0803.4477/ds2.tex
1: \documentclass[12pt,fleqn,a4paper]{article} % A4 size
2: \setlength{\topmargin}{-0.5in}              % 1.5 - 0.5 = 1(in)
3: \setlength{\textheight}{24.62cm}
4: \setlength{\oddsidemargin}{-0.25in}         % 1 - 0.25 = 0.75(in)
5: \setlength{\textwidth}{17.19cm} 
6: %
7: \usepackage{epsfig,graphics,graphicx}
8: \usepackage{clshan-math,clshan-dm-dd}
9: %
10: \def\lsim{\:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\:}
11: \def\gsim{\:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\:}
12: %
13: \begin{document}
14: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
15: 
16: \thispagestyle{empty}
17: \begin{flushright}
18: KIAS--P08026 \\
19: March 2008
20: \end{flushright}
21: \begin{center}
22: {\Large\bf
23:  Model--Independent Determination of the WIMP Mass      \\  \vspace{0.2cm}
24:   from Direct Dark Matter Detection Data}          \\
25: \vspace*{0.7cm}
26:  {\sc Manuel Drees}$^{1,2}$ and {\sc Chung-Lin Shan}$^1$ \\
27: \vspace*{0.3cm}
28: ${}^1$ {\it Physikalisches Inst. der Univ. Bonn, Nussallee 12, 53115 Bonn,
29:  Germany} \\
30: ${}^2$ {\it KIAS, School of Physics, 207--43 Cheongnyangni--dong, Seoul
31:  130--012, Republic of Korea}
32: \end{center}
33: \vspace{1cm}
34: 
35: \begin{abstract}
36:   Weakly Interacting Massive Particles (WIMPs) are one of the leading
37:   candidates for Dark Matter. We develop a model--independent method for
38:   determining the mass $m_\chi$ of the WIMP by using data (i.e., measured
39:   recoil energies) of direct detection experiments. Our method is independent
40:   of the as yet unknown WIMP density near the Earth, of the form of the WIMP
41:   velocity distribution, as well as of the WIMP--nucleus cross section.
42:   However, it requires positive signals from at least two detectors with
43:   different target nuclei. In a background--free environment, $m_\chi \sim
44:   50$ GeV could in principle be determined with an error of $\sim 35\%$ with
45:   only $2 \times 50$ events; in practice upper and lower limits on the recoil
46:   energy of signal events, imposed to reduce backgrounds, can increase the
47:   error. The method also loses precision if $m_\chi$ significantly exceeds the
48:   mass of the heaviest target nucleus used.
49: \end{abstract}
50: \clearpage
51: %
52: %
53: \section{Introduction}
54: 
55: First indications for the existence of Dark Matter were already found in the
56: 1930s \cite{evida}.  By now there is strong evidence \cite{evida}-\cite{bullet}
57: to believe that a large fraction (more than 80\%) of all matter in the
58: Universe is dark (i.e., interacts at most very weakly with electromagnetic
59: radiation and ordinary matter). The dominant component of this cosmological Dark Matter must be
60: due to some yet to be discovered, non--baryonic particles.  Weakly Interacting
61: Massive Particles (WIMPs) $\chi$ are one of the leading candidates for Dark
62: Matter.  WIMPs are stable particles which arise in several extensions of the
63: Standard Model of electroweak interactions. Typically they are presumed to
64: have masses between 10 GeV and a few TeV and interact with ordinary matter
65: only weakly (for reviews, see \cite{susydm}).
66: 
67: Currently, the most promising method to detect many different WIMP candidates
68: is the direct detection of the recoil energy deposited in a low--background
69: laboratory detector by elastic scattering of ambient WIMPs on the target
70: nuclei \cite{detaa, detab}. The recoil energy spectrum can be calculated
71: from an integral over the one--dimensional WIMP velocity distribution
72: $f_1(v)$, where $v$ is the absolute value of the WIMP velocity in the
73: laboratory frame. If this function is known, the WIMP mass $m_\chi$ can be
74: obtained from a one--parameter fit to the normalized recoil spectrum
75: \cite{green}, once a positive WIMP signal has been found. However, this
76: introduces a systematic uncertainty which is difficult to control. We remind
77: the reader that $N-$body simulations of the spatial distribution of Cold Dark
78: Matter (which includes WIMPs) seem to be at odds with observations at least
79: in the central region of galaxies \cite{crisis}; this may have ramifications
80: for $f_1$ as well. On the theory side, several modifications of the standard
81: ``shifted Maxwellian'' distribution \cite{susydm} have been suggested, ranging
82: from co-- or counter--rotating halos \cite{rotate} to scenarios where the
83: three--dimensional WIMP velocity distribution gets large contributions from
84: discrete ``streams'' with (nearly) fixed velocity \cite{sikivie,streams}.
85: 
86: The goal of our work is to develop model--independent methods which allow to
87: determine $m_\chi$ directly from (future) experimental data.  This builds on
88: our earlier work \cite{DMDD}, where we showed how to determine (moments of)
89: $f_1(v)$ from the recoil spectrum in direct WIMP detection experiments. In
90: this earlier analysis $m_\chi$ had been the only input required; one does not
91: need to know the WIMP--nucleus scattering cross section, nor the local WIMP
92: density. The fact that this method will give a result for $f_1$ for any
93: assumed value of $m_\chi$ already tells us that one will need at least two
94: different experiments, with different target nuclei, to model--independently
95: determine the WIMP mass from direct detection experiments. To do so, one
96: simply requires that the values of a given moment of $f_1$ determined by both
97: experiments agree. This leads to a simple expression for $m_\chi$, which can
98: easily be solved analytically; note that each moment can be used. An
99: additional expression for $m_\chi$ can be derived under the assumption that
100: the ratio of scattering cross sections on protons and neutrons is known. This
101: is true e.g., for spin--independent scattering of a supersymmetric neutralino,
102: which is the perhaps best motivated WIMP candidate \cite{susydm}; in this case
103: the spin--independent cross section for scattering on a proton is almost the
104: same as that for scattering on a neutron. Not surprisingly, the best result
105: obtains by combining measurements of several moments with that derived from
106: the assumption about the ratio of cross sections.
107: 
108: The remainder of this article is organized as follows. In Sec.~2 we review
109: briefly the methods for estimating the moments of the velocity distribution
110: function, paying special attention to experimentally imposed limits on the
111: range of allowed recoil energies. In Sec.~3 we will present the formalism for
112: determining the WIMP mass. Numerical results, based on Monte Carlo simulations
113: of future experiments, will be presented in Sec.~4. We conclude in Sec.~5.
114: Some technical details of our calculation will be given in an Appendix.
115: %
116: 
117: \section{Determining the moments of the velocity distribution of WIMPs}
118: 
119: In this section we review briefly the method for estimating the moments of the
120: one--dimensional velocity distribution function $f_1$ of WIMPs from the
121: elastic WIMP--nucleus scattering data. We first discuss the formalism, and
122: then describe how it can be implemented directly using (simulated, or future
123: real) data from direct WIMP search experiments.
124: 
125: \subsection{Formalism}
126: 
127: Our analysis starts from the basic expression for the differential rate for
128: elastic WIMP--nucleus scattering \cite{susydm}:
129: %
130: \beq \label{eqn2101}
131: \dRdQ = \calA \FQ \intvmin \bfrac{f_1(v)}{v} dv\, .
132: \eeq
133: %
134: Here $R$ is the direct detection event rate, i.e., the number of events per
135: unit time and unit mass of detector material, $Q$ is the energy deposited in
136: the detector, $F(Q)$ is the elastic nuclear form factor, and $v$ is the
137: absolute value of the WIMP velocity in the laboratory frame. The constant
138: coefficient $\calA$ is defined as
139: %
140: \beq \label{eqn2102}
141: \calA \equiv \frac{\rho_0 \sigma_0}{2 \mchi m_{\rm r,N}^2}\, ,
142: \eeq
143: %
144: where $\rho_0$ is the WIMP density near the Earth and $\sigma_0$ is the total
145: cross section ignoring the form factor suppression.  The reduced mass $m_{\rm
146:   r,N}$ is defined as
147: %
148: \beq \label{eqn2103}
149: m_{\rm r,N} \equiv \frac{\mchi \mN}{\mchi+\mN}\, ,
150: \eeq
151: %
152: where $\mchi$ is the WIMP mass and $\mN$ that of the target nucleus.  Finally,
153: $\vmin$ is the minimal incoming velocity of incident WIMPs that can deposit
154: the energy $Q$ in the detector:
155: %
156: \beq \label{eqn2104}
157: \vmin = \alpha \sqrt{Q}\, ,
158: \eeq
159: %
160: where we define
161: %
162: \beq \label{eqn2105}
163: \alpha \equiv \sfrac{\mN}{2 m_{\rm r,N}^2}\, .
164: \eeq
165:  
166: Eq.(\ref{eqn2101}) can be solved for $f_1$ \cite{DMDD}:
167: %
168: \beq \label{eqn2106} 
169: f_1(v) = \calN \cbrac{-2 Q \cdot \ddRdQoFQdQ}\Qva\, , 
170: \eeq 
171: %
172: where the normalization constant $\calN$ is given by
173: %
174: \beq \label{eqn2107} 
175: \calN = \frac{2}{\alpha} \cbrac{\intz \frac{1}{\sqrt{Q}} \bdRdQoFQ dQ}^{-1}
176: \, .  
177: \eeq 
178: %
179: Note that, first,
180: because $f_1(v)$ in Eq.(\ref{eqn2106}) is the {\em normalized} velocity distribution,
181: the normalization constant $\cal N$ here is {\em independent} of
182: the constant coefficient $\cal A$ defined in Eq.(\ref{eqn2102}).
183: Second, the integral here goes over the entire physically allowed range of
184: recoil energies, starting at $Q = 0$. The upper limit of the integral has been
185: written as $\infty$. However, it is usually assumed that the WIMP flux on
186: Earth is negligible at velocities exceeding the escape velocity \mbox{$v_{\rm esc}
187: \simeq 700$ km/s}. This leads to a kinematic maximum of the recoil energy,
188: %
189: \beq \label{qmaxkin}
190: Q_{\rm max,kin} = \frac {v^2_{\rm esc}} {\alpha^2}\, ,
191: \eeq
192: %
193: where $\alpha$ has been given in Eq.(\ref{eqn2105}). Eq.(\ref{eqn2107}) then
194: implies
195: %
196: \beq \label{fnorm}
197: \int_0^\infty f_1(v) \~ dv = \int_0^{v_{\rm esc}} f_1(v) \~ dv = 1\, .
198: \eeq
199: %
200: Using Eq.(\ref{eqn2106}), the moments of $f_1$ can be expressed as \cite{DMDD}
201: %
202: \beq \label{eqn2108} 
203: \expv{v^n} = \int_0^{v_{\rm esc}} v^n f_1(v) \~ dv
204: = \calN \afrac{\alpha^{n+1}}{2} (n+1) I_n\, , 
205: \eeq
206: %
207: which holds for all $n \geq 0$. Here the integral $I_n$ is given by
208: %
209: \beq \label{eqn2110} 
210: I_n = \int_0^{Q_{\rm max,kin}} Q^{(n-1)/2} \bdRdQoFQ dQ\, .  
211: \eeq 
212: %
213: In this notation, Eq.(\ref{eqn2107}) can be re--written as ${\cal N} = 2 /
214: (\alpha I_0)$. 
215: 
216: The results in Eqs.(\ref{eqn2106}) and (\ref{eqn2108}) depend on the WIMP mass
217: $m_{\chi}$ only through the coefficient $\alpha$ defined in
218: Eq.(\ref{eqn2105}). Evidently any (assumed) value of $m_\chi$ will lead to a
219: well--defined, normalized distribution function $f_1$ when used in
220: Eq.(\ref{eqn2106}). Hence $m_\chi$ can be extracted from a {\em single} recoil
221: spectrum {\em only if} one makes some assumptions about the velocity
222: distribution $f_1(v)$. 
223: 
224: A model--independent determination of $m_\chi$ thus requires that at least two
225: different recoil spectra, with two different target nuclei, have been
226: measured. As we will show in detail in the next section, $m_\chi$ can then be
227: obtained from the requirement that these two spectra lead to the same moments
228: of $f_1$.
229: 
230: Before coming to that, we have to incorporate the effects of a finite energy
231: acceptance of the detector. Any real detector will have a certain threshold
232: energy $Q_{\rm thre}$ below which it cannot register events. Off--line one may
233: need to impose a cut $Q > Q_{\rm min} > Q_{\rm thre}$ in order to suppress
234: (instrumental or physical) backgrounds. Similarly, background rejection may
235: require a maximum energy cut, $Q < Q_{\rm max} \leq Q_{\rm max,kin}$. In fact,
236: we will see below that, at least for smallish data samples, such a cut might
237: even be beneficial for the determination of $m_\chi$. We therefore now give
238: expressions for the case that only data with $Q_{\rm min} \leq Q \leq Q_{\rm
239:   max}$ are used in the analysis.
240: 
241: To that end, we introduce generalized moments of $f_1$:
242: %
243: \beqn \label{momnew}
244: \expv{v^n}(v_1, v_2) \= \int_{v_1}^{v_2} v^n f_1(v) \~ dv
245: \nonumber \\
246: \= {\cal N} \alpha^{n+1} \left[ \frac {Q_1^{(n+1)/2} r(Q_1)} {F^2(Q_1)} -
247: \frac {Q_2^{(n+1)/2} r(Q_2)} {F^2(Q_2)}+ \afrac {n+1} {2} I_n(Q_1,Q_2) 
248: \right] \, ,
249: \eeqn
250: %
251: where we have introduced the short--hand notation
252: %
253: \beq \label{r}
254: r(Q_i) = \left. \frac {dR} {dQ} \right|_{Q=Q_i}\, ,
255: ~~~~~~~~~~~~~~~~~~~~ % 20
256: i=1,~2\, ,
257: \eeq
258: %
259: and 
260: %
261: \beq \label{In}
262: I_n(Q_1,Q_2) = \int_{Q_1}^{Q_2} Q^{(n-1)/2} \bdRdQoFQ dQ \, .
263: \eeq
264: %
265: In order to arrive at the final expression in Eq.(\ref{momnew}) we used
266: Eq.(\ref{eqn2106}) and integrated by parts. The $Q_i$ in Eqs.(\ref{momnew})--(\ref{In})
267: are related to the original integration limits $v_i$ appearing
268: on the left--hand side of Eq.(\ref{momnew}) via Eq.(\ref{eqn2104}), i.e.,
269: %
270: \beq  \label{Qi}
271: Q_i = \frac {v_i^2} {\alpha^2}\, ,
272: ~~~~~~~~~~~~~~~~~~~~ % 20
273: i=1,~2\, .
274: \eeq
275: %
276: Of course, Eq.(\ref{momnew}) reduces to Eq.(\ref{eqn2108}) in the limit $v_1
277: \rightarrow 0, \, v_2 \rightarrow v_{\rm esc}$; note, however, that
278: Eq.(\ref{momnew}) is also applicable for the case $n = -1$, where the last
279: term on the right--hand side vanishes. The same restriction on the WIMP
280: velocity can also be introduced in the normalization constant ${\cal N}$ of
281: Eq.(\ref{eqn2107}), in which case $\int_{v_1}^{v_2} f_1(v) \~ dv$ is normalized
282: to unity. Formally this can be treated using Eqs.(\ref{momnew})--(\ref{Qi}) by
283: demanding $\expv{v^0}(v_1,v_2) = 1$.
284: 
285: \subsection{Experimental implementation}
286: 
287: In order to directly use our results for $f_1(v)$ and for its moments
288: $\expv{v^n}$ given in Eqs.(\ref{eqn2106}), (\ref{eqn2108}) and (\ref{momnew}),
289: one needs a functional form for the recoil spectrum $dR/dQ$. In practice this
290: results usually from a fit to experimental data. However, data fitting
291: can re--introduce some model dependence and makes the error analysis
292: more complicated.  Hence, expressions that allow to reconstruct $f_1(v)$ and
293: its moments directly from the data have been developed \cite{DMDD}. We
294: started by considering experimental data described by
295: %
296: \beq \label{eqn2201}
297:      {\T Q_n-\frac{b_n}{2}}
298:  \le \Qni
299:  \le {\T Q_n+\frac{b_n}{2}}\, ,
300:      ~~~~~~~~~~~~ %12
301:      i
302:  =   1,~2,~\cdots,~N_n,~
303:      n
304:  =   1,~2,~\cdots,~B.
305: \eeq
306: %
307: Here the total energy range has been divided into $B$ bins with central points
308: $Q_n$ and widths $b_n$. In each bin, $N_n$ events will be recorded. Note that
309: we assume that the sample to be analyzed only contains signal events, i.e., is
310: free of background, and ignore the uncertainty on the measurement of the
311: recoil energy $Q$. Active background suppression techniques \cite{pdg} should
312: make the former possible. The energy resolution of most existing detectors is
313: so good that its error will be negligible compared to the statistical
314: uncertainty for the foreseeable future.
315: 
316: Since the recoil spectrum $dR/dQ$ is expected to be approximately exponential,
317: we used the following ansatz for the spectrum in the $n-$th bin \cite{DMDD}:
318: %
319: \beq \label{eqn2202}
320:         \adRdQ_n
321:  \equiv \adRdQ_{Q \simeq Q_n}
322:  \simeq \trn \~ {\rm e}^{k_n (Q-Q_n)}
323:  \equiv \rn  \~ {\rm e}^{k_n (Q-Q_{s,n})}\, .
324: \eeq
325: %
326: Here $r_n$ is the standard estimator for $dR/dQ$ at $Q = Q_n$,
327: %
328: \beq \label{eqn2203}
329: r_n = \frac{N_n}{b_n}\, ,
330: \eeq
331: %
332: $\trn$ is the value of the recoil spectrum at the point $Q = Q_n$,
333: %
334: \beq \label{eqn2204}
335: \trn \equiv \adRdQ_{Q = Q_n}
336:  =      r_n \bfrac{k_n b_n/2}{\sinh (k_n b_n/2)}\, ,
337: \eeq
338: %
339: and $k_n$ is the logarithmic slope of the recoil spectrum in the $n-$th bin.
340: It can be computed numerically from the average $Q-$value in the $n-$th bin:
341: %
342: \beq \label{eqn2205}
343: \bQn = \afrac{b_n}{2} \coth\afrac{k_n b_n}{2}-\frac{1}{k_n}\, ,
344: \eeq
345: %
346: where
347: %
348: \beq \label{eqn2206}
349: \bQxn{\lambda} \equiv \frac{1}{N_n} \sumiNn \abrac{\Qni-Q_n}^{\lambda}\, .
350: \eeq
351: %
352: Finally, $Q_{s,n}$ is the shifted point at which the leading systematic error
353: due to the ansatz in Eq.(\ref{eqn2202}) is minimal \cite{DMDD},
354: %
355: \beq \label{eqn2207}
356: Q_{s,n} = Q_n+\frac{1}{k_n} \ln\bfrac{\sinh (k_n b_n/2)}{k_n b_n/2}\, .
357: \eeq
358: %
359: Note that $Q_{s,n}$ differs from the central point of the $n-$th bin, $Q_n$.
360: 
361: Using Eqs.(\ref{momnew}) and (\ref{r}) with $Q_1 = Q_{\rm min}$ and $Q_2 =
362: Q_{\rm max}$, the generalized $n-$th moment of the velocity distribution
363: function can be written as
364: %
365: \beq \label{eqn2208} 
366: \expv{v^n}(v(Q_{\rm min}), v(Q_{\rm max})) = \alpha^n \bfrac  {2 Q_{\rm
367:     min}^{(n+1)/2} r(Q_{\rm min}) / F^2(Q_{\rm min}) + (n+1) I_n(Q_{\rm min},
368:   Q_{\rm max})} {2 Q_{\rm min}^{1/2} r(Q_{\rm min}) / F^2(Q_{\rm min}) +
369:   I_0(Q_{\rm min}, Q_{\rm max})} \, ,  
370: \eeq
371: %
372: where $v(Q) = \alpha \sqrt{Q}$. Here we have implicitly assumed that $Q_{\rm
373:   max}$ is so large that terms $\propto r(Q_{\rm max})$ are negligible. We
374: will see later that this is not necessarily true for $n \geq 1$, since these
375: moments receive sizable contributions from large recoil energies \cite{DMDD}.
376: Nevertheless we will show in Sec.~3 that even in that case, Eq.(\ref{eqn2208})
377: can still be used for determining $m_\chi$. From the ansatz
378: Eq.(\ref{eqn2202}), the counting rate at $Q_{\rm min}$ can be expressed as
379: %
380: \beq \label{eqn2209}
381: r(Q_{\rm min}) = r_1 {\rm e}^{k_1 (Q_{\rm min} - Q_{s,1})}\, . 
382: \eeq
383: %
384: The integral $I_n(Q_{\rm min}, Q_{\rm max})$ defined in Eq.(\ref{In}) can be
385: estimated through the sum:
386: %
387: \beq \label{eqn2210} 
388: I_n(Q_{\rm min}, Q_{\rm max}) = \sum_a \frac{Q_a^{(n-1)/2}}{F^2(Q_a)}\, , 
389: \eeq
390: %
391: where the sum runs over all events in the data set that satisfy $Q_a \in
392: [Q_{\rm min}, Q_{\rm max}]$.
393: 
394: Since all $I_n$ are determined from the same data, they are correlated, with
395: \cite{DMDD}
396: %
397: \beq \label{eqn2211}
398: {\rm cov}(I_n,I_m) = \sum_a \frac{Q_a^{(n+m-2)/2}}{F^4(Q_a)}\, ,  
399: \eeq
400: %
401: where the sum again runs over all events with recoil energy between $Q_{\rm
402:   min}$ and $Q_{\rm max}$. 
403: 
404: On the other hand, the statistical error of $r(Q_{\rm min})$ can be obtained
405: from Eq.(\ref{eqn2209}) as
406: %
407: \beq \label{eqn2212}
408: \sigma^2(r(Q_{\rm min})) = r^2(Q_{\rm min}) 
409: \cbrac{ \frac{\sigma^2(r_1)}{r_1^2} + \bbrac{\frac{1}{k_1} - \afrac{b_1}{2} 
410: \left( 1 + \coth \afrac{b_1  k_1} {2} \right) }^2 \sigma^2(k_1)}\, .
411: \eeq
412: %
413: The error on $r_1$ follows directly from its definition in Eq.(\ref{eqn2203}):
414: %
415: \beq \label{eqn2213}
416: \sigma^2(r_n) = \frac{N_n}{b_n^2}\, .
417: \eeq
418: %
419: The error on the logarithmic slope $k_1$ can be computed from
420: Eq.(\ref{eqn2205}):
421: %
422: \beq \label{eqn2214}
423: \sigma^2(k_n) = k_n^2 \cbrac{1-\bfrac{k_n b_n/2}{\sinh (k_n b_n/2)}^2}^{-1}
424: \sigma^2\abrac{\bQn} \, ,
425: \eeq
426: %
427: with
428: %
429: \beq \sigma^2\abrac{\bQn} = \frac{1}{N_n-1} \bbigg{\bQQn-\bQn^2}\, .
430: \eeq
431: %
432: 
433: Finally, the correlation between the errors on $r(Q_{\rm min})$, which is
434: calculated entirely from the events in the first bin, and on $I_n$ is given by
435: \cite{DMDD}
436: %
437: \beqn \label{eqn2216}
438: \conti
439:    {\rm cov}(r(Q_{\rm min}),I_n)
440:    \non\\
441: \= r(Q_{\rm min}) \~ I_{n}(Q_{\rm min}, Q_{\rm min} + b_1)
442:    \non\\
443: \conti ~~~~ \times %4
444:    \cleft{\frac{\sigma^2(r_1)}{r_1^2} 
445:   +\bbrac{\frac{1}{k_1} - \afrac{b_1}{2} \left( 1 + \coth \afrac{b_1 k_1} {2} 
446: \right)   } }
447: \non\\
448: \conti ~~~~~~~~~~~~~~ \times % 14
449: \cright{\bbrac{\frac{I_{n+2}(Q_{\rm min}, Q_{\rm min}+b_1)}{I_{n}(Q_{\rm min},
450:     Q_{\rm min} + b_1)} - Q_1 + \frac{1}{k_1} - \afrac{b_1}{2} \coth \afrac{b_1
451:     k_1} {2}} \sigma^2(k_1)}\, ;
452: \eeqn
453: %
454: note that the integrals $I_i$ in Eq.(\ref{eqn2216}) only extend over the first
455: bin, which ends at $Q = Q_{\rm min} + b_1$.
456: 
457: \section{Determining the WIMP mass}
458: 
459: We are now ready to describe methods to extract the WIMP mass $m_\chi$ from
460: direct detection data. Recall that $m_\chi$ is an {\em input} in the
461: determination of $f_1$ and its moments as described in the previous
462: section. In particular, $m_\chi$ appears in the factor $\alpha^n$ on the
463: right--hand side of Eq.(\ref{eqn2208}) describing the experimental estimate
464: of the moments of $f_1$. A truly model--independent determination of $m_\chi$
465: from these data will therefore only be possible by requiring that two (or
466: more) experiments, using different target nuclei, lead to the same result for
467: $f_1$. 
468: 
469: Here we focus on the moments of the distribution function, rather than the
470: function itself, since non--trivial information about the former can already
471: be obtained with ${\cal O}(20)$ events \cite{DMDD}. We will also show how
472: $m_\chi$ can be estimated from the knowledge of the integral $I_0$ appearing
473: in the normalization of $f_1$, if the ratio of WIMP scattering cross sections
474: on protons and neutrons is known. For greater clarity, we will first discuss
475: the idealized situation where all signal events, irrespective of their recoil
476: energy $Q_a$, are included. In the second subsection we will introduce upper
477: and lower limits on $Q_a$. A third subsection is devoted to a discussion how
478: the different estimators for the WIMP mass can be combined using a $\chi^2$
479: fit. 
480: 
481: 
482: \subsection{Without cuts on the recoil energy}
483: 
484: If no cuts on the recoil energy need to be applied, we can use the original
485: moments, or equivalently, the generalized moments with $v_1 \rightarrow 0$ and
486: $v_2 \rightarrow v_{\rm esc}$; recall that the latter is the same as allowing
487: $v_2 \rightarrow \infty$, since by assumption $f_1(v) = 0$ for $v > v_{\rm
488:   esc}$. Eq.(\ref{eqn2208}) then simplifies to
489: %
490: \beq \label{eqn3101}
491: \expv{v^n} = \alpha^n (n+1) \afrac{I_n}{I_0}\, ,
492: \eeq
493: %
494: where $I_n$ and $I_0$ can be estimated as in Eq.(\ref{eqn2210}) with $Q_{\rm
495:   min} \rightarrow 0$ and $Q_{\rm max} \rightarrow \infty$ (or, equivalently,
496: $Q_{\rm max} \rightarrow Q_{\rm max,kin}$).
497: 
498: Suppose $X$ and $Y$ are two target nuclei. We denote their masses by $m_X$,
499: $m_Y$, and their form factors as $F_X(Q)$, $F_Y(Q)$. Similarly, we define
500: $\alpha_{X,Y}$ as in Eq.(\ref{eqn2105}), with $m_N \rightarrow m_{X,Y}$.
501: Eq.(\ref{eqn3101}) then implies
502: %
503: \beq \label{eqn3102}
504: \alpha_X^n \afrac{\InX}{\IzX} = \alpha_Y^n \afrac{\InY}{\IzY}\, .
505: \eeq
506: %
507: Note that the form factors in the estimates of $\InX$ and $\InY$ using
508: Eq.(\ref{eqn2210}) are different. Eq.(\ref{eqn3102}) can be solved for
509: $m_\chi$ using Eqs.(\ref{eqn2105}) and (\ref{eqn2103}):
510: %
511: \beq \label{eqn3103}
512: \mchi = \frac {\sqrt{\mX \mY} - \mX \calRn} {\calRn - \sqrt{\mX/\mY}}\, , 
513: \eeq
514: %
515: where we have defined
516: %
517: \beq \label{eqn3104} 
518: \calRn \equiv \frac{\alpha_Y}{\alpha_X} =
519: \abrac{\frac{\InX}{\IzX} \cdot \frac{\IzY}{\InY}}^{1/n}\, ,
520: ~~~~~~~~~~~~~~~~~~~~   n \ne    0,~-1.
521: \eeq
522: %
523: 
524: Using standard Gaussian error propagation, the statistical error on $m_\chi$
525: estimated from Eq.(\ref{eqn3103}) is
526: %
527: \beqn \label{eqn3105}
528:  \left.\sigma(\mchi)\right|_{\Expv{v^n}} 
529: \=  \frac{\sqrt{\mX/\mY} \vbrac{\mX-\mY}} {\abrac{\calRn -
530:     \sqrt{\mX/\mY}}^2} \cdot \sigma(\calRn)
531: \non\\
532:  \= \frac{\calRn \sqrt{\mX/\mY} \vbrac{\mX-\mY}} {\abrac{\calRn -
533:      \sqrt{\mX/\mY}}^2} 
534:         \non\\
535:  \conti ~~~~~~ \times %6
536:  \frac{1}{|n|} \bbrac{ \frac{\sigma^2(\InX)}{\InX^2}
537:  + \frac{\sigma^2(\IzX)}{\IzX^2}
538:  - \frac{2 {\rm cov}(\IzX,\InX)}{\IzX \InX}
539:  + (X \lto Y)}^{1/2}\, ,
540: \eeqn
541: %
542: where $\sigma^2(\InX) = {\rm cov}(\InX,\InX)$ and ${\rm cov}(\IzX,\InX)$ and
543: so on can be estimated from Eq.(\ref{eqn2211}).
544: 
545: A second method for determining $m_\chi$ starts directly from
546: Eq.(\ref{eqn2101}), plus an assumption about the relative strength for WIMP
547: scattering on protons p and neutrons n. The simplest such assumption is that the
548: scattering cross section is the same for both nucleons. This is in fact an
549: excellent approximation for the spin--independent contribution to the cross
550: section of supersymmetric neutralinos \cite{susydm}, and for all WIMPs which
551: interact primarily through Higgs exchange. Writing the ``pointlike'' cross
552: section $\sigma_0$ of Eq.(\ref{eqn2102}) as
553: %
554: \beq \label{sigma0}
555: \sigma_0 = \Afrac{4}{\pi} m_{\rm r,N}^2 A^2 |f_{\rm p}|^2\, ,
556: \eeq
557: %
558: where $f_{\rm p}$ is the effective $\chi \chi {\rm p p}$ 4--point coupling, $m_{\rm r,N}$ is the
559: reduced mass defined in Eq.(\ref{eqn2103}) and $A$ is the number of nucleons in the nucleus, we have from
560: Eqs.(\ref{eqn2101}), (\ref{eqn2102}) and the first Eq.(\ref{momnew}):
561: %
562: \beq \label{rQmin}
563: r(Q_{\rm min}) = \frac {\rho_0} {2 m_\chi} \Afrac{4}{\pi} A^2 |f_{\rm p}|^2 F^2(Q_{\rm min}) 
564: \expv{v^{-1}}(v(Q_{\rm min}), v_{\rm esc}) \, .
565: \eeq
566: %
567: Using the second Eq.(\ref{momnew}) we see that the counting rate at $Q_{\rm
568:   min}$ in fact drops out, and we are left with
569: %
570: \beq \label{one}
571: 1 = \frac { 4 \sqrt{2}}{\pi} \afrac{{\cal E} \rho_0 A^2 |f_{\rm p}|^2}{I_0}
572: \afrac{m_{\rm r,N}}{m_\chi \sqrt{\mN}}\, ,
573: \eeq
574: %
575: where we have assumed $Q_{\rm min} = 0$. Recall that the rate $dR/dQ$ is
576: defined as rate per unit mass and observation time interval, i.e., we need to
577: divide the actual event rate by the exposure ${\cal E} = M \tau$, where $M$ is
578: the (fiducial) mass of the detector and $\tau$ the observation time. In our
579: previous discussion this factor always dropped out in the end, due to the
580: appearance of the normalization ${\cal N}$. This is true also for the
581: right--hand side of Eq.(\ref{rQmin}), but not for the $1/{\cal E}$ factor on
582: the left--hand side of this equation; hence a factor ${\cal E}$ appears in the
583: numerator of Eq.(\ref{one}). Note that ${\cal E}$ is dimensionless in natural
584: units. On the other hand, the unknown factor $\rho_0 |f_{\rm p}|^2$ appearing in
585: Eq.(\ref{one}) will cancel out when we use this identity for two different
586: targets $X$ and $Y$, leading to the final result
587: %
588: \beq \label{msigma}
589: m_\chi =
590:  \frac{\abrac{\mX/\mY}^{5/2} \mY-\mX \calR_{\sigma}}{\calR_{\sigma}-\abrac{\mX/\mY}^{5/2}}\, .
591: \eeq
592: %
593: Here we have assumed $m_{X,Y} \propto A_{X,Y}$, and introduced the quantity
594: %
595: \beq \label{Rsigma}
596: {\cal R}_\sigma = \frac{{\cal E}_Y}{{\cal E}_X} \afrac{I_{0,X}}{I_{0,Y}}\, .
597: \eeq
598: %
599: 
600: Recall that, even though the derivation started from the expression for the
601: counting rate at $Q_{\rm min}$, this rate actually dropped out when going from
602: Eq.(\ref{rQmin}) to Eq.(\ref{one}). The final expression only depends on the
603: quantity $R_{\sigma}$, which is estimated from {\em all} the events in both
604: samples using Eq.(\ref{eqn2210}). The error on $m_\chi$ computed from
605: Eq.(\ref{msigma}) is therefore comparable to that for $m_\chi$ derived from a
606: moment of $f_1$. Still keeping $Q_{\rm min} = 0$, we have
607: %
608: \beq \label{sigsig}
609: \left. \sigma(m_\chi) \right|_\sigma
610:  = \frac{\calR_{\sigma} \abrac{\mX/\mY}^{5/2} \vbrac{\mX-\mY}}
611:         {\bbrac{\calR_{\sigma}-\abrac{\mX/\mY}^{5/2}}^2}
612:    \bbrac{\frac{\sigma^2\abrac{\IzX}}{\IzX^2}+\frac{\sigma^2\abrac{\IzY}}{\IzY^2}}^{1/2}\, .
613: \eeq
614: %
615: Ultimately the ${\cal R}_n,\, {\cal R}_\sigma$ and the errors in
616: Eqs.(\ref{eqn3105}) and (\ref{sigsig}) should be estimated from the data
617: directly. In the meantime it is instructive to note that the final expressions
618: for the statistical errors of our estimators for $m_\chi$ decompose into two
619: factors. The expectation value of the first factor does not depend on $f_1$;
620: it can be computed entirely from the masses of the involved particles:
621: %
622: \beqn \label{kappa}
623: \frac { {\cal R}_n \sqrt{m_X / m_Y} \left| m_X - m_Y \right|}
624: { \left( {\cal R}_n - \sqrt{ m_X / m_Y } \right)^2 } 
625: \= \frac{\calR_{\sigma} \abrac{\mX/\mY}^{5/2} \vbrac{\mX-\mY}}
626:         {\bbrac{\calR_{\sigma}-\abrac{\mX/\mY}^{5/2}}^2}
627:    \non\\
628: \= \frac { \left( m_\chi + m_X \right) \left( m_\chi + m_Y \right) } 
629: { \left| m_X - m_Y \right| }\non\\
630:  \eqnequiv \kappa \, .
631: \eeqn
632: %
633: Here we have made use of the identities ${\cal R}_n = \alpha_Y /\alpha_X$ in Eq.(\ref{eqn3104})
634: and
635: \beq
636: {\cal R}_\sigma = \left( \frac {m_X} {m_Y} \right)^{5/2} 
637: \afrac { m_\chi + m_Y } { m_\chi + m_X }\, ,
638: \eeq
639: which hold for the expectation values of these quantities as can be seen from
640: Eqs.(\ref{eqn3103}) and (\ref{msigma}).  Remarkably, the expectation values of
641: the factors in front of the expressions in square brackets are in fact the
642: same in Eqs.(\ref{eqn3105}) and (\ref{sigsig}), apart from the factor $1/|n|$
643: appearing in the former equation.  On the other hand, the expressions inside
644: these square parentheses do depend on $f_1(v)$, as well as on the involved
645: masses and form factors.
646: 
647: It is nevertheless instructive to study the behavior of the factor $\kappa$
648: for different target nuclei $X$ and $Y$. This largely determines what choice
649: of targets is optimal, i.e., minimizes the statistical errors of our estimators
650: of $m_\chi$. It is easy to see that the ratio $\kappa/m_\chi$, which will
651: appear in the relative error $\sigma(m_\chi) / m_\chi$, only depends on the
652: dimensionless ratios $m_X / m_\chi$ and $m_Y / m_\chi$. In
653: Fig.~1 %\ref{fig:contour} 
654: we therefore show contours of $\kappa/m_\chi$ in the
655: plane spanned by these two ratios, using the ordering $m_Y > m_X$. We note
656: first of all that $\kappa$ diverges as $m_Y \rightarrow m_X$. This is clear
657: from the fact that the denominator in the final expression in Eq.(\ref{kappa})
658: vanishes in this limit, while the numerator is finite. Physically this simply
659: means that performing two scattering experiments with the same target will not
660: allow one to determine $m_\chi$.
661: 
662: \begin{figure}[t!] \label{fig:contour}
663: \begin{center}
664: \vspace*{-1cm}
665: \rotatebox{-90}{\includegraphics[width=13.5cm]{fcont.ps}}
666: \end{center}
667: \vspace*{-1cm}
668: \caption{The five thick lines show contours of constant $\kappa/m_\chi$ in the
669:   plane spanned by $m_X / m_\chi$ and $m_Y / m_\chi$, where we have taken $m_Y
670:   > m_X$ without loss of generality. Here $\kappa$, has been defined in
671:   Eq.(\ref{kappa}); recall that the final relative error on $m_\chi$ is
672:   directly proportional to $\kappa / m_\chi$. The lower thin line indicates
673:   the end of the physical region, $m_Y = m_X$, whereas the upper thin line
674:   shows $m_Y = (76/28) m_X$, corresponding to Silicon and Germanium targets.}
675: \end{figure}
676: 
677: Slightly less trivially, we also see that $\kappa/m_\chi$, and hence the
678: relative error on $m_\chi$, becomes very large if both target nuclei are
679: either much heavier or both much lighter than the WIMP. This can be understood
680: from the fact that $m_\chi$ only enters via the reduced mass defined in
681: Eq.(\ref{eqn2103}).\footnote{Eq.(\ref{rQmin}) seems to have additional
682:   $m_\chi$ dependence. However, this comes from the factor $n_\chi = \rho_0 /
683:   m_\chi$, which drops out when the ratio of two targets is considered.} If
684: $m_\chi \gg \mN, \ m_{\rm r,N} \rightarrow \mN$ becomes completely independent of
685: the WIMP mass. In the opposite extreme, $m_\chi \ll \mN,\ m_{\rm r,N} \rightarrow
686: m_\chi$ becomes independent of the mass of the target nucleus, i.e., one is
687: effectively back in the situation where both experiments are performed with
688: the same target.
689: 
690: These considerations favor chosing the two targets to be as different as
691: possible. However, there are limits to this. On the one hand, taking a very
692: light target nucleus will lead to a low event rate for this experiment, and
693: hence very large statistical errors. Since the errors of both experiments are
694: added in quadrature inside the expressions in square parenthesis in
695: Eqs.(\ref{eqn3105}) and (\ref{sigsig}), this would lead to a large overall
696: error on $m_\chi$. In fact, if the total event number is held fixed, the final
697: error on $m_\chi$ will be minimal if both experiments contain approximately
698: the same number of events.
699: 
700: On the other hand, taking a very heavy target nucleus also leads to problems.
701: Heavy nuclei are large, which means they have quite soft form factors. For
702: example, the Woods--Saxon prediction \cite{FQb} for the form factor for
703: ${}^{136}$Xe, which is the target in some existing experiments \cite{pdg}, has
704: a zero at $Q \simeq 95$ keV. For our default choice $v_{\rm esc} = 700$ km/s,
705: this is below $Q_{\rm max,kin}$ of Eq.(\ref{qmaxkin}) for all $m_\chi > 45$
706: GeV. This is a serious problem, since the form factor, and hence the event
707: rate, remain (very) small even beyond this zero. Experiments with ${}^{136}$Xe
708: therefore effectively impose a cut $Q_{\rm max} \leq 95$ GeV.\footnote{Note
709:   also that while the integrals $I_n$ remain finite where the form factor
710:   vanishes, the estimates for the errors will diverge there, due to the factor
711:   $1/F^4$ appearing in Eq.(\ref{eqn2211}).}
712: 
713: From these considerations it seems that chosing $X =$ Si, $Y =$ Ge might be a
714: good compromise. This corresponds to points on the upper, thin solid straight
715: line in Fig.~1. %\ref{fig:contour}. 
716: We see that in this case $\kappa / m_\chi \geq
717: 4$ for all values of $m_\chi$. Not surprisingly, $\kappa/m_\chi$ reaches its
718: minimum when $m_\chi$ lies in between the masses of the two target nuclei, i.e
719: for the case at hand, for $m_\chi \simeq 50$ GeV. Note also that
720: $\kappa/m_\chi$ is unfortunately always well above unity. One will thus have
721: to get fairly accurate estimates of the relevant integrals $I_i$ if one wishes
722: to determine $m_\chi$ to better than a factor of two.
723: 
724: Before discussing the statistical uncertainty in more detail, we proceed to
725: include non--trivial upper and lower cuts on the recoil energy in our analysis.
726: 
727: \subsection{Incorporating cuts on the recoil energy}
728: 
729: As noted earlier, real experiments will probably have to impose both lower and
730: upper limits on the recoil energy, partly for instrumental reasons, and partly
731: to suppress backgrounds. This led us to introduce generalized moments of $f_1$
732: in Eq.(\ref{momnew}). These moments determined by two different detectors will
733: still be the same, if either the integration limits are identical, \mbox{$v_{i,X} =
734: v_{i,Y}$}, $i = 1,~2$, or the cuts are so loose that the contribution from WIMPs
735: with $v < v_1$ or $v > v_2$ is negligible; a combination of these two
736: possibilities can also occur, e.g., small but not necessarily equal values of
737: $v_{1,X}$ and $v_{1,Y}$, and $v_{2,X} = v_{2,Y}$ significantly below $v_{\rm
738:   esc}$. 
739: 
740: The crucial observation that forms the basis for our analysis of $m_\chi$ as
741: estimated from the generalized moments is that in fact {\em all three} terms
742: in the final expression in Eq.(\ref{momnew}) will agree {\em separately}
743: between two targets, as long as both integration limits coincide. This can be
744: shown by replacing the $r(Q_i)$ in Eq.(\ref{momnew}) via Eq.(\ref{eqn2101})
745: and using Eq.(\ref{Qi}). In principle we could therefore simply equate the
746: last (integral) terms between the two targets. This would lead to expressions
747: very similar to those derived in the previous subsection, the only difference
748: being that all integrals or sums over recoil energies would run over limited
749: ranges. 
750: 
751: The problem with this approach is that the velocities $v_i$ appearing in
752: Eq.(\ref{momnew}) are not directly observable. The recoil energies are.
753: However, the values of $Q_{\rm min}$ and $Q_{\rm max}$ would need to be {\em
754:   different} for the two targets, if they are to correspond to the same values
755: of $v_1$ and $v_2$. Worse, Eq.(\ref{Qi}) shows that the ratios $Q_{{\rm
756:     min},X} / Q_{{\rm min},Y}$ and $Q_{{\rm max},X} / Q_{{\rm max},Y}$ that
757: one has to chose in order to achieve $v_{i,X} = v_{i,Y}$ depend on
758: $m_\chi$. We are thus faced with the situation that, in the presence of
759: significant cuts on the recoil energy, the quantity we wish to determine is
760: needed as an input at an early stage of the analysis!
761: 
762: This statement is, strictly speaking, true; we see no way around this, if
763: significant cuts on the recoil energy are indeed needed. However, we can
764: alleviate the situation. To begin with, we noticed that the systematic error
765: one introduces by simply setting $Q_{{\rm min},X} = Q_{{\rm min},Y}$ is
766: reduced greatly if one keeps the sum of the first and third terms in
767: Eq.(\ref{momnew}). Moreover, we will show that, to some extent at least, one
768: can do the matching of two $Q_{\rm max}$ values for the two targets directly
769: from the data. In principle we could reduce the systematic error associated
770: with the choice of $Q_{\rm max}$ even more by also including the second term
771: in Eq.(\ref{momnew}). However, with limited statistics the error on $r(Q_{\rm
772:   max})$ will always be very large, so that keeping this term will not help
773: significantly.
774: 
775: Our determination of $m_\chi$ from generalized moments of $f_1$ is thus based
776: on equating the sum of the first and third terms in Eq.(\ref{momnew}) for two
777: different targets. The resulting expression for $m_\chi$ can still be cast in
778: the form of Eq.(\ref{eqn3103}), but ${\cal R}_n$ is now given by
779: %
780: \beq \label{eqn3201}
781: \calRn(Q_{\rm min}) = \bfrac{2 Q_{{\rm min},X}^{(n+1)/2} 
782: r_X(Q_{{\rm min},X})/F_X^2(Q_{{\rm min},X}) + (n+1) \InX}
783: {2 Q_{{\rm min},X}^{1/2} r_X(Q_{{\rm min},X})/ F_X^2(Q_{{\rm min},X}) + \IzX}^{1/n}
784:  (X \lto Y)^{-1}\, .
785: \eeq
786: %
787: Here $n \ne 0$
788: and $r_{(X,Y)}(Q_{{\rm min},(X,Y)})$ refer to the counting rate for detectors
789: $X$ and $Y$ at the respective lowest recoil energies included in the analysis.
790: They can be estimated from Eq.(\ref{eqn2209}); of course, the values of $r_1$,
791: $k_1$, $Q_{s,1}$, and $b_1$ will in general differ for the two targets. Note
792: that the integrals $I_n, \, I_0$ are now to be evaluated with $Q_{\rm min}$ as
793: lower and $Q_{\rm max}$ as upper limits, as in Eq.(\ref{eqn2210}).
794: 
795: \begin{figure}[b!] \label{fig:mrec_qmin}
796: \begin{center}
797: \vspace*{-1cm}
798: \rotatebox{-90}{\includegraphics[width=13.5cm]{mrec_qmin.ps}}
799: \end{center}
800: \vspace*{-1cm}
801: \caption{Ratio of reconstructed to input WIMP mass for imperfect $Q_{\rm min}$
802:   matching, for two different values of $Q_{\rm min}$. We have taken Si and Ge
803:   targets and assumed infinite statistics and (effectively) no upper cut on
804:   the recoil energy. The solid (dashed) curves have been computed from
805:   Eqs.(\ref{eqn3103}) and (\ref{eqn3201}) with $n=1$ including (neglecting)
806:   all terms $\propto r(Q_{\rm min})$. See the text for further details.}
807: \end{figure}
808: 
809: Before discussing the statistical uncertainty of this estimate of the WIMP
810: mass, we wish to demonstrate that keeping the terms $\propto r(Q_{\rm min})$
811: in Eq.(\ref{eqn3201}) indeed reduces the systematic error.  This is
812: demonstrated in Fig.~2, %\ref{fig:mrec_qmin}, 
813: which shows the ratio of the
814: reconstructed to input WIMP mass for experiments with ``infinite'' statistics.
815: Here we have chosen $X = {}^{28}$Si and $Y = {}^{76}$Ge, and we have set
816: $Q_{\rm max} = \infty$ in order to isolate the systematic effect due to
817: imperfect matching of the $Q_{\rm min}$ values in this figure.\footnote{More
818:   exactly, for $m_\chi > 100$ GeV we have set $Q_{\rm max, \, Ge} = 250$ keV,
819:   in order to avoid complications due to the zero of the form factor of
820:   ${}^{76}$Ge which occurs at $Q \simeq 270$ keV. We have then matched $Q_{\rm
821:     max, \, Si}$, so that the only systematic deviation of the reconstructed
822:   WIMP mass is indeed due to imperfect $Q_{\rm min}$ matching.} Here the black
823: (red) curves have been obtained with $Q_{\rm min, \, Ge} = Q_{\rm min, \, Si}
824: = 10 \ (3)$ keV. The solid lines include the terms $\propto r(Q_{\rm min})$,
825: while the dashed ones do not. Clearly including these terms is beneficial: for
826: $m_\chi > 20$ GeV, the systematic shift with these terms included and $Q_{\rm
827:   min} = 10$ keV is smaller than that without those terms with the much
828: smaller $Q_{\rm min} = 3$ keV. We also see that the systematic effect becomes
829: large at small $m_\chi$. This is not surprising, since a small $m_\chi$
830: implies a small $Q_{\rm max,kin}$, so that the lower cut on $Q$ removes a
831: correspondingly larger fraction of the total spectrum. The systematic error
832: vanishes for $m_\chi = \sqrt{m_X m_Y} = 43$ GeV, since for this value of
833: $m_\chi$ we have $\alpha_X = \alpha_Y$, i.e., $v_{1,X} = v_{1,Y}$ if $Q_{{\rm
834:     min},X} = Q_{{\rm min}, Y}$. For larger $m_\chi$ the systematic effect
835: changes sign, i.e., one now underestimates the true value of $m_\chi$.
836: However, if the terms $\propto r(Q_{\rm min})$ are included, the shift remains
837: relatively small even for $Q_{\rm min} = 10$ keV. If values $Q_{\rm min} \lsim
838: 3$ keV can be achieved, which should be possible \cite{low_th}, the systematic
839: effect due to imperfect $Q_{\rm min}$ matching will be a concern only for very
840: light WIMPs, $m_\chi < 20$ GeV.
841: 
842: Using Eq.(\ref{eqn3201}) in Eq.(\ref{eqn3103}) also leads to a lengthier
843: expression for the statistical error on our estimate of $m_\chi$. The first
844: equation in (\ref{eqn3105}) still holds, but the $r(Q_{\rm min})$ terms
845: give additional contributions to the final expression:
846: %
847: \beqn \label{sigmom}
848: \left. \sigma(m_\chi)\right|_{\Expv{v^n}} \= 
849: \frac{\sqrt{\mX/\mY} \vbrac{\mX-\mY}} {\abrac{\calRn -
850:      \sqrt{\mX/\mY}}^2}
851: \non\\
852: \conti ~~~~~~~~~~~~ \times %12
853:  \left[ \sum_{i,j=1}^3 \afrac {\partial {\cal R}_n}
854: {\partial c_{i,X}} \afrac {\partial {\cal R}_n} {\partial c_{j,X} } {\rm
855:   cov}(c_{i,X}c_{j,X}) + (X \lto Y) \right]^{1/2}\, .
856: \eeqn
857: %
858: Here we have introduced a short--hand notation for the six quantities on which
859: the estimate of $m_\chi$ depends:
860: %
861: \beq \label{ci}
862: c_{1,X} = I_{n,X}\, ; ~~~~~~~~~~~~ %12
863: c_{2,X} = I_{0,X}\, ; ~~~~~~~~~~~~ %12
864: c_{3,X} = r_X(Q_{{\rm min},X})\, ,
865: \eeq
866: %
867: and similarly for the $c_{i,Y}$. Estimators for ${\rm cov}(c_i, c_j)$ have
868: been given in Eqs.(\ref{eqn2211}) and (\ref{eqn2216}). Explicit expressions
869: for the derivatives of ${\cal R}_n$ with respect to these six quantities are
870: collected in the Appendix. Note that a factor ${\cal R}_n$ can be factored
871: out of all these derivatives. With this factor moved out of the square
872: brackets, the prefactor in Eq.(\ref{sigmom}) is identical to that in
873: Eq.(\ref{eqn3105}), i.e., its expectation value will again be given by $\kappa$
874: of Eq.(\ref{kappa}). Of course, Eq.(\ref{sigmom}) reduces to
875: Eq.(\ref{eqn3105}) in the limit $Q_{{\rm min},X}\, , Q_{{\rm min},Y}
876: \rightarrow 0$. Note finally that Eq.(\ref{sigmom}) also holds for $n = -1$,
877: if the derivatives with respect to $c_{1,(X,Y)}$ are neglected, since in this
878: case $\InX$ and $\InY$ do not contribute to ${\cal R}_n$ given in
879: Eq.(\ref{eqn3201}).
880: 
881: Finite lower energy cuts $Q_{{\rm min},(X,Y)}$ can also easily be incorporated
882: in the quantity ${\cal R}_\sigma$ appearing in Eq.(\ref{msigma}):
883: %
884: \beq \label{Rsigma1}
885: R_\sigma = \frac {{\cal E}_Y} {{\cal E}_X}
886: \bfrac{2Q_{{\rm min},X}^{1/2} r_X(Q_{{\rm min},X}) / F_X^2(Q_{{\rm min},X}) + \IzX}
887: {2 Q_{{\rm min},Y}^{1/2} r_Y(Q_{{\rm min},Y}) / F_Y^2(Q_{{\rm min},Y}) + \IzY} 
888: \,.
889: \eeq
890: %
891: Correspondingly, Eq.(\ref{sigsig}) changes to
892: %
893: \beqn \label{sigsig1}
894: \left. \sigma(m_\chi) \right|_\sigma \=
895: \frac{\abrac{\mX/\mY}^{5/2} \vbrac{\mX-\mY}}
896:      {\bbrac{\calR_{\sigma}-\abrac{\mX/\mY}^{5/2}}^2}
897: \non\\
898: \conti ~~~~~~~~~~~~ \times %12
899:  \left[ \sum_{i,j=2}^3 \afrac
900:   {\partial {\cal R}_\sigma} {\partial c_{i,X}}
901:   \afrac {\partial {\cal R}_\sigma}  {\partial c_{j,X} } 
902: {\rm cov}(c_{i,X}c_{j,X}) + (X \lto Y) \right]^{1/2}\, ,
903: \eeqn
904: %
905: where we have again used the short--hand notation of Eq.(\ref{ci}); note that
906: $c_{1(X,Y)}$ does not appear here. Expressions for the derivatives of ${\cal
907:   R}_\sigma$ are also given in the Appendix.
908: 
909: \subsection{Combined fit}
910: 
911: In the next step we wish to combine our estimators (\ref{eqn3103}) for
912: different $n$ with each other, and with the estimator (\ref{msigma}). This
913: could be done via an overall covariance matrix describing the errors of these
914: estimators and their correlations. The diagonal entries of this covariance
915: matrix are given by Eqs.(\ref{sigmom}) and (\ref{sigsig1}); the off--diagonal
916: entries can be computed analogously. This would yield the overall best--fit
917: value of $m_\chi$ as well as its Gaussian error.
918: 
919: Here we pursue a slightly different procedure, based on a $\chi^2$ fit. This
920: will yield the same best--fit value of $m_\chi$, which we denote by $m_{\chi,
921:   {\rm rec}}$, but it has two advantages. First, $\chi^2(m_{\chi,{\rm rec}})$
922: can be used as a measure of the quality of the fit, which in turn can be used
923: to match the $Q_{\rm max}$ values of the two experiments at least
924: approximately. Secondly, it allows to determine asymmetric error intervals.
925: Fig.~1 %\ref{fig:contour} 
926: implies that the errors should indeed be asymmetric.
927: For example, if the true $m_\chi$ is (much) larger than the masses of both
928: target nuclei, the experimental upper bound one can derive will be quite large
929: or even infinite, but one should still get a meaningful lower bound; the
930: opposite is true if $m_\chi$ lies well below the mass of both target nuclei.
931: 
932: We begin by defining fit functions
933: %
934: \cheqna
935: \beq \label{fia}
936: f_{i,X} = \alpha_X^i \bfrac  {2 Q_{{\rm min},X}^{(i+1)/2} 
937: r_X(Q_{{\rm min},X}) /
938:   F_X^2(Q_{{\rm min},X}) + (i+1) I_{i,X}} {2
939:   Q_{{\rm min},X}^{1/2} r_X(Q_{{\rm min},X}) / F_X^2(Q_{{\rm min},X}) + 
940:   I_{0,X}}
941:  \afrac {300 \ {\rm km}} {\rm s}^{-i} \, ,
942: \eeq
943:  for $i=-1,~1,~2,~\dots,~n_{\rm max}$;
944:  and
945: \cheqnb
946: \beq \label{fib}
947: f_{n_{\rm max}+1,X} = {\cal E}_X \bfrac {A_X^2} 
948: {2 Q_{{\rm min},X}^{1/2} r_X(Q_{{\rm min},X}) / F_X^2(Q_{{\rm min},X}) + \IzX}
949: \afrac{\sqrt{\mX}}{\mchi+\mX}
950: \, ;
951: \eeq
952: \cheqn
953: %
954: we analogously define $n_{\rm max} + 2$ functions $f_{i,Y}$.  Here $n_{\rm
955:   max}$ determines the highest (generalized) moment of $f_1$ that is included
956: in the fit. The $f_i$ are normalized such that they are dimensionless and very
957: roughly of order unity; this alleviates numerical problems associated with the
958: inversion of their covariance matrix. The first $n_ {\rm max}+1$ functions
959: $f_i$ are basically our estimators (\ref{eqn2208}) of the generalized moments
960: defined in Eq.(\ref{momnew}) with the term $\propto r(Q_{\rm max})$ omitted;
961: as discussed earlier, the error on this quantity is likely to be so large that
962: including this term will not be helpful. The last function is essentially the
963: ratio appearing in Eq.(\ref{one}). It is important to note that $m_\chi$ in
964: Eqs.(\ref{fia}) and (\ref{fib}) is a fit parameter, not the true (input) value of the WIMP
965: mass. Recall also that our estimator (\ref{eqn2210}) for the integrals $I_n$
966: appearing in Eqs.(\ref{fia}) and (\ref{fib}) is independent of $m_\chi$.
967: Hence the first $n_{\rm max} + 1$ fit functions depend on $m_\chi$
968: only through the overall factor
969: $\alpha^i$.
970: 
971: The $f_i$ allow us to introduce a $\chi^2$ function:
972: %
973: \beq \label{chisq}
974: \chi^2 = \sum_{i,j} \left( f_{i,X} - f_{i,Y} \right) {\cal C}^{-1}_{ij}
975: \left( f_{j,X} - f_{j,Y} \right) \, .
976: \eeq
977: %
978: Here ${\cal C}$ is the total covariance matrix. Since the $X$ and $Y$
979: quantities are statistically completely independent, ${\cal C}$ can be written
980: as a sum of two terms:
981: %
982: \beq \label{C}
983: {\cal C}_{ij} = {\rm cov}\left( f_{i,X}, f_{j,X} \right) 
984: + {\rm cov}\left( f_{i,Y}, f_{j,Y} \right)\,.
985: \eeq
986: %
987: The entries of this matrix involving only the moments of the WIMP velocity
988: distribution can be read off Eq.(82) of Ref.~\cite{DMDD}, with an obvious
989: modification due to the normalization factor in Eq.(\ref{fia}). Since
990: the last $f_i$ can be computed from the same basic quantities, i.e., the
991: counting rates at $Q_{\rm min}$ and the integrals $I_0$, the entries of the
992: covariance matrix involving this last fit function can also be computed
993: straightforwardly, using Eqs.(\ref{eqn2211})--(\ref{eqn2216}). Of course,
994: Eq.(\ref{chisq}) can also be used to compute asymmetric error intervals from a
995: single moment, by restricting the sum to a single term.
996: 
997: \section{Numerical results}
998: 
999: We are now ready to present some numerical results for the reconstructed WIMP mass.
1000: These results are based on Monte Carlo simulations of direct detection
1001: experiments. We assume that the scattering cross section is dominated by
1002: spin--independent interactions, and use the Woods--Saxon form for the elastic
1003: form factors $F(Q)$ \cite{FQb}. We describe the WIMP velocity distribution by
1004: a sum of a shifted Gaussian contribution \cite{susydm} and a ``late infall''
1005: component \cite{sikivie}; as in Ref.~\cite{DMDD}, for simplicity we model the
1006: latter as a $\delta-$function, keeping the normalization $N_{\rm l.i.}$ of
1007: this component as free parameter:
1008: %
1009: \beq \label{f1}
1010: f_1(v) = \frac {\left( 1 - N_{\rm l.i.} \right)} {\sqrt{\pi}} \afrac {v} {v_e
1011:   v_0} \left[ {\rm e}^{-(v-v_e)^2/v_0^2} - {\rm e}^{-(v+v_e)^2/v_0^2} \right]
1012: + N_{\rm l.i.} \delta(v - v_{\rm esc}) \,.
1013: \eeq
1014: %
1015: We take $v_0 = 220$ km/s, $v_e = 1.05 \~ v_0$\footnote{Strictly speaking, $v_e$
1016:   should oscillate around $1.05 \~ v_0$ with a period of one year \cite{susydm};
1017:   we ignore this time dependence here.}, and $v_{\rm esc} = 700$ km/s. 
1018: 
1019: \begin{figure}[t!] \label{fig:nocut}
1020: \begin{center}
1021: \vspace*{-.25cm}
1022: \rotatebox{-90}{\includegraphics[width=11.5cm]{m50_sige_nocut.ps}}
1023: \end{center}
1024: \vspace*{-1cm}
1025: \caption{Median values of ``$1\, \sigma$'' upper and lower bounds on the
1026:   reconstructed WIMP mass in $2 \times 5,000$ simulated experiments with Si
1027:   and Ge targets, as a function of the true value of $m_\chi$. We generated on
1028:   average 50 events per experiment. The short--dashed (black), dotted (green)
1029:   and long--dashed (violet) curves show the upper and lower bounds on $m_\chi$
1030:   as determined from moments with $n=-1, \, 1$ and 2, respectively; the
1031:   dash-dotted (blue) curves labeled $\sigma$ show the corresponding limits derived
1032:   from our assumption of equal cross sections for scattering on protons and
1033:   neutrons. Finally, the solid (red) curves show the upper and lower bound as
1034:   well as the median reconstructed $m_\chi$ for the global fit based on
1035:   minimizing $\chi^2$ of Eq.(\ref{chisq}) with $n_{\rm max} = 2$. These
1036:   results are for our standard halo, i.e., no late infall component, without
1037:   any cuts on the recoil energy.}
1038: \end{figure}
1039: 
1040: In Fig.~3 %\ref{fig:nocut} 
1041: we show upper and lower bounds on the reconstructed
1042: WIMP mass, calculated from the requirement that $\chi^2$ exceeds its minimum
1043: by 1, for the idealized scenario where no cuts on $Q$ have been applied. We
1044: took our default set of parameters, i.e., vanishing late infall component, and
1045: target nuclei $X = {}^{28}$Si, $Y = {}^{76}$Ge. This figure is based on
1046: simulating \mbox{2 $\times$ 5,000} experiments, where each experiment contains an
1047: expected 50 events; the actual number of events is Poisson--distributed around
1048: this expectation value.  As mentioned earlier, taking equal numbers of events
1049: in both experiments minimizes the statistical error for fixed total number of
1050: events. As discussed in Ref.~\cite{DMDD}, the error on the (high) moments is
1051: not quite Gaussian; the deviation becomes larger for smaller samples. The
1052: reason is that the high moments receive large contributions from the region of
1053: high $Q$, where on average very few events will lie; even the region where on
1054: average only a fraction of an event lies can contribute significantly. As a
1055: result, most simulated experiments will underestimate these moments, while a
1056: few (rare) experiments will overestimate them. In order to alleviate this
1057: problem, we only include moments up to $n_{\rm max} = 2$ in our fit. Moreover,
1058: we always show median, rather than mean, values for the (bounds on the)
1059: reconstructed WIMP mass.
1060: 
1061: We see that the minus--first moment by itself leads to very poor bounds on
1062: $m_\chi$. This is not surprising, since its error is dominated by the error on
1063: the counting rate at $Q_{\rm min}$, which is determined only from the events
1064: in the first bin. The higher moments lead to increasingly tighter bounds.
1065: However, the higher moments are very strongly correlated. Also, the systematic
1066: effect due to the limited event samples discussed in the previous paragraph
1067: becomes larger for larger $n$; for example, for $n=2$ and a true $m_\chi = 1$
1068: TeV, the median upper end of the (nominal) $1 \, \sigma$ range lies well below
1069: 1 TeV.  The lower bound on $m_\chi$ derived from the assumption of equal
1070: scattering cross sections on protons and neutrons is similar to that derived
1071: from the first moment, but the corresponding upper bound is significantly
1072: worse.  Nevertheless, this estimator of $m_\chi$ helps in narrowing down the
1073: error of the total fit, described by the upper and lower solid (red) curves.
1074: As expected from Fig.~1 %\ref{fig:contour} 
1075: the relative error on $m_\chi$ is
1076: minimal for $m_\chi = \sqrt{\mX \mY}$, although the increase towards smaller
1077: $m_\chi$ is less than expected from the behavior of the kinematical factor
1078: $\kappa$ alone.
1079: 
1080: Unfortunately we also see that the median reconstructed $m_\chi$ starts to
1081: deviate from the input value if $m_\chi \gsim 80$ GeV. This is a direct
1082: consequence of the fact that the median value of the estimators of the higher
1083: moments is too small, as discussed above. For very large $m_\chi$ the median
1084: reconstructed WIMP mass even becomes independent of its true value; this is
1085: true also for the upper end of the error band. This systematic shift presents
1086: another argument in favor of imposing an upper cut $Q_{\rm max}$ on $Q$,
1087: chosen sufficiently low that an average experiment will still have a few
1088: events not too far below $Q_{\rm max}$.
1089: 
1090: \begin{figure}[p!] \label{fig:5050}
1091: \begin{center}
1092: \vspace*{-1cm}
1093: \rotatebox{-90}{\includegraphics[width=11.cm]{m50_sige_5050.ps}}
1094: \end{center}
1095: \vspace*{-1cm}
1096: \caption{As in Fig.~3, %\ref{fig:nocut}, 
1097:   except that we have imposed the cut
1098:   $Q_{\rm max} = 50$ keV in both experiments. Note that the average of 50
1099:   events per experiment refers to the entire $Q$ range, i.e., the number of
1100:   events after cuts is smaller if $Q_{\rm max,kin} > Q_{\rm max}$.}
1101: \end{figure}
1102: \begin{figure}[p!] \label{fig:opt}
1103: \begin{center}
1104: \vspace*{-.3cm}
1105: \rotatebox{-90}{\includegraphics[width=11.cm]{m50_sige_opt.ps}}
1106: \end{center}
1107: \vspace*{-1cm}
1108: \caption{As in Fig.~4, %\ref{fig:5050}, 
1109:   except that we have imposed the cut
1110:   $Q_{\rm max} = 50$ keV for the Ge target. The value of $Q_{\rm max}$ for the
1111:   Si target has been chosen such that it corresponds to the same WIMP
1112:   velocity.}
1113: \end{figure}
1114: 
1115: For the case at hand, with rather small event samples, this would imply
1116: $Q_{\rm max} \simeq 50$ keV. Unfortunately Fig.~4 %\ref{fig:5050} 
1117: shows that
1118: simply imposing the same cut on $Q_{\rm max}$ in both targets makes the
1119: situation significantly {\em worse}. Now the median reconstructed $m_\chi$
1120: starts to undershoot the input value already at $m_\chi \simeq 60$ GeV, and the
1121: asymptotic reconstructed $m_\chi$ for large input WIMP mass lies well below
1122: 100 GeV. Clearly we need to choose different $Q_{\rm max}$ values for the two
1123: experiments if we want to get reliable results also for larger WIMP masses.
1124: 
1125: Fig.~5 %\ref{fig:opt} 
1126: indicates that this should be possible, at least in
1127: principle. Here we have again applied a fixed upper cut $Q_{\rm max} = 50$ keV
1128: for the Ge experiment, but matched the cut on $Q_{\rm max}$ for the Si
1129: experiment such that it corresponds to the same WIMP velocity:
1130: %
1131: \beq \label{match}
1132: Q_{\rm max,\,Si} = \left( \frac {\alpha_{\rm Ge}} {\alpha_{\rm Si}} \right)^2
1133: Q_{\rm max,\,Ge}\,,
1134: \eeq
1135: %
1136: where $\alpha$ has been given in Eq.(\ref{eqn2105}). We see that now the
1137: median reconstructed $m_\chi$ indeed tracks the input value even for very
1138: large WIMP masses. It should be noted that these results still only use 50
1139: events on average per experiment {\em before cuts}. For example, for $m_\chi =
1140: 500$ GeV Eq.(\ref{match}) with $Q_{\rm max,\,Ge} = 50$ keV implies $Q_{\rm
1141:   max,\,Si} = 21.7$ keV, meaning that the expected number of events in the Si
1142: experiments is only about 21. This explains why the error bands at large
1143: values of $m_\chi$ are significantly wider here than in Fig.~3 %\ref{fig:nocut}
1144: where no cuts on the recoil energy have been imposed. Of course, the
1145: systematic deviation of the reconstructed $m_\chi$ from its true value makes
1146: the error bands in Fig.~3 %\ref{fig:nocut} 
1147: largely meaningless for $m_\chi \gsim 100$
1148: GeV.
1149: 
1150: \begin{figure}[b!] \label{fig:inout}
1151: \begin{center}
1152: \rotatebox{-90}{\includegraphics[width=9cm]{minout.ps}} \hspace*{-3.7cm}
1153: \rotatebox{-90}{\includegraphics[width=9cm]{chisqinout.ps} }
1154: \end{center}
1155: \vspace*{-1cm}
1156: \caption{Sensitivity of the $\chi^2$ fit to the value of $m_\chi$ that is used
1157:   as input in the $Q_{\rm max}$ matching condition (\ref{match}), assuming a
1158:   true WIMP mass of 100 GeV, for our default choice of targets and $f_1(v)$.
1159:   The solid (black) lines are for $2 \times 50$ events and $Q_{\rm max} = 50$
1160:   keV, while the dashed (red) curves assume $2 \times 500$ events and $Q_{\rm
1161:     max} = 100$ keV; here $Q_{\rm max}$ stands for the bigger of the values
1162:   used for the two targets, the second one being fixed by Eq.(\ref{match}).
1163:   The curves terminate at $m_{\chi,{\rm in}} = 25$ GeV since for even smaller
1164:   WIMP masses, $Q_{\rm max,kin} < 50$ keV, making matching superfluous. The
1165:   left frame shows the median reconstructed WIMP mass from a $\chi^2$ fit with
1166:   $n_{\rm max} = 2$, and the right frame shows the corresponding mean value of
1167:   $\chi^2(m_{\chi,{\rm rec}}).$}
1168: \end{figure}
1169: 
1170: While the results in Fig.~5 %\ref{fig:opt} 
1171: look quite impressive, they suffer
1172: from the fact that optimal $Q_{\rm max}$ matching, as in Eq.(\ref{match}), is
1173: only possible if $m_\chi$ is already known. The left frame of
1174: Fig.~6 %\ref{fig:inout} 
1175: shows that inputting the wrong value of $m_\chi$ into
1176: Eq.(\ref{match}) will usually lead to a reconstructed WIMP mass in between
1177: this input value and the true value. Note that the median $m_{\chi,{\rm rec}}$
1178: as a function of $m_{\chi,{\rm in}}$ has a slope less than unity. This means
1179: that an iteration, where one starts with some input value of $m_\chi$ and uses
1180: the corresponding $m_{\chi,{\rm rec}}$ as new input value into
1181: Eq.(\ref{match}), will converge ``on average''. Unfortunately our Monte Carlo
1182: simulations show that for any one experiment, this procedure does not
1183: necessarily converge to a well--defined $m_{\chi,{\rm rec}}$; rather, one
1184: often ends up in an endless loop over several values of $m_{\chi,{\rm rec}}$.
1185: 
1186: An alternative is to try an algorithm for $Q_{\rm max}$ matching which is
1187: based on the minimum value of $\chi^2$ obtained in the fit; recall that this
1188: minimum defines the reconstructed WIMP mass. Unfortunately the right frame in
1189: Fig.~6 %\ref{fig:inout} 
1190: indicates that this may be difficult, at least with the
1191: initial small statistics expected. This figure shows that the mean value of
1192: $\chi^2(m_{\chi,{\rm rec}})$ is almost independent of the value of $m_\chi$
1193: input into Eq.(\ref{match}) if one has only $2 \times 50$ events in the
1194: sample. The situation looks considerably more promising with $2 \times 500$
1195: events, and larger allowed $Q_{\rm max}$: in this case $\chi^2(m_{\chi,{\rm
1196:     rec}})$ has a clear, if rather broad, minimum where $m_{\chi,{\rm in}}$
1197: equals the true WIMP mass, taken to be 100 GeV in this example. 
1198: \begin{figure}[b!] \label{fig:real}
1199: \begin{center}
1200: \vspace*{-1cm}
1201: \rotatebox{-90}{\includegraphics[width=13.5cm]{m50_sige_real.ps}}
1202: \end{center}
1203: \vspace*{-1cm}
1204: \caption{As in Fig.~4, %\ref{fig:5050}, 
1205:   except that we have imposed the fixed cut
1206:   $Q_{\rm max} = 50$ keV only on the Ge experiment and determined $Q_{\rm max,
1207:   Si}$ by minimizing $\chi^2(m_{\chi,{\rm rec}})$.}
1208: \end{figure}
1209: 
1210: Note that this minimum has mean $\chi^2(m_{\chi,{\rm rec}}) / {\rm d.o.f.}$
1211: well below unity; the same is true for the entire solid curve. Here the number
1212: of degrees of freedom is three, since we fit four quantities with one free
1213: parameter. This indicates that our numerical procedure on average
1214: over--estimates the true errors somewhat. In fact, following Ref.~\cite{DMDD},
1215: we add the ``error on the error'' to the diagonal entries of ${\rm
1216:   cov}(I_n,I_m)$, in order to tame non--Gaussian tails in the distribution of
1217: measured moments of $f_1$.
1218: 
1219: One possibility is to choose the $Q_{\rm max}$ values such that
1220: $\chi^2(m_{\chi,{\rm rec}})$ is minimal. Note that this implies a double
1221: minimization: for fixed values of $Q_{{\rm max}, X}$ and $Q_{{\rm max},Y}$,
1222: $m_{\chi, {\rm rec}}$ is the WIMP mass that minimizes $\chi^2(m_{\chi})$. In
1223: an outer loop, $Q_{\rm max}$ is varied. We found that varying both $Q_{\rm
1224:   max}$ values leads to quite misleading results, especially for larger (true)
1225: WIMP masses and/or limited statistics. On the other hand, Fig.~4 %\ref{fig:5050}
1226: shows that for small $m_\chi$, $Q_{\rm max}$ matching is not really necessary;
1227: if $m_{\chi} > \sqrt{ m_X m_Y}$, the matching condition (\ref{match}) implies
1228: that $Q_{\rm max}$ for the lighter target should be smaller than that for the
1229: heavier target.
1230: %
1231: \begin{figure}[b!] \label{fig:m50_100}
1232: \begin{center}
1233: \vspace*{-1cm}
1234: \rotatebox{-90}{\includegraphics[width=13.5cm]{m50_100_sige.ps}}
1235: \end{center}
1236: \vspace*{-1cm}
1237: \caption{Results for the median value of the reconstructed WIMP mass as well
1238:   as the ends of its error interval. All results are based on the combined
1239:   fit, using Eq.(\ref{chisq}) with $n_{\rm max} = 2$. We assume Si and Ge
1240:   targets with on average 50 events each before cuts, and took our default
1241:   ansatz for $f_1$. The dotted (green) lines show results for $Q_{\rm max, Si}
1242:   = Q_{\rm max, Ge} = 100$ keV, whereas the solid (black) lines have been
1243:   obtained using Eq.(\ref{match}) with the bigger of the two $Q_{\rm max}$
1244:   values fixed to 100 keV. Finally, the dashed (red) lines are for the case
1245:   that $Q_{\rm max, Ge} = 100$ keV, whereas $Q_{\rm max, Si}$ has been chosen
1246:   such that $\chi^2(m_{\chi,{\rm rec}})$ is minimal.
1247: }
1248: \end{figure}
1249: 
1250: In Fig.~7 %\ref{fig:real} 
1251: we have therefore minimized $\chi^2(m_{\chi,{\rm
1252:     rec}})$ {\em only} for the lighter target, here Silicon. Note that we
1253: always require $Q_{\rm max, Si} \leq Q_{\rm max, Ge}$ in this procedure; in
1254: Fig.~7 %\ref{fig:real} 
1255: the latter is fixed to 50 keV. We see that this leads to a
1256: systematic over--estimation of $m_\chi$ for small WIMP masses. For heavier
1257: WIMPs the results are somewhat better than for the case where both $Q_{\rm
1258:   max}$ values are simply taken to be equal, see Fig.~4. %\ref{fig:5050}. 
1259: However,
1260: this algorithm clearly still fails quite badly for $m_\chi \gsim 100$ GeV.
1261: This is partly due to the low maximal value of $Q_{\rm max}$ assumed here. The
1262: fact that the kinematic upper bound $Q_{\rm max,kin}$ of Eq.(\ref{qmaxkin})
1263: increases with $m_\chi$ indicates that the region of large $Q$ is more
1264: sensitive to the value of $m_\chi$ if the WIMP mass exceeds the mass of the
1265: heaviest target nucleus.
1266: 
1267: In Fig.~8 %\ref{fig:m50_100} 
1268: we have therefore increased the cut $Q_{\rm max}$ to
1269: 100 keV; of course, this cut only becomes effective once $Q_{\rm max, kin} >
1270: 100$ keV. Unlike in the previous figures, we here show results only for the
1271: final fit; the values of $m_\chi$ derived from single observables are no
1272: longer shown. This allows us to show results for three different choices of
1273: $Q_{\rm max, Si}$ and $Q_{\rm max, Ge}$.
1274: The dotted (green) curves show the median reconstructed WIMP mass and its
1275: ``$1\sigma$'' upper and lower bounds for the case where both $Q_{\rm max}$
1276: values have been fixed to 100 keV. Due to the higher $Q_{\rm max}$ chosen
1277: here, this works for considerably larger WIMP masses than in the corresponding
1278: Fig.~4, %\ref{fig:5050}, 
1279: where both $Q_{\rm max}$ values had been fixed to 50 keV.
1280: However, for large $m_\chi$ the median reconstructed WIMP mass is still only
1281: slightly above 100 GeV.
1282: The solid (black) lines show results for the case that perfect $Q_{\rm max}$
1283: matching has been applied, using Eq.(\ref{match}). Comparison with
1284: Fig.~5 %\ref{fig:opt} 
1285: shows that increasing $Q_{\rm max}$ actually slightly increases the width of
1286: the errors on the reconstructed WIMP mass. The reason is that $Q_{\rm max}$ is
1287: now so large that the median values of the estimators for the (generalized)
1288: moments of $f_1$ fall somewhat below the true values. Nevertheless the results
1289: for this optimal $Q_{\rm max}$ matching remain very encouraging.
1290: 
1291: Most importantly, our simple algorithm of fixing $Q_{\rm max, Ge}$, in this
1292: case to 100 keV, and determining $Q_{\rm max, Si}$ by minimizing
1293: $\chi^2(m_{\chi,{\rm rec}})$ now also seems to work reasonably well for WIMP
1294: masses up to $\sim 500$ GeV. For $m_\chi \lsim 100$ GeV the median WIMP mass
1295: determined in this way again over--estimates its true value by 15 to 20\%;
1296: however, the median value of the ``$1\sigma$'' lower bound lies below the true
1297: WIMP mass for all values of $m_\chi$. Similarly, the median ``$1\sigma$''
1298: upper bound now lies well above the true WIMP mass even for $m_\chi = 1$ TeV,
1299: in sharp contrast to the results shown in Fig~7. %\ref{fig:real}.
1300: 
1301: \begin{figure}[t!] \label{fig:arxe}
1302: \begin{center}
1303: \vspace*{-1cm}
1304: \rotatebox{-90}{\includegraphics[width=11.5cm]{m50_arxe_7575.ps}}
1305: \end{center}
1306: \vspace*{-1cm}
1307: \caption{As in Fig.~8, %\ref{fig:m50_100}, 
1308:   except for ${}^{40}$Ar and ${}^{136}$Xe
1309:   targets, and with both $Q_{\rm max}$ values restricted to not exceed 75 keV.
1310: }
1311: \end{figure}
1312: \begin{figure}[h!] \label{fig:li}
1313: \begin{center}
1314: \rotatebox{-90}{\includegraphics[width=11.5cm]{m50_100_sige_li.ps}}
1315: \end{center}
1316: \vspace*{-1cm}
1317: \caption{As in Fig.~8, %\ref{fig:m50_100}, 
1318:   except that we allowed a 25\% late
1319:   infall component in the local WIMP flux, by setting $N_{\rm l.i.} = 0.25$ in
1320:   Eq.(\ref{f1}).  }
1321: \end{figure}
1322: 
1323: We had argued earlier, based on the results of Fig.~1, %\ref{fig:contour}, 
1324: that Si
1325: and Ge is likely to be close to the optimal choice among the target nuclei
1326: currently being employed. In Fig.~9 %\ref{fig:arxe} 
1327: we back this up by showing
1328: results analogous to those of Fig.~8, %\ref{fig:m50_100}, 
1329: but with Argon and Xenon
1330: targets. Since the elastic form factor of ${}^{136}$Xe has a zero near 95 keV,
1331: we have lowered the upper bound on $Q_{\rm max}$ to 75 keV; larger values
1332: would start to probe the very sparsely populated region near the zero, leading
1333: to too small median values of the estimated moments, whereas smaller values
1334: would lead to even worse sensitivity to large WIMP masses. We see that, except
1335: for small WIMP masses, the results are clearly worse than those shown in
1336: Fig.~8. %\ref{fig:m50_100}.
1337: 
1338: 
1339: The starting point of our discussion was that we wanted to devise ways to
1340: estimate the WIMP mass which are independent of any assumptions on the WIMP
1341: velocity distribution $f_1$. In order to test this, we have simulated Si and
1342: Ge experiments, allowing 25\% of the local WIMP flux to come from a ``late
1343: infall'' component, i.e., we fixed $N_{\rm l.i.} = 0.25$ in Eq.(\ref{f1}).  The
1344: results are shown in Fig.~10. %\ref{fig:li}. 
1345: We see that the results are very
1346: similar to those with $N_{\rm l.i.}=0$ shown in Fig.~8 %\ref{fig:m50_100} 
1347: if
1348: optimal $Q_{\rm max}$ matching (\ref{match}) is used. On the other hand, the
1349: results for non--optimal choices of $Q_{\rm max}$ get somewhat worse. This is
1350: true both for the simple choice $Q_{\rm max,Si} = Q_{\rm max, Ge}$, where
1351: significant deviations set in for lower values of the WIMP mass, and for our
1352: algorithm of determining $Q_{\rm max, Si}$ by minimizing $\chi^2(m_{\chi,{\rm
1353:     rec}})$. In the latter case, the systematic difference between median
1354: reconstructed and true WIMP mass is larger than for $N_{\rm l.i.}=0$, both for
1355: small and for large $m_\chi$. The reason for this degradation is that
1356: introducing a large late--infall component, corresponding to WIMPs with
1357: velocity about three times larger than the mean velocity of the shifted
1358: Gaussian component, increases the number of events at large recoil energy.
1359: Hence correct $Q_{\rm max}$ matching becomes more important. Note, however,
1360: that the true $m_\chi$ always lies within the median limits of the
1361: ``$1\sigma$'' error interval estimated from our algorithmic $Q_{\rm max}$
1362: matching.
1363: 
1364: \begin{figure}[t!] \label{fig:500}
1365: \begin{center}
1366: \vspace*{-1cm}
1367: \rotatebox{-90}{\includegraphics[width=13.5cm]{m500_sige.ps}}
1368: \end{center}
1369: \vspace*{-1cm}
1370: \caption{As in Fig.~8, %\ref{fig:m50_100}, 
1371: except for $2 \times 500$ events before
1372: cuts. }
1373: \end{figure}
1374: 
1375: So far we have assumed that each experiment ``only'' has an exposure
1376: corresponding to 50 events before cuts. In Fig.~11 %\ref{fig:500} 
1377: we raise this
1378: number by a factor of 10. Not surprisingly, all error bars shrink by a factor
1379: $\gsim~3$ compared to the situation of Fig.~8. %\ref{fig:m50_100}. 
1380: The error
1381: interval also remains approximately symmetric out to much larger values of
1382: $m_\chi$. However, the larger number of events does not significantly change
1383: the median reconstructed $m_\chi$ if one simply fixes both $Q_{\rm max}$
1384: values to 100 keV. This is not surprising: for large $m_\chi$ this implies
1385: that $v_{2,{\rm Si}}$ and $v_{2,{\rm Ge}}$ in the definition (\ref{momnew}) of
1386: the generalized moments, and hence the moments themselves, are quite
1387: different. Hence the estimators of these moments will not agree if the true
1388: WIMP mass is used. On the other hand, our algorithm for fixing $Q_{\rm max,
1389:   Ge}$ now seems to perform very well over the entire range of WIMP masses
1390: shown. In particular, the median reconstructed WIMP mass now overshoots its
1391: true value by only a few percent for $m_\chi \lsim 100$ GeV, and remains close
1392: to the true value even at $m_\chi = 1$ TeV. Unfortunately we will see shortly
1393: that for large WIMP masses our algorithm still has some problems even in this
1394: case.
1395: 
1396: The problem lies in the distribution of the reconstructed WIMP masses in the
1397: simulated experiments. This distribution is supposed to be characterized by
1398: the error intervals shown in Figs.~3--5 %\ref{fig:nocut}--\ref{fig:opt} 
1399: and 7--11. %\ref{fig:real}--\ref{fig:500}. 
1400: In order to see how well this works, we
1401: introduce the quantity
1402: %
1403: \beq \label{delm}
1404: \renewcommand{\arraystretch}{2.5}
1405: \delta m = \left\{ \begin{array}{l c l}
1406: \displaystyle
1407: 1 + \frac {m_{\chi, {\rm lo1}} - m_\chi } 
1408: {m_{\chi, {\rm lo1}} - m_{\chi, {\rm lo2} } }\, , & ~~~~~~ &
1409: {\rm if} \ m_\chi \leq m_{\chi,{\rm lo1}}\, ; \\
1410: %
1411: \displaystyle
1412: \frac {m_{\chi, {\rm rec}} - m_\chi } 
1413: {m_{\chi, {\rm rec}} - m_{\chi, {\rm lo1} } }\, , & &
1414: {\rm if} \ m_{\chi,{\rm lo1}} < m_{\chi} < m_{\chi, {\rm rec}}\, ; \\ 
1415: %
1416: \displaystyle
1417: \frac {m_{\chi, {\rm rec}} - m_\chi } 
1418: {m_{\chi, {\rm hi1}} - m_{\chi, {\rm rec} } }\, , & &
1419: {\rm if} \ m_{\chi,{\rm rec}} < m_{\chi} < m_{\chi, {\rm hi1}}\, ; \\ 
1420: %
1421: \displaystyle
1422: \frac {m_{\chi, {\rm hi1}} - m_\chi } 
1423: {m_{\chi, {\rm hi2}} - m_{\chi, {\rm hi1} } } - 1 \, , & &
1424: {\rm if} \ m_\chi \geq m_{\chi,{\rm hi1}}\, .
1425: \end{array} \right.
1426: \eeq
1427: %
1428: \begin{figure}[b!] \label{fig:del50}
1429: \begin{center}
1430: \vspace*{-1cm}
1431: \rotatebox{-90}{\includegraphics[width=13.5cm]{del50_50_sige.ps}}
1432: \end{center}
1433: \vspace*{-1cm}
1434: \caption{Normalized distribution of the variable $\delta m$ defined in
1435:   Eq.(\ref{delm}) for 5,000 simulated experiments, for true WIMP mass
1436:  $m_\chi = 50$ GeV. The other parameters, are as in Fig.~8. %\ref{fig:m50_100}.
1437: The solid (black) histogram shows results for perfect $Q_{\rm max}$ matching
1438:  with $Q_{\rm max} \leq 100$ keV, based on Eq.(\ref{match}), whereas the
1439:  dotted (green) histogram is for fixed equal $Q_{\rm max}$ values of 100
1440:  keV. The dashed (red) histogram shows results when $Q_{\rm max, Si}$ is
1441:  determined by minimizing $\chi^2(m_{\chi,{\rm rec}})$.
1442:  }
1443: \end{figure}
1444: %
1445: Here $m_\chi$ is the true WIMP mass, $m_{\chi,{\rm rec}}$ its reconstructed
1446: value, $m_{\chi, {\rm lo1(2)}}$ is the ``$1\ (2)\, \sigma$'' lower bound
1447: satisfying $\chi^2(m_{\chi,{\rm lo(1,2)}}) = \chi^2(m_{\chi,{\rm rec}}) + 1\ 
1448: (4)$, and $m_{\chi, {\rm hi1(2)}}$ are the corresponding ``$1\ (2)\,\sigma$''
1449: upper bounds.  In the limit of purely Gaussian errors, where $\chi^2$ of
1450: Eq.(\ref{chisq}) is simply a parabola, $(\delta m)^2$ would itself be a
1451: $\chi^2$ variable, measuring the difference between the true and the
1452: reconstructed WIMP mass in units of the error of the reconstruction. However,
1453: we saw earlier that the error intervals are often quite asymmetric. Similarly,
1454: the distance between the ``$2\sigma$'' and ``$1\sigma$'' limits can be quite
1455: different from the distance between the ``$1\sigma$'' limit and the central
1456: value. The definition (\ref{delm}) takes these differences into account, and
1457: also keeps track of the sign of the deviation: if the reconstructed WIMP mass
1458: is larger (smaller) than the true one, $\delta m$ is positive (negative).
1459: Moreover, $|\delta m| \leq 1 \ (2)$ if and only if the true WIMP mass lies between the
1460: ``experimental'' $1 \ (2)\ \sigma$ limits.
1461: 
1462: In Fig.~12 %\ref{fig:del50} 
1463: we show the distribution of $\delta m$ calculated from
1464: 5,000 simulated experiments, assuming a rather small WIMP mass, $m_\chi = 50$
1465: GeV. The other parameters have been fixed as in Fig.~8. %\ref{fig:m50_100}. 
1466: In
1467: this case simply fixing both $Q_{\rm max}$ values to 100 keV still works fine,
1468: since the kinematic maximum values of $Q$ lie only slightly above 100 keV (at
1469: 122 keV for Si and 132 keV for Ge). The distributions for fixed $Q_{\rm max}$,
1470: or for optimal $Q_{\rm max}$ matching, look somewhat lopsided, since the error
1471: interval is already asymmetric, with $m_{\chi,{\rm hi1}} - m_{\chi,{\rm rec}}
1472: > m_{\chi,{\rm rec}} - m_{\chi,{\rm lo1}}$. As a result, negative values of
1473: $\delta m$ have a larger denominator than positive values, hence the
1474: distribution is narrower for $\delta m < 0$. These distributions also indicate
1475: that our errors are indeed over--estimated, since nearly 90\% of the simulated
1476: experiments have $|\delta m| \leq 1$; we remind the reader that a usual $1
1477: \sigma$ interval should only contain some 68\% of the experiments. 
1478: 
1479: We saw in Fig.~8 %\ref{fig:m50_100} 
1480: that our algorithm for determining $Q_{\rm max,
1481:   Si}$ tends to overestimate the WIMP mass if the latter is small. This is
1482: reflected by the dashed (red) histogram in Fig.~12, %\ref{fig:del50}, 
1483: which has
1484: significantly more entries at positive values than at negative values.
1485: Moreover, this histogram is rather flat between $\delta m = 1 $ and $2$. Since
1486: our algorithm is based on a double minimization of $\chi^2$ defined in
1487: Eq.(\ref{chisq}), it is not very surprising that the resulting final
1488: $\chi^2(m_{\chi,{\rm rec}})$ values are distributed quite differently from
1489: what one finds after a single minimization step. Nevertheless it is
1490:   reassuring that some 70\% of the simulated experiments lead to $|\delta m|
1491:   \leq 1$.
1492: 
1493: \begin{figure}[b!] \label{fig:del200}
1494: \begin{center}
1495: \rotatebox{-90}{\includegraphics[width=9cm]{del50_200_sige.ps}}
1496: % \hspace*{-3.7cm}
1497: \rotatebox{-90}{\includegraphics[width=9cm]{del500_200_sige.ps} }
1498: \end{center}
1499: \vspace*{-1cm}
1500: \caption{Distribution of $\delta m$ defined in Eq.(\ref{delm}) calculated from
1501: 5,000 simulated experiments. Parameters and notations are as in
1502: Fig.~12, %\ref{fig:del50}, 
1503: except that we have increased the WIMP mass to 200
1504: GeV. In the right frame we have in addition increased the average number of
1505: events (before cuts) in each experiment from 50 to 500. Note that the bins at
1506: $\delta m = \pm 5$ are overflow bins, i.e., they also contain all experiments
1507: with $|\delta m| > 5$.}
1508: \end{figure}
1509: 
1510: Unfortunately Fig.~13 %\ref{fig:del200} 
1511: shows that the situation becomes much less
1512: favorable at the larger WIMP mass of 200 GeV. Here we show results with 50
1513: (left) and 500 (right) events per experiment (before cuts), with the other
1514: parameters as in Fig.~12. %\ref{fig:del50}. 
1515: We see that for optimal $Q_{\rm max}$
1516: matching a large majority of the simulated experiments still satisfy $|\delta
1517: m| \leq 1$. The fact that this fraction decreases from nearly 90\% for the
1518: smaller events samples to about 85\% for the larger samples indicates that our
1519: error estimates become a bit more reliable for larger event
1520: numbers.\footnote{Assuming, unrealistically, that there are 50,000 events in
1521:   each experiment, we find that only about 75\% of all experiments have
1522:   $|\delta m| \leq 1$, indicating a very slow approach to the Gaussian value
1523:   of about 68\%.} Note also that the secondary peak at $\delta m \simeq -1$
1524: is due to the change of denominator in the definition (\ref{delm}).
1525: 
1526: We already saw in Figs.~8 %\ref{fig:m50_100} 
1527: and 11 %\ref{fig:500} 
1528: that setting
1529: $Q_{\rm max, Si} = Q_{\rm max, Ge} = 100$ keV significantly under--estimates
1530: the true WIMP mass if the latter exceeds 100 GeV. This is borne out by
1531: Figs.~13. %\ref{fig:del200}. 
1532: Since the statistical error decreases with increasing
1533: number of events, $\delta m$ is much smaller in the right frame; in fact, most
1534: of the simulated experiments now fall in the ``underflow bin'' $\delta m =
1535: -5$, which also contains all experiments giving even smaller values. In other
1536: words, most of the simulated experiments would give a reconstructed WIMP mass
1537: more than five estimated standard deviations below the true value, if no
1538: $Q_{\rm max}$ matching is used.
1539: 
1540: Unfortunately the error estimates resulting from our algorithm of determining
1541: $Q_{\rm max, Si}$ by minimizing $\chi^2(m_{\chi,{\rm rec}})$ are also not very
1542: reliable in this case. For the smaller event sample, we find that about 58\%
1543: of the simulated experiments yield $|\delta m| \leq 1$, while nearly all
1544: experiments give $|\delta m| \leq 2$. While these numbers are not so different
1545: from the corresponding Gaussian predictions, the distribution of $\delta m$ is
1546: clearly highly non--Gaussian in this case. Just as in the scenario without
1547: $Q_{\rm max}$ matching the error estimates actually become less reliable with
1548: increasing event samples. If both experiments contain on average 500 events
1549: each, less than 40\% of the experiments have $|\delta m| \leq 1$, while more
1550: than 30\% of the simulations yield $|\delta m| \geq 2$. In fact, the most
1551: likely value of $\delta m$ is now close to 3. In view of these observations,
1552: the fact that the median $\delta m$ is close to zero, so that the median
1553: reconstructed WIMP mass is close to the true value as already shown in
1554: Fig.~11, %\ref{fig:500}, 
1555: seems almost accidental. Recall also that the nominal ``$1
1556: \sigma$'' uncertainty of the reconstructed WIMP mass still amounts to about 40
1557: GeV in this case. This means that the most likely value of $m_{\chi,{\rm
1558:     rec}}$ predicted by our algorithm exceeds the true value by more than 100
1559: GeV. This clearly leaves some room for improvements. The fact that optimal
1560: $Q_{\rm max}$ matching continues to give good results for both the
1561: reconstructed WIMP mass and its error indicates that better data--based
1562: algorithms might very well exist.
1563: 
1564: \section{Summary and Conclusions}
1565: 
1566: In this paper we described methods to determine the mass of a Weakly
1567: Interacting Massive Particle detected ``directly'', i.e., through the recoil
1568: energy deposited in a detector by the recoiling nucleus after a WIMP scattered
1569: elastically off this nucleus. Our methods are model--independent in the sense
1570: that they do not need any assumption about the WIMP velocity distribution. The
1571: price one has to pay for this is that one will need positive signals in at
1572: least two different detectors, employing different target nuclei. 
1573: 
1574: Our methods are based on our earlier work \cite{DMDD} on reconstructing the
1575: WIMP velocity distribution, which we briefly reviewed in Sec.~2. In this
1576: earlier work, which was based on results from a single (simulated) experiment,
1577: the WIMP mass $m_\chi$ was an input. In Sec.~3 we showed how one can determine
1578: $m_\chi$ by equating results obtained by different experiments. Here the
1579: moments of the velocity distribution function are particularly useful, since
1580: all events in the sample contribute to any given moment, leading to relatively
1581: low statistical uncertainties. We also described a method for determining
1582: $m_\chi$ that can be used if the ratio of WIMP scattering cross sections on
1583: protons and neutrons is known; this is true, for example, for the
1584: spin--independent scattering of supersymmetric neutralinos, where these two
1585: cross sections are nearly equal. We also showed how to combine these methods
1586: using a $\chi^2$ fitting procedure.
1587: 
1588: Sec.~4 was devoted to a detailed numerical analysis of our methods. We saw
1589: that, assuming the sizes of the event samples are fixed, the statistical
1590: errors will be smaller for larger mass difference between the two target
1591: nuclei. In practice experiments with heavier targets will accumulate more
1592: events, assuming equal exposure, at least if the spin--independent
1593: contribution to the scattering cross section dominates. However, the number of
1594: useful events (after cuts, preferably in an almost background--free energy
1595: range) also depends on other factors, besides the masses of the target nuclei.
1596: 
1597: In our discussion we saw that, for WIMP masses exceeding 100 GeV or so, the
1598: maximal recoil energy $Q_{\rm max}$ of accepted signal events plays a crucial
1599: role. Existing experiments have $Q_{\rm max} \leq 100$ keV. If both targets
1600: used fixed $Q_{\rm max}$ values of this order or even smaller, a significant
1601: systematic error on the extracted WIMP mass results. In principle this problem
1602: can be solved by matching the $Q_{\rm max}$ values of the two experiments. The
1603: problem is that perfect matching requires prior knowledge of the WIMP mass. We
1604: tried two algorithms to overcome this problem. Determining $Q_{\rm max}$ of
1605: one experiment iteratively should converge ``on average'', but in a given
1606: experiment often leads to an endless loop, rather than a specific value of
1607: $Q_{\rm max}$; this problem is particularly severe for small event samples. On
1608: the other hand, determining $Q_{\rm max}$ of the experiment with the lighter
1609: target nucleus by minimizing $\chi^2$ also with respect to this quantity
1610: over--estimates the WIMP mass if it is small, and leads to unreliable error
1611: estimates if the WIMP mass is larger, the problem becoming worse with
1612: increasing event samples. However, the fact that optimal $Q_{\rm max}$
1613: matching works well in all cases, for both the median reconstructed WIMP mass
1614: and its error (which tends to be over--estimated by our expressions), gives us
1615: hope that a better algorithm for $Q_{\rm max}$ matching can be found which
1616: only relies on the data. One possibility that might be worth exploring is to
1617: employ a combination of an iterative procedure and a second $\chi^2$
1618: minimization, where the latter is used only if the former does not converge to
1619: a well--defined value of the reconstructed WIMP mass. 
1620: 
1621: We also found that imposing a cut $Q_{\rm max}$ may actually be beneficial for
1622: small event samples. This is related to our earlier observation \cite{DMDD}
1623: that a typical experiment will under--estimate the higher moments of $f_1$,
1624: which receive significant contributions from recoil energies where only a
1625: fraction of an event is expected to occur in a given experiment. This problem
1626: becomes more acute for heavier target nuclei, since they have softer form
1627: factors. In particular, using Xenon rather than Germanium does not improve the
1628: determination of $m_\chi$, since the spin--independent elastic form factor of
1629: Xenon as predicted by the Woods--Saxon ansatz vanishes for $Q\simeq 95$ keV.
1630: In contrast, the lower (threshold) energy of the experiment does not seem to
1631: be very important, if it can be pushed down to values near 3 keV or less.
1632: 
1633: Our analysis is idealized in that we ignore backgrounds, systematic
1634: uncertainties as well as the finite energy resolution. The relative error on
1635: the recoil energy in existing experiments is small compared to our most
1636: optimistic relative WIMP mass error estimates even with 500 events per
1637: experiment, so ignoring it should be a good approximation. Modern methods of
1638: discriminating between nuclear recoils and other events, combined with muon
1639: veto and good shielding, hold out the possibility of keeping (some)
1640: experiments nearly background--free also in future. 
1641: 
1642: In our numerical analysis we have ignored the expected annual modulation of
1643: the WIMP flux. In practice this can be done if one simply sums all events over
1644: (at least) one full calendar year. In principle one can also use our methods
1645: for subsets of data collected during specific times of the year. However, at
1646: least if the standard ``shifted Gaussian'' velocity distribution is
1647: approximately correct, we do not expect the small annual modulation to play a
1648: significant role, even if one compares experiments taken during different
1649: parts of the year.
1650: 
1651: We saw that our methods work best if the WIMP mass lies in between the masses
1652: of the two target nuclei. Even in that case the error will likely be
1653: significantly larger than the error on $m_\chi$ from collider experiments, if
1654: the WIMP is part of a well--motivated extension of the Standard Model of
1655: particle physics, e.g., if it is the lightest neutralino \cite{susydm} or the
1656: lightest $T-$odd particle in ``Little Higgs'' models \cite{little}. It will
1657: nevertheless be crucial to determine the WIMP mass from direct and/or indirect
1658: Dark Matter experiments as precisely as possible, in order to make sure that
1659: the particle produced at colliders is indeed the WIMP detected by these
1660: experiments. Once one is confident of this identification, one can use further
1661: collider measurements to constrain the WIMP couplings. This in turn will allow
1662: to calculate the WIMP--nucleus scattering cross section. Together with the
1663: determination of the WIMP velocity distribution \cite{DMDD}, this will then
1664: yield a determination of the local WIMP number density via the total counting
1665: rate in direct detection experiments. Knowledge of the WIMP couplings will
1666: also permit prediction of the WIMP annihilation cross section. Together with
1667: the Dark Matter density inferred from cosmological observations, this will
1668: allow to test our understanding of the early universe \cite{early}. A
1669: determination of the WIMP mass from Dark Matter detection experiments is thus
1670: a crucial ingredient in many analyses that shed light on the dark sector of
1671: the universe.
1672: 
1673: \subsubsection*{Acknowledgments}
1674: 
1675: This work was partially supported by the Marie Curie Training Research Network
1676: ``UniverseNet'' under contract no. MRTN-CT-2006-035863, as well as by the
1677: European Network of Theoretical Astroparticle Physics ENTApP ILIAS/N6 under
1678: contract no. RII3-CT-2004-506222.
1679: 
1680: %
1681: \appendix
1682: \setcounter{equation}{0}
1683: \renewcommand{\theequation}{A\arabic{equation}}
1684: %
1685: %
1686: \section{Derivatives needed in the error analysis}
1687: 
1688: At the end of Sec.~2 we gave the covariance matrix of the quantities appearing
1689: in the definition of the ${\cal R}_n$ as well as ${\cal R}_\sigma$. In
1690: Eqs.(\ref{sigmom}) and (\ref{sigsig1}) we also gave expressions relating the
1691: errors on the reconstructed WIMP masses to the errors on ${\cal R}_n$ and
1692: ${\cal R}_\sigma$. The only missing ingredients in the calculation of the
1693: errors on our various estimators of $m_\chi$ are the first derivatives of
1694: ${\cal R}_n$ and ${\cal R}_\sigma$.
1695: 
1696: We begin with the former. From Eq.(\ref{eqn3201}), it can be found directly
1697: that 
1698: %
1699: \cheqnXa{A}
1700: \beqn
1701: \Pp{\calRn}{r_X(Q_{{\rm min},X})} \=     \frac{2}{n}
1702: \bfrac{  Q_{{\rm min},X}^{(n+1)/2} \IzX-(n+1) Q_{{\rm min},X}^{1/2} \InX}
1703:       {2 Q_{{\rm min},X}^{(n+1)/2} r_X(Q_{{\rm min},X})+(n+1)
1704:  \InX F^2_X(Q_{{\rm min},X}) }
1705:         \non\\
1706:  \conti ~~~~~~~~~~~~~~~~ \times %16
1707:   \bfrac{F^2_X(Q_{{\rm min},X})} {2 Q_{{\rm min},X}^{1/2} r_X(Q_{{\rm min},X})
1708: + \IzX F^2_X(Q_{{\rm min},X})} \calRn\, ,
1709: \eeqn
1710: %
1711: \cheqnXb{A}
1712: \beqn
1713: \Pp{\calRn}{\InX} \= \frac{n+1}{n}
1714: \bfrac{F^2_X(Q_{{\rm min},X})} {2 Q_{{\rm min},X}^{(n+1)/2}
1715: r_X(Q_{{\rm min},X}) + (n+1) \InX F^2_X(Q_{{\rm min},X})} \calRn\, ,
1716:     \non\\
1717: \eeqn
1718: %
1719: and
1720: %
1721: \cheqnXc{A}
1722: \beq
1723: \Pp{\calRn} {\IzX} =-\frac{1}{n} \bfrac{F^2_X(Q_{{\rm min},X})}
1724: {2 Q_{{\rm min},X}^{1/2} r_X(Q_{{\rm min},X}) 
1725: + \IzX F^2_X(Q_{{\rm min},X})} \calRn\, .
1726: \eeq
1727: \cheqnX{A}
1728: %
1729: By first exchanging $Q_{{\rm min},X}^{(n+1)/2}$ and $(n+1) \InX$ with $Q_{{\rm
1730:     min},X}^{1/2}$ and $\IzX$, respectively, and then replacing $X$ by $Y$,
1731: one finds
1732: %
1733: \cheqnXa{A}
1734: \beqn
1735: \Pp{\calRn} {r_Y(Q_{{\rm min},Y})} \=    -\frac{2}{n}
1736:  \bfrac{ Q_{{\rm min},Y}^{(n+1)/2} \IzY - (n+1) Q_{{\rm min},Y}^{1/2} \InY}
1737: {2 Q_{{\rm min},Y}^{(n+1)/2} r_Y(Q_{{\rm min},Y}) 
1738: + (n+1) \InY F^2_Y(Q_{{\rm min},Y})}
1739:         \non\\
1740: \conti ~~~~~~~~~~~~ \times %12
1741: \bfrac{F^2_Y(Q_{{\rm min},Y})} {2 Q_{{\rm min},Y}^{1/2} r_Y(Q_{{\rm min},Y})
1742: + \IzY F^2_Y(Q_{{\rm min},Y})} \calRn \, ,
1743: \eeqn
1744: %
1745: \cheqnXb{A}
1746: \beqn
1747: \Pp{\calRn}{\InY} \=-\frac{n+1}{n}
1748: \bfrac{F^2_Y(Q_{{\rm min},Y})} {2 Q_{{\rm min},Y}^{(n+1)/2} 
1749: r_Y(Q_{{\rm min},Y})+(n+1) \InY F^2_Y(Q_{{\rm min},Y})} \calRn\, ,
1750:     \non\\
1751: \eeqn
1752: %
1753: and
1754: %
1755: \cheqnXc{A}
1756: \beq
1757: \Pp{\calRn} {\IzY} = \frac{1}{n} \bfrac{F^2_Y(Q_{{\rm min},Y})}
1758: {2 Q_{{\rm min},y}^{1/2} r_Y(Q_{{\rm min},Y}) + \IzY F^2_Y(Q_{{\rm min},Y})} 
1759: \calRn\, .
1760: \eeq
1761: \cheqnX{A}
1762: %
1763: Note that a factor ${\cal R}_n$ appears in all these expressions; this allows
1764: to cast the final result for the error on $m_\chi$ estimated using moments of
1765: $f_1$ into a form analogous to that in Eq.(\ref{eqn3105}) even in the presence
1766: of non--trivial cuts on the recoil energy, with the same prefactor. Moreover,
1767: all the $\IzX, \, \IzY, \, \InX, \, \InY$ should be understood to be computed
1768: according to Eq.(\ref{In}) or its discretization (\ref{eqn2210}) with
1769: integration limits $Q_{\rm min}$ and $Q_{\rm max}$ specific for that target.
1770: 
1771: Similarly, the derivatives of ${\cal R}_\sigma$ can be computed from
1772: Eq.(\ref{Rsigma1}): 
1773: %
1774: \cheqnXa{A}
1775: \beq \label{Rsigmadifa}
1776: \Pp{{\cal R}_\sigma}{r_X(Q_{{\rm min},X})} = \bfrac {2 Q_{{\rm min},X}^{1/2}} 
1777: {2 Q_{{\rm min},X}^{1/2} r_X(Q_{{\rm min},X}) + \IzX F_X^2(Q_{{\rm min},X})}
1778: {\cal R}_\sigma\, ,
1779: \eeq
1780:  and
1781: \cheqnXb{A}
1782: \beq \label{Rsigmadifb}
1783: \Pp{{\cal R}_\sigma}{\IzX} = \bfrac {F_X^2(Q_{{\rm min},X})}
1784: {2 Q_{{\rm min},X}^{1/2} r_X(Q_{{\rm min},X}) + \IzX F_X^2(Q_{{\rm min},X})}
1785: {\cal R}_\sigma\, ,
1786: \eeq
1787: \cheqnX{A}
1788: %
1789: $\!\!\!$
1790: The derivatives with respect to the $Y$ variables can be obtained from
1791: Eqs.(\ref{Rsigmadifa}) and ({\ref{Rsigmadifb}}) by simply changing $X \lto Y$ everywhere
1792: and changing the overall plus signs to minus signs.
1793: 
1794: 
1795: \begin{thebibliography}{99}
1796: %
1797: % Evidence for the existence of Dark Matter
1798: % Velocities of galaxies
1799: \bibitem{evida}
1800:  F. Zwicky, {\it Helv. Phys. Acta} {\bf 6}, 110 (1933);
1801:  S. Smith, {\it Astrophys. J.} {\bf 83}, 23 (1936).
1802: % Rotation curves of spiral galaxies
1803: \bibitem{evidb}
1804:  V. C. Rubin and W. K. Ford, {\it Astrophys. J.} {\bf 159}, 379 (1970);
1805:  S. M. Faber and J. S. Gallagher, {\it Annu. Rev. Astron. Astrophys.} {\bf
1806:  17}, 135 (1979); 
1807:  V. C. Rubin, W. K. Ford, and N. Thonnard, {\it Astrophys. J.} {\bf 238}, 471
1808:  (1980); 
1809:  K. G. Begeman, A. H. Broeils, and R. H. Sanders,
1810:  {\it Mon. Not. R. Astron. Soc.} {\bf 249}, 523 (1991);
1811:  R. P. Olling and M. R. Merrifield, {\it Mon. Not. R. Astron. Soc.} {\bf 311},
1812:  361 (2000). 
1813: % Escape velocity from the Milky Way
1814: \bibitem{evidc}
1815:  M. Fich and S. Tremaine, {\it Annu. Rev. Astron. Astrophys.} {\bf 29}, 409
1816:  (1991). 
1817: %
1818: % WMAP three year data analysis
1819: % ``Wilkinson Microwave Anisotropy Probe (WMAP) Three Year Results:
1820: % Implications for Cosmology''
1821: \bibitem{WMAP3}
1822:  WMAP Collab., D. N. Spergel {\it et al.}, {\it Astrophys. J. Suppl.} {\bf 170}, 377 (2007).
1823: %  {\tt arXiv:astro-ph/0603449}.}
1824: %
1825: % Bullet cluster, collision of two clusters of galaxies
1826: \bibitem{bullet}
1827:  D. Clowe {\it et al.}, {\it Astrophys. J. Lett.} {\bf 648}, L109 (2006);
1828: %  {\tt arXiv:astro-ph/0608407};
1829:  D. Clowe, S. W. Randall, and M. Markevitch, {\it Nucl. Phys. Proc. Suppl.} {\bf 173}, 28 (2007).
1830: %  {\tt arXiv:astro-ph/0611496}.
1831: %
1832: % Theories of Dark Matter
1833: % Supersymmetric Dark Matter
1834: \bibitem{susydm}
1835:  G. Jungman, M. Kamionkowski, and K. Griest, {\it Phys. Rep.} {\bf 267}, 195
1836:  (1996); 
1837:  G. Bertone, D. Hooper, and J. Silk, {\it Phys. Rep.} {\bf 405}, 279 (2005).
1838: %
1839: % Detections of Dark Matter
1840: % Elastic scattering from nucleus
1841: \bibitem{detaa}
1842:  M. W. Goodman and E. Witten, {\it Phys. Rev.} {\bf D 31}, 3059 (1985);
1843:  I. Wassermann, {\it Phys. Rev.} {\bf D 33}, 2071 (1986);
1844:  K. Griest, {\it Phys. Rev.} {\bf D 38}, 2357 (1988);
1845:  P. F. Smith and J. D. Lewin, {Phys. Rep.} {\bf 187}, (1990) 203.
1846: % Elastic scattering from nucleus - annual modulation
1847: \bibitem{detab}
1848:  A. K. Drukier, K. Freese, and D. N. Spergel, {\it Phys. Rev.} {\bf D 33},
1849:  3495 (1986);
1850: %
1851: %\bibitem{detac}
1852:  K. Freese, J. Frieman, and A. Gould, {\it Phys. Rev.} {\bf D 37}, 3388 (1988).
1853: %
1854: \bibitem{green}
1855: A. M. Green, {\it J. Cosmol. Astropart. Phys.} {\bf 0708}, 022 (2007),
1856:  {\tt arXiv:hep-ph/0703217}.
1857: %
1858: \bibitem{crisis}
1859: See e.g., A. Tasitsiomi, {\it Int. J. Mod. Phys.} {\bf D 12}, 1157 (2003).
1860: %{\tt astro-ph/0205464}.
1861: %
1862: \bibitem{rotate}
1863: M. Kamionkowski and A. Kinkhabwala, {\it Phys. Rev.} {\bf D 57}, 3256 (1998).
1864: %{\tt hep-ph/9710337}.
1865: %
1866: \bibitem{sikivie}
1867: P. Sikivie and J. R. Ipser, {\it Phys. Lett.} {\bf B 291}, 288 (1992);
1868: P. Sikivie, {\it Phys. Lett.} {\bf B 567}, 1 (2003); %, {\tt astro-ph/0109296};
1869: A. Natarajan and P. Sikivie, {\it Phys. Rev.} {\bf D 73}, 023510 (2006),
1870: %{\tt astro-ph/0510743},
1871:  and {\it Phys. Rev.} {\bf D 76}, 023505 (2007),
1872: {\tt arXiv:0705.0001 [astro-ph]}.
1873: %
1874: \bibitem{streams}
1875: K. Freese, P. Gondolo, H. J. Newberg, and M. Lewis, {\it Phys. Rev. Lett.} {\bf 92},
1876: 111301 (2004). %, {\tt astro-ph/0310334}.
1877: %
1878: \bibitem{DMDD} 
1879:  M. Drees and C. L. Shan, {\it J. Cosmol. Astropart. Phys.} {\bf 0706}, 011
1880:  (2007), {\tt arXiv:astro-ph/0703651}.
1881: %
1882: \bibitem{pdg}
1883: For a brief review on direct Dark Matter detection techniques, see the
1884: contribution by M. Drees and G. Gerbier to the Particle Data Booklet, 
1885: W.-M. Yao {\it et al.}, {\it J. Phys.} {\bf G 33}, 1 (2006).
1886: %
1887: %
1888: % Form factors
1889: % Exponential form factor
1890: %\bibitem{FQa}
1891: % S. P. Ahlen {\it et al.}, {\it Phys. Lett.} {\bf B 195}, 603 (1987).
1892: 
1893: % Woods--Saxon form factor
1894: \bibitem{FQb}
1895: J. Engel, {\it Phys. Lett.} {\bf B 264}, 114 (1991).
1896: %
1897: \bibitem{low_th}   %Low-threshold detector
1898: H. T. Wong, {\tt arXiv:0803.0033 [hep-ex]} (2008).
1899: 
1900: \bibitem{little}
1901: See e.g., A. Birkedal, A. Noble, M. Perelstein, and A. Spray, {\it Phys. Rev.} {\bf
1902:   D 74}, 035002 (2006). %, {\tt hep-ph/0603077}.
1903: 
1904: \bibitem{early}
1905: See e.g., M. Drees, H. Iminniyaz, and M. Kakizaki, {\it Phys. Rev.} {\bf D 76}, 103524
1906: (2007), {\tt arXiv:0704.1590 [hep-ph]};
1907: D. J. H. Chung, L. L. Everett, K. Kong, and K. T. Matchev, {\it J. High Energy. Phys.} {\bf
1908:   0710}, 016 (2007), {\tt arXiv:0706.2375 [hep-ph]}.
1909: 
1910: \end{thebibliography}
1911: 
1912: \end{document}
1913: