0808.0704/dm.tex
1: %\documentclass[prd,nofootinbib,preprint]{revtex4}
2: \documentclass[prd,nofootinbib,preprint,superscriptaddress]{revtex4}
3: %\documentclass[twocolumn,preprintnumbers,nofootinbib,prl]{revtex4}
4: %showpacs
5: 
6: \usepackage{graphicx}
7: \usepackage{amsmath}
8: \usepackage{amssymb}
9: 
10: % \usepackage{srcltx}
11: \usepackage[hypertex]{hyperref}
12: % \usepackage{hyperref}
13: % \usepackage{times}
14: \newcommand{\vev}[1]{\langle {#1} \rangle}
15: \newcommand{\lsim}{\lesssim}
16: \newcommand{\ord}[1]{\mathcal{O}{(#1)}}
17: \newcommand{\co}{\mathcal{O}}
18: \newcommand{\gsim}{\gtrsim}
19: %\newcommand{\z3}{Z^{\ell}_3}
20: %\newcommand{\z9}{Z^q_9}
21: \newcommand{\beq}{\begin{equation}}
22: \newcommand{\eeq}{\end{equation}}
23: \newcommand{\be}{B_\oplus}
24: \newcommand{\re}{R_\oplus}
25: \newcommand{\unit}[1]{\,\mathrm{#1}}
26: \newcommand{\vect}[1]{\boldsymbol{\rm #1}}
27: 
28: %-- command (re)definitions -----------------------------------------
29: 
30: \newcommand{\capdef}{}
31: \newcommand{\mycaption}[2][\capdef]{\renewcommand{\capdef}{#2}%
32:        \caption[#1]{{\footnotesize #2}}}
33: \makeatletter
34: \renewcommand{\fnum@table}{\textbf{\tablename~\thetable}}
35: \renewcommand{\fnum@figure}{\textbf{\figurename~\thefigure}}
36: \makeatother
37: 
38: \hyphenation{pa-ra-meter pa-ra-meters}
39: 
40: \renewcommand{\baselinestretch}{1.05}
41: 
42: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
43: \begin{document}
44: % page numbers bottom-center
45: \pagestyle{plain}
46: 
47: \vspace*{1cm}
48: \preprint{CERN-PH-TH/2008-171}
49: 
50: \title{Spin-independent elastic WIMP scattering and\\ 
51: the DAMA annual modulation signal\vspace*{1cm}}
52: 
53: \author{\textbf{Malcolm Fairbairn}}
54: \email{malc_at_cern.ch}
55: \affiliation{Physics Department, Theory Division, CERN,
56: 1211 Geneva 23, Switzerland}
57: 
58: \affiliation{King's College London, Strand, WC2R 2LS, UK\vspace*{5mm}}
59: 
60: \author{\textbf{Thomas Schwetz}\vspace*{3mm}}
61: \email{schwetz_at_cern.ch}
62: \affiliation{Physics Department, Theory Division, CERN,
63: 1211 Geneva 23, Switzerland}
64: 
65: 
66: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67: 
68: \begin{abstract}
69:   \vspace*{5mm} We discuss the interpretation of the annual modulation
70:   signal seen in the DAMA experiment in terms of spin-independent
71:   elastic WIMP scattering. Taking into account channeling in the
72:   crystal as well as the spectral signature of the modulation signal
73:   we find that the low-mass WIMP region consistent with DAMA data is
74:   confined to WIMP masses close to $m_\chi \simeq 12$~GeV, in
75:   disagreement with the constraints from CDMS and XENON. We conclude
76:   that even if channeling is taken into account this interpretation of
77:   the DAMA modulation signal is disfavoured. There are no overlap
78:   regions in the parameter space at 90\%~CL and a consistency test
79:   gives the probability of $1.2\times 10^{-5}$. We study the
80:   robustness of this result with respect to variations of the WIMP
81:   velocity distribution in our galaxy, by changing various parameters
82:   of the distribution function, and by using the results of a
83:   realistic $N$-body dark matter simulation. We find that only by
84:   making rather extreme assumptions regarding halo properties can we
85:   obtain agreement between DAMA and CDMS/XENON.
86: \end{abstract}
87: \maketitle
88: 
89: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
90: 
91: \section{Introduction}
92: 
93: The DAMA collaboration has collected an impressive amount of data in
94: their search for the scattering of weakly interacting dark matter
95: particles (WIMPs) off Sodium Iodine. The combined data from DAMA/NaI
96: (7 annual cycles) and DAMA/LIBRA (4 annual cycles) amounts to a total
97: exposure of 0.82~ton~yr~\cite{Bernabei:2008yi}, in a field where
98: exposure is measured in units of kg~days. DAMA/LIBRA has now provided
99: further evidence for an annual modulation of the event rate in the
100: energy range between 2 and 6~keVee, the claimed statistical confidence
101: of the positive signal being $8.2\sigma$~\cite{Bernabei:2008yi}. The
102: phase of the observed modulation (with maximum on day $144\pm8$) is in
103: striking agreement with the expectation for a modulation in a WIMP
104: scattering signal due to the rotation of the Earth around the Sun,
105: (expected maximum day 152, June 2nd), see e.g.,~\cite{Jungman:1995df}
106: for a review.
107: %
108: An interpretation of this effect in terms of spin-independent
109: interactions of conventional WIMPs with masses $m_\chi \gtrsim 50$~GeV
110: is in direct conflict with the constraints from several experiments
111: looking for direct WIMP detection, most notably with the data from
112: CDMS~\cite{Ahmed:2008eu} and XENON10~\cite{Angle:2007uj}, which
113: exclude the WIMP cross section consistent with the DAMA modulation for
114: $m_\chi\sim 50$~GeV by many orders of magnitude.  In light of this,
115: several alternative explanations of the DAMA annual modulation have
116: been proposed, for example spin-dependent
117: interactions~\cite{Ullio:2000bv, Savage:2004fn}, light WIMPs with
118: $\lesssim 10$~GeV masses~\cite{Bottino:2003iu, Bottino:2003cz,
119:   Gondolo:2005hh}, keV scale axion-like dark
120: matter~\cite{Bernabei:2005ca} (see however, \cite{Pospelov:2008jk,
121:   Gondolo:2008dd}), dark matter interacting only with
122: electrons~\cite{Bernabei:2007gr}, inelastic WIMP
123: scattering~\cite{TuckerSmith:2001hy, Chang:2008gd} and mirror dark
124: matter~\cite{Foot:2008nw}.
125: 
126: In this work we reconsider the possibility of spin-independent elastic
127: scattering of light WIMPs with $\lesssim 10$~GeV
128: masses~\cite{Bottino:2003iu, Bottino:2003cz, Gondolo:2005hh},
129: see~\cite{Bottino:2007qg, Bottino:2008mf, Petriello:2008jj,
130: Feng:2008dz} for recent studies. The original idea is that light dark
131: matter scattering on the relatively light Sodium nuclei in DAMA could
132: deposit enough energy in the detector to give a signal, whereas the
133: scattering of light halo particles off heavier nuclei, such as for
134: example Ge in CDMS or Xe in XENON would lead to energy depositions
135: below the threshold of those detectors. Recently the importance of the
136: so-called channeling effect~\cite{Drobyshevski:2007zj} in the crystal
137: structure of the experiment has also been
138: emphasized~\cite{Petriello:2008jj, Bottino:2008mf}.
139: %
140: Specific models for WIMPs with $m_\chi \sim 10$~GeV have been studied
141: for example in~\cite{Bottino:2002ry, Bottino:2007qg, Barger:2005hb,
142: Gunion:2005rw}. Here we do not discuss theoretical implications but
143: focus on the phenomenology of direct detection experiments in a
144: model-independent way by assuming that such light WIMPs can provide the
145: correct relic abundance while any direct collider constraints can be
146: evaded.
147: 
148: In this region of WIMP masses several
149: experiments~\cite{Altmann:2001ax, Akerib:2003px, Lin:2007ka,
150: Aalseth:2008rx} exclude WIMP--nucleon scattering cross sections in the
151: range $\sigma_p \gtrsim 10^{-40}$~cm$^2$. As we will see in the next
152: pages, once we have included channeling as well as the spectral shape
153: of the DAMA modulation signal, the allowed region of our interest is
154: obtained at much small cross sections, around $\sigma_p \sim
155: 10^{-41}$~cm$^2$ and $m_\chi \sim 10$~GeV. In this region the most
156: relevant constraints come from XENON~\cite{Angle:2007uj}, the 2008
157: Germanium data from CDMS~\cite{Ahmed:2008eu}, and the 2005 CDMS data
158: on Silicon~\cite{Akerib:2005kh}. Indeed, as we will discuss, the
159: spectral shape of the DAMA annual modulation restricts $m_\chi$ and
160: $\sigma_p$ to a region excluded by these experiments.
161: 
162: In our study we elaborate on this result and discuss how robust it is
163: with respect to different assumptions about the dark matter halo of
164: our galaxy. The impact of non-standard halo properties on dark matter
165: direct detection experiments has been discussed by many authors, see
166: for example~\cite{Belli:2002yt, Fornengo:2003fm, Green:2002ht,
167:   Vergados:2007nc}.  At a qualitative level, one would expect that
168: smaller velocity dispersions or truncated velocity distributions would
169: seem to favour the dark matter interpretation of the DAMA signal, as
170: they could lead to more events above the low energy threshold of DAMA
171: but below that of other experiments.  Furthermore, anisotropies in the
172: velocity dispersion could amplify annual modulation signals.
173: 
174: The outline of our work is as follows. In Sec.~\ref{sec:analysis} we
175: briefly summarise the phenomenology of elastic WIMP scattering in
176: direct detection experiments and give some technical details on our
177: analysis of DAMA, CDMS and XENON data. The results for a standard dark
178: matter halo are presented in Sec.~\ref{sec:std-halo}.  In
179: Sec.~\ref{sec:nonstd-halo} we consider deviations from the standard
180: assumptions made about the WIMP velocity distribution: we use results
181: from the Via Lactea $N$-body dark matter
182: simulation~\cite{Diemand:2006ik}, we vary several parameters of the
183: Maxwellian distribution and consider asymmetric velocity profiles.
184: Sec.~\ref{sec:conclusions} contains our conclusions.  In
185: Appendix~\ref{sec:dama-mod} we comment on the DAMA fit using the
186: annual modulation energy spectrum, and in
187: Appendix~\ref{app:comparison} we briefly compare our results to the
188: ones from other authors.
189: 
190: 
191: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
192: \section{The WIMP signal in direct detection experiments}
193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
194: \label{sec:analysis}
195: 
196: In this section we briefly summarise the phenomenology of WIMP
197: scattering and describe our analysis of DAMA, CDMS and XENON data.
198: 
199: \subsection{The event spectrum from elastic WIMP scattering}
200: 
201: The differential event spectrum for WIMP scattering in counts per unit
202: mass of a given nucleus per unit exposure time and per unit energy as
203: a function of the recoil energy $E_R$ is given by the expression (see e.g.,
204: \cite{Jungman:1995df})
205: %
206: \begin{equation}\label{eq:spectrum}
207: R(E_R) = \frac{\rho \, \sigma_p A^2 F^2(q)}{2 m_\chi \mu^2_p} \, \eta(E_R, t) \,.
208: \end{equation}
209: %
210: Here $\rho$ is the local WIMP energy density for which we adopt the
211: canonical value $\rho = 0.3$~GeV/cm$^3$, $\sigma_p$ is the WIMP
212: scattering cross section on a proton\footnote{Note that only the
213: product of $\rho \times \sigma_p$ is relevant for the scattering
214: rate. Therefore, whenever we use the symbol $\sigma_p$ the cross
215: section is implicitly normalised to the value of $\rho =
216: 0.3$~GeV/cm$^3$.}, $A$ is the mass number of the target nucleus,
217: $\mu_p = m_\chi m_p/(m_\chi + m_p)$ is the reduced WIMP--proton mass
218: and we use the common Helm form factor $F(q) = 3 e^{-q^2 s^2/2}
219: [\sin(qr)-qr\cos(qr)] / (qr)^3$, with $s = 1$~fm, $r = \sqrt{R^2 - 5
220: s^2}$, $R = 1.2 A^{1/3}$~fm, $q = \sqrt{2 M E_R}$, with $M$ being the
221: nucleus mass. The function $\eta$ contains the integral over the WIMP
222: velocity distribution:
223: %
224: \begin{equation}\label{eq:eta}
225: \eta(E_R, t) = \int d\Omega_{\vect{v}} \int_{v_\mathrm{min}(E_R)}^\infty 
226: dv\, v \, f_\oplus(\vect{v},t) \,,
227: \end{equation}
228: %
229: where $v_\mathrm{min} = \sqrt{M E_R/2 \mu_M^2}$ is the minimum
230: velocity of a WIMP to produce a recoil energy $E_R$, and $v =
231: |\vect{v}|$. The WIMP velocity distribution in the Earth rest frame
232: $f_\oplus(\vect{v},t)$ is obtained from the distribution in the
233: galactic rest frame $f_\mathrm{gal}(\vect{v})$ by
234: %
235: \begin{equation}
236: f_\oplus(\vect{v},t) = f_\mathrm{gal}
237: (\vect{v} + \vect{v}_\odot +  \vect{v}_\oplus(t)) \,.
238: \end{equation}
239: %
240: In the coordinate system in which $x$ points towards the galactic
241: center, $y$ towards the direction of galactic rotation, and $z$
242: towards the galactic north pole, we use for the velocity of the Sun
243: $\vect{v}_\odot = (0,220,0) + (10, 13, 7)$~km/s~\cite{Gelmini:2000dm}
244: (including the local Keplerian velocity of 220~km/s~\cite{pdg} as well
245: as the Sun's peculiar velocity, see also~\cite{Green:2003yh} and
246: references therein). To describe the motion of the Earth around the
247: Sun we use the parametrisation of~\cite{Gelmini:2000dm}:
248: $\vect{v}_\oplus(t) = v_\oplus(\vect{e}_1 \sin\lambda - \vect{e}_2
249: \cos\lambda)$, with $v_\oplus = 2\pi \mathrm{A.U./yr} = 29.8$~km/s,
250: $\vect{e}_1 = (-0.0670, 0.4927, -0.8676)$, $\vect{e}_2 = (-0.9931,
251: -0.1170, 0.01032)$, and $\lambda(t) = 2\pi(t - 0.218)$.
252: 
253: The ``standard halo model'' assumes for the DM distribution an
254: isotropic isothermal sphere, which leads to a Maxwellian velocity
255: distribution in the galactic frame, truncated at the escape velocity
256: $v_\mathrm{esc}$:
257: %
258: \begin{equation}\label{eq:std-halo}
259: f_\mathrm{gal}(\vect{v}) = \left\{
260: \begin{array}{l@{\qquad}l}
261: N \, \left[\exp\left(-v^2 / \bar v^2 \right) - 
262: \exp\left(-v_\mathrm{esc}^2 / \bar v^2 \right)\right]
263: & v < v_\mathrm{esc} \\
264: 0 & v > v_\mathrm{esc}
265: \end{array}\right. \,,
266: \end{equation}
267: %
268: where we adopt as default values $\bar v = 220$~km/s and
269: $v_\mathrm{esc} = 650$~km/s. Here and throughout the paper we use the notation $\bar v^2 = 2 (<v^2>-<v>^2)=2\sigma^2$.  In order to properly take into account
270: the impact of the finite escape velocity as well as allowing for
271: non-standard halos deviating from Eq.~\ref{eq:std-halo} we perform the
272: integral in Eq.~\ref{eq:eta} numerically. In Sec.~\ref{sec:std-halo}
273: we first consider the standard halo model, whereas in
274: Sec.~\ref{sec:nonstd-halo} we go beyond these default assumptions by
275: varying the parameters of the velocity distribution as well as
276: changing its shape.
277: 
278: \subsection{On quenching and channeling}
279: 
280: In the analysis of DAMA data the effects of quenching and channeling
281: are important~\cite{Drobyshevski:2007zj, Bernabei:2007hw}. For
282: quenched events the recoiling nucleus loses its energy both
283: electromagnetically as well as via nuclear force interactions, where
284: the light yield in the scintillator comes mainly from the
285: electromagnetic part. To take this effect into account the event
286: energy is measured in equivalent electron energy (in keVee), defined
287: by $q \times E_R$ for the total nuclear recoil energy $E_R$ in
288: keV. For the elements in DAMA one has $q_\mathrm{Na} = 0.3$ and
289: $q_\mathrm{I} = 0.09$.
290: %
291: However, due to the crystalline structure of the target, for certain angles and
292: energies of the particles no nuclear force interactions happen and the
293: entire energy is lost electromagnetically. Hence, for these so-called
294: channeled events one has $q\approx 1$, see~\cite{Drobyshevski:2007zj,
295: Bernabei:2007hw}. For the fraction $f$ of channeled events relevant
296: for DAMA we use the parameterisation
297: %
298: \begin{equation}
299: f_\mathrm{Na}(E_R)
300: \approx \frac{e^{-E_R/18}}{1+0.75\,E_R} 
301: \,,\qquad
302: f_\mathrm{I}(E_R) \approx
303: \frac{e^{-E_R/40}}{1+0.65\,E_R}
304: \end{equation}
305: %
306: for $E_R$ in keV. These expressions reproduce to good accuracy the
307: curves shown in figure~4 of~\cite{Bernabei:2007hw}.  Departing from
308: Eq.~\ref{eq:spectrum}, the predicted spectrum in DAMA (in units of
309: counts/kg/day/keVee) is obtained by
310: %
311: \begin{equation}\label{eq:spect-DAMA}
312: R_\mathrm{DAMA}(E) = \sum_{X =\rm Na, I} 
313: \frac{M_X}{M_\mathrm{Na} + M_\mathrm{I}}
314: \left\{[1-f_X(E/q_X)] R_X(E/q_X) + f_X(E) R_X(E) \right\} \,,
315: \end{equation}
316: %
317: where the first term in the curled bracket corresponds to quenched
318: events and the second to channeled (and therefore unquenched) events.
319: 
320: Channeling does not occur in liquid Nobel gases like in the XENON
321: experiment. Since no information on channeling in Germanium and
322: Silicon is available for us, we do not take into account channeling in
323: CDMS. Note, however, that CDMS requires the coincidence of signals in
324: phonons and ionisation and hence, since channeled events would not
325: give a phonon signal they would not look like a WIMP signal defined by
326: the coincidence. Therefore, the fraction of channeled events
327: corresponds effectively to an efficiency factor reducing the effective
328: exposure. Hence, if channeling was indeed relevant for CDMS the final
329: exclusion limits would be somewhat weaker. 
330: 
331: In conclusion, channeling is an important effect for the
332: interpretation of data from direct detection experiments and we stress
333: the need of reliable information (probably requiring dedicated
334: measurements) on this effect for any solid DM detector.
335: 
336: \subsection{Fitting DAMA/LIBRA data}
337: 
338: For the model-independent analysis of DAMA data the signal as a
339: function of energy and time is parametrised as
340: %
341: \begin{equation}\label{eq:DAMAsignal}
342: S(E,t) = S_0(E) + A(E)\cos\omega(t-t_0) \,,
343: \end{equation}
344: %
345: with $\omega = 2\pi/1$~yr, $t_0 = 152$~day. For our analysis we use
346: the data on the modulation amplitude $A(E)$ for the full 0.82~ton~yr
347: DAMA exposure\footnote{Here and in the following we use the acronym
348: ``DAMA'' to denote the combined DAMA/NaI + DAMA/LIBRA data, except
349: where explicitly noted otherwise.} given in figure~9
350: of~\cite{Bernabei:2008yi} in 36 bins from 2 to 20~keVee. As we will
351: see the spectral shape of the signal is quite important for
352: constraining the WIMP parameters. The prediction for the modulation
353: amplitude in an energy bin $i$ from $E_i^-$ to $E_i^+$ is obtained
354: from Eq.~\ref{eq:spect-DAMA} by
355: %
356: \begin{equation}
357: A^\mathrm{pred}_i = \int dE \frac{1}{2}
358: [R_\mathrm{DAMA}(E, t = 152) - R_\mathrm{DAMA}(E, t = 335)]
359: \, \int_{E_i^-}^{E_i^+} dE' G(E, E') \,
360: \end{equation}
361: %
362: where $G(E, E')$ is a Gaussian energy resolution function
363: with width~\cite{Bernabei:2008yh}
364: %
365: \begin{equation}\label{eq:Eres}
366: \sigma_E^\mathrm{DAMA} / E = 0.45/\sqrt{E \,[{\rm keVee}]} + 0.0091 \,.
367: \end{equation}
368: %  
369: Then we construct a $\chi^2$ function
370: %
371: \begin{equation}\label{eq:chisq}
372: \chi^2_\mathrm{DAMA}(m_\chi, \sigma_p) = \sum_{i=1}^{36} \left(
373: \frac{A^\mathrm{pred}_i(m_\chi, \sigma_p) - A^\mathrm{obs}_i}{\sigma_i}
374: \right)^2 \,,
375: \end{equation}
376: %
377: using the experimental data points $A^\mathrm{obs}_i$ and their errors
378: $\sigma_i$ from figure~9 of~\cite{Bernabei:2008yi}.\footnote{Fig.~10
379:   of \cite{Bernabei:2008yi} shows that the $A^\mathrm{obs}_i$ are
380:   consistent with being Gaussian distributed, justifying the $\chi^2$
381:   adopted in Eq.~\ref{eq:chisq}.}  We find the best fit point for
382: the WIMP mass and the scattering cross section by minimising
383: Eq.~\ref{eq:chisq} with respect to $m_\chi$ and $\sigma_p$. Allowed
384: regions in the $(m_\chi, \sigma_p)$ plane at a given CL are obtained
385: by looking for the contours $\chi^2(m_\chi, \sigma_p) =
386: \chi^2_\mathrm{min} + \Delta\chi^2({\rm CL})$, where
387: $\Delta\chi^2({\rm CL})$ is evaluated for 2 degrees of freedom (dof),
388: e.g., $\Delta\chi^2(90\%) = 4.6$ or $\Delta\chi^2(99.73\%) = 11.8$.
389: 
390: In general the constant part of the spectrum, $S_0(E)$,
391: will consist of a time-averaged dark matter contribution $\langle R\rangle$
392: plus an un-identified background $B$:
393: %
394: \begin{equation}
395: S_0(E) = \langle R(E)\rangle + B(E) \,.
396: \end{equation}
397: %
398: In a given model such as for example WIMP scattering, the annual
399: modulation amplitude $A(E)$ and the averaged signal $\langle
400: R(E)\rangle$ are not independent. Hence, for a given fit to the data
401: on $A(E)$, the expected constant signal $\langle R(E)\rangle$ can be
402: predicted by using Eq.~\ref{eq:spect-DAMA}. In order to take this
403: additional information into account we use the data from figure~1
404: of~\cite{Bernabei:2008yi}, which shows the constant signal $S_0$ in 32
405: energy bins from 2 to 10~keVee for the DAMA/LIBRA detectors. For each
406: pair of $(m_\chi, \sigma_p)$ we calculate the expected signal from
407: WIMP scattering $\langle R\rangle$ in each of these energy
408: bins. Whenever $\langle R\rangle$ exceeds the observed rate in one of
409: the bins that particular values of $(m_\chi, \sigma_p)$ are not
410: consistent with the data and have to be excluded. Note that for event
411: rates of order 1~count/kg/day/keVee and the DAMA/LIBRA exposure of
412: 0.53~t~yr statistical errors are negligible for this purpose.
413: 
414: 
415: \subsection{Analysis of CDMS and XENON}
416: 
417: In our analysis we include the constraints from CDMS 2005 data using
418: Silicon (CDMS-Si)~\cite{Akerib:2005kh}, which, despite the relatively
419: low exposure of 12~kg~day, provides good sensitivity to the low-mass
420: WIMP region because of the light mass of the target nucleus
421: ($M_\mathrm{Si} \simeq 26$~GeV) and the low analysis threshold of
422: 7~keV. Furthermore, we include CDMS 2008 data on Germanium
423: (CDMS-Ge)~\cite{Ahmed:2008eu} with a threshold of 10~keV. We use the
424: energy dependent efficiency from Fig.~2 of \cite{Ahmed:2008eu} which
425: reduces the total exposure of 398.7~kg~day to an effective exposure of
426: about 121.3~kg~day. For both, CDMS-Si and CDMS-Ge, no event has been
427: observed. We calculate the expected number of events $N^\mathrm{pred}
428: $ as a function of $m_\chi$ and $\sigma_p$ by integrating
429: Eq.~\ref{eq:spectrum} over the relevant energy range and scaling with
430: the exposure. A $\chi^2$ is constructed using the common expression
431: for Poisson distributed data~\cite{pdg}, which for zero observed
432: events simply becomes
433: %
434: \begin{equation}\label{eq:chisq-cdms}
435: \chi^2_\mathrm{CDMS} = 2 N^\mathrm{pred} \,. 
436: \end{equation}
437: %
438: Exclusion contours are defined by the standard $\Delta\chi^2$ cuts for
439: 2~dof with respect to the minimum, which of course occurs for
440: $N^\mathrm{pred} = 0$. Conceptually this prescription differs from the
441: usual way to set a limit on $\sigma_p$ for fixed $m_\chi$ by requiring
442: $N^\mathrm{pred} < 2.3$ for a 90\%~CL limit. However, by accident,
443: since $\Delta\chi^2(90\%) = 4.6$ for 2~dof, in practice our $\chi^2$
444: definition in Eq.~\ref{eq:chisq-cdms} leads to the same exclusion
445: contour as the more conventional method of setting a limit on
446: $\sigma_p$.
447: 
448: For the analysis of data from the XENON10
449: experiment~\cite{Angle:2007uj} (XENON for brevity) we proceed in the
450: following way. Using the 7 bins in nuclear recoil energy from 4.5 to
451: 26.9~keV of table~1 of~\cite{Angle:2007uj} the predicted number of
452: events in bin $i$, $N^\mathrm{pred}_i(m_\chi, \sigma_p)$ can be
453: calculated by integrating Eq.~\ref{eq:spectrum} and scaling with the
454: exposure 316~kg~day as well as the bin dependent efficiencies
455: $\epsilon_{c}$ and $A_{nr}$ given in table~1 of~\cite{Angle:2007uj}.
456: %
457: After the publication of Ref.~\cite{Angle:2007uj} the so called
458: parameter $\mathcal{L}_\mathrm{eff}$ relevant for the nuclear recoil
459: energy scale in XENON was remeasured~\cite{Sorensen:2008ec}. Whereas
460: in~\cite{Angle:2007uj} a constant value $\mathcal{L}_\mathrm{eff} =
461: 0.19$ was used, Fig.~3 of~\cite{Sorensen:2008ec} shows an energy
462: dependent deviation of $\mathcal{L}_\mathrm{eff}$ from that value. We
463: use the information from this figure to correct nuclear recoil
464: energies in XENON. This leads to a somewhat higher energy threshold of
465: about 5.5~keV (instead of 4.5), which shifts the bound on DM
466: parameters to slightly higher values of $m_\chi$.
467: 
468: XENON observes 10 candidate events whose
469: recoil energies can be inferred from figure~3
470: of~\cite{Angle:2007uj}. They are distributed over the 7 bins as $(D_i)
471: = (1, 0, 0, 0, 3, 2, 4)$, with an expected background $(B_i) =
472: (0.2, 0.3, 0.2, 0.8, 1.4, 1.4, 2.7)$. We use the $\chi^2$ for
473: Poisson distributed data~\cite{pdg}:
474: %
475: \begin{equation}\label{eq:chisq_xenon}
476: \chi^2_\mathrm{XENON} = 2 \sum_{i=1}^7 \left[ 
477: N^\mathrm{pred}_i + B_i - D_i + 
478: D_i \log\left(\frac{D_i}{N^\mathrm{pred}_i + B_i}\right) \right] \,, 
479: \end{equation}
480: %
481: where the second term in the square bracket is zero if $D_i = 0$.
482: Again we define the exclusion curve in the $(m_\chi, \sigma_p)$ plane
483: by $\Delta\chi^2({\rm CL})$ contours for 2~dof with respect to the
484: minimum.
485: 
486: In both cases, CDMS and XENON we include an energy resolution of 
487: $20\% / \sqrt{E_R \, [{\rm keV}]}$, and an uncertainty in the energy
488: scale of 10\%, which is added to the $\chi^2$ definitions
489: Eqs.~\ref{eq:chisq-cdms} and \ref{eq:chisq_xenon} with the help of
490: nuisance parameters. Note that CDMS and XENON report their results
491: directly in terms of the recoil energy already corrected by the
492: quenching factor. In contrast to DAMA, here this is possible
493: because only a single element is used as target.
494: 
495: 
496: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
497: \section{Standard halo results}
498: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
499: \label{sec:std-halo}
500: 
501: 
502: Figure~\ref{fig:regions} summarises our results assuming standard halo
503: properties, showing the allowed region from DAMA together with the
504: constraints from CDMS-Si, CDMS-Ge and XENON. First we discuss the fit
505: to DAMA data alone (without constraints from CDMS and XENON).  We find
506: two islands in the $(m_\chi, \sigma_p)$ plane where DAMA can be
507: accommodated. The best fit point is obtained at
508: %
509: \begin{equation}\label{eq:dama-bfp}
510: m_\chi = 12\,{\rm GeV}\,,\quad
511: \sigma_p = 1.3\times 10^{-41}\,{\rm cm}^2 \,,\qquad
512: \chi^2_\mathrm{DAMA,min} = 36.8/34\,{\rm dof} \,,
513: \end{equation}
514: %
515: with an excellent goodness of fit of 34\%. There is also a local
516: minimum at $m_\chi = 51$~GeV with $\chi^2_\mathrm{local} = 47.9$. This
517: solution is disfavoured with respect to the best fit point at about
518: $3\sigma$ for 2~dof ($\Delta\chi^2 = 11.1$). The allowed regions
519: around $m_\chi \simeq 50$~GeV shown in figure~\ref{fig:regions} are
520: defined with respect to the local minimum. The low and high WIMP-mass
521: solutions correspond to channeled and quenched scatterings on Iodine,
522: respectively. In contrast to the situation when all events are assumed
523: to be quenched~\cite{Gondolo:2005hh}, it turns out that scattering on
524: Sodium is not relevant once channeling of Iodine events takes
525: place. The reason is that quenched events on Sodium require a similar
526: WIMP mass as channeled events on Iodine (i.e., $m_\chi \simeq 10$~GeV)
527: but a much larger cross section $\sigma_p$ (due to the $A^2$
528: dependence of the total cross section on the nucleus), and therefore,
529: are highly suppressed once channeled scattering on Iodine takes
530: place. In principle there would be also a solution from channeled
531: events on Na, around $m_\chi \simeq 5$~GeV. However, it turns out that
532: in this case the un-channeled events on Na still contribute to the signal,
533: and indeed prevent fitting the data with the channeled Na events. Note
534: that the solution around $m_\chi \simeq 50$~GeV is excluded by some
535: orders of magnitude by XENON and CDMS-Ge, and therefore we focus in
536: the following on the low-mass region $m_\chi \simeq 10$~GeV.
537: 
538: \begin{figure}
539: \centering \includegraphics[width=0.7\textwidth]{regions}
540:   \mycaption{\label{fig:regions} Allowed regions at 90\% and
541:   99.73\%~CL for WIMP mass and scattering cross section on nucleon for
542:   DAMA, and exclusion contours for CDMS-Si, CDMS-Ge and XENON at
543:   90\%~CL. We also display the limit from CoGeNT extracted from figure~2
544:   of~\cite{Aalseth:2008rx}.  The global best fit for DAMA is marked
545:   with a star, the allowed region around $m_\chi \simeq 50$~GeV is
546:   defined with respect to the local minimum, which is marked with a
547:   dot. For DAMA we show the regions obtained from using only the
548:   modulation amplitude for 2--6~keVee (gray curves) and from using the
549:   spectral shape of the modulation signal (shaded regions). For
550:   parameters above the dashed curve the predicted number of events in
551:   DAMA/LIBRA is larger than the observed number of events.}
552: \end{figure}
553: 
554: 
555: The gray contours in figure~\ref{fig:regions} correspond to an
556: alternative method of fitting DAMA. Instead of using the detailed
557: spectral information of the annual modulation, we fit the time
558: dependence of their signal integrated over energy. In figure~6
559: of~\cite{Bernabei:2008yi} data on the residual rate $(S(t) - S_0)$
560: (c.f., Eq.~\ref{eq:DAMAsignal}) is given in 7~time bins of one single
561: annual cycle. For the gray contours in figure~\ref{fig:regions} we use
562: these data for the energy intervals 2 to 6~keVee and 6 to 14~keVee,
563: where in the latter interval data are consistent with no annual
564: variation. These results are very similar to the ones
565: of~\cite{Gondolo:2005hh} (if channeling is neglected, not shown in the
566: figure) and~\cite{Petriello:2008jj} (including channeling), where only
567: two data points for the modulation amplitude below and above 6~keVee
568: have been used.
569: 
570: \begin{figure}
571: \centering 
572: \includegraphics[height=0.45\textwidth]{spectrum-mod} \quad
573: \includegraphics[height=0.45\textwidth]{spectrum-rate}
574: %
575:   \mycaption{\label{fig:spectrum} Left: Energy distribution of the
576:   annual modulation amplitude from DAMA/NaI and DAMA/LIBRA data
577:   extracted from figure~9 of~\cite{Bernabei:2008yi} (points with error
578:   bars), together with the prediction for three examples of WIMP
579:   masses and scattering cross sections (curves). Right: Energy
580:   distribution of the time averaged rate observed in DAMA/LIBRA
581:   extracted from figure~1 of~\cite{Bernabei:2008yi} (points), together
582:   with the prediction for two examples of WIMP masses and scattering
583:   cross sections (thick curves) as well as the corresponding
584:   un-identified background (thin curves). The data are corrected for
585:   the energy dependent efficiency.}
586: \end{figure}
587: 
588: We observe from figure~\ref{fig:regions} that the two methods of
589: analysing DAMA data are consistent with each other (as it should
590: be), but also that using the spectral information gives significantly
591: stronger constraints on the allowed region. This is illustrated in
592: figure~\ref{fig:spectrum} (left), showing the 36 data points on the
593: modulation amplitude $A_i$ used in our default analysis. 
594: %
595: While the prediction from the best fit point of Eq.~\ref{eq:dama-bfp}
596: nicely follows the data (solid curve), moving to smaller WIMP masses
597: leads to a modulation signal more peaked at the lowest
598: energies. Therefore, although it is still possible to obtain the
599: integrated signal in the interval from 2 to 6~keVee, the spectral
600: shape is clearly inconsistent with data, as illustrated for $m_\chi =
601: 6$~GeV by the dashed curve.\footnote{The value for the cross section
602: $\sigma_p = 2.8\times 10^{-41}$~cm$^2$ formally gives the best fit to
603: the data shown in figure~\ref{fig:spectrum} (left) for $m_\chi =
604: 6$~GeV. However, the value required to obtain the integrated
605: modulation amplitude for this $m_\chi$ is about a factor 2 larger,
606: $\sigma_p = 6\times 10^{-41}$~cm$^2$, as can be seen from the gray
607: contours shown in figure~\ref{fig:regions}.}
608: 
609: Finally we mention the implication of the data on the time averaged
610: rate observed in DAMA. Parameter values above the dashed curve in
611: figure~\ref{fig:regions} are excluded because they would lead to a
612: higher event rate than observed. This leads to additional constraints
613: for the high-mass solution. In figure~\ref{fig:spectrum} (right) we show
614: the observed rate together with the predictions for the two local
615: minima. Note that for the DAMA/LIBRA exposure of 0.53~t~yr statistical
616: errors are not visible at the scale of the plot. Clearly, solutions
617: predicting a relatively large rate require that the un-identified
618: background drops rapidly in order to give space for the WIMP
619: signal. In particular, the solution at $m_\chi = 51$~GeV requires that
620: the background drops to zero in the first energy bin.  Although this
621: cannot be excluded a priori, at least such a background shape seems
622: somewhat unlikely. The issue is less severe for the best fit point at
623: $m_\chi = 12$~GeV, since the ratio of modulation amplitude to average
624: rate increases for decreasing WIMP mass. However, any point close to
625: the dashed line in figure~\ref{fig:regions} is affected by this problem.
626: 
627: \bigskip %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
628: 
629: From figure~\ref{fig:regions} we find that the parameters allowed by
630: DAMA data at 90\%~CL are excluded by the 90\%~CL limits of CDMS-Si,
631: CDMS-Ge, and XENON. If all data are combined by adding the individual
632: $\chi^2$ functions,
633: %
634: \begin{equation}
635: \chi^2_\mathrm{global} = \chi^2_\mathrm{DAMA} + \chi^2_\mathrm{CDMS-Ge}
636: + \chi^2_\mathrm{CDMS-Si} + \chi^2_\mathrm{XENON} \,,
637: \end{equation}
638: %
639: we find the minimum at $m_\chi = 9.5$~GeV and $\sigma_p = 1.2\times
640: 10^{-41}$~cm$^2$ with $\chi^2_\mathrm{global,min} = 59.3/(45-2)\,\rm
641: dof$, which corresponds to a 5\% goodness of fit.  
642: 
643: 
644: Let us note that the goodness of fit test based on
645: $\chi^2_\mathrm{min}/$dof often is not very sensitive to tensions in
646: the fit, especially in case of a large number of data points. This can
647: happen if there are many data points which actually are not sensitive
648: to the relevant parameters, and hence, allow to ``hide'' the problem
649: in the fit, see for example the discussion in~\cite{Maltoni:2003cu}.
650: Because of this the goodness of fit depends also on the way of binning
651: the data. To circumvent this weakness of the standard goodness of fit
652: test the so-called Parameter Goodness of fit (PG) can be
653: used~\cite{Maltoni:2003cu}. Whereas the standard test measures the
654: probability that all individual data points are fitted by an
655: hypothesis, the PG tests the consistency of different data sets under
656: an hypotesis. It is based on the $\chi^2$ function
657: %
658: \begin{equation} \label{eq:PG}
659:     \chi^2_\text{PG} =
660:     \chi^2_\text{global,min} - \sum_i \chi^2_{i,\text{min}} \,,
661: \end{equation}
662: %
663: where $\chi^2_\text{global,min}$ is the $\chi^2$ minimum of all data
664: sets combined and $\chi^2_{i,\text{min}}$ is the minimum of the data
665: set $i$. This $\chi^2$ function measures the ``price'' one has to pay
666: by the combination of the data sets compared to fitting them
667: independently. It follows a standard $\chi^2$ distribution and should
668: be evaluated for the number of dof corresponding to the number of
669: parameters in common to the data sets, see~\cite{Maltoni:2003cu} for a
670: precise definition.
671: 
672: To apply this method we consider the two data sets DAMA versus all the
673: other data showing no evidence. Hence, we combine CDMS-Ge, CDMS-Si,
674: and XENON into one data set which we denote by NEV. Then we find
675: $\chi^2_\text{PG} = 22.6$. Evaluating this for 2~dof (corresponding to
676: the two parameters $m_\chi$ and $\sigma_p$ in common to both data
677: sets) one finds that DAMA and NEV data are consistent only at a
678: probability of $1.2\times 10^{-5}$. This corresponds roughly to the
679: probability of a $2.9\sigma$ fluctuation in both data sets at the same
680: time.\footnote{\label{ft:gof} Note that the standard
681:   $\chi^2_\mathrm{min}$/dof probability of 5\% tests the fit of all
682:   individual data points at the best fit solution, whereas the PG of
683:   $1.2\times 10^{-5}$ reflects the compatibility of the DAMA and NEV
684:   data sets. These are different questions and therefore the two
685:   probabilities are consistent with each other. Our DAMA analysis is
686:   based on 36 data points on the modulation amplitude, where most of
687:   them (above 6~keVee) are always fitted perfectly, irrespectively of
688:   the disagreement with CDMS/XENON. This is an example of the above
689:   mentioned ``dilution'' of the standard goodness of fit test.} We
690: conclude that the explanation of DAMA results in terms of
691: spin-independent elastic scattering of WIMPs with standard halo
692: properties is strongly disfavoured by XENON and CDMS data. Next we
693: investigate the stability of this result with respect to modifications
694: of the velocity distribution of the WIMPs in the halo of our galaxy.
695: 
696: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
697: \section{Non-standard halos}
698: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
699: \label{sec:nonstd-halo}
700: 
701: The precise limits on the cross section and mass of a dark matter
702: candidate which are obtained from a particular
703: observation/non-observation of a signal in a direct detection experiment
704: depend upon the velocity dispersion of dark matter around the detector, and consequently ultimately upon astrophysical assumptions.  The same set of astrophysical assumptions are
705: normally made by different experiments so that their results can be
706: compared with each other, the first of which being that the dark matter halo of the
707: galaxy is an isothermal sphere which means a spherically symmetric density distribution of the form $\rho\propto r^{-2}$.  For such a density distribution the Keplerian velocity is independent of radius and the value of this velocity normally assumed for dark matter studies is
708: a radius independent Keplerian velocity of 220 kms$^{-1}$.  It is also
709: assumed that the velocity dispersion of the dark matter profile is everywhere
710: isotropic and Gaussian, the width of the Gaussian distribution
711: corresponding to the Keplerian velocity of the profile.  It turns out
712: that we do not actually expect any of these assumptions to hold true
713: for a realistic dark matter halo.
714: 
715: Over the past decade, $N$-body simulations of increasingly large
716: numbers of dark matter particles have allowed us to obtain more
717: information about the kind of dark matter halos that one would expect
718: to form in an expanding universe (see e.g.~\cite{Springel:2005nw}).
719: These simulations show that one might expect a dark matter density
720: that decreases more steeply with radius at large radii rather than
721: have the same power law at all radii as in an isothermal profile
722: \cite{Navarro:1995iw}.  Furthermore, the orbits taken by dark matter
723: particles in a realistic simulation are usually rather radial, resulting in an anisotropic
724: velocity dispersion \cite{Gustafsson:2006gr}.  
725: 
726: Note that one can also obtain an anisotropic velocity dispersion in a halo where the density distribution is not spherically symmetric, but rather triaxial as in \cite{Green:2002ht}.  In this work however, we only consider dark matter halos where the density distribution is spherically symmetric although the velocity dispersion does not have to be.
727: %
728: Also, in a non-extensive ideal gas where there is a long
729: range attractive force between the particles such as we have here in
730: the form of gravity, one generically expects deviations from Gaussian
731: velocity distribution \cite{Vergados:2007nc}.  Furthermore, if the dark matter is still not completely virialised but is still coming into equilibrium,  there will be a superposition of multiple dark matter populations at any given place in the halo.  This effect would also lead to deviations from a Gaussian distribution of velocities.
732: 
733: \begin{figure}
734: \centering 
735: \includegraphics[height=0.345\textwidth]{aplot} \quad
736: \includegraphics[height=0.345\textwidth]{cplot}
737: %
738:   \mycaption{\label{fig:vl} The parameters $\alpha_i$ and $f_i$
739:   explained in the text fitted to the radial and tangential velocity
740:   dispersions of dark matter at different radii from the centre of the
741:   Via Lactea simulation.  The vertical lines indicate the position of
742:   the Sun at $r=8.5$~kpc. The velocity dispersions are clearly
743:   non-Gaussian as one approaches the centre of the galaxy.}
744: \end{figure}
745: 
746: In 2006, the results from a Milky Way size dark matter halo simulation
747: called Via Lactea containing 234 million particles were published
748: \cite{Diemand:2006ik}.  We have looked at this data
749: to see how much the dark matter distribution experienced by an
750: observer at the Solar radius within this simulation would vary from
751: the normal assumptions stated above for observers on Earth.
752: %
753: For each particle in the simulation there is a position vector $x_i$
754: and velocity vector $v_i$, plus the local gravitational potential per
755: unit mass in units of velocity squared $U(x_i)$.  We express the
756: velocities in terms of the square root of their local potential
757: $\tilde{v}_i=v_i/\sqrt{-U(x_i)}$.
758: %
759: Next we work out the angle between
760: the radial direction and the overall velocity vector.  We then use
761: this to decompose the velocity into a radial part and a part
762: perpendicular to that which we call tangential.  For each radius we
763: bin the tangential and the radial velocities obtaining two
764: distributions.  We fit the one dimensional radial distribution using
765: the following expression which we find to be a better fit than the
766: Tsallis distributions (designed to fit non-extensive or multiple temperature distributions) used in \cite{Vergados:2007nc}
767: %
768: \begin{equation}
769: \frac{1}{N_R}\exp\left[-\left(\frac{\tilde{v}_R^2}{f_R^2}\right)^{\alpha_R}\right] \,.
770: \end{equation}
771: %
772: Because we have rescaled the velocities with respect to $\sqrt{-U(x_i)}$, $f_R$ and $\alpha_R$ are dimensionless constants of order one.  The naive assumption is that the width of the velocity distribution at a given radius is simply the Keplerian velocity at that radius.  While information about the mass distribution of dark matter is required to go from the potential to the Keplerian velocity, The parameter $f_R$ is an indication of how badly this assumption is broken.  $\alpha$ encodes the deviation from
773: Gaussianity ($\alpha=1$ corresponding to a Gaussian).  For the one
774: dimensional case, the normalisation is analytic, $N_R = 2f_R
775: \Gamma(1+1/2\alpha_R)$.  We perform the same fitting procedure for the
776: tangential velocity, fitting
777: %
778: \begin{equation}
779: \frac{2\pi v_T}{N_T}
780: \exp\left[{-\left(\frac{\tilde{v}_T^2}{f_T^2}\right)^{\alpha_T}}\right]
781: \end{equation}
782: %
783: and while we are not aware of an analytic expression for $N_T$ it is
784: trivial to obtain it numerically. Note, in terms of the two dimensions
785: perpendicular to the radial direction $R$,
786: $\tilde{v}_T^2=\tilde{v}_\theta^2+\tilde{v}_\phi^2$.
787: 
788: Using the data from the Via Lactea simulation, we have fitted for
789: values of $f_i$ and $\alpha_i$ as a function of radius from the centre
790: of the galaxy.  The results, which can be seen in figure \ref{fig:vl},
791: show that there is a considerable deviation from Gaussianity in the
792: velocity dispersion of the galaxy.  Both the deviation from
793: Gaussianity, the anisotropy of the velocity dispersion and the change
794: in the relationship between the width of the dispersion and the local
795: Keplerian velocity will change the ratio between the expected
796: modulation in the DAMA experiment and the total expected events at
797: XENON and CDMS.  We have calculated these changes and
798: attempted to see if they can increase the likelihood of the results
799: from both experiments being compatible, the results of the new fits
800: can be seen in figure~\ref{fig:regions-nonstd}(a).
801: 
802: \begin{figure}
803: \centering 
804: \includegraphics[height=0.4\textwidth]{regions-vl} \qquad
805: \includegraphics[height=0.4\textwidth]{regions-110} \\
806: \includegraphics[height=0.4\textwidth]{regions-asym} \qquad
807: \includegraphics[height=0.4\textwidth]{regions-v_esc} 
808: %
809:   \mycaption{\label{fig:regions-nonstd} Allowed regions at 90\% and
810:   99.73\%~CL for DAMA, and exclusion contours for CDMS-Si, CDMS-Ge and
811:   XENON at 90\%~CL for the DM halo obtained in the Via Lactea
812:   simulation (a), an isotropic Maxwellian halo with dispersion $\bar v
813:   = 110$~km/s (b), an asymmetric Maxwellian halo with dispersion $\bar
814:   v_R = 142$~km/s in the radial direction and $\bar v_T = 63$~km/s in
815:   the tangential direction (c), and an isotropic Maxwellian halo with
816:   dispersion $\bar v = 220$~km/s and escape velocity $v_\mathrm{esc} =
817:   450$~km/s (d). The best fit for DAMA is marked with a star. In the
818:   panels (b) and (c) we show also the 90\% and 99.73\%~CL regions
819:   for the global data combining all experiments, as well as the global
820:   best fit point (marked with a dot). }
821: \end{figure}
822: 
823: It turns out that the deviation from the assumptions of the isothermal
824: sphere which are predicted by the Via Lactea simulation are not
825: sufficient to bring the region in parameter space favoured by DAMA
826: away from the region disfavoured by XENON and CDMS.  The numbers
827: associated with these regions are provided in table~\ref{tab:fits} and
828: show that using the velocity dispersion predicted by Via Lactea leads
829: to only a very small reduction in $\chi^2$ and the goodness of fit is
830: still unacceptably small.
831: 
832: \begin{table}
833: \begin{tabular}{l|cc|cc@{\quad}c@{\quad}cc}
834: \hline\hline
835: halo model & $\chi^2_\mathrm{DAMA,min}$ & $m_{\chi,\rm best}^{\rm DAMA}$ &
836:              $\chi^2_\mathrm{glob,min}$ & GOF &
837:              $\chi^2_\mathrm{PG}$ & PG & $m_{\chi,\rm best}^{\rm glob}$ \\
838: \hline
839: default analysis    & 36.8 & 12 & 59.3 & 0.05 & 22.6 & $1\times 10^{-5}$ & 9.5 \\
840: Via Lactea simulation
841:                     & 35.1 & 16 & 56.7 & 0.08 & 21.6 & $2\times 10^{-5}$ & 13.9  \\
842: Maxwellian 
843: $\bar v = 110$~km/s & 32.9 & 108& 46.8 & 0.32 & 13.5 & $1\times 10^{-3}$ & 16  \\
844: $\bar v_R = 142$~km/s, 
845: $\bar v_T = 63$~km/s& 32.7 & 18 & 39.6 & 0.62 & 6.5  & 0.04 & 18  \\
846: $v_\mathrm{esc} = 450$~km/s
847:                     & 36.5 & 12 & 51.6 & 0.17 & 15.1  & $5\times 10^{-4}$ & 11 \\
848: \hline\hline
849: \end{tabular}
850: %
851:   \mycaption{\label{tab:fits} Summary of the fits to DAMA data and
852:   global data (DAMA, CDMS-Ge, CDMS-Si, XENON) for different WIMP
853:   halos.  We give the best fit $\chi^2$ values, the goodness of fit
854:   (assuming 43~dof), the PG testing the consistency of DAMA with all
855:   other data, as well as the best fit WIMP masses (in GeV).}
856: \end{table}
857: 
858: It is therefore interesting to ask what kind of halo parameters could
859: lead to a better fit to the data, and how realistic would such
860: parameters be?  There are examples in the literature of the use of a
861: stream of dark matter to boost the annual modulation signal
862: \cite{Gondolo:2005hh}.  In this work, we choose to retain the
863: spherical symmetry of the halo density and instead of adding a stream, vary both the
864: width of the velocity dispersion and the anisotropy parameter $\beta_{vel}$
865: defined as
866: %
867: \begin{equation}
868: \beta_{vel}=1-\frac{\bar{v}_T^2}{\bar{v}_R^2} \,.
869: \end{equation}
870: 
871: A reduction in the width of the velocity
872: dispersion of dark matter alone helps reconcile the two data sets without the need to introduce anisotropy.  If
873: we assume an isotropic distribution of dark matter ($\beta_{vel}=0$)
874: and reduce $\bar{v}$ to 110 km/s which is half of the Keplerian
875: velocity at the solar radius, the goodness of fit increases
876: dramatically (see table~\ref{tab:fits}).  A further improvement in the
877: fit is made if one assumes a velocity dispersion lower than Keplerian,
878: but also highly anisotropic such that $\bar{v}_R=142$ km/s and
879: $\bar{v}_T=63$ km/s. As visible in figure~\ref{fig:regions-nonstd}~(c)
880: in this cases the entire 90\%~CL region of DAMA is consistent with the
881: 90\%~CL bounds from CDMS and XENON. For the asymmetric velocity
882: distribution the global $\chi^2$ drops by about 20 units compared to
883: the default analysis and provides an excellent goodness of fit of
884: 62\%.  The PG test gives compatibility of DAMA with NEV data with a
885: probability of 4\%, due to the remaining constraint from XENON.
886: 
887: A valid question is then whether such low and anisotropic values of
888: the velocity dispersions are at all realistic. In order to check on
889: the feasibility of such values, one needs to think about particular
890: dark matter halos and see if the solutions of the (Maxwell-) Jeans equations
891: allow simultaneously both a high velocity anisotropy ($\beta_{vel}
892: \sim 0.8$ ) and low velocity distribution at the location of the Sun.
893: 
894: In order to solve the Jeans equations, we will need to understand the distribution of mass in the galaxy.  Integration of a spherically symmetric dark matter profile with a well defined functional form is trivial.  However, at the radius of the Sun, it is important to consider not only dark
895: matter but also the presence of baryons, which make up most of the
896: mass in the central regions of the galaxy.  To model the Milky Way
897: baryon density we assume cylindrical symmetry and ignore any spiral
898: arms or bars. For the central bulge of stars we assume a density of
899: the form $\rho\propto r^{-\gamma}e^{-r/\lambda}$ while for the disk we
900: assume a (Kuzmin) delta function of matter in the $z$ direction ($z$
901: is the coordinate perpendicular to the disk) with a surface density
902: $\sigma_{\mathrm{disk}}(r) =
903: \frac{cM_{\mathrm{disk}\infty}}{2\pi\left(r^2+c^2\right)^{\frac{3}{2}}}$. We
904: choose the parameters of the model to match observations of the Milky
905: Way: $\gamma=1.85$, $\lambda=1 \,\mathrm{kpc}$, $c=5 \,\mathrm{kpc}$
906: and with the total disk and bulge mass $M_{\mathrm{disk}\infty} = 5
907: M_{\mathrm{bulge}}=6.5\times10^{10}M_{\odot}$ \cite{Zhao:1995qh,
908: Dehnen:1996fa, Klypin:2001xu, Kent:1991me}.  We assume that the disk
909: comes to an end at a radius of 15~kpc.
910: 
911: In order to parametrise our dark matter density profile we will
912: consider a profile which assumes two asymptotic radial power law
913: behaviors at both small ($\gamma$) and large ($\beta$) radii\footnote{This $\beta$ in the density profile should not be confused with the velocity dispersion anisotropy parameter $\beta_{vel}$.}. In this
914: profile, known as the '$\alpha\beta\gamma$' profile (or the Zhao
915: profile), the density as a function of radius is given by the
916: expression
917: %
918: \begin{equation}
919: \rho(r)=\frac{\rho_0}{(r/a)^\gamma\left[1+(r/a)^\alpha\right]^{\frac{\beta-\gamma}{\alpha}}}
920: \label{eq:abg}
921: \end{equation}
922: %
923: where $\alpha$ governs the radial rate at which the profile
924: interpolates between the asymptotic powers $-\gamma$ and $-\beta$.
925: The parameter $a$ is a characteristic scale radius determining the
926: location dividing the two regions described by a single power law.
927: 
928: Having assumed a value for $\alpha,\beta,\gamma$ and $a$ we then solve
929: for $\rho_0$ in order to get the correct value of the Keplerian
930: velocity at the solar radius.  This also determines the location of
931: the virial radius $r_{vir}$ which in this work is defined to be the
932: radius of the sphere within which the average density is 250 times the
933: critical density of the universe (we assume $h=0.7$).  The ratio
934: between $r_{vir}$ and $a$ is referred to as the concentration of the dark matter
935: halo.
936: 
937: Once we are in possession of these parameters, we can proceed to solve
938: the Jeans equation for the radial velocity dispersion
939: \cite{1987gady.book.....B}
940: %
941: \begin{equation}
942: \frac{1}{\rho}\frac{d\left(\rho\bar{v}_R^2\right)}{dr}+\frac{2\beta_{vel}\bar{v}_R^2}{r}=-\frac{d\phi}{dr}=-\frac{V_c^2}{r} \,,
943: \end{equation}
944: %
945: where $\phi(r)$ and $V_c(r)$ are the potential and Keplerian velocity
946: at a given radius.  We integrate this equation inwards from a large
947: radius several times the magnitude of $r_{vir}$ where we assume that
948: $\rho\bar{v}_R^2=0$.  We have checked that the result at $r=8.5$ kpc
949: is independent of the exact radius at which this boundary condition is
950: applied.  We have also assumed that, in the absence of a better
951: approximation, the anisotropy parameter $\beta_{vel}$ is a constant
952: throughout the halo.
953: 
954: \begin{figure}
955: \centering \includegraphics[height=0.5\textwidth]{conc}
956:   \mycaption{\label{fig:jeans} Here we plot the radial velocity
957:   dispersion $\bar v_R$ at the solar radius $r=8.5$~kpc as a function
958:   of the concentration of the dark matter halo $r_{vir}/a$ for two
959:   different dark matter profiles.  We have assumed that the velocity
960:   dispersion anisotropy parameter $\beta_{vel}=0.8$ and is a constant
961:   with respect to radius.  The horizontal line corresponds to the
962:   value of $\bar{v}_R$ which helps explain the discrepancy.  It
963:   appears that only for sets of halo parameters such as
964:   $(\alpha,\beta,\gamma)=(1,4,1.5)$ can one reconcile such a high
965:   value of $\beta$ with a low enough radial velocity dispersion to
966:   help explain the discrepancy between DAMA and XENON/CDMS.}
967: \end{figure}
968: 
969: The results are plotted in figure \ref{fig:jeans} and show that for a
970: NFW profile where the parameters are chosen such that
971: $(\alpha,\beta,\gamma)=(1,3,1)$ it seems to be rather difficult to
972: imagine that such a large value of the velocity anisotropy $\beta_{vel}$
973: could be consistent with low enough values of the velocity dispersion
974: to match the data.
975: %
976: We also look at a non-standard halo with
977: $(\alpha,\beta,\gamma)=(1,4,1.5)$.  The inner slope of such a halo is
978: quite steep, but even larger values of $\gamma$ than this may be
979: expected in dark matter halos where adiabatic contraction due to the
980: presence of baryons has occurred \cite{Gustafsson:2006gr}.  The rate
981: at which density decreases at larger radii is also larger than what is normally assumed.
982: 
983: If we are willing to accept such parameters for the dark matter
984: profile, it seems that the highly anisotropic value of $\beta_{vel}\sim 0.8$
985: that we require as one ingredient to make the DAMA data more
986: consistent with XENON and CDMS is not completely inconsistent with the
987: very low velocity dispersions that form the other ingredient.  The
988: analysis presented here is meant only as a suggestion of the magnitude
989: of possible effects.  If such explanations of the DAMA data were to be
990: taken seriously, a much deeper analysis of the Jeans equations should
991: be undertaken.  
992: 
993: Finally we mention that if one assumes a very low dark
994: matter escape velocity at the solar radius then one would remove many
995: of the fastest moving dark matter particles which would leave the
996: halo.  This would also result in more accord between DAMA and other
997: experiments but obtaining a large enough effect is difficult -- as
998: expressed in table~\ref{tab:fits}, lowering the escape velocity to 450~km/s
999: would only marginally make the fit more acceptable.  This, however,
1000: must be considered an unrealistic solution, since the escape velocity
1001: at the Solar radius is already 440~km/s even if there were no more
1002: matter in the Galaxy at larger radii.
1003: 
1004: To summarise this section, it seems that without the use of streams
1005: but rather by considering highly anisotropic velocity dispersions with
1006: magnitudes far below the local Keplerian velocity at the radius of the
1007: Sun would it be possible to reduce the conflict between DAMA and
1008: XENON/CDMS.
1009: 
1010: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1011: \section{Conclusions}
1012: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1013: \label{sec:conclusions}
1014: 
1015: Prompted by recent results from DAMA/LIBRA which establish the annual
1016: modulation of their event rate at the 8.2$\sigma$ level, we have
1017: studied the interpretation of this signal in terms of spin-independent
1018: elastic WIMP scattering. We have shown that the energy spectrum of the
1019: modulation signal strongly restricts the region of WIMP masses below
1020: 10~GeV, confining WIMP masses consistent with the DAMA data close to
1021: $m_\chi \simeq 12$~GeV. This region is excluded by the limits from
1022: CDMS and XENON, and therefore we conclude that even if channeling is
1023: taken into account this interpretation of the DAMA modulation signal
1024: is disfavoured. Applying a stringent test to evaluate the
1025: consistency of DAMA with null-result experiments we find consistency
1026: only with a formal probability of $10^{-5}$.
1027: 
1028: We have studied how robust this result is with respect to variations
1029: of the WIMP velocity distribution in our galaxy by changing various
1030: parameters of the distribution function. We find that decreasing the
1031: dispersion of the distribution can somewhat reduce the tension in
1032: the fit. Adopting in addition an asymmetric WIMP velocity profile with
1033: a larger dispersion in the radial direction than tangential improves
1034: the fit considerably.  We conclude that in principle it is possible to
1035: reconsile DAMA in the considered framework, at the price of rather
1036: exotic properties of the DM halo. The question remains whether such
1037: halo properties can be realistic at all. We have checked that a WIMP
1038: velocity distribution based on the Via Lactea $N$-body dark matter
1039: simulation does not improve the fit considerably with respect to the
1040: standard Maxwellian halo model.
1041: 
1042: Finally we mention that the negative conclusion on the compatibility
1043: of DAMA with CDMS and XENON relies crucially on the energy threshold
1044: of the latter two. In particular, a shift in the nuclear recoil energy
1045: scale in these experiments may change the conclusion. Indeed, the new
1046: measurements of the $\mathcal{L}_\mathrm{eff}$ parameter in
1047: XENON~\cite{Sorensen:2008ec} (which has not been implemented in the
1048: first arXiv version of this work) made the disagreement between DAMA
1049: and XENON somewhat less sever.
1050: 
1051: 
1052: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1053: \acknowledgments
1054: 
1055: We thank Graciela Gelmini for discussions in the initial stage of this
1056: work and are very grateful to J\"urg Diemand for providing us with the
1057: Via Lactea data. We thank the anonymous referee for pointing out the
1058: new $\mathcal{L}_\mathrm{eff}$ measurement relevant for the XENON10
1059: data analysis to us, and we acknowledge Laura Baudis for useful
1060: correspondence on this issue.
1061: 
1062: \appendix
1063: 
1064: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1065: \section{Comments on the DAMA spectral information}
1066: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1067: \label{sec:dama-mod}
1068: 
1069: Our results are largely based on the fact that DAMA spectral
1070: information excludes the low-mass WIMP region below
1071: 10~GeV. Obviously any effect which affects the spectral shape of the
1072: signal will have an impact on this conclusion. First, the smearing due
1073: to the energy resolution of the detector is important. We have checked
1074: this by artificially increasing the width of the energy resolution
1075: function given in Eq.~\ref{eq:Eres}~\cite{Bernabei:2008yh} by a factor
1076: of two. The global fit improves by roughly 7 units in $\chi^2$, but
1077: the tension between DAMA and NEV data persists at the level of
1078: $8\times 10^{-4}$, compare table~\ref{tab:fits-mod} and
1079: figure~\ref{fig:regions-mod-DAMA} (left).
1080: 
1081: 
1082: \begin{table}
1083: \begin{tabular}{l|cc|cc@{\quad}c@{\quad}cc}
1084: \hline\hline
1085:            & $\chi^2_\mathrm{DAMA,min}$ & $m_{\chi,\rm best}^{\rm DAMA}$ &
1086:              $\chi^2_\mathrm{glob,min}$ & GOF &
1087:              $\chi^2_\mathrm{PG}$ & PG & $m_{\chi,\rm best}^{\rm glob}$ \\
1088: \hline
1089: default analysis    & 36.8 & 12 & 59.3 & 0.05 & 22.6 & $1\times 10^{-5}$ & 9.5 \\
1090: \hline
1091: double $\sigma_E^\mathrm{DAMA}$ 
1092:                     & 37.9 & 11 & 52.1 & 0.16 & 14.3 & $8\times 10^{-4}$ & 9.0 \\
1093: w/o $1^\mathrm{st}$ DAMA data point
1094:                     & 30.3 & 10 & 44.4 & 0.37 & 14.1 &  $9\times 10^{-4}$ & 8.6 \\
1095: \hline\hline
1096: \end{tabular}
1097: %
1098:   \mycaption{\label{tab:fits-mod} Summary of fits to DAMA data and
1099:   global data (DAMA, CDMS-Ge, CDMS-Si, XENON) for the two ad-hoc
1100:   modifications of the DAMA analysis of
1101:   figure~\ref{fig:regions-mod-DAMA}.  We give the best fit $\chi^2$
1102:   values, the goodness of fit (assuming 42~dof for the last row and
1103:   43~dof otherwise), the PG testing the consistency of DAMA with all
1104:   other data, as well as the best fit WIMP masses (in GeV).}
1105: \end{table}
1106: 
1107: \begin{figure}
1108: \centering 
1109: \includegraphics[height=0.4\textwidth]{regions-e_res} \qquad
1110: \includegraphics[height=0.4\textwidth]{regions-1st-bin}
1111: %
1112:   \mycaption{\label{fig:regions-mod-DAMA} Allowed regions at 90\% and
1113:   99.73\%~CL for DAMA, and exclusion contours for CDMS-Si, CDMS-Ge and
1114:   XENON at 90\%~CL for two ad-hoc modifications of the DAMA
1115:   analysis. Left: we artificially assume an energy resolution in DAMA
1116:   a factor two worse than the value given
1117:   in~\cite{Bernabei:2008yh}. Right: omitting the lowest energy bin of
1118:   the annual modulation spectrum between 2 and 2.5~keVee.  The best
1119:   fit for DAMA is marked with a star. In the right panel we show also
1120:   the 90\% and 99.73\%~CL regions for the global data combining all
1121:   experiments, as well as the global best fit point (marked with a
1122:   dot). }
1123: \end{figure}
1124: 
1125: From figure~\ref{fig:spectrum} (left) it follows that the somewhat low
1126: data point in the first energy bin is very important in constraining
1127: the WIMP mass. We have repeated the analysis by excluding this bin
1128: from the fit, using only the data on the modulation signal above
1129: 2.5~keVee. In this case the DAMA allowed region extends to lower
1130: values of the WIMP mass, and once the NEV data are added the globally
1131: allowed region includes values of $m_\chi \sim 4$~GeV and $\sigma_p
1132: \sim 10^{-39}$~cm$^2$ at 90\%~CL. This region originates from
1133: channeled events on Sodium which now can accommodate the spectrum
1134: without the first bin, despite the contribution of un-channeled Na
1135: events. Let us note, however, that in this region constraints from
1136: other experiments, like CRESST-I~\cite{Altmann:2001ax},
1137: TEXONO~\cite{Lin:2007ka}, or CoGeNT~\cite{Aalseth:2008rx} are
1138: relevant. The global best fit point has an excellent
1139: $\chi^2_\mathrm{glob,min} = 44.4/42$~dof.\footnote{The improvement of
1140:   the consistency of NEV and DAMA data is only partially visible in
1141:   the PG value of about 0.1\%, which is still rather low.  The reason
1142:   is that also the fit of DAMA alone improves from
1143:   $\chi^2_\mathrm{DAMA,min} = 36.8$ to 30.3 by dropping the first bin,
1144:   which compensates partially the improvement in the global fit.}
1145: 
1146: Furthermore, we remark that any systematical uncertainty affecting the
1147: low energy spectrum may be relevant. For example, figure~26
1148: of~\cite{Bernabei:2008yh} shows that the efficiency for DAMA
1149: single-hit events starts to deviate from 1 below about 8~keVee, just
1150: in the signal region. Therefore, a possible uncertainty on this low
1151: energy efficiency may affect the exclusion of the light WIMP
1152: region. In the absence of detailed information on possible
1153: energy-dependent systematic uncertainties we neglect such effects in
1154: our analysis.
1155: %
1156: Let us note that the {\it ratio} of the signals in June and December
1157: would be less affected by systematics, since any multiplicative
1158: uncertainty (even energy dependent) would cancel, whereas the rate
1159: {\it difference} published by the DAMA collaboration is affected by
1160: such uncertainties.
1161: 
1162: Finally, we mention that the so-called Migdal effect could lead to
1163: modifications of the predicted energy spectrum in DAMA,
1164: see~\cite{Bernabei:2007jz} for a discussion and references. An
1165: investigation of this effect is beyond the scope of this work.
1166: 
1167: 
1168: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1169: \section{Comparison with other studies}
1170: \label{app:comparison}
1171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1172: 
1173: In this appendix we compare the results of our work to some studies
1174: from other authors. The authors of~\cite{Petriello:2008jj} come to
1175: a positive conclusion on the consistency between DAMA and constraints
1176: from other experiments, since they do not include the information on
1177: the spectral shape of the DAMA signal. In a work which appeared after
1178: ours on the preprint server \cite{Hooper:2008cf}, the same authors
1179: performed also an analysis including the spectrum which is in
1180: agreement with our results.
1181: 
1182: Our work appeared on the preprint server basically at the same time as
1183: \cite{Chang:2008xa}, with similar results. As in our study
1184: these authors emphasize the importance of the spectral information of
1185: the DAMA annual modulation and the constraint from the total
1186: unmodulated rate. Whereas~\cite{Chang:2008xa} discusses DM streams,
1187: our work considers non-standard halo models.
1188: 
1189: A similar study has been performed in \cite{Savage:2008er}, stressing
1190: also the relevance of spectral information and extending the analysis
1191: to spin-dependent cross sections. The general results for the
1192: spin-independent case are in quantitative agreement with us,
1193: though in some cases the authors draw different conclusions. In
1194: particular, they use a variety of statistical tests complementary to
1195: ours. Whereas our methods are largely based on parameter estimation
1196: ($\Delta\chi^2$ values with respect to the best fit point), these
1197: authors show also contours of probabilities from a goodness-of-fit test
1198: based on absolute $\chi^2$ values. As mentioned at the end of
1199: section~\ref{sec:std-halo} this method is often not very sensitive to
1200: a tension between different data sets, see also
1201: footnote~\ref{ft:gof}. Furthermore, by showing contours up to the
1202: 5 and even 7$\sigma$~CL they do find overlap regions. 
1203: %
1204: Let us also comment on the best fit point for DAMA, obtained at
1205: $m_\chi = 80$~GeV in Tab.~IV of~\cite{Savage:2008er}, compared to our
1206: result $m_\chi = 12$~GeV from Eq.~\ref{eq:dama-bfp}. The reason why
1207: we disfavour the fit in the large DM mass region is the inclusion of
1208: the constraint from the unmodulated rate in DAMA, which cuts away
1209: large part of this region, including also the best fit point of
1210: \cite{Savage:2008er}, compare figure~\ref{fig:regions}, and shifts the
1211: global best fit point to the low mass region.
1212: 
1213: 
1214: \bibliographystyle{apsrev}
1215: \bibliography{./dm}
1216: 
1217: \end{document}
1218: