainties in the sparticle spectrum
1222:   calculation. \label{tab:regions2}} 
1223: \begin{tabular}{|c|c|}
1224:   \hline Region & likelihood \\ \hline
1225: $h^0$ pole & 0.03$\pm$0.01 \\
1226: $A^0$ pole & 0.41$\pm$0.05 \\
1227: co-annihilation & 0.26$\pm$0.08 \\ \hline
1228: \end{tabular}
1229: \end{table}
1230: Theoretical uncertainties in the sparticle mass predictions have been shown to
1231: produce non-negligible effects in fits to data~\cite{Allanach:2003jw},
1232: including fits to the relic
1233: density~\cite{Allanach:2004jh,Allanach:2004xn,Belanger:2005jk}. 
1234: In this section, we perform a second MCMC analysis taking theoretical
1235: uncertainty into account in order to estimate the size of its effect.
1236: {\tt SOFTSUSY} performs the Higgs potential minimisation then 
1237: calculates sparticle pole masses at a scale $M_{SUSY}=\sqrt{m_{{\tilde t}_1}
1238:   m_{{\tilde t}_2}}$. This scale is chosen because it is hoped that loop
1239: corrections to the pole mass corrections that are not yet taken into account
1240: (typically two-loop corrections) are small at this scale. 
1241: In order to estimate the size of theoretical uncertainties, we vary this
1242: scale by a factor of two in either direction (but it is always constrained 
1243: to be greater than $M_Z$)\footnote{A more complete estimate would be to 
1244:  calculate the Higgs
1245: potential minimisation conditions and the sparticle
1246: masses all at different scales, varying each independently by a factor of
1247: two. However, such a prescription is impractical here due to CPU time
1248: constraints.}. 
1249: Implementation of the uncertainty in the MCMC algorithm is
1250: simple: we simply add an input
1251: parameter $x$ which is bounded between 0.5 and 2, giving the factor by which
1252: $M_{SUSY}$ is to be multiplied. The MCMC is then re-run as before and explores
1253: the  full 8d parameter space (including $x$) accordingly. 
1254: 
1255: Such a
1256: procedure automatically takes into account the correlations in predictions due
1257: to correlated theoretical uncertainties in the sparticle mass predictions.
1258: The likelihoods of the three mSUGRA regions are shown in
1259: Table~\ref{tab:regions2}. The likelihoods are approximately equal to those for
1260: the 7d case, as a comparison to Table~\ref{tab:regions} indicates.
1261: In fact, comparing
1262: results and plots produced with and without theoretical uncertainties, 
1263: the results are generally very similar. 
1264: This indicates that the theoretical uncertainties don't
1265: make a huge difference to the 1d and 2d marginalisations
1266: compared to those coming from the data.
1267: The decay chain ${{\tilde q}_L} \rightarrow
1268: {\chi_2^0} \rightarrow {\tilde l}_R \rightarrow {\chi_1^0}$ has a likelihood
1269: of $0.22 \pm 0.08$, not significantly 
1270: different to the case when theoretical uncertainties were not taken into
1271: account. 
1272: 
1273: We show
1274: two of 
1275: the  2d marginalised likelihoods in
1276: Figs.~\ref{fig:uncer}a,\ref{fig:uncer}b. 
1277: Comparing with Fig.~\ref{fig:like7}, we see that Fig.~\ref{fig:uncer} shows no
1278: significant effects deriving from theoretical errors.
1279: %% two main effects: the Higgs pole region at $m_0 \sim M_{1/2}$ becomes somewhat
1280: %% reduced in importance and the co-annihilation region dominates the likelihood
1281: %% over the Higgs pole region more. 
1282: %% Fig.~\ref{fig:uncer}b also displays the 
1283: %% enhanced low $m_0$ co-annihilation region when compared to
1284: %% Fig.~\ref{fig:like7}c.  
1285: Marginalising mass likelihood distributions down to 1d (as in
1286: Fig.~\ref{fig:hists}a for example), one obtains distributions that are
1287: identical to  the 7d case except for small statistical fluctuations. 
1288: \begin{figure*}
1289: \twographst{scan8m0m12mod}{scan8m0A0}
1290: \caption{Likelihood maps of mSUGRA parameter space including theoretical
1291:   uncertainty. 
1292: The graphs show the
1293: likelihood distributions sampled from
1294:   8d parameter space and marginalised down to two.
1295: The likelihood (relative to the likelihood in the highest bin) 
1296: is displayed by reference to the bar on the right hand side of each plot.
1297: The contours show the 68$\%$ and $95\%$ confidence level limits.
1298: \label{fig:uncer}}
1299: \end{figure*}
1300: 
1301: \section{Other Sources of Dark Matter\label{sec:other}}
1302: For completeness, we may ask how robust our results are with respect to the
1303: assumption that non thermal-neutralino contributions to $\Omega_{DM}$ are
1304: negligible.  
1305: Thus, we allow the predicted amount of thermal neutralino relic density
1306: $p_{\Omega_{DM} h^2}$ (assuming some mSUGRA point $s$) to be some fraction of
1307: the total relic predicted relic density $\Omega_{DM}^{tot} h^2$:
1308: \begin{equation}
1309: \lambda = \frac{p_{\Omega_{DM} h^2}}{\Omega_{DM}^{tot} h^2}, \qquad 0 \leq
1310: \lambda \leq 1.
1311: \end{equation}
1312: Thus, the total amount of dark matter predicted is 
1313: $p_{\Omega_{DM} h^2} / \lambda$ and,
1314: assuming a flat pdf for $\lambda$ as shown in Fig.~\ref{fig:prior}a,
1315: we obtain the likelihood penalty for a given SUSY dark matter prediction
1316: \begin{widetext}
1317: \begin{equation}
1318: {\mathcal L}_{\Omega_{DM} h^2} = 
1319: \int^1_0d \lambda \ 
1320: \frac{1}{\sqrt{2 \pi} s_{\Omega_{DM} h^2}(\lambda)}
1321: \exp \left[ \frac{-(m_{\Omega_{DM} h^2} - 
1322: p_{\Omega_{DM} h^2} / \lambda)^2}{2 s_{\Omega_{DM} h^2}(\lambda)^2}
1323: \right], \label{lowOm}
1324: \end{equation}
1325: \end{widetext}
1326: where 
1327: \begin{eqnarray}
1328: s_{\Omega_{DM} h^2}(\lambda) &\equiv& 
1329: 0.0081\ \theta(p_{\Omega_{DM} h^2} /
1330: \lambda - m_{\Omega_{DM} h^2}) +\nonumber \\
1331: &&
1332: 0.0091\ \theta(-p_{\Omega_{DM} h^2} /
1333: \lambda + m_{\Omega_{DM} h^2})
1334: \end{eqnarray}
1335: in accordance with the asymmetric errors in Eq.~\ref{omConst}.
1336: $\theta(x)$ is the Heavisde step function, $\theta(x)=1$ for all $x\geq 0$,
1337: $\theta(x)=0$ for all $x<0$.
1338: We calculate Eq.~\ref{lowOm} numerically and display it in
1339: Fig.~\ref{fig:prior}b, where it is contrasted with the old dark matter
1340: likelihood penalty that assumes that all dark matter is of thermal neutralino
1341: origin. 
1342: \begin{figure*}
1343: \twographs{prior}{l}
1344: \caption{(a) $p(\lambda | m_{\Omega_{DM} h^2})$, the probability distribution
1345:   assumed for $\lambda$,  
1346: the ratio of predicted SUSY
1347: dark matter to total dark matter, 
1348: given some measured value $m_{\Omega_{DM} h^2}$. 
1349: (b) Comparison of the likelihood     penalty 
1350:   ${\mathcal L}_{\Omega_{DM} h^2}$
1351:     paid for a prediction of SUSY dark matter $p_{\Omega_{DM} h^2}$ with 
1352: (``extra DM allowed'') and
1353:       without (``no extra DM allowed'') 
1354:       the allowance of non thermal-neutralino dark matter. \label
1355:     {fig:prior}}
1356: \end{figure*}
1357: The figure shows that if the relic density is too {\em high}, a severe
1358: likelihood penalty applies (similar to the ``no extra DM allowed'' case)
1359: but a much less severe penalty applies
1360: if the prediction is below the central value
1361: of the observed WMAP value in Eq.~\ref{omConst}. 
1362: The additional contribution to $\Omega h^2$ is assumed 
1363: to be provided by some non thermal-neutralino source (late decays or hidden
1364:     sector dark matter for example)~\cite{leszek}. 
1365: The rest of the analysis proceeds exactly as in section~\ref{sec:Lmaps} (i.e.\
1366: without simultaneously taking theoretical uncertainty into account). 
1367: \begin{table}
1368: \caption{Likelihood of being in a certain region of mSUGRA parameter
1369:   space including possible an additional contribution from
1370:   non thermal-neutralino dark matter. \label{tab:regions3}}  
1371: \begin{tabular}{|c|c|}
1372: \hline Region & likelihood \\ \hline
1373: $h^0$ pole & 0.04$\pm$0.01 \\
1374: $A^0$ pole & 0.52$\pm$0.02 \\
1375: co-annihilation & 0.14$\pm$0.02 \\ \hline
1376: \end{tabular}
1377: \end{table}
1378: 
1379: The MCMC algorithm turns out to be much more efficient once we drop the
1380: assumption that the cold dark matter consist only of neutralinos:
1381: 19.9$\%$ efficiency was achieved compared to 4.1$\%$. One consequence of this
1382: is that statistical fluctuations in the results are smaller. 
1383: The likelihoods in each region are shown in Table~\ref{tab:regions3}. 
1384: A comparison of Tables~\ref{tab:regions},\ref{tab:regions3} shows that
1385: annihilation through $A^0$ pole has acquired a significantly larger likelihood
1386: through allowing for other forms of dark matter. The higgs-pole and
1387: co-annihilation regions are still, within statistics, compatible with their
1388: previous likelihoods. All of the listed uncertainties have decreased, due to
1389: the additional efficiency.
1390: 
1391: \begin{figure*}
1392: \twographst{scan7lm0m12}{scan7lm0A0}
1393: \caption{Likelihood maps of mSUGRA parameter space allowing for non
1394:   thermal-neutralino contributions to the dark matter relic density. 
1395: The graphs show the
1396: likelihood distributions sampled from
1397:   7d parameter space and marginalised down to two.
1398: The likelihood (relative to the likelihood in the highest bin) 
1399: is displayed by reference to the bar on the right hand side of each plot.
1400: The contours show the 68$\%$ and $95\%$ confidence level limits.
1401: \label{fig:noLow}}
1402: \end{figure*}
1403: Fig.~\ref{fig:noLow}a shows the likelihood map marginalised to the
1404: $M_{1/2}-m_0$ plane. Comparing it to Fig.~\ref{fig:like7}a, we see a very
1405: similar picture except for the fact that the higher volume of likelihood in
1406: the $A^0$-pole region is evident. The same comment can be made of all of the
1407: plots analogous to the ones in to Fig.~\ref{fig:like7}: we show the
1408: likelihood marginalised to the $m_0-A_0$ plane in Fig.~\ref{fig:noLow} as an
1409: example.  There are no other qualitative changes in any of the plots, and
1410: indeed the 
1411: likelihood distributions marginalised to the $A_0-\tan \beta$ and
1412: $M_{1/2}-A_0$ planes (analogous to Figs.~\ref{fig:like7}e,f) 
1413: are identical by eye, except for being smoother due to
1414: the increased efficiency. Likelihoods of sparticle masses also look the same,
1415: except for the spike in the gluino mass, which has twice as much
1416: likelihood. As mentioned before, this spike is due mainly 
1417: to the light $h^0$-pole
1418: region which is subject to relatively large fluctuations, as
1419: Tables~\ref{tab:regions},\ref{tab:regions3} illustrate. We cannot conclude
1420: that the $h^0$-pole region obtains more integrated
1421: likelihood by admitting non thermal-neutralino components to the relic
1422: density because the statistics in the MCMC algorithm are not high enough. 
1423: 
1424: One distribution that does significantly change shape is that of $BR(B_s
1425: \rightarrow \mu^+ \mu^-)$. We show its marginalised distribution after
1426: dropping the lower likelihood penalties on $\Omega_{DM} h^2$ from the MCMC
1427: algorithm procedure in Fig.~\ref{fig:BsmumuNo} as the histogram
1428:  marked ``extra DM allowed''. 
1429: For the purpose of comparison, the default calculation where we assume all
1430: dark matter to be thermal neutralinos from 
1431: Fig.~\ref{fig:hists}c is also displayed, being marked ``no extra DM allowed''.
1432: Comparing the two distributions, 
1433: we see a broader distribution due to the enhanced
1434: $A^0$-pole  annihilation region when additional components are allowed in the
1435: fit. The $A^0$-pole region has higher $\tan \beta$, and therefore higher
1436: values for the branching ratio.
1437: \begin{figure}
1438: \epsfig{file=BsmumuNo, width=\wth}
1439: \caption{Investigation on the effects of allowing for a non
1440:   thermal-neutralino component of dark matter in the 
1441:   branching ratios for the decay $B_s \rightarrow \mu^+
1442:   \mu^-$ in mSUGRA. The current
1443:   Tevatron upper limit is displayed by a vertical line.
1444: \label{fig:BsmumuNo}} 
1445: \end{figure}
1446: The estimated amount of likelihood that could be covered by Tevatron
1447: measurements with 8 fb$^{-1}$ of integrated luminosity increases by 6$\%$ to 
1448: 35$\%$ of the currently allowed density due to the presence of non
1449: thermal-neutralino dark matter contributions. 
1450: 
1451: \section{Conclusions \label{sec:conc}}
1452: 
1453: Previous studies of mSUGRA in the context of dark matter and particle physics
1454: constraints have tended to not use the full dimensionality of the parameter
1455: space, and to have put hard 95$\%$  (C.L.) limits on
1456: predictions. Here, in the full dimensionality of parameter space, we include
1457: all of the information in a likelihood fit, so that violating one constraint
1458: slightly might be traded against fitting another one better in a consistent
1459: manner. 
1460: Although there is plenty of qualitative information about possible dark matter
1461: annihilation regions in mSUGRA in 
1462: the literature, this paper gives a quantitative calculation of the
1463: likelihood distributions in the full dimensionality of the parameter
1464: space. However, the 
1465: most important contribution of this paper lies in the implications of the
1466: results to particle physics. 
1467: 
1468: We have successfully employed an MCMC algorithm to provide 
1469: likelihood maps of the full 7d input parameter space of mSUGRA\@. 
1470: By using a statistical test, we have shown that the likelihood distributions
1471: have achieved good convergence before a total of 9$\times 10^6$ samplings of
1472: the likelihood. 
1473: We have presented the likelihood marginalised down to each 2d mSUGRA parameter
1474: pair. Such plots provide the totality of the
1475: current information we have about the model given the experimental
1476: constraints and are quantitative results. Theoretical
1477: uncertainties in the  
1478: sparticle spectrum calculation broaden a couple of the distributions a little
1479: but do not change them radically. 
1480: 
1481: The main new contribution of this paper is to our knowledge of what current
1482: constraints on mSUGRA mean for particle physics in a quantitative sense. 
1483: Marginalised 1d likelihood distributions of quantities such as sparticle
1484: masses (or mass differences) 
1485: already show some significant structure from the data, providing
1486: interesting information for future collider searches. In particular, the
1487: likelihood of the ``golden cascade'' 
1488: ${\tilde q}_L \rightarrow \chi_2^0 \rightarrow {\tilde l}_R
1489: \rightarrow \chi_1^0$ being kinematically allowed is
1490:   24$\pm 4\%$. $m_{{\tilde \tau}_1} - m_{\chi_1^0}$ is peaked below 10 GeV,
1491: implying that stau reconstruction at hadron colliders could be
1492: problematic at the LHC\@. Integrating the likelihood density, we find the
1493: likelihood of $m_{{\tilde \tau}_1} - m_{\chi_1^0}<10$ GeV is 19.9$\%$.
1494: Our likelihood distributions for $BR(B_s \rightarrow \mu^+ \mu^-)$ corroborate
1495: the 
1496: conclusions of ref.~\cite{Ellis:2005sc}: that current bounds upon the
1497: branching ratio do not yet place significant constraints upon mSUGRA once
1498: other constraints have been taken into account, but any improvement on the
1499: upper bounds constrain the currently available parameter space.  
1500: The quantitative results on $BR(B_s \rightarrow \mu^+ \mu^-)$ are particularly
1501: important: if the Tevatron experiments can reach down to 2$\times 10^{-8}$,
1502: they will cover 29$\%$ of the mSUGRA likelihood, or 35$\%$ if we allow the
1503: possibility of additional contributions to the relic density other than 
1504: thermal neutralinos.
1505: 
1506: We have shown that the correlation between $BR(B_s \rightarrow \mu^+ \mu^-)$
1507: and $\delta 
1508: (g-2)_\mu$ noticed in Ref.~\cite{Dedes:2001fv} is much diluted once
1509: simultaneous variations
1510: of all mSUGRA parameters are taken into account. 
1511: The $A^0$-pole annihilation region
1512: and the stau co-annihilation region each have approximately
1513: an order of magnitude more likelihood than the $h^0$-pole region.
1514: Stop co-annihilation is highly 
1515: disfavoured compared to these other regions due to
1516: more restrictive Higgs mass constraints coming from a lower value of $m_t$, 
1517: as well as the $BR(B_s \rightarrow \mu^+ \mu^-)$ and $\delta 
1518: (g-2)_\mu$ predictions. 
1519:   The light $h^0-$pole region just survives the LEP2 Higgs
1520:   mass constraints despite the new reduced 
1521:   top mass value albeit with a reduced likelihood (in the usual frequentist
1522:   language, it's outside the 95$\%$ confidence level but not the 99$\%$ one).
1523:   The light $h^0$-pole region has more likelihood if one allows additional
1524:   non thermal-neutralino components of dark matter. 
1525: 
1526: The analysis of
1527:   Ref.~\protect\cite{Ellis:2004tc} 
1528: includes $M_W$ and $\sin^2 \theta_{eff}^l$ in the $\chi^2$ statistic, excluding
1529: $M_{1/2}>1500(600)$ GeV at the 90$\%$ confidence level for $m_t=178$ GeV and
1530:   $\tan \beta=50(10)$ respectively. These numbers are not exactly reproduced
1531:   in the present analysis for several reasons: we use a more up-to-date value
1532:   of $m_t$, we vary $m_t$, $m_b$, $\alpha_s (M_Z)$ and $\tan \beta$
1533:   simultaneously with the other mSUGRA parameters and we do {\em not}\/ include
1534:   $M_W$, $\sin^2 \theta_{eff}^l$ in 
1535:   the fit. Indeed, one may wonder about introducing a posteriori bias as a result
1536:   of only picking these two precision electroweak observables, since they are
1537:   the two that show a preference for a SUSY contribution. Other observables
1538:   would presumably prefer heavier SUSY particles. 
1539:   However, $M_W$ and $\sin^2 \theta_{eff}^l$ do show a preference for lower
1540:   $M_{1/2}$ for $m_t=178$ GeV. Having said that, our results are not wildly
1541:   different, as an examination of Fig.~\ref{fig:like7}a shows. 
1542: 
1543: We suspect that the MCMC techniques exemplified here could be found 
1544: extremely useful in SUSY fitting programs such as {\tt
1545:   FITTINO}~\cite{Bechtle:2004pc} and {\tt SFITTER}~\cite{Lafaye:2004cn} in
1546: order to provide a likelihood profile of the 
1547: parameter space, including secondary local minima. These programs are designed
1548: to fit more general MSSM models than mSUGRA to data, with an associated
1549: increase in the number of free input parameters, so the linear calculating
1550: time of MCMCs ought to be very useful.
1551: 
1552: While our results presented for mSUGRA  are in themselves interesting, it is
1553: obvious that the method will be applicable in a much wider range of
1554: circumstances. Once new observables become relevant, such as some LHC
1555: end-points, for example, it would be trivial to include them into the
1556: likelihood and re-perform the MCMC~\cite{whitey}. 
1557: The method should be equally applicable to
1558: other models, and provided enough CPU power is to hand, could provide
1559: likelihood maps for models with even more parameters and/or detailed
1560: electroweak fits.
1561: 
1562: \acknowledgments
1563: We would like to thank other members of the Cambridge SUSY Working Group,
1564: W de Boer, 
1565: P Gondolo, P H\"afliger, B Heinemann, S Heinemeyer, G Manca, A Peel, M Rauch
1566: and T Plehn 
1567: for helpful conversations. This work has been partially supported by PPARC. 
1568: 
1569: \appendix
1570: \section{Toy Models and Sampling}
1571: \label{sec:badsampling}
1572: 
1573: By definition, a sampler able to sample correctly from a pdf
1574: $p({\mathbf x})$ must generate a list of ${\mathbf x}$ values
1575: whose local  
1576: density, at large step-numbers, is proportional to the probability density
1577: $p({\mathbf x})$ at each part of the space (in the preceding parts of the
1578: paper we have set this pdf to be the likelihood).
1579: The value of the constant of proportionality between the probability
1580: density and the local density of ${\mathbf x}$ values is unimportant, but the
1581: key point is that it is {\em constant}\/ across the whole space.
1582: 
1583: It can sometimes be hard to implement a good sampler for a given
1584: probability distribution.  In fact it is often easier to invent an
1585: algorithm which generates a sequence of ${\mathbf x}$ values, and which may
1586: superficially resemble a sampler, {\em but which lacks constancy of
1587: proportionality over the space}.  We might call such algorithms
1588: pseudo-samplers as their output can sometimes resemble that of a
1589: true sampler, provided that the variation in proportionality is not
1590: too great across the space considered.  Pseudo-samplers are sometimes
1591: useful (e.g.\ as a means of exploring a multidimensional space
1592: in which case sample density may be neither interesting nor the
1593: end product of the analysis).
1594: 
1595: In the present and in many other papers, however, sample density represents
1596: confidence in some particular part of parameter space, and {\em is}\/
1597: the final product of the analysis.  Extreme care, then, must be taken
1598: to ensure that any creative modifications to established Markov
1599: Chain sampling techniques do not break the principle of
1600: detailed balance -- the test which ensures that the algorithm remains a
1601: true sampler rather than a pseudo-sampler.
1602: 
1603: It is often desirable for Metropolis-Hastings type samplers to have an
1604: efficiency of about 25\% for the acceptance of newly proposed points.
1605: Efficiencies much smaller than this may suggest that the proposal
1606: distribution is too wide and is too often proposing jumps to
1607: undesirable locations far away from the present point, leading to large
1608: statistical fluctuations in the result.  Efficiencies
1609: much larger than this can be indicative of proposal distributions
1610: which are too narrow and may take too many steps to be practical
1611: to random-walk 
1612: from one side of a region of high probability to the other.
1613: It is tempting, therefore, to adapt the present step size
1614: (i.e.\ proposal distribution width) on the basis of recent efficiency.
1615: With a couple of toy examples, however, we will illustrate that 
1616: this is a dangerous path to follow, and we will demonstrate that it break the
1617: principle of detailed balance, and thus turns the Markov Chain algorithm from a
1618: sampler into a pseudo-sampler.
1619:  
1620: We take as our example the method used by Baltz and
1621:   Gondolo~\cite{Baltz:2004aw} which is 
1622:   designed to keep the target efficiency of the authors' Markov Chain
1623:   close to 25\%:
1624: \begin{enumerate}
1625: \item
1626: Double the current step size if the last
1627: three proposed points were all accepted
1628: \item
1629: Halve the current step size of the last seven proposed points were all rejected.\end{enumerate}
1630: We will refer the the above method as ``the adaptive algorithm'' and show
1631: that it fails to sample correctly from some simple distributions, a signature
1632: of detailed balance being broken.
1633: 
1634: %% \FIGURE{
1635: %% \twographschris{goodPathology}{badPathology}
1636: %% \caption{Part (a) and (b) show two different attempts to sample two
1637: %%   million points from the ``double top hat'' $p_{dTH}(x)$ of
1638: %%   equation~(\ref{eq:blocdist}).  The sampling in part (a) is done
1639: %%   correctly.  The sampling in part (b) is incorrect and shows what can
1640: %%   go wrong when a property of the proposal distribution (in this case
1641: %%   its width) is varied on the basis of the rejection/acceptance
1642: %%   history of the Markov Chain.  See also figure~\ref{fig:cauchy} for
1643: %%   an alternative example.
1644: %% \label{fig:discrete}
1645: %% }}
1646: \begin{figure*}
1647: \twographschris{doubGau_1.0_0.5_good.eps}{doubGau_1.0_0.5_bad.eps}
1648: \caption{Binned samples of the double Gaussian distribution $p_{dGau}(x)$. The
1649:   normalisation is arbitrary and has no relevance   here. 
1650:   (a) uses a Metropolis-Hastings algorithm and yields a good approximation
1651:   whereas (b) uses the adaptive algorithm. 
1652: %%   $p_{dGau}(x)$ defined in equation~(\ref{eq:dGau}).  The sampling in
1653: %%   part (a) is done correctly.  The sampling in part (b) is incorrect
1654: %%   and shows what can go wrong when a property of the proposal
1655: %%   distribution (in this case its width) is varied on the basis of the
1656: %%   rejection/acceptance history of the Markov Chain.
1657: \label{fig:dGau}}
1658: \end{figure*}
1659: Take, for example, the 1-dimensional double Gaussian probability distribution
1660: $p_{dGau}(x)$ 
1661: defined by
1662: \begin{equation}
1663: p_{dGau}(x) \propto g(x, 0, 1) + g(x, 10, 1/2),\label{eq:dGau}
1664: \end{equation}
1665: where $g(x, m, \sigma)=\exp(-(x-m)^2/(2\sigma^2))$ is a Gaussian distribution
1666: of unit height and width $\sigma$ centred on $x=m$.
1667: %\begin{cases}
1668: %1/3 & \mbox {if }x>0 \mbox{ and } x<2, \\
1669: %1/3 & \mbox {if }x>9\mbox{ and }x<10, \\
1670: %0 & \mbox{otherwise}.
1671: %\end{cases}
1672: %\label{eq:blocdist}
1673: %\end{eqnarray}
1674: Note that the Gaussian near the origin is wider than the
1675: Gaussian near $x=10$.  This pdf mocks up the approximate situation along the
1676: $M_{1/2}$ direction in mSUGRA for large $m_0$, as Fig.~\ref{fig:like7}a
1677: shows. The narrow Gaussian would then correspond to the light $h^0$-pole
1678: region, which is quite disconnected from the wider $A^0$-pole
1679: region. 
1680: A Metropolis-Hastings algorithm (see section~\ref{sec:mcmc})
1681: with a fixed Gaussian proposal pdf $Q(x)$ of width 5 was
1682: run for 2 000 000 steps and the binned result is shown in
1683: Figure~\ref{fig:dGau}a. It reproduces the target distribution very well.
1684: In contrast, Figure~\ref{fig:dGau}b shows what happens when
1685: the adaptive algorithm is run on $p_{dGau}(x)$.  Clearly the result
1686: is very different to that in Figure~\ref{fig:dGau}a and is thus
1687: very wrong.  The narrow Gaussian has been sampled many times more frequently
1688: than it should have been relative to the wider one near the origin.
1689: 
1690: The adaptive algorithm ensures that whenever the current
1691: point is in one of the two Gaussian regions, the step size is adjusted
1692: to be proportional to the width of that region.  This adaptation is
1693: not immediate (seven successive rejections must occur before the step
1694: size is halved) but suppose for the moment that adaptation were to
1695: take place almost immediately. For the moment, let us also neglect 
1696: random fluctuations of the step size.
1697: In such a limit, when the current
1698: point is in the left-hand Gaussian region, the step size is double what
1699: it is when the current point is in the right-hand region.  Making a
1700: proposal for a jump from the region on the left to the region on the
1701: right, therefore, is something like a $10$-sigma event.  In contrast,
1702: making the reverse proposal (from right to left) is more like a
1703: $20$-sigma event.
1704: %\footnote{In the left-hand top-hat region the adaptive
1705: %algorithm usually selected a step size of 2, but opted for the larger
1706: %4 about 20\% of the time.  This larger step size of 4 has the largest
1707: %effect on the flow of traffic from the left region to the right
1708: %region.  Similarly, in the right-hand top-hat region the adaptive
1709: %algorithm tended to select a step size of 1, but about 20\% of the
1710: %time chose the larger value of 2, and this larger value determines the
1711: %rate of flow of traffic from the right-hand to the left-hand region.}
1712: In this limit, it is thus $e^{((20)^2 - (10)^2)/2}=e^{150}$
1713: times more likely that 
1714: jumps to the right get proposed than 
1715: jumps to the left.  
1716: This is a vast overestimate of the bias toward
1717: the narrow region for two reasons.  Firstly, step size adaptation is
1718: not immediate (even when the step size is half of what it should be,
1719: quite a few steps occur before
1720: seven successive rejections).  Secondly and more importantly,
1721: even when settled in one of the two regions, the adaptive step size
1722: makes excursions about its mean value.  Excursions to very high step
1723: sizes (double or quadruple the average step size) are infrequent
1724: but still occur.  When they do, they elevate the chance of proposing a
1725: jump from one region to another, and help to equilibrate between the
1726: two regions.  Both of these effects reduce the bias favouring the narrow
1727: region, but still break detailed balance, and overall the right hand
1728: Gaussian is still sampled about 10 times more frequently than it
1729: should be.
1730: 
1731: %The fact that the regions are completely disconnected is not a
1732: %necessary part of the example, though it helps to highlight the
1733: %effect.  As an example of the effect in connected regions, consider
1734: %the double gaussian distribution defined by:
1735: %\begin{eqnarray}
1736: %p_{dGau}(x) \propto g(x, 0, 1) + g(x, 10, 1/2),\label{eq:dGau}
1737: %\end{eqnarray}
1738: %where $g(x, m, \sigma)=\exp(-(x-m)^2/(2\sigma^2))$ is a gaussian distribution
1739: %of unit height and width $\sigma$ centred on $x=m$.
1740: 
1741: %Good and bad samplings from $p_{dGau}(x)$ are shown in
1742: %figure~\ref{fig:dGau}.  Once again the exponential suppression of
1743: %the broad regions is seen relative to the narrow region.
1744: 
1745: %In general, broad regions are vacated and undersampled (relative to
1746: %narrow regions) by the adaptive sampler.  The effect is made more
1747: %pronounced as the separation of the regions increases, or as the ratio
1748: %of their widths increases.  In both cases the bias toward oversampling
1749: %the narrow region grows exponentially.
1750: 
1751: \begin{figure*}
1752: \twographschris{goodCauchy}{badCauchy}
1753: \caption{Binned sampling of a Cauchy distribution $1/(1+x^2)$ (shown in solid
1754:   red line) truncated at   $x=\pm   500$.
1755:   (a) shows the result of direct sampling and yields a good
1756:   approximation whereas (b) uses the adaptive algorithm. 
1757: %is incorrect
1758: %  and shows what can go wrong when a property of the proposal
1759: %  distribution (in this case its width) is varied on the basis of the
1760: %  rejection/acceptance history of the Markov Chain.  
1761: \label{fig:cauchy}}
1762: \end{figure*}
1763: For quite a different example, consider the truncated Cauchy
1764: distribution defined by
1765: \begin{eqnarray}
1766: p_{cauchy}(x)\propto
1767: \begin{cases}
1768: 1/(1+x^2)& \mbox {if~}x>-500 \mbox{~and~} x<500, \\
1769: 0 & \mbox{otherwise}.
1770: \end{cases}
1771: \label{eq:cauchydist}
1772: \end{eqnarray}
1773: A faithful sampling from this distribution is shown in
1774: Figure~\ref{fig:cauchy}(a), and a mis-sampling using the adaptive
1775:   algorithm is shown for comparison in Figure~\ref{fig:cauchy}(b).
1776: Although there is better correspondence between the samples
1777: generated by the two methods than was the case earlier, it is
1778: nevertheless evident that there are large differences between the
1779: degrees to which the two methods have sampled the tails of the
1780: distribution.  As a consequence, samples from the adaptive method
1781: have a root-mean-squared value of 7.6 compared to the value of 17.9 obtained
1782: by the true 
1783: sampling.  The cause of the discrepancy is again due to the very
1784: different scales in the cauchy distribution.  Its core is very narrow
1785: with a width of order 1, but there are significant parts of probability also
1786: lodged in the tails many orders of magnitude away.  The adaptive
1787: method has to raise and lower the step size frequently to sample from
1788: the whole dynamic range, and in doing so it has to break detailed
1789: balance many times, resulting in the cumulative effect of an overall
1790: order of magnitude bias in the tails.
1791: 
1792: 
1793: 
1794: 
1795: \providecommand{\href}[2]{#2}\begingroup\raggedright\begin{thebibliography}{10}
1796: 
1797: \bibitem{Martin:1997ns}
1798: S.~P. Martin, 
1799: {\tt hep-ph/9709356}.
1800: 
1801: \bibitem{Ellis:2003dn}
1802: J.~R. Ellis, K.~A. Olive, Y.~Santoso, and V.~C. Spanos, {Phys.\ Lett.\ } {\bf
1803:   B588} (2004) 7--16, 
1804:  {\tt hep-ph/0312262}.
1805: 
1806: %\cite{Roszkowski:2004jd}
1807: \bibitem{Roszkowski:2004jd}
1808: L.~Roszkowski, R.~Ruiz de Austri and K.-Y.~Choi,
1809: JHEP {\bf 0508} (2005) 080
1810: [arXiv:hep-ph/0408227];
1811: D.G.~Cerdeno {\em et al}, {\tt hep-ph/0509275}.
1812: %%CITATION = HEP-PH 0408227;%%
1813: 
1814: %\cite{Covi:2004rb}
1815: \bibitem{Covi:2004rb}
1816: L.~Covi, J.~E.~Kim, L.~Roszkowski
1817: { Phys. Rev. Lett.} {\bf 82} (1999) 4180-4183, {\tt hep-ph/9905212};
1818: L.~Covi, H.~ B.~ Kim, J.~E.~ Kim, L.~ Roszkowski
1819: { JHEP} {\bf 0105} (2001) 033, {\tt hep-ph/0101009};
1820: L.~Covi, L.~Roszkowski, R.~Ruiz de Austri and M.~Small,
1821: JHEP {\bf 0406} (2004) 003
1822: {\tt hep-ph/0402240}.
1823: %%CITATION = HEP-PH 0402240;%%
1824: 
1825: \bibitem{Ellis:1983ew}
1826: J.~R. Ellis, J.~S. Hagelin, D.~V. Nanopoulos, K.~A. Olive, and M.~Srednicki,
1827:     { Nucl. Phys.} {\bf B238}
1828:   (1984) 453--476.
1829: 
1830: \bibitem{Baer:2002gm}
1831: H.~Baer {\em et.~al.},   { JHEP} {\bf 07} (2002) 050,
1832:   {\tt hep-ph/0205325}.
1833: 
1834: \bibitem{Ellis:2003cw}
1835: J.~R. Ellis, K.~A. Olive, Y.~Santoso, and V.~C. Spanos,   { Phys. Lett.} {\bf B565} (2003) 176--182,
1836:   {\tt hep-ph/0303043}.
1837: 
1838: \bibitem{Battaglia:2003ab}
1839: M.~Battaglia {\em et.~al.},   { Eur. Phys. J.} {\bf C33} (2004) 273--296,
1840:   {\tt hep-ph/0306219}.
1841: 
1842: \bibitem{Ellis:2003si}
1843: J.~R. Ellis, K.~A. Olive, Y.~Santoso, and V.~C. Spanos,   { Phys. Rev.} {\bf D69} (2004)
1844:   095004, {\tt
1845:   hep-ph/0310356}.
1846: 
1847: %\cite{Gomez:2004ek}
1848: \bibitem{Gomez:2004ek}
1849: M.~E.~Gomez, T.~Ibrahim, P.~Nath and S.~Skadhauge,
1850: Phys.\ Rev.\ D {\bf 70}, 035014 (2004)
1851: [arXiv:hep-ph/0404025].
1852: 
1853: \bibitem{Ellis:2004tc}
1854: J.~R. Ellis, S.~Heinemeyer, K.~A. Olive, and G.~Weiglein,   { JHEP} {\bf 02} (2005)
1855:   013, {\tt hep-ph/0411216}.
1856: 
1857: \bibitem{Armstrong:1994it}
1858: {\bf ATLAS} Collaboration, W.~W. Armstrong {\em et.~al.},  CERN-LHCC-94-43.
1859: 
1860: \bibitem{Spergel:2003cb}
1861: D.~N. Spergel {\em et.~al.},   {
1862:   Astrophys. J. Suppl.} {\bf 148} (2003) 175,
1863:   {\tt astro-ph/0302209}.
1864: 
1865: \bibitem{Bennett:2003bz}
1866: C.~L. Bennett {\em et.~al.},   {
1867:   Astrophys. J. Suppl.} {\bf 148} (2003) 1,
1868:   {\tt astro-ph/0302207}.
1869: 
1870: \bibitem{Allanach:2004xn}
1871: B.~C. Allanach, G.~Belanger, F.~Boudjema, and A.~Pukhov, 
1872:   { JHEP} {\bf 12} (2004) 020,
1873:   {\tt hep-ph/0410091}.
1874: 
1875: \bibitem{Griest:1990kh}
1876: K.~Griest and D.~Seckel,   { Phys. Rev.} {\bf D43} (1991) 3191--3203.
1877: 
1878: \bibitem{Drees:1992am}
1879: M.~Drees and M.~M. Nojiri,   { Phys. Rev.} {\bf D47} (1993) 376--408,
1880:   {\tt hep-ph/9207234}.
1881: 
1882: \bibitem{Arnowitt:1993mg}
1883: R.~Arnowitt and P.~Nath,   { Phys. Lett.} {\bf B299} (1993) 58--63,
1884:   {\tt hep-ph/9302317}.
1885: 
1886: \bibitem{Djouadi:2005dz}
1887: A.~Djouadi, M.~Drees, and J.-L. Kneur, 
1888: {\tt hep-ph/0504090}.
1889: 
1890: \bibitem{Feng:1999mn}
1891: J.~L. Feng, K.~T. Matchev, and T.~Moroi,   { Phys. Rev. Lett.} {\bf 84} (2000) 2322--2325,
1892:   {\tt hep-ph/9908309}.
1893: 
1894: \bibitem{Feng:1999zg}
1895: J.~L. Feng, K.~T. Matchev, and T.~Moroi,   { Phys. Rev.} {\bf D61} (2000) 075005,
1896:   {\tt hep-ph/9909334}.
1897: 
1898: \bibitem{Feng:2000gh}
1899: J.~L. Feng, K.~T. Matchev, and F.~Wilczek,   { Phys. Lett.} {\bf B482} (2000) 388--399,
1900:   {\tt hep-ph/0004043}.
1901: 
1902: \bibitem{Boehm:9911}
1903: C.~Boehm, A.~Djouadi and M.~Drees,  { Phys. Rev.} {\bf D62} (2000) 035012
1904:   {\tt hep-ph/9911496}.
1905: 
1906: \bibitem{arnie}
1907:  R.~Arnowitt, B.~ Dutta and Y.~Santoso,  { Nucl. Phys.} {\bf B606} (2001) 59
1908:   {\tt hep-ph/0102181}.
1909: 
1910: \bibitem{Ellis:2001nx}
1911: J.~R. Ellis, K.~A. Olive and Y.~Santoso,  { Astropart. Phys.} {\bf 18} (2003) 395,
1912:   {\tt hep-ph/0112113}.
1913: 
1914: \bibitem{Brhlik:2000dm}
1915: M.~Brhlik, D.~J.~H. Chung, and G.~L. Kane,   { Int. J. Mod. Phys.} {\bf D10} (2001) 367,
1916:   {\tt hep-ph/0005158}.
1917: 
1918: \bibitem{Drees:2000he}
1919: M.~Drees { et.~al.},   {
1920:   Phys. Rev.} {\bf D63} (2001) 035008,
1921:   {\tt hep-ph/0007202}.
1922: 
1923: \bibitem{Polesello:2004qy}
1924: G.~Polesello and D.~R. Tovey,   { JHEP} {\bf 05} (2004) 071,
1925:   {\tt hep-ph/0403047}.
1926: 
1927: \bibitem{Bambade:2004tq}
1928: P.~Bambade, M.~Berggren, F.~Richard, and Z.~Zhang, 
1929: {\tt hep-ph/0406010}.
1930: 
1931: \bibitem{Allanach:2004jh}
1932: B.~C. Allanach, G.~Belanger, F.~Boudjema, A.~Pukhov, and W.~Porod, 
1933: {\tt hep-ph/0402161}.
1934: 
1935: \bibitem{Moroi:2005nc}
1936: T.~Moroi, Y.~Shimizu, and A.~Yotsuyanagi, 
1937: {\tt hep-ph/0505252}.
1938: 
1939: \bibitem{Bourjaily:2005ax}
1940: J.~L. Bourjaily and G.~L. Kane, 
1941: {\tt hep-ph/0501262}.
1942: 
1943: \bibitem{Roszkowski:2001sb}
1944: L.~Roszkowski, R.~Ruiz~de Austri, and T.~Nihei,   { JHEP} {\bf 08} (2001) 024,
1945:   {\tt hep-ph/0106334}.
1946: 
1947: \bibitem{Baer:2002fv}
1948: H.~Baer, C.~Balazs, and A.~Belyaev,   { JHEP} {\bf 03} (2002) 042,
1949:   {\tt hep-ph/0202076}.
1950: 
1951: \bibitem{Baer:2003yh}
1952: H.~Baer and C.~Balazs,   {
1953:   JCAP} {\bf 0305} (2003) 006,
1954:   {\tt hep-ph/0303114}.
1955: 
1956: \bibitem{Chattopadhyay:2003xi}
1957: U.~Chattopadhyay, A.~Corsetti, and P.~Nath,   { Phys. Rev.}
1958:   {\bf D68} (2003) 035005, {\tt
1959:   hep-ph/0303201}.
1960: 
1961: \bibitem{Lahanas:2003yz}
1962: A.~B. Lahanas and D.~V. Nanopoulos,   { Phys. Lett.} {\bf B568} (2003) 55--62,
1963:   {\tt hep-ph/0303130}.
1964: 
1965: \bibitem{Belanger:2004hk}
1966: G.~Belanger, F.~Boudjema, A.~Cottrant, A.~Pukhov, and A.~Semenov, 
1967: {\tt hep-ph/0412309}.
1968: 
1969: %\cite{Roszkowski:2004jc}
1970: \bibitem{leszek}
1971: L.~Roszkowski,
1972: Pramana {\bf 62} (2004) 389
1973: {\em hep-ph/0404052}.
1974: %%CITATION = HEP-PH 0404052;%%
1975: 
1976: \bibitem{Herndon:2004tk}
1977: {\bf CDF and D0} Collaboration, M.~Herndon,  To appear in the proceedings of 32nd
1978:   International Conference on High-Energy Physics (ICHEP 04), Beijing, China,
1979:   16-22 Aug 2004.
1980: 
1981: \bibitem{Ellis:2005sc}
1982: J.~R. Ellis, K.~A. Olive, and V.~C. Spanos, 
1983: {\tt hep-ph/0504196}.
1984: 
1985: \bibitem{Profumo:2004at}
1986: S.~Profumo and C.~E. Yaguna,   { Phys. Rev.} {\bf D70} (2004) 095004,
1987:   {\tt hep-ph/0407036}.
1988: 
1989: \bibitem{Stark:2005mp}
1990: L.~S. Stark, P.~Hafliger, A.~Biland, and F.~Pauss, 
1991: {\tt hep-ph/0502197}.
1992: 
1993: \bibitem{Baltz:2004aw}
1994: E.~A. Baltz and P.~Gondolo,   { JHEP} {\bf 10}
1995:   (2004) 052, {\tt
1996:   hep-ph/0407039}.
1997: 
1998: \bibitem{MacKay}
1999: D.~MacKay, {\em Information Theory, Inference, and Learning Algorithms}.
2000: \newblock Cambridge University Press, 2003.
2001: 
2002: \bibitem{Allanach:2004my}
2003: B.~C. Allanach, D.~Grellscheid, and F.~Quevedo,   {JHEP} {\bf 07} (2004) 069,
2004:   {\tt hep-ph/0406277}.
2005: 
2006: \bibitem{Brein:2004kh}
2007: O.~Brein, 
2008: {\tt hep-ph/0407340}.
2009: 
2010: \bibitem{Allanach:2000ii}
2011: B.~C. Allanach, J.~P.~J. Hetherington, M.~A. Parker, and B.~R. Webber, 
2012:   { JHEP} {\bf 08} (2000) 017,
2013:   {\tt hep-ph/0005186}.
2014: 
2015: \bibitem{Allanach:2001kg}
2016: B.~C. Allanach,   { Comput. Phys. Commun.} {\bf 143} (2002) 305--331,
2017:   {\tt hep-ph/0104145}.
2018: 
2019: \bibitem{PDBook}
2020: S.~{Eidelman} {\em et.~al.},   { Phys.
2021:   Lett. B} {\bf 592} (2004) 1--end.
2022: 
2023: \bibitem{Allanach:2004rh}
2024: B.~C. Allanach, A.~Djouadi, J.~L. Kneur, W.~Porod, and P.~Slavich,   { JHEP}
2025:   {\bf 09} (2004) 044, {\tt
2026:   hep-ph/0406166}.
2027: 
2028: \bibitem{Barate:2003sz}
2029: {\bf ALEPH} Collaboration, R.~Barate {\em et.~al.},   { Phys. Lett.} {\bf B565} (2003)
2030:   61--75, {\tt
2031:   hep-ex/0306033}.
2032: 
2033: \bibitem{Degrassi:2002fi}
2034: G.~Degrassi, S.~Heinemeyer, W.~Hollik, P.~Slavich, and G.~Weiglein,   { Eur.
2035:   Phys. J.} {\bf C28} (2003) 133--143,
2036:   {\tt hep-ph/0212020}.
2037: 
2038: \bibitem{Skands:2003cj}
2039: P.~Skands {\em et.~al.},   { JHEP} {\bf
2040:   07} (2004) 036, {\tt
2041:   hep-ph/0311123}.
2042: 
2043: \bibitem{Belanger:2001fz}
2044: G.~Belanger, F.~Boudjema, A.~Pukhov, and A.~Semenov,   { Comput. Phys. Commun.}
2045:   {\bf 149} (2002) 103--120,
2046:   {\tt hep-ph/0112278}.
2047: 
2048: \bibitem{Belanger:2004yn}
2049: G.~Belanger, F.~Boudjema, A.~Pukhov, and A.~Semenov,   {\tt hep-ph/0405253}.
2050: 
2051: \bibitem{Bennett:2004pv}
2052: {\bf Muon g-2} Collaboration, G.~W. Bennett {\em et.~al.},   { Phys. Rev.
2053:   Lett.} {\bf 92} (2004) 161802,
2054:   {\tt hep-ex/0401008}.
2055: 
2056: \bibitem{Passera:2004bj}
2057: M.~Passera,   { J. Phys.} {\bf G31} (2005) R75--R94,
2058:   {\tt hep-ph/0411168}.
2059: 
2060: \bibitem{deTroconiz:2004tr}
2061: J.~F. de~Troconiz and F.~J. Yndurain,   { Phys. Rev.} {\bf D71} (2005)
2062:   073008, {\tt
2063:   hep-ph/0402285}.
2064: 
2065: \bibitem{Allanach:2005yq}
2066: B.~C. Allanach, A.~Brignole, and L.~E. Ibanez,   { JHEP} {\bf 05} (2005) 030,
2067:   {\tt hep-ph/0502151}.
2068: 
2069: \bibitem{Gambino:2004mv}
2070: P.~Gambino, U.~Haisch, and M.~Misiak,   { Phys. Rev. Lett.} {\bf 94} (2005)
2071:   061803, {\tt
2072:   hep-ph/0410155}.
2073: 
2074: \bibitem{hfag}
2075: {\bf Heavy Flavour Averaging Group} Collaboration.\
2076:   http://www.slac.stanford.edu/xorg/hfag.
2077: 
2078: \bibitem{Group:2005cc}
2079: {\bf CDF and D0} Collaboration,  the Tevatron Electroweak Working Group, 
2080: {\tt hep-ex/0507091}.
2081: 
2082: \bibitem{stat}
2083: A.~Gelman and D.~Rubin,   { Stat. Sci.} {\bf 7} (1992) 457--472.
2084: 
2085: \bibitem{Abazov:2005ku}
2086: {\bf D0} Collaboration, V.~M. Abazov {\em et.~al.}, 
2087: {\tt hep-ex/0504032}.
2088: 
2089: \bibitem{future}
2090: {\bf CDF and D0} Collaboration. See URL
2091:   http://www-cdf.fnal.gov/physics/projections/.
2092: 
2093: \bibitem{Nojiri:1996fp}
2094: M.~M. Nojiri, K.~Fujii, and T.~Tsukamoto,   { Phys. Rev.} {\bf D54} (1996) 6756--6776,
2095:   {\tt hep-ph/9606370}.
2096: 
2097: \bibitem{Dedes:2004yc}
2098: A.~Dedes and B.~T. Huffman,   { Phys. Lett.} {\bf B600}
2099:   (2004) 261--269, {\tt
2100:   hep-ph/0407285}.
2101: 
2102: \bibitem{Acosta:2004xj}
2103: {\bf CDF} Collaboration, D.~Acosta {\em et.~al.},   { Phys. Rev. Lett.} {\bf 93}
2104:   (2004) 032001, {\tt
2105:   hep-ex/0403032}.
2106: 
2107: \bibitem{Abazov:2004dj}
2108: {\bf D0} Collaboration, V.~M. Abazov {\em et.~al.},  { Phys.
2109:   Rev. Lett.} {\bf 94} (2005) 071802,
2110:   {\tt hep-ex/0410039}.
2111: 
2112: \bibitem{Dagu}
2113: {\bf CDF and D0} Collaboration, S.~Dugad,  To
2114:   appear in the proceedings of the Hadron Collider Physics Symposium, Les
2115:   Diablerets, Switzerland, 3-9 Jul 2005.
2116: 
2117: \bibitem{Dedes:2001fv}
2118: A.~Dedes, H.~K. Dreiner, and U.~Nierste,   { Phys. Rev.
2119:   Lett.} {\bf 87} (2001) 251804,
2120:   {\tt hep-ph/0108037}.
2121: 
2122: \bibitem{Allanach:2000kt}
2123: B.~C. Allanach, C.~G. Lester, M.~A. Parker, and B.~R. Webber,   {
2124:   JHEP} {\bf 09} (2000) 004,
2125:   {\tt hep-ph/0007009}.
2126: 
2127: \bibitem{Weiglein:2004hn}
2128: {\bf LHC/LC Study Group} Collaboration, G.~Weiglein {\em et.~al.}, 
2129: {\tt hep-ph/0410364}.
2130: 
2131: \bibitem{Allanach:2003jw}
2132: B.~C. Allanach, S.~Kraml, and W.~Porod,   { JHEP} {\bf 03}
2133:   (2003) 016, {\tt
2134:   hep-ph/0302102}.
2135: 
2136: \bibitem{Belanger:2005jk}
2137: G.~Belanger, S.~Kraml, and A.~Pukhov, 
2138: {\tt hep-ph/0502079}.
2139: 
2140: \bibitem{Bechtle:2004pc}
2141: P.~Bechtle, K.~Desch, and P.~Wienemann, 
2142: {\tt hep-ph/0412012}.
2143: 
2144: \bibitem{Lafaye:2004cn}
2145: R.~Lafaye, T.~Plehn, and D.~Zerwas,   {\tt
2146:   hep-ph/0404282}.
2147: 
2148: \bibitem{whitey}
2149: M.~White, C.~Lester, and A.~Parker,   {\tt hep-ph/0508143}.
2150: 
2151: \end{thebibliography}\endgroup
2152: 
2153: %\bibliography{map}
2154: 
2155: \end{document}
2156: