1: \cleardoublepage{}
2: \chapter{Enhanced Charge Dividing Circuits}
3: \label{ch:enhanced-charge-dividing-circuits}
4:
5: \chapterquote{%
6: Character is like a tree and reputation like a shadow.
7: The shadow is what we think of it; the tree is the real thing.%
8: }{%
9: Abraham Lincoln, $\star$ 1809 -- $\dagger$ 1865
10: }
11:
12: \PARstart{T}{he} first time that the center of gravity algorithm (CoG) was used for
13: position estimation in nuclear medicine was in 1953, when Hal Anger
14: developed the first scintillation camera \mycite{Anger}{}{1958} (refer
15: to figure~\ref{fig:anger-patent} of chapter
16: \ref{ch:introduction}). As a consequence of its subtle and easy
17: electronic implementation, it has nowadays become very useful
18: for the position estimation of $\gamma$-ray imaging detectors
19: even though it introduces systematic errors.
20: The method is often called {\em centroiding} and the measured position
21: estimates for the $x$ and $y$ spatial directions are given the name
22: {\em centroid}. Actually, only in rare cases can one avoid using the
23: center of gravity algorithm for the analysis of the signal
24: distribution. High-energy physics experiments show a
25: dramatic increase in the detector segmentation and therefore a
26: tremendous amount of data have to be processed. As a consequence,
27: the use of the center of gravity
28: algorithm for position determination has become very widespread for
29: scientific and practical applications (Lauterjung {\em et al.}\
30: \cite{Lauterjung:1963}, Kuhlmann {\em et al.}\ \cite{Kuhlmann:1966a},
31: Bock {\em et al.}\ \cite{Bock:1966}, McDicken {\em et al.}\ \cite{McDicken:1967},
32: Doehring {\em et al.}\ \cite{Doehring:1969} and Landi \cite{Landi:2002}).
33:
34: In this section, the mathematical properties of the
35: center of gravity
36: algorithm and especially the consequences that result from its
37: electronic implementation are reviewed in detail. However, the special
38: emphasis is put on the analysis of how the existing charge dividing
39: circuits can be enhanced to measure the second moment in addition
40: to the centroid and the total charge.
41:
42:
43: \section{Introduction, Conventions and General Considerations}
44:
45: \subsection{Statistical Estimates}
46: \label{ch:stat-estimates}
47:
48: By analogy with the geometric centroid, which
49: represents the center of gravity\footnote{Occasionally, the method is
50: also called {\em center of mass} algorithm. In an uniform gravitational
51: field, the center of mass and the center of gravity are identical.}
52: of a body, the generalized functional centroid is defined as
53: \begin{equation}
54: \label{eq:func-centroid}
55: \langle x\rangle = \frac{\int_\omega x \;\varphi(x) dx}{\int_\omega \varphi(x) dx},
56: \end{equation}
57: provided that the integrals exist. In equation~(\ref{eq:func-centroid})
58: $\varphi(x)$ is an integrable function and $\omega \subseteq
59: \Omega \subseteq \mathbb{R}$ is the support within the domain $\Omega$ of $\varphi(x)$.
60: If $\varphi(x)$ is replaced by an arbitrary probability density function
61: $\mathcal{P}(x)$, equation~(\ref{eq:func-centroid}) reduces to
62: \begin{equation}
63: \label{eq:expec-val}
64: \mathcal{E}(x) = \int_\omega x \mathcal{P}(x) dx,
65: \end{equation}
66: with $\int_\omega \mathcal{P}(x) dx = 1$, since $\mathcal{P}(x)$
67: is a probability density function. $\mathcal{E}(x)$ is called the
68: {\em mean}, or more generally, the expectation value of $x$ respect to
69: $\mathcal{P}(x)$.
70: The mean (\ref{eq:expec-val}) is not to be confused with the {\em
71: average probability} given by
72: \begin{equation}
73: \label{eq:average-prob}
74: \bar{\mathcal{P}} = \frac{\int_\omega \mathcal{P}(x) dx}{\int_\omega dx}.
75: \end{equation}
76: Equation~(\ref{eq:expec-val}) can be
77: generalized to yield the expectation value $\mathcal{E}[g(x)]$
78: for an arbitrary function $g(x)$ respect to the probability density
79: function $\mathcal{P}(x)$:
80: \begin{equation}
81: \label{eq:expec-val-gen}
82: \mathcal{E}[g(x)] = \int_\omega g(x) \mathcal{P}(x) dx,
83: \end{equation}
84: whenever the integral is absolutely convergent. The same
85: generalizations can be made for equation~(\ref{eq:func-centroid}).
86: Of special statistical interest is the class of monomials $g(x)=x^k$
87: with $k\in\mathbb{N}$. Plugging in this class into
88: equation~(\ref{eq:func-centroid}) for the functional centroid, one
89: obtains
90: \begin{gather}
91: \label{eq:func-moments}
92: \mu_k=\frac{\int_\omega x^k\varphi(x)dx}{\int_\omega\varphi(x)
93: dx}\quad\mbox{and}\\
94: \mu_k'=\frac{\int_\omega (x-\mu_1)^k \varphi(x) dx}{\int_\omega \varphi(x) dx}.
95: \end{gather}
96: The sequences ${\mu_k}$ and ${\mu_k'}$ of real or complex numbers are called the
97: normalized {\em moments} and the normalized {\em central moments} of
98: $\varphi(x)$ respectively being $\mu_0,\mu_0'\equiv 1$ and $\mu_1'\equiv 0$ by
99: definition. Special names are given to the lowest order central moments of
100: probability distribution functions. As already mentioned, the first
101: moment $\mu_1$ is called the mean. $\mu_2'$ is given the name of {\em
102: variance}. It is a measure of the spread or width of
103: $\mathcal{P}(x)$. The next two higher centered moments are called
104: skewness and kurtosis respectively. While the skewness is a measure
105: for the degree of antisymmetry of the function, the kurtosis
106: characterizes its relative flatness compared to a standard
107: distribution. Further details can be found in standard textbooks,
108: {\em e.g.}\ \cite{Lindley,Press:1992}.
109: If the moments are not normalized, then $\mu_0$ and $\mu_0'$ give the norm of
110: $\varphi(x)$ and normally represent physical quantities. It is also possible
111: (and straightforward) to define the {\em joint moments} and expectation values for
112: multivariate functions $\varphi(\bar{x})$, $g(\bar{x})$ and
113: $\mathcal{P}(\bar{x})$ defined on $\omega \subseteq \Omega \subseteq
114: \mathbb{R}^n$:
115: \begin{gather}
116: \label{eq:multivariate-moms-1}
117: \mu_{k_1,\ldots,k_n} = \frac{\int_\omega \prod_i^n x_i^{k_i}
118: \varphi(\bar{x}) d\bar{x}}{\int_\omega \varphi(\bar{x}) d\bar{x}},\\\label{eq:multivariate-moms-2}
119: {\mu'}_{k_1,\ldots,k_n} = \frac{\int_\omega \prod_i^n (x_i-\mu_{1_i})^{k_i}
120: \varphi(\bar{x}) d\bar{x}}{\int_\omega \varphi(\bar{x}) d\bar{x}}\mbox{ and}\\\label{eq:multivariate-moms-3}
121: \mathcal{E}[g(\bar{x})]=\int_\omega g(\bar{x}) \mathcal{P}(\bar{x}) d\bar{x}.
122: \end{gather}
123:
124: \begin{figure}[!t]
125: \centering
126: \subfigure[][Median, mode and mean of the non-symmetric example
127: distribution function
128: $\mathcal{P}_a(x)=\frac{xe^{10-x/2}}{4e^{10}-44}$. In general, the
129: three values are different.]{\label{subfig:stat-gen-func}%
130: \psfrag{median}{\it median}
131: \psfrag{C}{$\bar{\mathcal{P}}$}
132: \psfrag{A}{\rotatebox{-20}{\it mode}}
133: \psfrag{B}{\rotatebox{-20}{\it mean}}
134: \includegraphics[width=0.488\textwidth]{cdr/general-dist-stat}
135: }
136: \subfigure[][Median, mode and mean of the axially symmetric example
137: distribution function
138: $\mathcal{P}_b(x)=\frac{e^{-(x-10)^2}}{\sqrt{\pi}\erf{10}}$. The
139: three estimators give the same result for this special case.]{\label{subfig:stat-gauss-func}%
140: \psfrag{median}{\it median}
141: \psfrag{B}{$\bar{\mathcal{P}}$}
142: \psfrag{A}{\rotatebox{-20}{\it mode}}
143: \psfrag{C}{\rotatebox{-20}{\it mean}}
144: \includegraphics[width=0.488\textwidth]{cdr/gauss-dist-stat}
145: }
146: \caption[Mean, mode , median and average probability for
147: two unimodal PDF]{Location of mean, mode, median and average probability for
148: two special cases of unimodal probability distribution functions in
149: a closed interval. The average probability of both distributions is also drawn.}
150: \label{fig:stat-descript}
151: \end{figure}
152:
153: Alternative estimators for probability distribution functions are the
154: {\em median} and the {\em mode}. The mode is the value $\bar{x}$,
155: where $\mathcal{P}(\bar{x})$ takes on a maximum. Clearly there exist
156: {\em multi-modal} probability distribution functions. The median of
157: $\mathcal{P}(x)$ is the value $x_\mathit{med}$ for which larger and
158: smaller values are equally probable, {\em i.e.}\ the value for which the
159: area below $\mathcal{P}(x)$ to the left from $x_\mathit{med}$ is equal
160: to the area below $\mathcal{P}(x)$ to the right from $x_\mathit{med}$.
161: Figures \ref{fig:stat-descript} show these values for the two
162: different probability distributions
163: $\mathcal{P}_a(x)=\frac{xe^{10-x/2}}{4e^{10}-44}$ and
164: $\mathcal{P}_b(x)=\frac{e^{-(x-10)^2}}{\sqrt{\pi}\erf{10}}$ on the
165: support interval $\omega=[0,20]$. Note that for general probability
166: distribution functions, all mentioned estimators take on different
167: values, as can be seen for $\mathcal{P}_a(x)$ in figure
168: \ref{subfig:stat-gen-func}, while there are classes of distribution
169: functions for which some or even all of the mentioned estimators give the same results (figure
170: \ref{subfig:stat-gauss-func}). Actually the single condition of only
171: considering axially symmetric univariate functions obviously makes the
172: median coincide with the mean. For single mode axially symmetric
173: univariate functions these values also coincide with the mode. This
174: and the fact that numeric computation of the different estimators are
175: of different computational effort, are the reasons for occasionally taking one
176: of these estimators as the approximation for another. One has to take
177: special care with these approximations if one is dealing with asymmetric
178: distribution functions, since the introduced errors can become significant.
179:
180: \subsection{Signal Characteristics of Photomultiplier Tubes}
181: \label{subsec:pmt-as-ideal--current-source}
182:
183: The anode current of a photomultiplier tube operated in the saturated
184: current region depends on the incident scintillation light flux but
185: not on the load that is connected to the anode \mycite{Flyckt}{{\em et al.}\
186: }{2002}. One can treat the anodes of the photomultiplier
187: tubes like ideal current sources, since they are more or less
188: sophisticated collection electrodes with necessarily low internal impedance
189: (Güttinger {\em et al.}\ \cite{Guettinger:1976}, McHose \cite{McHose:1989-tp133}).
190:
191: \begin{figure}[!t]
192: \centering
193: \subfigure[][Anode circuit with load resistance $R_L$ and capacitance $C_L$]
194: {\label{subfig:anode-connect}
195: \psfrag{dn}{$\mathrm{D_n}$}
196: \psfrag{dn1}{$\mathrm{D_{n-1}}$}
197: \psfrag{dn2}{$\mathrm{D_{n-2}}$}
198: \psfrag{A}{$\mathrm{A}$}
199: \psfrag{Rl}{$\mathrm{R_L}$}
200: \psfrag{Cl}{$\mathrm{C_L}$}
201: \includegraphics[height=0.27\textwidth]{cdr/anode-connection}
202: }\hspace*{0.04\textwidth}
203: \subfigure[][Wave forms of the output voltages for different time
204: constants of the anode load. The $\kappa$-values for the 5 different
205: graphs are 0, 0.2, 0.4, 0.6, 0.8 and 1.]
206: {\label{subfig:voltage-pulses}
207: \psfrag{J}{\hspace*{-5em}$\mathcal{V}_\tincaps{PMT}(t)/\mathcal{V}_\tincaps{PMT}(0)\big|_{\kappa=0}$}
208: \psfrag{t}{$t/\tau_\tincaps{Scint}$}
209: \psfrag{k1}{\rotatebox{-70}{$\kappa=1$}}
210: \psfrag{k0}{\rotatebox{75}{$\kappa=0$}}
211: \includegraphics[height=0.32\textwidth]{cdr/voltage-pulses}
212: }
213: \caption[Anode circuit and pulse shapes for anode
214: loads]{Typical anode circuit (left) and pulse shapes for anode
215: loads of different $\tau_\tincaps{RC}=R_L C_L$ (right).
216: }
217: \end{figure}
218:
219: The temporal dependence of the scintillation light emission
220: $\mathcal{L}(t)$ that was induced at time $t_0$ by a detected
221: $\gamma$-photon can be approximated by an exponential decay law %{\red X}
222: \begin{equation}
223: \label{eq:scint-time-depend}
224: \mathcal{L}(t)=\frac{\bar{n}_\tincaps{Ph}(\mathcal{E})}{\tau_\tincaps{Scint}}e^{-\frac{t}{\tau_\tincaps{Scint}}},
225: \end{equation}
226: $\tau_\tincaps{Scint}$ being the decay time constant of the scintillator
227: used and $\bar{n}_\tincaps{Ph}(\mathcal{E})$ the average number of scintillation
228: photons set free from the particle with energy $\mathcal{E}$.
229: The resulting number of photoelectrons is given by the convolution of
230: $\mathcal{L}(t)$ with the photomultiplier's pulse response
231: $\mathcal{R}_\tincaps{PMT}(t)$. This response is mainly determined by the transit
232: time that requires the electron avalanche to pass the
233: multiplying stages and to reach the anode.
234: In particular, the difference in this time for each single
235: photoelectron leads to temporal impulse broadening \mycite{Kume}{}{1994}. If one assumes a
236: $\mathcal{R}_\tincaps{PMT}(t)$ of Gaussian shape,
237: the current at the anodes becomes
238: \begin{equation}
239: \label{eq:scint-pmt-response}
240: \begin{split}
241: \mathcal{J}_\tincaps{PMT}(t) & = \mathcal{L}_\tincaps{Scint}(t)*\mathcal{R}_\tincaps{PMT}(t)\\
242: & = \frac{e^-G_\tincaps{PMT}\bar{n}_\tincaps{PE}(\mathcal{E})}{2\tau_\tincaps{Scint}}%
243: \exp{\left[\frac{\tau_\tincaps{PMT}^{2}}{2\tau_\tincaps{Scint}^{2}}%
244: -\frac{t}{\tau_\tincaps{Scint}}\right]}\left(1+\erf\left[\frac{t}{\sqrt{2}\tau_\tincaps{PMT}}%
245: -\frac{\tau_\tincaps{PMT}}{\sqrt{2}\tau_\tincaps{Scint}}\right]\right),
246: \end{split}
247: \end{equation}
248: where $e^-$ is the elementary charge, $G_\tincaps{PMT}$ the gain of the
249: photomultiplier tube, $\tau_\tincaps{PMT}$ the electron transit time
250: spread, $\bar{n}_\tincaps{PE}(\mathcal{E})$ the average number of
251: photoelectrons and $\erf$ the error function. Increasing the supply voltages
252: of the dynode system leads to a higher electric field strength between
253: the dynode stages and consequently leads to a higher electron speed and
254: shorter transit times. The geometric design type of the dynode
255: system and the total size of the photomultiplier are also of importance
256: for the timing resolution. After all, photomultiplier tubes are,
257: with rise times between $\mathrm{0.7\,ns}$ and $\mathrm{7\,ns}$, photodetectors of exceptionally
258: fast response (Kume {\em et al.}\ \cite{Kume:1988}, McHose \cite{McHose:1989-tp114}).
259: The response pulse width $\tau_\tincaps{PMT}$ is therefore normally
260: negligible compared to the decay time $\tau_\tincaps{Scint}$ of the
261: scintillator and equation \ref{eq:scint-pmt-response} can be approximated by
262: \begin{equation}
263: \label{eq:approx-scint-pmt-response}
264: \mathcal{J}_\tincaps{PMT}(t)\approx\frac{\bar{q}_a(\mathcal{E})}%
265: {2\tau_\tincaps{Scint}}e^{-t/\tau_\tincaps{Scint}},
266: \end{equation}
267: where $e^-G_\tincaps{PMT}\bar{n}_\tincaps{PE}(\mathcal{E})$ has been
268: substituted by the average charge $\bar{q}_a(\mathcal{E})$ collected
269: at the anode. For proper signal collection, the anode is maintained at
270: ground potential via a load resistance $R_L$. This is shown in figure
271: \ref{subfig:anode-connect}. The parallel capacitance $C_L$
272: can be required, or is just a parasitic one and builds, together with
273: the load resistance $R_L$, an R-C network. As can be easily
274: verified by Fourier-analysis, the pulse response of this R-C network
275: is a decaying exponential with the time constant
276: $\tau_\tincaps{RC}=R_LC_L$. It has to be convolved with the
277: photomultiplier pulse given in equation~(\ref{eq:approx-scint-pmt-response})
278: leading to the following voltage pulse
279: \begin{equation}
280: \label{eq:approx-voltage-pulse}
281: \mathcal{V}_\tincaps{PMT}(t)\approx\frac{\bar{q}_a(\mathcal{E})}%
282: {C_L}\frac{\tau_\tincaps{RC}}{\tau_\tincaps{Scint}-\tau_\tincaps{RC}}%
283: \left(e^{-t/\tau_\tincaps{Scint}}-e^{-t/\tau_\tincaps{RC}}\right).
284: \end{equation}
285: The rise time of the voltage pulse is the time from the beginning of
286: the pulse at $t=0$ until it reaches its maximum value, and is
287: given\footnote{In order to cope with unavoidable noise, the anode
288: signal rise time is often defined as the elapsed time between 10\%
289: and 90\% amplitude of the leading edge of the anode current pulse.} by
290: the position of the only extremum of (\ref{eq:approx-voltage-pulse}):
291: \begin{equation}
292: \label{eq:tmax-value}
293: t_\mathit{max}=\frac{\kappa\,\tau_\tincaps{Scint}}{\kappa-1}\log\kappa,
294: \end{equation}
295: being $\kappa=\tau_\tincaps{RC}/\tau\tincaps{Scint}$. For the ideal case of
296: $\tau_\tincaps{RC}=0$, the rise-time $t_\mathit{max}$ in equation~(\ref{eq:tmax-value}) becomes zero. However, this is due to the
297: approximation made in (\ref{eq:approx-scint-pmt-response}). The real
298: rise-time in this case would be given by the maximum position of (\ref{eq:scint-pmt-response}).
299: As $\kappa$ is switched on, the rise-time will be rapidly dominated by
300: the time constant $\tau_\tincaps{RC}$ of the anode circuit. This result is of
301: particular interest for the study of charge dividing circuits made in
302: this chapter. Since $\tau_\tincaps{RC}$ is proportional to $R_L$ as well as to
303: $C_L$, any change in their values leads to a variation in the pulse
304: shape of the anode signal. As will be seen, the resistor
305: $R_L$ especially is required to feature a large variation depending on the
306: position of the anode-segment of the position-sensitive
307: photomultiplier tubes used. This will lead to a strong variation of the
308: rise-time of the anode-signal depending on the impact position of the
309: $\gamma$-ray. If this signal were used to generate the trigger
310: signal for the coincidence detection, the varying pulse shape would
311: introduce a significant temporal jitter, lowering the temporal
312: resolution of the detector as a result. The use of the last dynode
313: instead avoids this problem completely and should be preferred when
314: a charge divider is used for position detection.
315:
316: While these last considerations are of major importance for the timing
317: characteristics of $\gamma$-ray imaging detectors, especially in
318: coincidence mode, they are of sparse interest for the positioning
319: algorithm itself. All versions of the charge dividing circuits discussed here
320: work without capacitors except for a few that are needed for
321: frequency and offset compensation of an operational amplifier.
322: However, they do not change the pulse shape and do not affect the
323: computation of the first order moments of the signal distribution.
324: These are obtained exclusively by using resistances, which are supposed
325: to behave in an ideal way, {\em i.e.}\ any frequency dependence,
326: temperature dependency, etc.\ is neglected. Consequently, analyzing the currents that
327: are the temporal derivatives of the charges is equivalent to analyzing
328: the charges itself. In this work, except for a few exceptions, the
329: currents will be studied.
330:
331:
332: \subsection{General Preamplifier Design}
333: \label{ch:general-preamplifier-configuration}
334:
335: \begin{figure}[t]
336: \centering
337: \subfigure[][Schematic for a current controlled voltage
338: source.]{\label{subfig:i-u-converter}%
339: \psfrag{R}{$R$}
340: \psfrag{OP}{$\mathrm{OP}$}
341: \psfrag{Ji}{$J_\mathit{In}$}
342: \psfrag{Ui}{$U_\mathit{In}$}
343: \psfrag{Ua}{$U_\mathit{Out}$}
344: \includegraphics[width=0.28\textwidth]{cdr/i-u-converter}}
345: \subfigure[][Schematic for a voltage controlled voltage
346: source.]{\label{subfig:electrometer-v}%
347: \psfrag{R1}{$R_1$}
348: \psfrag{R2}{$R_2$}
349: \psfrag{OP}{$\mathrm{OP}$}
350: \psfrag{Ui}{$U_\mathit{In}$}
351: \psfrag{Ua}{$U_\mathit{Out}$}
352: \includegraphics[width=0.28\textwidth]{cdr/electrometer-v}}
353: \subfigure[][Schematic for the preamplifier configuration to be used
354: with charge dividing circuits.]{\label{subfig:preamp-conf}%
355: \psfrag{R}{$R$}
356: \psfrag{OP1}{$\mathrm{OP1}$}
357: \psfrag{OP2}{$\mathrm{OP1}$}
358: \psfrag{Ji}{$J_\mathit{In}$}
359: \psfrag{Ua}{$U_\mathit{Out}$}
360: \includegraphics[width=0.42\textwidth]{cdr/preamp-conf}}
361: \caption[Elementary operational amplifier
362: configuration and current sensitive preamplifier]{Schematics of elementary operational amplifier
363: configuration and a current sensitive preamplifier for charge
364: dividing circuits.}
365: \end{figure}
366:
367: In the following section, three possible electronic implementations of
368: the center of gravity algorithm are presented. All three versions
369: have in common that they provide currents which have to be measured
370: as exactly as possible. A particular property of all presented charge dividing
371: circuits is that the stages that read and amplify these currents
372: necessarily need an input impedance $Z_\mathit{In}$ that is small
373: compared to the typical resistor values used for the charge
374: divider. Therefore, amplifier configurations with minimized
375: $Z_\mathit{In}$ are required so that the described circuits work predictably.
376:
377:
378: As input stage, normally the current controlled voltage source
379: shown in figure \ref{subfig:i-u-converter} is used. Expressions for its input
380: impedance $Z_\mathit{In}$, output voltage $U_\mathit{Out}$ and output
381: impedance $Z_\mathit{Out}$ can be found in standard textbooks, {\em e.g.}\
382: \cite{Tietze} and, when neglecting parasitic capacitances, are given by
383: \begin{equation}
384: \label{eq:u-i-wandler-chars}
385: \begin{array}{ccc}
386: Z_\mathit{In}=\ds\frac{R}{A_D}\mbox{, }&U_\mathit{Out}=-RJ_\mathit{In}\mbox{, }&Z_\mathit{Out}=\ds\frac{Z_0}{g}.
387: \end{array}
388: \end{equation}
389: Here, $A_D$ denotes the differential gain and $Z_0$ the output
390: impedance of the used operational amplifier OP. $g$ is the loop gain
391: of the circuit and for the inverting amplifier is given by $g=k_FA_D$,
392: with $k_F$ a factor depending only on the external feedback
393: network. In the case of the current controlled voltage source, $k_F$
394: becomes very small, while the differential gain $A_D$ for modern
395: operational amplifiers easily exceeds $\mathrm{10^6}$. As one can see
396: from the middle equation in (\ref{eq:u-i-wandler-chars}), it is easy
397: to reduce the input impedance to a very small value. However, the output
398: resistance $Z_0$ of the same operational amplifier will typically be
399: of a few hundred Ohms, so that the output impedance of circuit
400: \ref{subfig:i-u-converter} can get very small. The problem of the low
401: output impedance of the $I$-$U$ converter stage can be solved by a
402: downstream voltage controlled voltage source, shown in
403: figure~\ref{subfig:electrometer-v}. Its input impedance, output
404: voltage and output impedance are given by
405: \begin{equation}
406: \label{eq:electrometer-chars}
407: \begin{array}{ccc}
408: Z_\mathit{In}=R_\tincaps{CMR}\mbox{, }&U_\mathit{Out}=\ds\frac{U_\mathit{In}}{k_F}\mbox{, }&Z_\mathit{Out}=\ds\frac{Z_0}{g},
409: \end{array}
410: \end{equation}
411: where $R_\tincaps{CMR}$ denotes the common mode resistance of the used
412: amplifier, which normally is of a few $M\Omega$ and can reach in
413: special devices up to $1T\Omega$. For the non-inverting amplifier, the
414: open loop gain is also given by $g=k_FA_D$, with $k_F$ taking the
415: value $R_1/(R_2+R_1)$. For the maximum value of $k_F=1$, the output
416: voltage follows the input voltage. It can be realized by removing
417: $R_1$ from the circuit and shorting $R_2$. In this case, one also
418: obtains the minimum possible output impedance. The final configuration
419: is shown in figure \ref{subfig:preamp-conf} and represents a standard
420: configuration for charge dividing circuit readout. It combines the
421: advantages of both sub-circuits discussed, and is still inexpensive and
422: easy to implement. There are further proposals to optimize
423: operational parameters of preamplifier circuits for photomultipliers,
424: which is outside the scope of this work (refer to Kume
425: \cite{Kume:1994}, Flyckt \cite{Flyckt:2002} and Fabris {\em et al.}\ \cite{Fabris:1999}).
426:
427:
428: \section{Charge Dividing Circuits for Position Determination}
429: \label{ch:charge-div-circuits}
430:
431: The underlying principle of the positioning principle invented by
432: Anger for his scintillation camera \mycite{Anger}{}{1958} is based on
433: Kirchhoff's rules. It uses the fact that a current $J$ that is
434: injected into a configuration consisting of two resistors, $R_l$ and
435: $R_r$ (refer to figure~\ref{subfig:simplest-cdr}), is divided into two
436: partial currents $J_l$ and $J_r$ depending only on the ratio
437: $R_l/R_r$. As a corollary, it is possible to deduce this ratio from
438: both currents $J_l$ and $J_r$ for the notional case that although the
439: sum of both resistor values $R_l$ and $R_r$ is known, their ratio is not.
440: \begin{equation}
441: \renewcommand{\arraystretch}{1.3}
442: \label{eq:kirchoff-rules}
443: \begin{array}{c}
444: J=J_l+J_r\\R_rJ_r=R_lJ_l
445: \end{array}\Longrightarrow%
446: \begin{array}{c}
447: R_l=J_r\ds\frac{R_l+R_r}{J}\\
448: R_r=J_l\ds\frac{R_l+R_r}{J}
449: \end{array}
450: \end{equation}
451: This can be of special interest for position detection of the injected
452: current when there is a known correlation between the resistances
453: $R_l$ and $R_r$ and the position.
454: Consider the simplest case, when one
455: has the situation shown in figure \ref{subfig:wire-and-punctual-current}.
456: A steady current $J$ is injected at the position
457: \mbox{$x_0\in\,\,\rbrack\frac{-L}{2},\frac{L}{2}\lbrack$} into a wire
458: of length $L$. An infinitely small wire segment has the resistance
459: $dR=g(x)dx$ with $dR>0$ and \mbox{$g(x)=\rho(x)/A(x)$}, where $\rho(x)$
460: is the specific resistance of the wire and $A(x)$ its
461: cross-section. The total resistance
462: $R_\tincaps{W}$ of the wire from $-\frac{L}{2}$ to $\frac{L}{2}$
463: is obtained by integration of $g(x)$ over its length, while the
464: resistances to the left and to the right of the injection point $x_0$
465: are obtained by integrating over the corresponding wire-segment:
466: \begin{equation}
467: \label{eq:r-values}
468: R_\mathit{l}(x_0)=\int_{-\frac{L}{2}}^{x_0}g(x)\;dx\,,\quad%
469: R_\mathit{r}(x_0)=\int^{\frac{L}{2}}_{x_0}g(x)\;dx\,,\quad%
470: R_\tincaps{W}=R_\mathit{l}(x_0)+R_\mathit{r}(x_0)=\int_{-\frac{L}{2}}^{\frac{L}{2}}g(x)\;dx
471: \end{equation}
472:
473: The current $J$, which is injected at position $x_0$ into the wire, will be divided into two fractions,
474: $J_\mathit{l}$ and $J_\mathit{r}$, corresponding to the ratio of
475: $R_\mathit{l}$ and $R_\mathit{r}$ and thus on the injection position $x_0$.
476: Likewise, the parallel connection of $R_\mathit{l}$ and
477: $R_\mathit{r}$ in equation~(\ref{eq:r-values}) is given by
478: \begin{equation}
479: \label{eq:parallel-connection}
480: R_\mathit{l}(x_0)\parallel
481: R_\mathit{r}(x_0)=\frac{\int_{-\frac{L}{2}}^{x_0}%
482: g(x)\;dx\,\,\int^{\frac{L}{2}}_{x_0}%
483: g(x)\;dx}{\int_{-\frac{L}{2}}^{\frac{L}{2}}g(x)\;dx}.
484: \end{equation}
485: It is just the overall resistance that sees the current $J$.
486: Making use of Kirschoff's voltage law,
487: $J_\mathit{r}(x_0)R_\mathit{r}(x_0)=R_\mathit{l}(x_0)J_\mathit{l}(x_0)$,
488: the currents to the right and to the left of $x_0$, namely $J_\mathit{r}(x_0)$ and
489: $J_\mathit{l}(x_0)$, can easily be computed:
490: \begin{gather}
491: \label{eq:right-left-c-1}
492: J_\mathit{r}(x_0)=\frac{R_\mathit{l}(x_0)}{R_\mathit{r}(x_0)}\big(J-J_\mathit{r}(x_0)\big)\;=\;\frac{R_\mathit{l}(x_0)
493: }{R_\mathit{r}(x_0)+R_\mathit{l}(x_0)}J\\
494: \label{eq:right-left-c-2}
495: J_\mathit{l}(x_0)=\frac{R_\mathit{r}(x_0)}{R_\mathit{l}(x_0)}\big(J-J_\mathit{l}(x_0)\big)\;=\;\frac{R_\mathit{r}(x_0)}
496: {R_\mathit{r}(x_0)+R_\mathit{l}(x_0)}J
497: \end{gather}
498: where $J_\mathit{r}(x_0)+J_\mathit{l}(x_0)=J$ was used. With either of
499: the two equations~(\ref{eq:right-left-c-1})
500: and~(\ref{eq:right-left-c-2}), it is now possible to reconstruct exactly the position $x_0$ on the wire
501: where the current $J$ was injected. For this, the values of $J_\mathit{r}(x_0)$ and
502: $J_\mathit{l}(x_0)$ and the distribution $g(x)$ of the resistance along the wire must be known.
503: Normally one uses the difference of both currents for the computation:
504: \begin{figure}%
505: \centering
506: \subfigure[][Simplest case for a charge dividing
507: configuration.]{\label{subfig:simplest-cdr}%
508: \psfrag{J}{$J$}
509: \psfrag{Jl}{$J_l$}
510: \psfrag{Jr}{$J_r$}
511: \psfrag{Rr}{\hspace*{-0.5em}$R_r$}
512: \psfrag{Rl}{$R_l$}
513: \psfrag{GND}{$\mathrm{GND}$}
514: \includegraphics[width=0.19\textwidth]{cdr/angeraproach}}\hspace*{1em}
515: \subfigure[][A 'punctual' current $J\delta(x-x_0)$ is injected at the position $x_0$
516: into the wire.]{\label{subfig:wire-and-punctual-current}%
517: \psfrag{x}{$x_0$}
518: \psfrag{L/2}{$\frac{L}{2}$}
519: \psfrag{-L/2}{$\frac{-L}{2}$}
520: \psfrag{x}{$x_0$}
521: \psfrag{wire}{wire}
522: \psfrag{Jl}{$J_\mathit{l}$}
523: \psfrag{Jr}{$J_\mathit{r}$}
524: \psfrag{Jd(x-xo)}{$J\delta(x-x_0)$}
525: \psfrag{g(x)}{$g(x)$}
526: \includegraphics[width=0.37\textwidth]{cdr/delta_current}}\hspace*{1em}
527: \subfigure[][A current distribution $Jg(x)$ is injected along the
528: wire.]{\label{subfig:wire-and-extended-current}%
529: \psfrag{L/2}{$\frac{L}{2}$}
530: \psfrag{-L/2}{$\frac{-L}{2}$}
531: \psfrag{wire}{wire}
532: \psfrag{Jl}{$J_\mathit{l}$}
533: \psfrag{Jr}{$J_\mathit{r}$}
534: \psfrag{Jf(x)}{$J\varphi(x)$}
535: \psfrag{Rg(x)}{$g(x)$}
536: \includegraphics[width=0.37\textwidth]{cdr/current_dist}}
537: \caption[Graphical representation for different injected current distributions
538: along a conductor]{\label{fig:current-dists}%
539: Graphical representation for different injected current distributions
540: along a conductor. If the resistor values $R_l$ and $R_r$ in figure
541: \ref{subfig:simplest-cdr} are chosen in such a way that they are
542: proportional to a spatial direction, the
543: currents $J_l$ and $J_r$ can be used for the determination of the
544: position where $J$ is injected. }
545: \end{figure}
546: \begin{equation}
547: \label{eq:position-general}
548: \frac{J_\mathit{r}(x_0)-J_\mathit{l}(x_0)}{J_\mathit{r}(x_0)+J_\mathit{l}(x_0)}=%
549: \frac{R_\mathit{r}(x_0)-R_\mathit{l}(x_0)}{R_\mathit{r}(x_0)+R_\mathit{l}(x_0)}
550: \end{equation}
551: If one supposes a constant resistance distribution $g(x)\equiv g$
552: along the wire, the resistance to the left and to the
553: right are then given by $R_l=(x_0+L/2)\;g$ and $R_r=(L/2-x_0)\;g$,
554: where $x_0\in\,\,]\frac{-L}{2},\frac{L}{2}[$. Thus, from
555: equations~(\ref{eq:kirchoff-rules}) and (\ref{eq:r-values}), one obtains the
556: following expressions:
557: \begin{equation}
558: \label{eq:example-consant-res-wire}
559: \renewcommand{\arraystretch}{1.3}
560: \begin{array}{c}
561: (x_0+\frac{L}{2})\;g=J_r\frac{R_w}{J}\\
562: (\frac{L}{2}-x_0)\;g=J_l\frac{R_w}{J}
563: \end{array}\Longrightarrow%
564: \begin{array}{c}
565: x_0=(J_r-J_l)\frac{R_w}{2Jg}\\
566: L=(J_r+J_l)\frac{R_w}{Jg}
567: \end{array}\Longrightarrow%
568: x_0=\frac{(J_r-J_l)}{(J_r+J_l)}\frac{L}{2}.
569: \end{equation}
570: For this reason, one can easily deduce the injection point of the current $J$ from
571: $J_l$ and $J_r$.
572: Note that the equation on the far right of
573: (\ref{eq:example-consant-res-wire}) does not depend on the resistance
574: $g$ but only on the length of the wire.
575:
576: The statistical
577: uncertainties of this positioning method are obtained straightforwardly
578: by error propagation and are given by:
579: \begin{equation}
580: \label{eq:anger-logic-errors}
581: \delta x=\frac{L}{J^2}\sqrt{J_r^2\delta J_l^2+J_l^2\delta
582: J_r^2}\quad\mbox{and}\quad\frac{\delta x}{x}=2\frac{\sqrt{J_r^2\delta
583: J_l^2+J_l^2\delta J_r^2}}{J_r^2-J_l^2},
584: \end{equation}
585: with the limits
586: \begin{equation}
587: \label{eq:a-log-error-lims}
588: \frac{\delta x}{x}\Big|_{J_r\rightarrow0}=-\frac{2\sqrt{\delta J_{r}}}{J_l}\mbox{,
589: }\frac{\delta x}{x}\Big|_{J_l\rightarrow0}=\frac{2\sqrt{\delta
590: J_{l}}}{J_r}\quad\mbox{and}\quad\delta x\Big|_{J_{l,r}\rightarrow\frac{J}{2}}=\frac{L}{2J}\sqrt{\delta
591: J_l^2+\delta J_r^2}=\frac{L}{4}\sqrt{\left(\frac{\delta
592: J_l}{J_l}\right)^2+\left(\frac{\delta J_r}{J_r}\right)^2}
593: \end{equation}
594: Clearly, the relative error in the measurement of $x$ at the center
595: position $x=0$ diverges. The absolute error at this position is just
596: the geometric sum of the relative measurement errors in the currents
597: to the left and to the right, scaled by a quarter of the
598: wire-length. At both ends of the wire, the relative positioning error
599: depends only on the measurement error of the current extracted at the opposite
600: end, which does not necessarily equal zero, although the respective
601: current does.
602:
603: Now consider the case of a
604: current distribution $J\varphi(x)$, that is injected simultaneously
605: at all possible positions \mbox{$x_0\in\,\,]\frac{-L}{2},\frac{L}{2}[$}
606: along the wire as shown in figure \ref{subfig:wire-and-extended-current}.
607: It is supposed that the injected current distribution $\varphi(x)$ is normalized.
608: \begin{equation}
609: \label{eq:current-norm}
610: \int_{-\frac{L}{2}}^{\frac{L}{2}}\varphi(x)\;dx=1
611: \end{equation}
612: Since the superposition principle applies to electromagnetic fields,
613: the impulse response of the wire (equations~\ref{eq:right-left-c-1}
614: and~\ref{eq:right-left-c-2}) must be summed up for all possible
615: injection points $x_0$, weighted by the value of the function at
616: this point $\varphi(x_0)$. This will transform the currents $J_\mathit{l}$
617: and $J_\mathit{r}$ into the following functionals of $\varphi(x)$:
618: \begin{equation}
619: \label{eq:superposed}
620: J_\mathit{r}[\varphi]=\frac{J}{R_\tincaps{W}}\int_{-\frac{L}{2}}^{\frac{L}{2}}R_\mathit{l}(x)\varphi(x)\;dx\,,\quad%
621: J_\mathit{l}[\varphi]=\frac{J}{R_\tincaps{W}}\int_{-\frac{L}{2}}^{\frac{L}{2}}R_\mathit{r}(x)\varphi(x)\;dx,
622: \end{equation}
623: and leads to the equivalent of equation~(\ref{eq:position-general}),
624: \begin{equation}
625: \label{eq:centroid}
626: \frac{J_\mathit{r}[\varphi]-J_\mathit{l}[\varphi]}{J_\mathit{r}[\varphi]+J_\mathit{l}[\varphi]}=\frac{1}{R_\tincaps{W}}\int_{-\frac{L}{2}}^{\frac{L}{2}}%
627: \big(R_\mathit{r}(x)-R_\mathit{l}(x)\big)\varphi(x)\;dx,
628: \end{equation}
629: with\begin{equation}
630: \label{eq:energy}
631: J_\mathit{r}[\varphi]+J_\mathit{l}[\varphi]=\frac{J}{R_\tincaps{W}}\int_{-\frac{L}{2}}^{\frac{L}{2}}%
632: \big(R_\mathit{r}(x)+R_\mathit{l}(x)\big)\varphi(x)\;dx=\frac{J}{R_\tincaps{W}}\int_{-\frac{L}{2}}^{\frac{L}{2}}%
633: R_\tincaps{W}\varphi(x)\;dx=J.
634: \end{equation}
635:
636: Equations~(\ref{eq:centroid}) and (\ref{eq:energy}) are valid for any possible dependence
637: of the resistance-distribution $g(x)$ along the wire. Now
638: consider the simplest non-trivial case of a wire with a constant cross section $A(x)\equiv A$
639: and constant specific resistance $\rho(x)\equiv\rho$. One than has
640: $g(x)\equiv g=\mathit{const.}$
641: Setting $g(x)=g$ in equations~(\ref{eq:r-values}), the resistances to the
642: left, to the right, of the whole wire (equation~\ref{eq:r-values}) and the parallel connection in
643: equation~(\ref{eq:parallel-connection}) become
644: \begin{equation}
645: \label{eq:left-r-special}
646: R_\mathit{l}(x)=g\;\!L\left(\frac{1}{2}+\frac{x}{L}\right)\,,\quad%
647: R_\mathit{r}(x)=g\;\!L\left(\frac{1}{2}-\frac{x}{L}\right)\,,\quad%
648: R_\tincaps{W}(x)=g\;\!L,
649: \end{equation}
650: and
651: \begin{equation}
652: \label{eq:parallel-spec}
653: R_\mathit{r}(x)\parallel R_\mathit{l}(x)=g\;\!L\left(\frac{1}{4}-\frac{x^2}{L^2}\right).
654: \end{equation}
655: Equivalently, the currents to the left and to the right in
656: equation~(\ref{eq:superposed}) are given by the
657: simple relations
658: \begin{equation}
659: \label{eq:superposed-spec}
660: J_\mathit{r}[\varphi]=\frac{J}{2}+\frac{J}{L}\int_{-\frac{L}{2}}^{\frac{L}{2}}x\;\varphi(x)\;dx\quad\mbox{and}\quad%
661: J_\mathit{l}[\varphi]=\frac{J}{2}-\frac{J}{L}\int_{-\frac{L}{2}}^{\frac{L}{2}}x\;\varphi(x)\;dx,
662: \end{equation}
663: that directly leads to the first moment of the arbitrary distribution
664: $\varphi(x)$ as discussed in the previous section~\ref{ch:stat-estimates}.
665: \begin{equation}
666: \label{eq:centroid-spec}
667: \frac{J_\mathit{r}[\varphi]-J_\mathit{l}[\varphi]}{J_\mathit{r}[\varphi]+J_\mathit{l}[\varphi]}=\frac{2}{L}%
668: \int_{-\frac{L}{2}}^{\frac{L}{2}}x\;\varphi(x)\;dx,
669: \end{equation}
670: when the currents (equations~\ref{eq:superposed-spec}) are plugged into equation~(\ref{eq:centroid}).
671: The discretized version of equation~(\ref{eq:centroid-spec}) is widely used in the field of nuclear
672: medical imaging to compute the planar interaction position of the impinging $\gamma$-ray within the
673: scintillation crystal. Axial symmetry of the distribution $\varphi(x)$ and the existence of exactly
674: one maximum is required for this method to work, since only for axial
675: symmetric distributions does the first
676: moment coincide with the unique maximum of the distribution. If this
677: condition is not fulfilled, systematic errors
678: in real $\gamma$-ray detector setups are introduced. Since the length of the wire is always limited to a finite
679: value, in the most normal case the distribution $\varphi(x)$ is truncated on its ends at the left and the
680: right. This will destroy the symmetry of the distribution whenever the position of its maximum is
681: different from the center of the wire. These and other errors of the
682: center of gravity algorithm will be discussed briefly in
683: section~\ref{ch:errors-of-cog-and-cdr}.
684:
685: \section{Anger's Approach}
686: \label{ch:anger-approach}
687:
688: \begin{figure}[!t]
689: \centering
690: \psfrag{Ja}{$J^u$}
691: \psfrag{Jb}{$J^d$}
692: \psfrag{J1}{$J_1$}
693: \psfrag{J2}{$J_2$}
694: \psfrag{Ji}{$J_i$}
695: \psfrag{Jn}{$J_n$}
696: \psfrag{Jn1}{$J_{n-1}$}
697: \psfrag{x1}{$x_1$}
698: \psfrag{x2}{$x_2$}
699: \psfrag{xi}{$x_i$}
700: \psfrag{xn1}{$x_{n-1}$}
701: \psfrag{xn}{$x_n$}
702: \psfrag{x}{$X$}
703: \psfrag{rin}{$R_\mathit{In}$}
704: \psfrag{ru1}{$R_1^u$}
705: \psfrag{rd1}{$R_1^d$}
706: \psfrag{rui}{$R_i^u$}
707: \psfrag{rdi}{$R_i^d$}
708: \psfrag{ru2}{$R_2^u$}
709: \psfrag{rd2}{$R_2^d$}
710: \psfrag{run}{$R_n^u$}
711: \psfrag{rdn}{$R_n^d$}
712: \psfrag{run1}{$R_{n-1}^u$}
713: \psfrag{rdn1}{$R_{n-1}^d$}
714: \psfrag{d}{$\Delta_x$}
715: \includegraphics[width=0.55\textwidth]{cdr/anger-1D}
716: \caption[Circuit diagram showing the 1D-version of the anger
717: positioning circuit]{Circuit diagram showing the 1D-version of the Anger
718: positioning circuit. The resistances $R_{In}$ represent the input
719: impedance of the following amplifier stage and are supposed to be
720: connected to virtual ground because these impedances should be zero.}
721: \label{fig:true-anger}
722: \end{figure}
723:
724: For the Anger-type gamma camera one necessarily needs a discrete
725: positioning logic instead of wires with a constant resistivity since
726: commercially available photodetectors in general come with a finite
727: number of detector segments. The adaptation to that case can be achieved
728: by connecting a set of resistors to each photomultiplier or each
729: anode-segment with their values adjusted in such a way that there is a
730: linear correspondence between the currents and the position of the
731: photomultiplier tube. Figure~\ref{fig:true-anger} shows such a
732: configuration for one
733: dimension. For each position $x_i$, the injected current $J_i$ is
734: divided into $J^u_i$ and $J^d_i$ with a ratio that corresponds
735: uniquely to that position by choosing the values of resistors
736: $R^u_i$ and $R^d_i$ in an adequate way. Since electrical currents
737: are subjected to the superposition principle, the currents $J^u_i$ and
738: $J^d_i$ are summed up on the upper and lower bus giving rise
739: to $J^u$ and $J^d$ respectively. The resistances $R_{In}$ represent the
740: input impedance of the attached amplifier stage, which have to be as low as possible for optimal linearity of
741: the positioning network. Rewriting equation~(\ref{eq:kirchoff-rules})
742: for this case, one obtains
743: \begin{equation}
744: \label{eq:1D-anger-logik}
745: \renewcommand{\arraystretch}{1.3}
746: \begin{array}{c}
747: J^u_i=\ds\frac{R^d_i}{R^d_i+R^u_i}J_i\\[0.9em]
748: J^d_i=\ds\frac{R^u_i}{R^d_i+R^u_i}J_i
749: \end{array}\Longrightarrow
750: \begin{array}{c}
751: J^u_i-J^d_i=\ds\frac{R^d_i-R^u_i}{R^d_i+R^u_i}J_i\\[0.9em]
752: J^u_i+J^d_i=J_i
753: \end{array}\Longrightarrow
754: \frac{J^u_i-J^d_i}{J^u_i+J^d_i}=\frac{R^d_i-R^u_i}{R^d_i+R^u_i}.
755: \end{equation}
756: In the special case that there is only one single position $x_i$, with $J_i\ne0$
757: and all other $J_{j\ne i}=0$, one has $J^u\equiv J^u_i$ and $J^d\equiv
758: J^d_i$. For the general case, the superposition
759: principle has to be applied and all currents $J^u_i$ and $J^d_i$ must
760: be summed up to obtain $J^u$ and $J^d$ respectively.
761: \begin{equation}
762: \label{eq:sum-currents}
763: \begin{array}{c}
764: J^u = \sum_i^N J^u_i = \sum_i^N \frac{R^d_i}{R^d_i+R^u_i}J_i \\[0.7em]
765: J^d = \sum_i^N J^d_i = \sum_i^N \frac{R^u_i}{R^d_i+R^u_i}J_i,
766: \end{array}\Longrightarrow
767: \begin{array}{c}
768: J^u-J^d = \sum_i^N \frac{R^d_i-R^u_i}{R^d_i+R^u_i}J_i \\[0.7em]
769: J^d+J^d = \sum_i^N J_i,
770: \end{array}\Longrightarrow
771: \frac{J^u-J^d}{J^u+J^d}=\frac{\sum_i^N \frac{R^d_i-R^u_i}{R^d_i+R^u_i}J_i}{\sum_i^N J_i}
772: \end{equation}
773: where $N$ is the total number of injection points $x_i$. If the input
774: impedance of the downstream current amplifier cannot be neglected, the
775: centroid would read
776: \begin{equation}
777: \label{eq:sum-currents-with-input-impedance}
778: \frac{J^u-J^d}{J^u+J^d}=\frac{\sum_i^N \frac{R^d_i-R^u_i}{R^d_i+R^u_i+2R_\mathit{In}}J_i}{\sum_i^N J_i}.
779: \end{equation}
780: In order to determine the required
781: resistor values $R^d_i$ and $R^u_i$ a current $J$ is considered that is
782: injected at all positions $x_i$, but only at one position at a time.
783: One can then compute the positions from the currents $J^u$ and $J^d$.
784: For any arbitrary linear position encoding one just need to choose the
785: resistor pairs to fulfill
786: \begin{equation}
787: \label{eq:1D-anger-logik-lin}
788: \frac{R^d_i-R^u_i}{R^d_i+R^u_i}\stackrel{!}{=}ax_i+b.
789: \end{equation}
790:
791: However, there are infinitely many possible functional dependences for
792: $R^u_i$ and $R^d_i$ which fulfill this requirement. $R^u_i+R^d_i=R_i$
793: is just the total resistance and must not be the same for all
794: positions. Therefore, any pair of resistor values of the form
795: \begin{equation}
796: \label{eq:general-res-encoding}
797: \renewcommand{\arraystretch}{1.3}
798: \begin{array}{c}
799: R^u_i=\frac{R_i}{2}(1-b-ax_i)\\[0.7em]
800: R^d_i=\frac{R_i}{2}(1+b+ax_i)
801: \end{array}
802: \end{equation}
803: fulfill equation~(\ref{eq:1D-anger-logik-lin}) and an
804: additional constraint for $R_i$ is necessary. Choosing a constant $R_i\equiv
805: R=\mathit{const.}$, yields
806: the trivial solution to (\ref{eq:1D-anger-logik-lin}). The parameter
807: $a$ is the scale and $b$ defines, which position $i$ within the resistor
808: array maps to $x=0$. A very common choice is $b=0, a=1$ for the center of
809: the resistor array and a unity scale. For these values, the right hand
810: side of equation~(\ref{eq:sum-currents}) transforms to the discrete
811: version for the computation of the centroid (equation~\ref{eq:func-centroid}).
812:
813: The impedance seen by the source of the current $J_i$ is given by the
814: parallel connexion of each resistor pair,
815: \begin{equation}
816: \label{eq:1D-anger-parallel}
817: R^u_i\parallel R^d_i=\frac{R^u_i R^d_i}{R^u_i+R^d_i}=\frac{R_i}{4}\left(1-(ax_i+b)^2\right).
818: \end{equation}
819: As discussed in paragraph \ref{subsec:pmt-as-ideal--current-source},
820: photomultipliers act like an ideal current source and the
821: impedance~(\ref{eq:1D-anger-parallel}) does not affect the working of
822: the charge dividing circuit. The only restriction is given by the fact
823: that the anode currents $J_i$ lift the anode potential by the amount
824: $J_i(R^u_i\parallel R^d_i)$ with respect to the last dynode. When this
825: voltage becomes too large, the PMT will stop working correctly.
826:
827: For situations where the impedance of the different inputs of the
828: charge divider is important, equation~(\ref{eq:1D-anger-parallel}) can be
829: used to adjust it to the required value for each $x_i$. Of special interest is the constraint
830: $R^u_i\parallel R^d_i\equiv R^p=\mathit{const.}$\,, {\em i.e.}\ an input
831: impedance that is equal for all positions $x_i$. One then obtains for
832: the resistor values $R^u_i$, $R^d_i$ and $R_i$:
833: \begin{equation}
834: \label{eq:const-para-encoding}
835: R^u_i=\ds\frac{2R^p}{1+b+ax_i}\mbox{,}\quad R^d_i=\ds\frac{2R^p}{1-b-ax_i}\quad\mbox{and }\quad
836: R_i=\ds\frac{4R^p}{1-(b+ax_i)^2},
837: \end{equation}
838: with $x_i,i\in\mathbb{N}$ equidistant points along the $x$-axis.
839:
840: \begin{figure}[!t]
841: \centering
842: \subfigure[][The two-dimensional version of the positioning logic as
843: proposed by Anger for the scintillation camera.]{\label{fig:2D-anger-logic}%
844: \psfrag{j}{$J_{ij}$}
845: \psfrag{ru}{$R^u_{ij}$}
846: \psfrag{rl}{$R^l_{ij}$}
847: \psfrag{rr}{$R^r_{ij}$}
848: \psfrag{rd}{$R^d_{ij}$}
849: \psfrag{jl}{$J^l$}
850: \psfrag{jr}{$J^r$}
851: \psfrag{ju}{$J^u$}
852: \psfrag{jd}{$J^d$}
853: \psfrag{jli}{$J^l_{ij}$}
854: \psfrag{jl1}{$J^l_{ij-1}$}
855: \psfrag{jl2}{$J^l_{ij+1}$}
856: \psfrag{jri}{$J^r_{ij}$}
857: \psfrag{jr1}{$J^r_{ij-1}$}
858: \psfrag{jr2}{$J^r_{ij+1}$}
859: \psfrag{jui}{$J^u_{ij}$}
860: \psfrag{ju1}{$J^u_{i-1j}$}
861: \psfrag{ju2}{$J^u_{i+1j}$}
862: \psfrag{jdi}{$J^d_{ij}$}
863: \psfrag{jd1}{$J^d_{i-1j}$}
864: \psfrag{jd2}{$J^d_{i+1j}$}
865: \includegraphics[height=0.41\textwidth]{cdr/2D-anger-node}}
866: \hspace*{0.03\textwidth}\subfigure[][General case of $k$
867: dimensions. For each dimension an additional pair of resistors is
868: required for every injection point
869: $(x_{i_1},x_{i_2},\ldots,x_{i_k})$.]{\label{fig:k-D-anger-logic}%
870: \psfrag{j}{$J_{i_1,i_2,\ldots,i_k}$}
871: \psfrag{j1n1}{$J_a^{n_1}$}
872: \psfrag{j2n1}{$J_b^{n_1}$}
873: \psfrag{j1n2}{$J_a^{n_2}$}
874: \psfrag{j2n2}{$J_b^{n_2}$}
875: \psfrag{j1nk}{$J_a^{n_k}$}
876: \psfrag{j2nk}{$J_b^{n_k}$}
877: \psfrag{r1n1}{$R_{i_1i_2\ldots i_k}^{a,n_1}$}
878: \psfrag{r2n1}{$R_{i_1i_2\ldots i_k}^{b,n_1}$}
879: \psfrag{r1n2}{$R_{i_1i_2\ldots i_k}^{a,n_2}$}
880: \psfrag{r2n2}{$R_{i_1i_2\ldots i_k}^{b,n_2}$}
881: \psfrag{r1nk}{\hspace*{0.7em}$R_{i_1i_2\ldots i_k}^{a,n_k}$}
882: \psfrag{r2nk}{\hspace*{0.7em}$R_{i_1i_2\ldots i_k}^{b,n_k}$}
883: \includegraphics[height=0.41\textwidth]{cdr/anger-multi-dim}}
884: \caption[Anger positioning logic for higher dimensions]{Anger positioning logic for higher dimensions.}
885: \label{fig:multi-dim-anger}%
886: \end{figure}
887:
888: The generalization to the higher dimensional case is straightforward,
889: however applications of the CoG using charge dividing circuits
890: normally do not exceed 3 dimensions. One only has to connect
891: one more charge divider with its corresponding busses for each
892: additional spatial dimension that has to be measured.
893: This is shown in
894: figure \ref{fig:2D-anger-logic} for the two-dimensional case. On
895: obtains for the currents to the
896: left, right, up and down at each injection point $(x_i,y_j)$
897: \begin{equation}
898: \label{eq:2D-anger-logic-elementar-current}
899: \renewcommand{\arraystretch}{2.4}
900: \begin{array}{cc}
901: J^l_{ij}=\ds\frac{R^r_{ij}R^u_{ij}R^d_{ij}}{R_i}J_{ij}\,, &
902: J^r_{ij}=\ds\frac{R^l_{ij}R^u_{ij}R^d_{ij}}{R_i}J_{ij} \\
903: J^u_{ij}=\ds\frac{R^r_{ij}R^l_{ij}R^d_{ij}}{R_i}J_{ij}\,, &
904: J^d_{ij}=\ds\frac{R^l_{ij}R^u_{ij}R^r_{ij}}{R_i}J_{ij}
905: \end{array}
906: \end{equation}
907: and their equivalent bus currents $J^l$, $J^r$, $J^u$ and $J^d$
908: are obtained by summation over both position indexes $i$ and $j$
909: \begin{equation}
910: \label{eq:2D-anger-logic-sum-current}
911: \renewcommand{\arraystretch}{2.4}
912: \begin{array}{cc}
913: J^l=\ds\sum_{ij}\ts\frac{R^r_{ij}R^u_{ij}R^d_{ij}}{R_i}J_{ij}\,, &
914: J^r=\ds\sum_{ij}\ts\frac{R^l_{ij}R^u_{ij}R^d_{ij}}{R_i}J_{ij} \\
915: J^u=\ds\sum_{ij}\ts\frac{R^r_{ij}R^l_{ij}R^d_{ij}}{R_i}J_{ij}\,, &
916: J^d=\ds\sum_{ij}\ts\frac{R^l_{ij}R^u_{ij}R^r_{ij}}{R_i}J_{ij}
917: \end{array}
918: \end{equation}
919: with
920: $R_i=(R^d_{ij}+R^u_{ij})R^l_{ij}R^r_{ij}+(R^l_{ij}+R^r_{ij})R^d_{ij}R^u_{ij}$.
921:
922: In analogy to equation~(\ref{eq:sum-currents}) one obtains
923: \begin{equation}
924: \label{eq:2D-anger-positions}
925: \begin{array}{ccc}
926: \ds\frac{J^u-J^d}{J^u+J^d}=\frac{\ds\sum_{ij}^{N_x,N_y}\ds\frac{R^d_{ij}-R^u_{ij}}{R^d_{ij}+R^u_{ij}}J_{ij}}{\ds\sum_{ij}^{N_x,N_y}J_{ij}} &\mbox{and}&
927: \ds\frac{J^r-J^l}{J^r+J^l}=\frac{\ds\sum_{ij}^{N_x,N_y}\ds\frac{R^r_{ij}-R^l_{ij}}{R^r_{ij}+R^l_{ij}}J_{ij}}{\ds\sum_{ij}^{N_x,N_y}J_{ij}}.
928: \end{array}
929: \end{equation}
930: The impedance of the resistor network at point $(x_i,y_i)$ is given by
931: the parallel connexion of the four resistors $R^l_i$, $R^r_i$, $R^u_i$
932: and $R^d_i$, which is just the inverse of the sum of their inverse
933: values. Once again, there is no unique solution for the required
934: linear encoding for the current differences in
935: (\ref{eq:2D-anger-positions}) and further constraints can be
936: implemented on the positioning circuit. Reasonable ones are the
937: requirement of a constant input impedance for all positions
938: $(x_i,y_i)$ or, as will be shown later, an input impedance that
939: depends quadratically on the positions. It can be easily shown that
940: if the resistor values fulfill
941: \begin{equation}
942: \label{eq:2D-anger-const-imp}
943: \begin{array}{cccc}
944: R^l_{ij} \propto \frac{1}{1-x_i}\,, & R^r_{ij} \propto
945: \frac{1}{1+x_i}\,, & R^u_{ij} \propto \frac{1}{1+y_j}\mbox{ and } & R^d_{ij} \propto \frac{1}{1-y_j}
946: \end{array}
947: \end{equation}
948: a constant input impedance will be achieved. Likewise, one obtains the above-mentioned
949: quadratic encoding of the impedance if the following
950: resistor-value position dependence instead of
951: (\ref{eq:2D-anger-const-imp}) are chosen.
952: \begin{equation}
953: \label{eq:2D-anger-quad-imp}
954: \begin{array}{cccc}
955: R^l_{ij} \propto \frac{4(x_i^2+y_j^2)}{1-x_i}\,, & R^r_{ij} \propto
956: \frac{4(x_i^2+y_j^2)}{1+x_i}\,, & R^u_{ij} \propto \frac{4(x_i^2+y_j^2)}{1+y_j}\mbox{ and } & R^d_{ij} \propto \frac{4(x_i^2+y_j^2)}{1-y_j}.
957: \end{array}
958: \end{equation}
959: Equivalent results can be obtained for the $k$-dimensional case shown
960: in figure \ref{fig:k-D-anger-logic}.
961:
962: \section{Proportional Resistor Chains}
963: \label{ch:prop-res-chains}
964:
965: The second possible configuration of charge dividing circuits is based on
966: position-sensitive RC-line readouts for proportional gas chambers
967: \mycite{Borkowski}{{\em et al.}\ }{1970}. It was developed to determine
968: the position of ionizing events with large volume proportional
969: detectors. A discretized version designed for use with multi-anode
970: position-sensitive photomultiplier tubes was presented by Siegel et
971: al.\ \cite{Siegel:1996}. This version needs significantly fewer
972: resistors than the original Anger logic and also reduces the wiring
973: effort, exposing a nearly equivalent positioning behavior. This is
974: clearly a great advantage with respect to its implementation.
975: As for the charge dividing circuit by Anger, it is
976: possible to extend the circuit to arbitrary many dimensions, as long
977: as the level of electronic noise remains sufficiently low.
978:
979: The continuous case has already been treated in
980: section~\ref{ch:charge-div-circuits}. For the discrete case the wire
981: is replaced by a chain of a finite number of resistors $R_1\cdots
982: R_{n+1}$. Consider such a network (figure~\ref{fig:prop-r-net}) for
983: a number \mbox{$n\in\mathbb{N}$} of detector outputs. The anodes are
984: numbered by \mbox{$i\in\,[\frac{1-n}{2};\frac{n-1}{2}]$} and in
985: unit steps. In this way, one only has to multiply the index $i$ by
986: the distance between the centers of two adjacent anode
987: segments $\Delta x$ to get the true position $x$.
988: If now a current
989: $J_i$ is injected at an arbitrary position $i$, it is grounded by the resistances,
990:
991: \begin{figure}[!t]
992: \centering
993: \psfrag{i}{$i$:}
994: \psfrag{dx}{$\Delta x$}
995: \psfrag{x}{$x$}
996: \psfrag{x=0}{$x=0$}
997: \psfrag{xi}{$x_i$}
998: \psfrag{r1}{$R_1$}
999: \psfrag{rnp}{$R_{n+1}$}
1000: \psfrag{Jr}{$J_\mathit{right}$}
1001: \psfrag{Jl}{$J_\mathit{left}$}
1002: \psfrag{Ji}{$J_i$}
1003: \psfrag{Ui}{$U_i$}
1004: \psfrag{1-np/2}{$\frac{1-n}{2}$}
1005: \psfrag{np-1/2}{$\frac{n-1}{2}$}
1006: \psfrag{3/2}{$\frac{3}{2}$}
1007: \psfrag{1/2}{$\frac{1}{2}$}
1008: \psfrag{-3/2}{$-\frac{3}{2}$}
1009: \psfrag{-1/2}{$-\frac{1}{2}$}
1010: \psfrag{3/2}{$\frac{3}{2}$}
1011: \psfrag{a1}{\hspace*{-2em}anode 1}
1012: \psfrag{anp}{anode $n$}
1013: \includegraphics[width=0.95\textwidth]{cdr/prop-r-net}
1014: \caption[Normally used one-dimensional DPC-circuit for current PSPMTs]{\label{fig:prop-r-net}
1015: Normally used DPC-circuit for current PSPMTs in one dimension, especially multi-wire type PSPMTs,
1016: with $n$ anode-wires. The injected currents $J_i$ will cause the voltages $U_i$ at the
1017: interconnection points.}
1018: \end{figure}
1019:
1020: \begin{gather}
1021: \label{eq:resistances_l}
1022: R_\mathit{l}(i)=\left(\frac{n+1}{2}+i\right)R_d=\left(\frac{R}{2R_d}+i\right)R_d\\
1023: \label{eq:resistances_r}
1024: R_\mathit{r}(i)=\left(\frac{n+1}{2}-i\right)R_d=\left(\frac{R}{2R_d}-i\right)R_d
1025: \end{gather}
1026: to the left (equation~\ref{eq:resistances_l}) and to the right
1027: (equation~\ref{eq:resistances_r}) from $i$ respectively.
1028: \mbox{$R=(n+1)R_d$} is the sum of all resistances used
1029: in the network. Plugging equations~(\ref{eq:resistances_l}) and
1030: (\ref{eq:resistances_r}) into the second theorem of Kirchoff
1031: \mbox{$R_\mathit{l}(i)J_\mathit{l}(i)=R_\mathit{r}(i)J_\mathit{r}(i)$}
1032: and applying its first theorem
1033: \mbox{$J(i)=J_\mathit{l}(i)+J_\mathit{r}(i)$},
1034: one obtains the following currents at the two ends of the chain
1035: \begin{gather}
1036: \label{eq:current-1}
1037: J_\mathit{l}(i)=\left(\frac{1}{2}-\frac{R_d}{R}i\right)J_i\quad\mbox{and} %
1038: \quad J_\mathit{r}(i)=\left(\frac{1}{2}+\frac{R_d}{R}i\right)J_i.
1039: \end{gather}
1040: They depend linearly on the injection position $i$. In the case of more than one injected
1041: current at two or more different positions, one obtains the resulting currents by superposition of all
1042: $n$ different currents $J_\mathit{l}(i)$ and $J_\mathit{r}(i)$:
1043: \begin{gather}
1044: \label{eq:current-2}
1045: J_\mathit{l}=\frac{1}{2}\sum_iJ_i-\frac{R_d}{R}\sum_iiJ_i\quad\mbox{and} %
1046: \quad J_\mathit{r}=\frac{1}{2}\sum_iJ_i+\frac{R_d}{R}\sum_iiJ_i.
1047: \end{gather}
1048: Equation~(\ref{eq:current-2}) directly leads to the relations normally used for the sum of
1049: currents and the centroid of the index:
1050: \begin{gather}
1051: \label{eq:anger_result}
1052: J=\sum_iJ_i=J_\mathit{l}+J_\mathit{r}\quad\mbox{and}\quad%
1053: \frac{J_\mathit{r}-J_\mathit{l}}{J_\mathit{r}+J_\mathit{l}}=\frac{2R_d}{R}\frac{\sum_iiJ_i}{\sum_iJ_i}.
1054: \end{gather}
1055: In order to get the centroid in the position space, one simply uses the
1056: fact that $x_i=i \Delta x$ and thus $i=x_i/\Delta x.$
1057: \begin{equation}
1058: \label{eq:1D-pos-centroid}
1059: \langle x\rangle = \frac{\sum_ix_iJ_i}{\sum_iJ_i} =
1060: \frac{R}{R_d}\frac{\Delta x}{2}\frac{J_\mathit{r}-J_\mathit{l}}{J_\mathit{r}+J_\mathit{l}}.
1061: \end{equation}
1062: The current $J_i$ at the point $x_i$ will see the impedance
1063: \begin{equation}
1064: \label{eq:dis-1D-r-chain-imp}
1065: R_l(i)\parallel R_r(i)=\frac{R_d}{n+1}\left(\frac{(n+1)^2}{4}-i^2\right).
1066: \end{equation}
1067:
1068: \subsection{2D Proportional Resistor Network}
1069: \label{subsection:2D-prop-net}
1070:
1071: \begin{figure}[!t]
1072: \centering
1073: \psfrag{Rd1}{${\Ss R_{v}}$}
1074: \psfrag{Rd2}{${\Ss R_{h}}$}
1075: \psfrag{Rv}{${R_{v}}$}
1076: \psfrag{Rh}{${R_{h}}$}
1077: \psfrag{Rh1}{${R_{h_1}}$}
1078: \psfrag{Rh2}{${R_{h_2}}$}
1079: \psfrag{Rhm}{${R_{h_m}}$}
1080: \psfrag{Rhm1}{${R_{h_{m-1}}}$}
1081: \psfrag{Ja}{${\ts J_A}$}
1082: \psfrag{Jb}{${\ts J_B}$}
1083: \psfrag{Jc}{${\ts J_C}$}
1084: \psfrag{Jd}{${\ts J_D}$}
1085: \psfrag{J11}{${J_{i_1j_1}}$}
1086: \psfrag{J12}{${J_{i_2j_1}}$}
1087: \psfrag{J21}{${J_{i_1j_2}}$}
1088: \psfrag{J22}{${J_{i_2j_2}}$}
1089: \psfrag{Jnm}{${J_{i_nj_m}}$}
1090: \psfrag{Jij}{${J_{ij}}$}
1091: \psfrag{Jl1}{${J^l_1}$}
1092: \psfrag{Jl2}{${J^l_2}$}
1093: \psfrag{Jlm}{${J^l_m}$}
1094: \psfrag{Jlm1}{${J^l_{m-1}}$}
1095: \psfrag{Jr1}{\hspace*{0.6em}${J^r_1}$}
1096: \psfrag{Jr2}{\hspace*{0.6em}${J^r_2}$}
1097: \psfrag{Jrm}{\hspace*{0.6em}${J^r_m}$}
1098: \psfrag{Jrm1}{\hspace*{0.3em}${J^r_{m-1}}$}
1099: \includegraphics[width=0.77\textwidth]{cdr/prop-r-net-2d}
1100: \caption[Two-dimensional proportional resistor network]{\label{fig:bidim-schem}%
1101: Two-dimensional proportional resistor network for a PSPMT
1102: with a $n\times m$ matrix of anode pads. The position indexes are
1103: given by $i_l=l-\frac{1+n}{2},\, l\in[1,2,\ldots,n]$ and
1104: $j_l=l-\frac{1+m}{2},\, l\in[1,2,\ldots,m]$.
1105: }
1106: \end{figure}
1107:
1108: It was shown that the generalization to higher dimensions in the case
1109: of Anger's solution is straightforward. For the proportional
1110: resistor chain approach however, one faces a situation where the
1111: electronic implementation of higher-dimensional detector-readouts is
1112: easy, but it is very difficult to give explicit expressions for the
1113: centroids as functions of the used resistor values and number of anode
1114: segments. To achieve positional sensitivity for two dimensions
1115: Borkowski {\em et al.}\ \cite{Borkowski:1970} proposed a circuit
1116: configuration that was later discretized by Siegel {\em et al.}\
1117: \cite{Siegel:1996} and which is shown in figure~\ref{fig:bidim-schem}. In this
1118: configuration, the currents from the different sources are injected
1119: into the interconnection points of $m$ (horizontal) 1D-proportional
1120: resistor-chains, where $m$ is the number of anode-segments along the $y$
1121: spatial direction. Equivalently, $n$ denotes the number of anode
1122: segments along the $x$ spatial direction. The currents along one
1123: horizontal resistor chain are divided and superposed according to
1124: equations~(\ref{eq:current-1}-\ref{eq:current-2}) of section
1125: \ref{ch:prop-res-chains}. Now one has $2m$ such currents $J^l_1,J^l_2,\ldots,
1126: J^l_m$ and $J^r_1,J^r_2,\ldots,J^r_m$ that are injected into the two vertical
1127: resistor chains where they are divided and superposed as in the 1D case.
1128: However, contrary to this case, the horizontal sum currents $J^l_1,J^l_2,\ldots,
1129: J^l_m$ and $J^r_1,J^r_2,\ldots,J^r_m$ do not see the same impedance to
1130: ground. According to (\ref{eq:dis-1D-r-chain-imp}), they will see approximately
1131: \begin{equation}
1132: \label{eq:lateral-impedance}
1133: R_\mathit{Imp}(j)\approx R_u(j)\parallel R_d(j)=\frac{R_v}{m+1}\left(\frac{(m+1)^2}{4}-j^2\right)
1134: \end{equation}
1135: instead, if $R_v\ll R_h$. Therefore, the
1136: equations~(\ref{eq:current-1}-\ref{eq:current-2}) have to be modified
1137: in order to correctly describe the actual charge division that takes into
1138: account the lateral impedance $R_u(j)\parallel R_d(j)$:
1139: \begin{gather}
1140: \label{eq:imp-corrected-horiz-div}
1141: J_\mathit{l}(i,j)=\left(\frac{1}{2}-\frac{iR_d}{(n-1)R_d+2R_\mathit{Imp}(j)}\right)J_i\,\mbox{ and }\,
1142: J_\mathit{r}(i,j)=\left(\frac{1}{2}+\frac{iR_d}{(n-1)R_d+2R_\mathit{Imp}(j)}\right)J_i.
1143: \end{gather}
1144: The $j$ dependence in equation~(\ref{eq:imp-corrected-horiz-div}) will
1145: propagate through the rest of the derivation for the centroids and
1146: finally results in a nonlinear positioning behavior. Fortunately, one
1147: can re-linearize the behavior of the network by varying the
1148: value of the lateral horizontal resistors
1149: $R_{h_1},R_{h_2},\ldots,R_{h_m}$ in figure~\ref{fig:bidim-schem}
1150: \mycite{Siegel}{{\em et al.}\ }{1996}. However, the fact that there are
1151: closed loops within this circuitry adds a new order of
1152: complexity to the problem and makes a general description for $N$
1153: dimension, by explicit expression for the centroids as a function of
1154: the number of anode-segments per dimension and the position index
1155: rather complicated. The solution can be found by network
1156: analysis using the {\em branch current method}. In this method, one
1157: establishes a set of equations that describes the relationship of the
1158: currents and voltages to each other through Kirchhoff's current law and
1159: Ohm's law. This set of algebraic equations can be solved, giving the
1160: exact currents and voltages at each node of the network.
1161:
1162: \begin{figure}[t]
1163: \centering
1164: \psfrag{Rv}{$R_v$}
1165: \psfrag{Rh}{$7R_h$}
1166: \psfrag{R1}{$R_{h_1}$}
1167: \psfrag{R2}{$R_{h_2}$}
1168: \psfrag{R3}{$R_{h_3}$}
1169: \psfrag{R4}{$R_{h_4}$}
1170: \psfrag{Rl}{$R_l$}
1171: \psfrag{Rr}{$R_r$}
1172: \psfrag{Ja}{${\ts J_A}$}
1173: \psfrag{Jb}{${\ts J_B}$}
1174: \psfrag{Jc}{${\ts J_C}$}
1175: \psfrag{Jd}{${\ts J_D}$}
1176: \psfrag{J}{${\ts J}$}
1177: \psfrag{Ui}{${\ts U}$}
1178: \psfrag{Ur1}{${\ts U^r_1}$}
1179: \psfrag{Ur2}{${\ts U^r_2}$}
1180: \psfrag{Ur3}{${\ts U^r_3}$}
1181: \psfrag{Ur4}{${\ts U^r_4}$}
1182: \psfrag{Ur5}{${\ts U^r_5}$}
1183: \psfrag{Ur6}{${\ts U^r_6}$}
1184: \psfrag{Ur7}{${\ts U^r_7}$}
1185: \psfrag{Ur8}{${\ts U^r_8}$}
1186: \psfrag{Ul1}{${\ts U^l_1}$}
1187: \psfrag{Ul2}{${\ts U^l_2}$}
1188: \psfrag{Ul3}{${\ts U^l_3}$}
1189: \psfrag{Ul4}{${\ts U^l_4}$}
1190: \psfrag{Ul5}{${\ts U^l_5}$}
1191: \psfrag{Ul6}{${\ts U^l_6}$}
1192: \psfrag{Ul7}{${\ts U^l_7}$}
1193: \psfrag{Ul8}{${\ts U^l_8}$}
1194: \includegraphics[width=0.82\textwidth]{cdr/lateral-r-compute}
1195: \caption[Electric circuit with naming convention for the computation
1196: of the lateral resistors]{Electric circuit with naming convention
1197: for the computation of the lateral resistors.}
1198: \label{fig:comp-lat-res}
1199: \end{figure}
1200:
1201: As a first step, one has to compute the lateral horizontal resistor values
1202: $R_{h_1},R_{h_2},\ldots,R_{h_m}$ in order to recover the linearity of
1203: the 2D-proportional positioning network. Since the position-sensitive
1204: multi-anode photomultiplier tube H8500 \cite{data:H8500} from Hamamatsu Photonics Co.\
1205: was used for the experimental
1206: validation of the findings in this section, it is done here for the spacial case of
1207: $n=m=8$ (refer to figure \ref{fig:comp-lat-res}).
1208: Consider the case where, of all 64 anode-segments, only one is active,
1209: and that the active is one of the first row in figure
1210: \ref{fig:bidim-schem} (or the first column in figure
1211: \ref{fig:comp-lat-res} respectively). Then the following set of equations is obtained:
1212: \newcommand{\setsolv}{\stackrel{!}{=}}
1213: \begin{equation}
1214: \label{eq:lateral-r-equations}
1215: \renewcommand{\arraystretch}{2.4}
1216: \begin{array}{ccc}
1217: \ds\frac{U-U^r_1}{R_r}+\frac{U-U^l_1}{R_l}&\setsolv&J \\
1218: \ds\frac{U^r_1}{R_v}+\frac{U^r_1-U^r_2}{R_v}-\frac{U-U^r_1}{R_r}&\setsolv&0 \\
1219: \ds\frac{U^l_1}{R_v}-\frac{U^l_1-U^l_2}{R_v}-\frac{U-U^l_1}{R_l}&\setsolv&0 \\
1220: \ds\frac{U^l_{j+1}-U^l_j}{R_v}+\frac{U^l_{j+1}-U^r_{j+1}}{7R_h+2R_{h_{j+1}}}+
1221: \frac{U^l_{j+1}-U^l_{j+2}}{R_v}&\setsolv&0 \\
1222: \ds\frac{U^r_{j+1}-U^r_j}{R_v}-\frac{U^l_{j+1}-U^r_{j+1}}{7R_h+2R_{h_{j+1}}}+
1223: \frac{U^r_{j+1}-U^r_{j+2}}{R_v}&\setsolv&0 \\
1224: \ds\frac{U^l_8-U^l_7}{R_v}+\frac{U^l_8-U^r_8}{7R_h+2R_{h_8}}+\frac{U^l_8}{R_v}&\setsolv&0 \\
1225: \ds\frac{U^r_8-U^r_7}{R_v}-\frac{U^l_8-U^r_8}{7R_h+2R_{h_8}}+\frac{U^r_8}{R_v}&\setsolv&0
1226: \end{array}
1227: \qquad\mbox{with}\qquad
1228: \begin{array}{c}
1229: j\in[1,2,\ldots,6]\\
1230: R_{h_8} = R_{h_1}\\
1231: R_{h_7} = R_{h_2}\\
1232: R_{h_6} = R_{h_3}\\
1233: R_{h_5} = R_{h_4}
1234: \end{array}
1235: \end{equation}
1236: The equations $R_{h_{9-j}} = R_{h_j},\,j\in[1,2,3,4]$ are valid
1237: due to the symmetry of the charge dividing circuit. If the
1238: system~(\ref{eq:lateral-r-equations}) is solved, the
1239: four currents $J_A$, $J_B$, $J_C$, and $J_D$ can be expressed as functions of the different
1240: resistor values and $J$. The centroids are computed from these
1241: currents using equations~(\ref{eq:anger_result}) that can be naturally
1242: adapted to the 2D case as follows:
1243: \begin{equation}
1244: \label{eq:bidim-cents}
1245: \langle
1246: j\rangle=c_j\frac{J_A+J_B-(J_C+J_D)}{J_A+J_B+J_C+J_D}\quad\mbox{and}\quad\langle i\rangle=c_i\frac{J_B+J_C-(J_A+J_D)}{J_A+J_B+J_C+J_D},
1247: \end{equation}
1248: where the constants $c_i$ and $c_j$ still have to be determined. If one
1249: further constrains this set of equations with $R_r=R_{h_1}$ and
1250: $R_l=7R_h+R_{h_1}$, one gets the constant value of 7/9 for $\langle
1251: j\rangle$ independent of any resistor value. For the
1252: expectation value $\langle i \rangle$, a quotient of
1253: polynomials in $R_v$, $R_{h_1}-R_{h_4}$ and $R_{1}$ is obtained. However, this
1254: value is supposed to give the same value as $\langle j\rangle$ owing
1255: to the symmetry of the network and the requirement that it behaves in
1256: the same manner for both spatial directions. Repeating this procedure
1257: three times with currents only at the positions $(i,j)=(5/2,5/2)$,
1258: $(i,j)=(3/2,3/2)$, and $(i,j)=(1/2,1/2)$, one gets a new set of 4
1259: equations, whose solution gives the values for the lateral horizontal
1260: resistors.
1261: \begin{equation}
1262: \label{eq:solution-laterals}
1263: \begin{array}{cc}
1264: R_{h_1}=R_{h_8}=R_h-4R_v, & R_{h_2}=R_{h_7}=R_h-7R_v\\
1265: R_{h_3}=R_{h_6}=R_h-9R_v, & R_{h_4}=R_{h_5}=R_h-10R_v
1266: \end{array}
1267: \end{equation}
1268: They can be parameterized in the following way:
1269: \begin{equation}
1270: \label{eq:solution-laterals-param-a}
1271: R^n_{h_l}=\frac{l}{2}\left(l-(n+1)\right)\,,\mbox{ with }l\in[1,2,\ldots,n]
1272: \end{equation}
1273: or
1274: \begin{equation}
1275: \label{eq:solution-laterals-param-b}
1276: R^n_{h_l}=\frac{1}{2}\left(l^2-\frac{(n+1)^2}{4}\right)\,,\mbox{ with }l\in[\frac{1-np}{2},\ldots,\frac{n-1}{2}].
1277: \end{equation}
1278: Equations~(\ref{eq:solution-laterals-param-a}) and
1279: (\ref{eq:solution-laterals-param-b}) have been verified for the cases
1280: $n=m$ and $n,m\in[2,4,6,8,10]$. Once the 2D-proportional
1281: resistor network has been linearized, one can determine the two proportionality constants
1282: $c_i$ and $c_j$. This can be done in exactly the same way as the
1283: linearization of the network, but replacing the resistances seen by
1284: the injected current to its left and right by
1285: $R_l=(7/2+i)R_h+R_{h_j}$ and $R_r=(7/2-i)R_h+R_{h_j}$ in
1286: equations~(\ref{eq:lateral-r-equations}) and using the
1287: results~(\ref{eq:solution-laterals}). In analogy to
1288: (\ref{eq:1D-pos-centroid}), for the expectation values
1289: of $x$ and $y$ one obtains
1290: \begin{equation}
1291: \label{eq:2D-centroids}
1292: \renewcommand{\arraystretch}{1.4}
1293: \begin{array}{cc}
1294: \ds\langle x\rangle =\Delta
1295: x\frac{9}{2}\frac{J_B+J_C-(J_A+J_D)}{J_A+J_B+J_C+J_D}, &
1296: \ds\langle y\rangle =\Delta
1297: y\frac{9}{2}\frac{J_A+J_B-(J_C+J_D)}{J_A+J_B+J_C+J_D}.
1298: \end{array}
1299: \end{equation}
1300: Note that the factors $9/2$ coincide with $(n+1)/2$ for
1301: the case $n=8$. However, the general validity of this expression for
1302: arbitrary $n$ has not been proved here.
1303:
1304: The set of equations~(\ref{eq:lateral-r-equations}) can equally be used to
1305: determine the impedance of the network of each of its 64 inputs. For
1306: this, one solves equations (\ref{eq:lateral-r-equations}) and its three equivalents when
1307: the current is injected in rows 2,3 and 4 for the voltage $U$ at the
1308: injection point. By virtue of Ohm's law and together with the injected
1309: current $J$, this is equivalent to the impedance. The other 4 rows are
1310: given implicitly due to the symmetry of the network.
1311:
1312: One obtains 4 equations for the impedances at the different
1313: interconnection points along the horizontal resistor chains which are
1314: given in Appendix \ref{app:exact_solutions}. An explicit
1315: parameterization in the position indexes $i$ and $j$ can be given for
1316: the case reported by \mycite{Siegel}{{\em et al.}\ }{1996} with $R_h=10
1317: R_v$:
1318: \begin{equation}
1319: \label{eq:exact-parametrization}
1320: R_\mathit{Imp}(i,j)=\frac{5}{18} \left(81-4 i^2\right) R_v+\left(a_2 j^2+a_0+i^2 \left(b_6 j^6+b_4 j^4+b_2 j^2+b_0\right)\right) R_v,
1321: \end{equation}
1322: with the parameter values
1323: \begin{equation}
1324: \label{eq:exact-parametrization-constants}
1325: \begin{array}{llllll}
1326: a_0=-\frac{63}{16}, & a_2=-\frac{7}{36}, & b_0\approx0.24, & b_2\approx0.015,& b_4\approx0.19\cdot10^{-3},& b_6\approx0.19\cdot10^{-5}.
1327: \end{array}
1328: \end{equation}
1329: Equation~(\ref{eq:exact-parametrization}) shows that the quadratic dependence
1330: of $R_\mathit{Imp}(i,j)$ is reproduced only for the $x$ spatial
1331: direction, {\em i.e.}\ $j=\mathit{const.}$, while the $y$ spatial
1332: direction includes $\mathcal{O}(j^4)$, manifesting in this way the
1333: antisymmetry of the network in both spatial directions. Note that the
1334: behavior is not due to the linearization of the position response by
1335: varying the lateral horizontal resistor values $R_{h_j}$. If all $R_{h_j}$
1336: are set to $R_h$, one will obtain the following parameter values
1337: \begin{equation}
1338: \label{eq:exact-parametrization-constants-b}
1339: \begin{array}{llllll}
1340: a_0=\frac{9}{8}, & a_2=-\frac{1}{18}, & b_0\approx0.048,& b_2\approx0.0021, & b_4\approx0.15\cdot10^{-4}, & b_6\approx0.47\cdot10^{-7},
1341: \end{array}
1342: \end{equation}
1343: that still contain higher than quadratic orders in $j$. Note that
1344: approximate values of the parameters $b_0,b_2,b_4$ and $b_6$ are given
1345: here for clearity although they can be computed exactly. They are
1346: given in Appendix~\ref{app:exact_solutions}.
1347:
1348:
1349: \section{Hybrid Solution}
1350:
1351: The third and last possible implementation of the CoG
1352: algorithm using charge divider circuits is the combination of both
1353: previous methods \mycite{Siegel}{{\em et al.}\ }{1996}. Clearly there can
1354: be no hybrid version for one dimension. Instead, the different
1355: previously mentioned approaches are applied to different spatial
1356: dimensions in the circuit. This is shown in
1357: figure~\ref{fig:hybrid-solution} for the 2D-case.
1358:
1359: \begin{figure}[t]
1360: \centering
1361: \psfrag{Rd1}{${\Ss R^d_1}$}
1362: \psfrag{Rd2}{${\Ss R^d_2}$}
1363: \psfrag{Ru1}{${\Ss R^u_1}$}
1364: \psfrag{Ru2}{${\Ss R^u_2}$}
1365: \psfrag{Rh}{${R_{h}}$}
1366: \psfrag{Rum}{${R^u_m}$}
1367: \psfrag{Rdm}{${R^d_m}$}
1368: \psfrag{Rum1}{${R^u_{m-1}}$}
1369: \psfrag{Rdm1}{${R^d_{m-1}}$}
1370: \psfrag{Ja}{${\ts J_A}$}
1371: \psfrag{Jb}{${\ts J_B}$}
1372: \psfrag{Jc}{${\ts J_C}$}
1373: \psfrag{Jd}{${\ts J_D}$}
1374: \psfrag{J11}{${J_{i_1j_1}}$}
1375: \psfrag{J12}{${J_{i_2j_1}}$}
1376: \psfrag{J21}{${J_{i_1j_2}}$}
1377: \psfrag{J22}{${J_{i_2j_2}}$}
1378: \psfrag{Jnm}{${J_{i_nj_m}}$}
1379: \psfrag{Jij}{${J_{ij}}$}
1380: \psfrag{Jl1}{${J^l_1}$}
1381: \psfrag{Jl2}{${J^l_2}$}
1382: \psfrag{Jlm}{${J^l_m}$}
1383: \psfrag{Jlm1}{${J^l_{m-1}}$}
1384: \psfrag{Jr1}{${J^r_1}$}
1385: \psfrag{Jr2}{${J^r_2}$}
1386: \psfrag{Jrm}{${J^r_m}$}
1387: \psfrag{Jrm1}{${J^r_{m-1}}$}
1388: \includegraphics[width=0.8\textwidth]{cdr/hybrid-solution}
1389: \caption[Circuit for a two-dimensional hybrid charge dividing readout]
1390: {Circuit for the 2D hybrid charge dividing readout proposed
1391: by \mycite{Siegel}{{\em et al.}\ }{1996}. It uses $m$ horizontal
1392: resistor chains for the $x$-centroid and Anger current scaling for
1393: the other spatial direction.}
1394: \label{fig:hybrid-solution}
1395: \end{figure}
1396:
1397:
1398: As discussed in section~\ref{ch:general-preamplifier-configuration},
1399: it is possible to read the currents $J_A$, $J_B$,
1400: $J_C$ and $J_D$ with shunt feedback transresistance amplifiers that
1401: have an input resistance $Z_\mathit{In}$ of virtually $\mathrm{0\,
1402: \Omega}$. In this case, the horizontal resistor chains in figure
1403: \ref{fig:hybrid-solution} decouple among each other and can be treated
1404: separately according to section~\ref{ch:prop-res-chains}. It was also
1405: shown that all resistors, including both at the chain
1406: ends, need to have the same value in order to obtain correct results
1407: from equations~(\ref{eq:anger_result}-\ref{eq:dis-1D-r-chain-imp}).
1408: This can be assured using equations~(\ref{eq:const-para-encoding}) of
1409: section~\ref{ch:anger-approach}, where the positioning
1410: properties for the vertical component in
1411: figure~\ref{fig:hybrid-solution} were also discussed.
1412: Choosing the following values for $R^u_j$ and $R^d_j$ of
1413: the $y$ spatial direction charge dividers:
1414: \begin{equation}
1415: \label{eq:hybrid-lateral-values}
1416: R^u_j = \frac{R_h}{\frac{1}{2}-\frac{j}{m+1}}\quad\mbox{and}\quad
1417: R^d_j=\frac{R_h}{\frac{1}{2}+\frac{j}{m+1}},
1418: \end{equation}
1419: the expressions for the centroids along the $x$- and $y$-spatial
1420: directions give
1421: \begin{equation}
1422: \label{eq:2D-hybrid-centroids}
1423: \renewcommand{\arraystretch}{1.4}
1424: \begin{array}{cc}
1425: \ds\langle x\rangle =\Delta
1426: x\frac{n+1}{2}\frac{J_B+J_C-J_A-J_D}{J_A+J_B+J_C+J_D}, &
1427: \ds\langle y\rangle =\Delta
1428: y\frac{m+1}{2}\frac{J_A+J_B-J_C-J_D}{J_A+J_B+J_C+J_D}.
1429: \end{array}
1430: \end{equation}
1431: Along the $x$-spatial direction, the impedance of the
1432: hybrid-circuit inputs is the same as for the 1D proportional resistor
1433: chain given by equation~(\ref{eq:dis-1D-r-chain-imp}). It does not
1434: depend, however, on the index $j$ for the $y$-spatial direction, since this was
1435: the constraint for finding
1436: expression~(\ref{eq:hybrid-lateral-values}).
1437: Thus one obtains for the input impedance of the hybrid solution
1438: \begin{equation}
1439: \label{eq:hybrid-impedance}
1440: R_\mathit{In}(i,j)=\frac{R_h}{n+1}\left(\frac{(n+1)^2}{4}-i^2\right).
1441: \end{equation}
1442:
1443: Siegel {\em et al.}\ state that the positioning performance of
1444: the three circuits is quite similar. Therefore, the natural
1445: consequence of choosing one of the three versions is to consider the
1446: complexity of implementation and their cost as the most important
1447: criteria. In this case, the clear favorite is the multidimensional
1448: version of the proportional resistor chain. First, it is the version
1449: which can be managed with fewest resistors; only $m(n+1)+2(m+1)$
1450: components are required. Second, there are no crossing wires in the
1451: circuit which eases the layout considerably and allows for a compact
1452: design. The most expensive solution is the proposal from Anger. It
1453: requires $4nm$ resistors and a more complicated layout. The hybrid
1454: solution gets by with $m(n+1)+4m$ resistors.
1455:
1456:
1457: \section{Simultaneous Measurement of the Second Moment}
1458: \label{sec:sim-measurement-of-sec-mom}
1459:
1460: After the discussion of how the centroids of a distribution can be
1461: measured with different charge dividing circuits, the
1462: emphasis is now put on the second moment. The motivation for measuring this
1463: moment is justified by the fact that the square root of the centered second moment
1464: is an excellent measure for the width of a distribution function.
1465: It was discussed in detail in section \ref{ch:light-distribution} that
1466: there is a strong correlation between the $\gamma$-ray's depth of
1467: interaction within thick crystals and the induced scintillation light
1468: distribution observed at the sensitive area of a photodetector.
1469: Using this effect for depth of interaction determination has been
1470: proposed at various times and by
1471: different researchers (refer to section \ref{ch:doi-detectors}).
1472: However, the problem consists of how to implement this measurement
1473: in the detector electronics, while simultaneously meeting all typical requirements
1474: such as fast measurement, low computational effort, low costs
1475: and sufficiently good measurement.
1476: Observing the impedances seen by the currents fed into the charge
1477: dividing circuits (equations~\ref{eq:1D-anger-parallel},
1478: \ref{eq:dis-1D-r-chain-imp} and \ref{eq:hybrid-impedance}) a
1479: possible solution becomes quite obvious. Since the impedances are
1480: quadratically encoded with the injection position and owing to Ohm's
1481: law one has $U=RJ$, one already disposes of the correct weighting for the sampled
1482: distribution and the only step that remains is to do the summation
1483: over these voltages. This can be done with a standard configuration of
1484: a summation amplifier explained in electronics textbooks and shown in figure
1485: \ref{subfig:simple-adder}.
1486:
1487: \begin{figure}[!t]
1488: \centering
1489: \subfigure[][Standard configuration of an analogue adder circuit.]{\label{subfig:simple-adder}
1490: \psfrag{R1}{$R_1$}
1491: \psfrag{Ri}{$R_i$}
1492: \psfrag{Rn}{$R_n$}
1493: \psfrag{Rc}{$R_c$}
1494: \psfrag{Rf}{$R_f$}
1495: \psfrag{Cc}{$C_c$}
1496: \psfrag{Cf}{$C_f$}
1497: \psfrag{U1}{$U_1$}
1498: \psfrag{Ui}{$U_i$}
1499: \psfrag{Un}{$U_n$}
1500: \psfrag{Ua}{$U_\mathit{Out}$}
1501: \psfrag{OP}{$\mathrm{OP}$}
1502: \includegraphics[width=0.3\textwidth]{cdr/adder}}
1503: \subfigure[][Configuration with voltage follower for a high input
1504: impedance of the individual branches.]{\label{subfig:adder-with-buffer}
1505: \psfrag{R1}{$R_1$}
1506: \psfrag{Ri}{$R_i$}
1507: \psfrag{Rn}{$R_n$}
1508: \psfrag{Rc}{$R_c$}
1509: \psfrag{Rf}{$R_f$}
1510: \psfrag{Cc}{$C_c$}
1511: \psfrag{Cf}{$C_f$}
1512: \psfrag{U1}{$U_1$}
1513: \psfrag{Ui}{$U_i$}
1514: \psfrag{Un}{$U_n$}
1515: \psfrag{Ua}{$U_\mathit{Out}$}
1516: \psfrag{OP}{$\mathrm{OP}$}
1517: \includegraphics[width=0.38\textwidth]{cdr/adder-with-buffer}}
1518: \subfigure[][Block-diagram of the combined summation and charge dividing
1519: circuit.]{\label{subfig:block-sum-cdr}
1520: \psfrag{JA}{$J_A$}
1521: \psfrag{JB}{$J_B$}
1522: \psfrag{JC}{$J_C$}
1523: \psfrag{JD}{$J_D$}
1524: \psfrag{S}{$\sum$}
1525: \psfrag{J1}{$J_{1,1}$}
1526: \psfrag{Jn}{$J_{n,m}$}
1527: \psfrag{cdr}{CDR}
1528: \psfrag{PSPMT}{PSPMT}
1529: \psfrag{sumAmp}{adder}
1530: \includegraphics[width=0.28\textwidth]{cdr/cdrSumBlock}}
1531: \caption[Inverting amplifier configuration for use as summation amplifier]{\label{fig:analogue-adder}
1532: Inverting amplifier configuration used as summation amplifier. The
1533: contributions of the voltages $U_i$ to the output voltages can be
1534: adjusted with the resistors $R_i$ and $R_f$. $C_f$ is required for
1535: phase correction and $C_c||R_c$ for offset correction.}
1536: \end{figure}
1537:
1538: The output voltage of this circuit is the sum of the input voltages
1539: $U_i$, weighted by $-R_f/R_i$:
1540: \begin{equation}
1541: \label{eq:adder-output-voltage}
1542: U_\mathit{Out}=-R_f\sum_{i=1}^n\frac{U_i}{R_i}.
1543: \end{equation}
1544: Theoretically there is no upper limit on the number of
1545: individual inputs $U_n$ when the electronic components behave as
1546: ideal ones. In practice, however, this limit is imposed by the thermal
1547: noise of the resistors $R_1,R_2,\ldots,R_n$ and the maximal output voltage
1548: swing of the operational amplifier used. The spectral density for the
1549: noise for each ohmic resistance $R_i$ in figure
1550: \ref{fig:analogue-adder} is given by
1551: $|U^\mathit{noise}_i(f)|^2=4k_BTR_i$, where $k_B$ is the Boltzmann constant
1552: and $T$ the absolute temperature.
1553: After integration over all frequencies, one obtains the measurable
1554: effective noise. For this, the parasitic capacitance $C^p_{R_i}$ of each
1555: resistance $R_i$ has to be taken into account which results in
1556: $|U^\mathit{noise}_i(f)|\rightarrow 0$ for $f\rightarrow\infty$,
1557: ensuring the convergence of the integral.
1558: \begin{equation}
1559: \label{eq:effective-adder-noise}
1560: U^\mathit{noise,eff}_i=\sqrt{\frac{k_BT}{C^p_{R_i}}}.
1561: \end{equation}
1562: Equation~(\ref{eq:effective-adder-noise}) holds for each resistance
1563: used in the configurations of figure \ref{fig:analogue-adder} and also
1564: the impedances (equations~\ref{eq:1D-anger-parallel},
1565: \ref{eq:dis-1D-r-chain-imp} and \ref{eq:hybrid-impedance}) of the
1566: centroid networks. For ambient temperature and a typical capacitance of
1567: $\mathrm{0.2 pF}$, equation~(\ref{eq:effective-adder-noise}) gives an effective
1568: noise of approximately $\mathrm{150\mu V}$, which is three orders of magnitude
1569: below the given signal level. A more important constraint is given by the
1570: maximal output voltage swing that can be delivered by the
1571: operational amplifier used. With an adequate choice of $R_f$, one has to
1572: ensure that $U_\mathit{Out}$ is always smaller than this device
1573: parameter, otherwise the operational amplifier saturates and distorts
1574: the signal.
1575:
1576: The impedance $R^\mathit{In}_i$ of the inputs of the summing amplifier
1577: is approximately given by their input resistance $R_i$. Since
1578: $R^\mathit{In}_i$ is connected in parallel to the impedances
1579: (\ref{eq:1D-anger-parallel}), (\ref{eq:dis-1D-r-chain-imp})
1580: and (\ref{eq:hybrid-impedance}) of the networks for the transverse
1581: positioning, the desired quadratic dependence gets distorted.
1582: In order to minimize this effect, one has to choose the individual
1583: $R^\mathit{In}_i$ very large compared to the impedances at the
1584: different inputs of the charge dividing circuits. Alternatively one
1585: can use the adder configuration of figure
1586: \ref{subfig:adder-with-buffer}. In this version, each input of the
1587: summation amplifier is buffered with a voltage follower described in
1588: section \ref{ch:general-preamplifier-configuration}. In this way, one is
1589: able to realize very high input impedances that are also independent of the weights
1590: of the adder. However, the large number of required operational
1591: amplifiers can pose thermal and power consumption problems on this solution.
1592:
1593: The resistance $R_c$ in figures \ref{fig:analogue-adder} is required
1594: for offset compensation (Tietze and Schenk \cite{Tietze}). Since the inverting
1595: terminal of the operational amplifier is connected to virtual ground
1596: when using the inverting configuration, the input current causes
1597: an offset voltage $U_\mathit{Off}=I_\mathit{In}R_f$ at the output of
1598: of the amplifier. This offset voltage can be compensated by a
1599: resistance $R_c$ of value
1600: \begin{equation}
1601: \label{eq:com-r-value}
1602: R_c=\left(\frac{1}{R_f}+\sum_{i=1}^n\frac{1}{R_i}\right)^{-1}
1603: \end{equation}
1604: and connected as displayed in figures~\ref{fig:analogue-adder}.
1605: In order not to introduce additional noise with this resistor,
1606: the capacity $C_c$ is connected in parallel to $R_c$, shortening by
1607: this means all signal components with non-vanishing frequencies. Similarly, $C_f$
1608: is not required for the correct working of the method of determining
1609: the second moments, but is needed for compensation of the non-ideal behavior of
1610: the amplifier configuration (Franco \cite{Franco}). It avoids
1611: oscillations of the operational amplifier caused by stray capacities
1612: at the inverting terminal.
1613:
1614:
1615: The summation amplifiers of figure~\ref{fig:analogue-adder} can be
1616: attached to any of the three discussed charge dividing circuits as
1617: shown in figure~\ref{subfig:block-sum-cdr}. One can use the values of
1618: the individual $R_i$ to adapt the summation amplifier to the
1619: particular requirements of each of the three charge dividing circuits.
1620: The value of $R_f$ can be used to adjust the amplification of the sum
1621: signal for avoiding saturations of the operational amplifier.
1622: Throughout the remaining part of this work, only implementations of
1623: the summation circuit shown in figure~\ref{subfig:simple-adder} will
1624: be considered. This configuration was preferred to
1625: the buffered version shown in figure~\ref{subfig:adder-with-buffer},
1626: which is complicated to realize in practice. The $n\times m$ voltage
1627: follower requires space and impedes the
1628: development of a simple circuit layout. On the other hand, the power
1629: consumption of each of the required $n\times m$ operational amplifiers
1630: imposes a serious design problem. A very low supply
1631: current of $\mathrm{15\,mA}$ already leads to an overall consumption of approximately
1632: $1 A$ required for the buffers when a position sensitive photodetector
1633: with $8\times8$ anode-segment is used. For this and other
1634: reasons, the solution proposed in figure \ref{subfig:adder-with-buffer}
1635: is more suitable for ASICs ({\em application-specific integrated
1636: circuits}).
1637:
1638: As an important disadvantage, one introduces a systematic
1639: error when computing the second moment using the circuit in figure
1640: \ref{subfig:simple-adder}, since the current coming from the
1641: anode-segment of the photodetector will see the input impedance of the
1642: summation amplifier in parallel to that of the charge dividing
1643: circuit. This leads not only to improper weights for the signal
1644: distribution but also extracts currents from the charge dividing
1645: circuit for the transverse positioning. In order to minimize this
1646: effect, the weighting resistors $R_i$ of the summation amplifier have
1647: to be chosen as large as possible. Here, the fabrication tolerance of
1648: 1\% of the standard SMD-resistors used can serve for a reasonable design
1649: criterion. If one chooses the weighting resistors $R_i$ at least 100
1650: times larger than the smallest input impedance of all the inputs of
1651: the charge dividing circuit, the errors introduced in the centroid and
1652: the second moment will be smaller than 1\% for a 1D resistor chain.
1653:
1654: \subsection{Anger Logic}
1655:
1656: In the case of the true Anger Logic, the application of the analogue
1657: summation circuit is straightforward. The specific design of this
1658: circuit, which uses an individual charge divider for each current from
1659: the anode-segments, produces a completely decoupled set of input
1660: voltages for the summation circuit since the output currents of the
1661: Anger logic are connected to virtual ground (refer to
1662: figure~\ref{fig:2D-anger-logic}). One now has two possibilities to
1663: obtain a quadratic weighting of the signal distribution. Choosing the
1664: resistor values for the charge divider according to
1665: equations~(\ref{eq:2D-anger-const-imp}), {\em i.e.}\ the same input impedance
1666: for all CDR
1667: inputs, implies the implementation of the quadratic dependence
1668: using the resistors $R_i$ of the summation circuit. By virtue of
1669: equation~(\ref{eq:adder-output-voltage}) and the constant input
1670: impedance
1671: $R_\mathit{In}^\mathit{CDR}$ of the Anger logic the relative values of the
1672: summation weights $R_{ij}$ are given by
1673: \begin{equation}
1674: \label{eq:anger-sum-const-input}
1675: R_{ij}\propto\frac{R_\mathit{In}^\mathit{CDR}R_f}{x_i^2+y_i^2}.
1676: \end{equation}
1677: Note that equation~(\ref{eq:anger-sum-const-input}) has a singularity
1678: at $(x_i,y_i) = (0,0)$, {\em i.e.}\ for this position one would obtain
1679: $R_{ij}=\infty$. In practice, this requirement can be implemented by
1680: omitting this resistor, that is to say, one does not include this value
1681: into the weighted sum in the same manner as this value is excluded
1682: in the mathematical definitions (\ref{eq:multivariate-moms-1})-(\ref{eq:multivariate-moms-3}) for
1683: $x_i = 0\,\forall\,i\in[1,n]$. In the majority of the cases,
1684: however, there is no reason for this consideration, since most
1685: position-sensitive photodetectors have an even number of detection
1686: segments. Also one subsequently has to invert the signal since
1687: summing amplifiers are realized by the inverting configuration of
1688: operational amplifiers. For the determination of the absolute values
1689: of the $R_{ij}$ one can use the above-mentioned design criterion
1690: of adjusting the minimal systematic error to the resistor tolerance.
1691:
1692: The second implementation configuration is given by equations
1693: (\ref{eq:2D-anger-quad-imp}). Here, the quadratic impedance variation
1694: seen by the anode-currents is inherently incorporated into the charge
1695: divider. One therefore has to give all summands the same weights with
1696: one single value $R_{ij}=R$. Once again the
1697: error-criterion for the determination of the absolute values can be used.
1698:
1699:
1700: \subsection{Proportional Resistor Chains}
1701: \label{sec:proportional-resistor-chains}
1702:
1703: In the case of the proportional resistor chains, the computation of
1704: the second moment using the described summation amplifier is not as
1705: trivial as in the case of the true Anger logic. The reason lies in the
1706: fact that all resistors are used to generate the weights for each
1707: position. Therefore, the charge dividing circuit does not decouple
1708: the voltages produced by the input currents over the impedances of the
1709: network from their neighborhood. Fortunately, this does not avert the
1710: successful computation of the second moment but makes the mathematical
1711: description more complicated. First the one-dimensional case is
1712: studied, since for this it is possible to find an explicit and
1713: exact expression for the summed voltage. Furthermore, it is of
1714: interest for the hybrid case. Based on this result, approximate
1715: solutions for the two-dimensional case can be found.
1716:
1717: \subsubsection{One-Dimensional Case}
1718: \label{subsection:dpc-with-2m-1D}
1719:
1720: \begin{figure}[!t]
1721: \centering
1722: \psfrag{i}{$i$:}
1723: \psfrag{dx}{$\Delta x$}
1724: \psfrag{x}{$x$}
1725: \psfrag{x=0}{$x=0$}
1726: \psfrag{xi}{$x_i$}
1727: \psfrag{r1}{$R_1$}
1728: \psfrag{rnp}{$R_{n+1}$}
1729: \psfrag{Jr}{$J_\mathit{right}$}
1730: \psfrag{Jl}{$J_\mathit{left}$}
1731: \psfrag{Ji}{$J_i$}
1732: \psfrag{Rd}{$R_d$}
1733: \psfrag{Ui}{$U_i$}
1734: \psfrag{1-np/2}{$\frac{1-n}{2}$}
1735: \psfrag{np-1/2}{$\frac{n-1}{2}$}
1736: \psfrag{3/2}{$\frac{3}{2}$}
1737: \psfrag{1/2}{$\frac{1}{2}$}
1738: \psfrag{-3/2}{$-\frac{3}{2}$}
1739: \psfrag{-1/2}{$-\frac{1}{2}$}
1740: \psfrag{3/2}{$\frac{3}{2}$}
1741: \psfrag{a1}{\hspace*{-2em}anode 1}
1742: \psfrag{anp}{anode $n$}
1743: \psfrag{Uij}{$U_{ij}$}
1744: \psfrag{Uii}{$U_{i,i}$}
1745: \includegraphics[width=0.85\textwidth]{cdr/prop-r-net-with-voltage}
1746: \caption[Schematic diagram of the resistor chain and naming conventions]{\label{fig:prop-r-net-with-voltage}
1747: Schematic diagram of the resistor chain and naming conventions for the
1748: computation of the sum of the voltage fraction induced by the injected
1749: current $J_i$.}
1750: \end{figure}
1751:
1752: First, the situation that induces a current $J_i$ injected
1753: at a single but arbitrary anode of index $i$ with
1754: \mbox{$i\in\,[\frac{1-n}{2};\frac{n-1}{2}]$} is considered. Owing to Ohm's law,
1755: the voltage $U_{i,i}$ seen at this same position is just the product of
1756: the current $J_i$ and the impedance to ground, {\em i.e.}\ $R_l(i)\parallel
1757: R_r(i)$ given in equation~(\ref{eq:dis-1D-r-chain-imp}):
1758: \begin{equation}
1759: \label{eq:ij-voltage}
1760: U_{i,i} =\frac{J_iR_h}{n+1}\left(\frac{(n+1)^2}{4}-i^2\right)=J_i\,R\,\left(\frac{1}{4}-\left(\frac{R_h}{R}\right)^2i^2\right),
1761: \end{equation}
1762: where $R=R_h(n+1)$ was used for the last step. As in the second case
1763: of the Anger logic, the required quadratic dependence is inherently incorporated
1764: into the charge divider. However, an attached summation amplifier
1765: would not give just the sum of all $U_{i,i}$ with
1766: \mbox{$i\in\,[\frac{1-n}{2};\frac{n-1}{2}]$} since the resistor
1767: chain in figure~\ref{fig:prop-r-net-with-voltage} would act as a voltage divider for
1768: the $U_{i,i}$. Therefore the connected analogue adder will see and add
1769: $U_{i,i}$ and all $U_{i,j}$ with $j\neq i;\,
1770: i,j\in\,[\frac{1-n}{2};\frac{n-1}{2}]$ that appear due to the
1771: current $J_i$ at all interconnection points $j\neq i$. Since all
1772: resistors of the chain have the same value, the voltage fractions
1773: $U_{j\leq i}$ and $U_{j>i}$ to the left and to the right of the
1774: injection position $i$ are given by
1775: \begin{equation}
1776: \label{eq:voltage-fracs}
1777: U_{j\leq i}=U_{i,i}\frac{j}{i}\qquad\mbox{and}\qquad U_{j>i}=U_{i,i}\frac{j}{n+1-i}.
1778: \end{equation}
1779: As a consequence of the superposition principle, the
1780: adder sums up all voltage fractions $U_{i,j}$ (including $U_{i,i}$) of
1781: all $J_i$. Additionally each summand will be amplified by the (same)
1782: weight $-R_f/R_s$. Therefore one has to sum over all $i$ and all $j$:
1783: \begin{equation}
1784: \label{eq:voltage-sum-all}
1785: U_\Sigma\approx-\frac{R_f}{R_s}\sum_{i,j}^{n} U_{i,j}.
1786: \end{equation}
1787: Equation~(\ref{eq:voltage-sum-all}) can only be an approximation
1788: since the finite impedance of the
1789: summing amplifier's inputs was not taken into account. However, this approximation is sufficiently
1790: good if the adder version of figure~\ref{subfig:adder-with-buffer}
1791: is used. If one considers for a moment the situation where the current
1792: $J_i$ is only injected at one position, then the sum over $i$ collapses. The
1793: remaining sum over $j$ reads
1794: \begin{equation}
1795: \label{eq:fraction-sum}
1796: U_i\approx-\frac{R_f}{R_s}\sum_jU_{i,j}=-\frac{R_f}{R_s} U_{i,i}\left(\frac{1}{i}\sum_{j=1}^{i}j+\frac{1}{n-i+1}%
1797: \sum_{j=1}^{n-i}j\right)=-U_{i,i}\frac{R_f}{R_s}\frac{n+1}{2} = -U_{i,i}\frac{R_fR}{2R_sR_h},
1798: \end{equation}
1799: where the sum of the arithmetic series was computed in the next to
1800: last step.
1801: Note that we did not impose any restriction on where the current is
1802: injected, so the contribution of this partial sum to the complete sum
1803: (\ref{eq:voltage-sum-all}) is the same constant factor for all
1804: injection points $i$. It is this particular detail that allows the
1805: computation of the second moment by attaching a summing amplifier to
1806: the proportional resistor chain and hybrid charge divider. Besides, it leads to an
1807: additional amplification of $(n+1)/2$ of the voltage signal $U_{i,i}$ since
1808: $n$ fractions are fed into the summing amplifier. For
1809: a signal composed of various $J_i$ with different $i$, the sum over $i$
1810: gives
1811: \begin{equation}
1812: \label{eq:voltage-sum-disc}
1813: U_\Sigma\approx\sum_iU_i=-\frac{R_fR^2}{2R_sR_h}\left(\frac{1}{4}\sum_iJ_i-\left(\frac{R_h}{R}\right)^2\sum_ii^2J_i\right).
1814: \end{equation}
1815: The second moment can be normalized by virtue of
1816: equations~(\ref{eq:anger_result}). However, one has to take into
1817: account that the electronic
1818: channel and amplifiers for the voltage sum $U_\Sigma$ and the currents
1819: $J_r$ and $J_l$ in general exhibit unequal amplifications $g_J$ and
1820: $g_\Sigma$ due to different design requirements. One finally gets
1821: \begin{equation}
1822: \label{eq:voltage-sum-disc-normal}
1823: \frac{U_\Sigma}{J}\approx\frac{|g_\Sigma|R_fR^2}{2|g_J|R_sR_h}\left(\frac{1}{4}-\left(\frac{R_h}{R}\right)^2\frac{\sum_ii^2J_i}{\sum_iJ_i}\right),
1824: \end{equation}
1825: which can be solved for the wanted second moment of the set of currents
1826: $J_i$:
1827: \begin{equation}
1828: \label{eq:1D-secmom-DPC}
1829: \frac{\sum_ii^2J_i}{\sum_iJ_i}\approx\frac{R^2}{4R_h^2}-\frac{2|g_J|R_s}{|g_\Sigma|R_fR_h}\,\frac{U_\Sigma}{J_r+J_l}
1830: \end{equation}
1831: and
1832: \begin{equation}
1833: \label{eq:1D-secmom-DPC-spatial}
1834: \frac{\sum_ix_i^2J_i}{\sum_iJ_i}\approx\Delta x^2\left(\frac{R^2}{4R_h^2}-\frac{2|g_J|R_s}{|g_\Sigma|R_fR_h}\,\frac{U_\Sigma}{J_r+J_l}\right)
1835: \end{equation}
1836: respectively.
1837:
1838: \begin{figure}[t]
1839: \centering
1840: \psfrag{Rh}{$R_h$}
1841: \psfrag{Rs}{$R_s$}
1842: \psfrag{Rf}{$R_f$}
1843: \psfrag{Rhl}{$R_h$}
1844: \psfrag{Ri1}{$R^\mathit{In}_1$}
1845: \psfrag{Ri2}{}
1846: \psfrag{Ri3}{$R^\mathit{In}_i$}
1847: \psfrag{Ri4}{}
1848: \psfrag{Ri5}{}
1849: \psfrag{Ri6}{}
1850: \psfrag{Ri7}{}
1851: \psfrag{Ri8}{$R^\mathit{In}_8$}
1852: \psfrag{OpAmp}{OpAmp}
1853: \includegraphics[width=0.85\textwidth]{cdr/r-line-with-adder}
1854: \caption[One dimensional proportional resistor chain for 8
1855: anodes]{One dimensional proportional resistor chain for 8
1856: anodes. For an ideal operational amplifier, signals that feed into
1857: its inverting terminal can be treated as if they were connected
1858: to ground (dashed line inside the amplifier). In this case one can
1859: compute the voltages at each connection of the resistor network.}
1860: \label{fig:r-chain-with-adder}
1861: \end{figure}
1862:
1863: For the simpler implementation version of the summing amplifier (refer
1864: to figure~\ref{subfig:simple-adder}), an exact solution for the
1865: voltage sum can be found, where {\em exact} means that all electronic
1866: components are considered as ideal. That is to say, the input
1867: impedance of each input is exactly given by $R_s$ (refer to
1868: figure~\ref{fig:r-chain-with-adder}). With the branch current method
1869: for analysis of the network (figure~\ref{fig:r-chain-with-adder}), an
1870: explicit solution for the input impedance at the connection $i$ can be
1871: found:
1872: \begin{equation}
1873: \label{eq:dis-1D-r-chain-imp-exact}
1874: R^\mathit{In}_i=\frac{R_d}{n+1}\left(\frac{(n+1)^2}{4}-i^2\right)\kappa_\tincaps{I}(i,\epsilon);
1875: \end{equation}
1876: \begin{equation}
1877: \label{eq:adder-imp-correction}
1878: \begin{split}
1879: \kappa_\tincaps{I}(i,\epsilon)&=a_{0}+a_{2}\;i^2+a_{4}\;i^4+a_{6}\;i^6;\\[1.4eX]
1880: a_0& =\frac{(\epsilon +2)
1881: (\epsilon (\epsilon +4) (\epsilon (\epsilon +4) (3637 \epsilon (\epsilon +4)+25504)+54528)+36864)}{8192 (\epsilon +1) (\epsilon +3)
1882: \left(\epsilon (\epsilon +3)^2+1\right) \left(\epsilon (\epsilon
1883: +3)^2+3\right)},\\[1.4eX]
1884: a_2&=\frac{\epsilon (\epsilon +2) (\epsilon +4) (\epsilon (\epsilon +4) (1731 \epsilon (\epsilon +4)+10304)+26880)}{71680
1885: (\epsilon +1) (\epsilon +3) \left(\epsilon (\epsilon
1886: +3)^2+1\right) \left(\epsilon (\epsilon +3)^2+3\right)},\\[1.4eX]
1887: a_4&= -\frac{\epsilon ^2 (\epsilon +2) (\epsilon +4)^2 (3
1888: \epsilon (\epsilon +4)-224)}{17920 (\epsilon +1) (\epsilon +3) \left(\epsilon (\epsilon +3)^2+1\right) \left(\epsilon (\epsilon
1889: +3)^2+3\right)},\\[1.4eX]
1890: a_6&= \frac{\epsilon ^3 (\epsilon +2) (\epsilon +4)^3}{4480 (\epsilon +1) (\epsilon +3) \left(\epsilon (\epsilon
1891: +3)^2+1\right) \left(\epsilon (\epsilon +3)^2+3\right)},
1892: \end{split}
1893: \end{equation}
1894: with $\epsilon=R_h/R_s$. As expected, $\kappa_\tincaps{I}(i,\epsilon)$ tends to
1895: unity for a vanishing $\epsilon$, and for $\epsilon\lesssim0.002$ one can
1896: make $\kappa_\tincaps{I}(i,\epsilon)\gtrsim99/100$. Similarly, an implicit and exact
1897: expression for the sum-voltage $U_i^\Sigma$ that is generated at the output of the
1898: operational amplifier by a single current $J_i$ at the position $i$
1899: can be computed.
1900: \begin{gather}
1901: \label{eq:1D-r-chain-with-adder-exact-sum}
1902: \frac{U_i^\Sigma}{J_i}=-\frac{R_fR_d}{2R_s}\left(\frac{(n+1)^2}{4}-i^2\right)\kappa_\tincaps{II}(i,\epsilon)\\[1.4eX]
1903: \kappa_\tincaps{II}(i,\epsilon)=\frac{\ds\frac{\left(64 i^6-48 i^4+6924 i^2+127295\right) \epsilon ^3}{1290240}+\ds\frac{\left(16 i^4+184
1904: i^2+3985\right) \epsilon ^2}{5760}+\ts\frac{1}{48}\ds\left(4 i^2+71\right) \epsilon +1}{(\epsilon +1)
1905: \left(\epsilon (\epsilon +3)^2+1\right)}
1906: \end{gather}
1907: Also $\kappa_\tincaps{II}(i,\epsilon)$ in
1908: equation~(\ref{eq:1D-r-chain-with-adder-exact-sum}) tends to unity if $\epsilon$
1909: vanishes. However, the errors in each input impedance get summed up
1910: and one now has to assure that $\epsilon\lesssim0.001$ if one requires a
1911: deviation smaller than 1\% from the ideal sum-voltage that would have
1912: been obtained with a summing amplifier with infinite input impedance.
1913: If a compound signal consisting of various $J_i$ at different
1914: positions is applied to the charge divider their contributions are
1915: superposed. Thus, the sum over $i$ has to be performed as follows:
1916: \begin{equation}
1917: \label{eq:1D-r-chain-with-adder-total-sum}
1918: U_\Sigma=-\frac{R_fR_d}{2R_s}\sum_i \left[J_i\left(\frac{(n+1)^2}{4}-i^2\right)\kappa_\tincaps{II}(i,\epsilon)\right].
1919: \end{equation}
1920:
1921: \subsubsection{Two-Dimensional Case}
1922: \label{subsec:prop-2D-case}
1923:
1924: For the two-dimensional case of the charge dividing circuit with
1925: proportional resistor chains it becomes rather difficult to find an
1926: explicit expression for the voltage sum. As for the one-dimensional
1927: case, a current $J_{ij}$ injected at an arbitrary input $(i,j)$ of
1928: the network causes not only a voltage $U_{ij}$\footnote{The notation for a `punctual'
1929: current $J_{ij}$ and of the voltage $U_{ij}$ for the two-dimensional cases
1930: of the charge divider are not to be confused with the voltages
1931: $U_{i,j}$ (equations~\ref{eq:ij-voltage}-\ref{eq:fraction-sum})
1932: that will arise at positions $j$ when the single current $J_i$ is
1933: injected into position $i$ of an one-dimensional resistor chain.}
1934: at this input but also
1935: well-defined fractions at all the other inputs since they are coupled
1936: among themselves through the weighting resistors for the first and
1937: second moments.
1938: This leads to a distortion of the desired quadratic
1939: weighting for the summing amplifier that is different for both spatial
1940: directions and that can be corrected only partially. An additional
1941: disadvantage is that the expressions that describe the dependency of
1942: the quantity of interest on the configuration of the charge divider
1943: are complex. Therefore it is a clear advantage of the true Anger logic that
1944: each input has their charge dividing resistors for its own and that an exact
1945: quadratic weighting can be implemented in this case.
1946:
1947: An important point is that the electronic configuration of the
1948: proportional resistor network breaks the symmetry with respect to
1949: rotations by 90\textdegree within the $x$-$y$-plane. It was shown
1950: in section~\ref{subsection:2D-prop-net} that one is
1951: able to correct this symmetry breaking for the determination of both
1952: centroids. The input impedance however, which is used for the
1953: determination of the second moment of the signal distribution, shows a
1954: rather different functional dependence along both spatial
1955: directions. Instead of the desired variation proportional to
1956: $(i^2+j^2)$, the position dependent impedance
1957: (\ref{eq:exact-parametrization}) was obtained. Unfortunately, one
1958: cannot adjust the different resistor values of the network in order to
1959: obtain the desired behavior as was done for the true anger
1960: logic, since this would affect the computation of the
1961: centroids. Clearly one also finds this unequal behavior in the
1962: voltage sum that is needed for the measurement of the
1963: second moment. Nevertheless, the summing amplifier itself gives us the
1964: possibility to nearly regain symmetric behavior of the voltage sum.
1965: If an analogue adder, as shown in figures~\ref{subfig:simple-adder}
1966: or \ref{subfig:adder-with-buffer}, is attached to the proportional
1967: resistor network of figure~\ref{fig:bidim-schem}, one can give each row
1968: of signals a different weight (see
1969: equation~\ref{eq:adder-output-voltage}). But, as shown in the previous section,
1970: along each row the weights of summing amplifier have to be the same
1971: since this is the spatial direction for which the 2D-proportional
1972: network intrinsically shows quadratic position encoding.
1973:
1974: First consider the bare 2D-proportional resistor network and suppose
1975: that an ideal summing amplifier is attached to it, {\em i.e.}\ with
1976: $R_{i,j}^\mathit{In}=\infty$. A network analysis now gives a coupled
1977: linear system of 76
1978: equations and 76 unknowns. A general solution depending on the
1979: different resistor values is very hard to find and the result would be
1980: probably much more complex than the exact solutions for the 1D-resistor
1981: chain (equations \ref{eq:dis-1D-r-chain-imp-exact} and
1982: \ref{eq:1D-r-chain-with-adder-exact-sum} in section
1983: \ref{subsection:dpc-with-2m-1D}). For this reason and within the scope
1984: of this work, only a solution for the specific case of the charge
1985: divider configuration reported in case B of Siegel {\em et al.}\
1986: \cite{Siegel:1996} is given; (that is: $8\times8$ anode-segments,
1987: $R_h=10 R_v$ and the corresponding values given in
1988: equation~\ref{eq:solution-laterals}).
1989: For other network configurations, a solution can be found in an
1990: analogous way.
1991:
1992: The voltage that can be observed at the
1993: output of the (ideal) summing amplifier of unity gain for all rows when the single current
1994: $J_{ij}$ is injected into the charge divider at the position $(i,j)$
1995: was found to be
1996: \begin{equation}
1997: \label{eq:sum-voltage-ideal-2D-DPC}
1998: U_{ij}^\Sigma=-5 \left(\frac{81}{4}-i^2\right) R_v J_{ij}.
1999: \end{equation}
2000: While the column index has the desired quadratic behavior, the row
2001: index $j$ does not even appear in equation
2002: (\ref{eq:sum-voltage-ideal-2D-DPC}). Furthermore, a detailed
2003: comparison between this result and equation
2004: (\ref{eq:voltage-sum-disc}) brings out that the only difference is a
2005: factor of ten. Figure \ref{fig:sumvoltage-native} gives a graphical
2006: representation of $U_{ij}^\Sigma/R_vJ_{ij}$. Once again one has to
2007: sum over $i$ and $j$ if instead of the single current a composite
2008: signal is applied.
2009:
2010: \begin{equation}
2011: \label{eq:SUM-voltage-ideal-2D-DPC}
2012: U^\Sigma=-5 R_v \sum_{i,j}\left(\frac{81}{4}-i^2\right) J_{ij}.
2013: \end{equation}
2014: \begin{figure}[t]
2015: \centering
2016: \psfrag{j}{\hspace*{2ex}$j$}
2017: \psfrag{i}{$i$}
2018: \psfrag{U}{\hspace*{-3ex}$\frac{U^\Sigma_{ij}}{J_{ij}}$}
2019: \includegraphics[width=0.47\textwidth]{cdr/sumvoltage-native}
2020: \caption[Dependence of the voltage sum on the position of the
2021: injected current]{Dependence of the voltage sum on the position of the
2022: injected current for the native 2D-proportional resistor
2023: network.}
2024: \label{fig:sumvoltage-native}
2025: \end{figure}
2026:
2027: Note that the last obtained result (\ref{eq:SUM-voltage-ideal-2D-DPC})
2028: already allows for sensible measurement of the signal distribution's
2029: second moment. This is a direct consequence of the rotational symmetry
2030: about an axis normal to the $x$-$y$-plane and through the
2031: photoconversion point of the signal distribution generated by the
2032: detected $\gamma$-photon which was commented on in section
2033: \ref{ch:light-distribution}. Owing to this symmetry, a supposed
2034: resistor network with attached unity-gain summing amplifier is able to
2035: successfully measure the second moment of the distribution since
2036: they are equal along both spatial directions. The motivation for
2037: searching for a symmetric equivalent to equation
2038: (\ref{eq:SUM-voltage-ideal-2D-DPC}) is founded on the possible
2039: maximization of the SNR of the measurement. An electronic
2040: implementation of a quadratic variation for the $j$ direction would
2041: lead to a highly increased signal while the statistical and
2042: electronic errors remain the same. Actually, one already performs the sum over
2043: all the 64 voltages, but because all rows have the same weight one will
2044: get no benefit from the sum over the index $j$. The summing amplifier
2045: itself allows us to implement different weights for each row of the
2046: network by varying the values of the input resistors $R_1,R_2,\ldots R_n$
2047: of the circuits shown in figures \ref{subfig:simple-adder} and
2048: \ref{subfig:adder-with-buffer}. As a criterion for the determination of
2049: these values, the symmetry behavior of the quadratic encoding of the
2050: anode-segments was used. For this, the voltage sum of the 64 inputs
2051: was computed as a function of the gains $g_1=-R_f/R_{S_1}$,
2052: $g_2=-R_f/R_{S_2}$, $g_3=-R_f/R_{S_3}$ and $g_4=-R_f/R_{S_4}$.
2053: Here, $R_{S_1},\ldots,R_{S_4}$ are the input resistances of the first
2054: and eighth row, the second and seventh row, the third and sixth row and
2055: the fourth and the fifth row. This choice already implements the
2056: desired symmetry with respect to the center of the network.
2057: For better clarity, the summed voltage at the different inputs of
2058: the enhanced charge divider network is shown in matrix form:
2059: \begin{equation}
2060: \label{eq:voltage-sum-withgains}
2061: U^\Sigma_{ij}=-JR_vG_{ij},
2062: \end{equation}
2063: where the gain matrix $G_{ij}$ is given by
2064: {\scriptsize
2065: \begin{equation}
2066: {\left(
2067: \begin{array}{c|c}
2068: \begin{array}{cccc}
2069: 28 g_1+4 \left(g_2+g_3+g_4\right) &
2070: 58 g_1+4 \left(g_2+g_3+g_4\right) &
2071: 78 g_1+4 \left(g_2+g_3+g_4\right) &
2072: 88 g_1+4 \left(g_2+g_3+g_4\right) \\
2073: 20 g_2+4 \left(g_1+2 g_3+2 g_4\right) &
2074: 50 g_2+4 \left(g_1+2 g_3+2 g_4\right) &
2075: 70 g_2+4 \left(g_1+2 g_3+2 g_4\right) &
2076: 80 g_2+4 \left(g_1+2 g_3+2 g_4\right) \\
2077: 16 g_3+4 \left(g_1+2 g_2+3 g_4\right) &
2078: 46 g_3+4 \left(g_1+2 g_2+3 g_4\right) &
2079: 66 g_3+4 \left(g_1+2 g_2+3 g_4\right) &
2080: 76 g_3+4 \left(g_1+2 g_2+3 g_4\right) \\
2081: 4 \left(g_1+2 g_2+3 g_3\right)+16 g_4 &
2082: 4 \left(g_1+2 g_2+3 g_3\right)+46 g_4 &
2083: 4 \left(g_1+2 g_2+3 g_3\right)+66 g_4 &
2084: 4 \left(g_1+2 g_2+3 g_3\right)+76 g_4
2085: \end{array} & \cdots \\&\\\hline \vdots & \ddots
2086: \end{array}\right).}\nonumber
2087: \end{equation}}
2088: \vspace*{-0.7eX}
2089: \begin{equation}\label{eq:voltage-sum-gain-matrix}\end{equation}
2090:
2091: The quadrants of the matrix which are not displayed have the same
2092: elements as the one shown but in a different order and according to the
2093: symmetry of the network. If one wants a second moment with same weighting
2094: behavior along both directions $i$ and $j$, one has to make the matrix
2095: $G_{ij}$ symmetric. Unfortunately, this is not
2096: possible and one have to settle for an approximate symmetry between
2097: the $i$ and the $j$ index. As possible Ansatz for finding an almost symmetric
2098: weighting matrix, one can require that at least one row-column pair of
2099: (\ref{eq:voltage-sum-gain-matrix}) has the desired symmetry. This
2100: gives a linear system of 4 equations which can be solved exactly and
2101: as a function of $g_4$. One needs at least one degree of freedom left in
2102: order to adjust the minimum input impedance of all 64 inputs to a
2103: value that ensures the correct working of the centroid
2104: determination.
2105:
2106: Since one is free in the choice of the column-row pair, there are 4
2107: such sets of equations and for all exist unique but suboptimal
2108: solutions that differ from each other. To find an optimum
2109: solution to the posed problem, one can measure the asymmetry
2110: $\mathcal{S}[\mathbf{G}]$ of the matrix as follows:
2111: \begin{equation}
2112: \label{eq:matrix-unsymmetry}
2113: \mathcal{S}[\mathbf{G}]=\sqrt{\sum_{i,j}\left[\frac{G_{ij}-G_{ji}}{G_{ij}}\right]^2}.
2114: \end{equation}
2115: The following values,
2116: \begin{equation}
2117: \label{eq:opt-gain-vals}
2118: g_1 = 0.26972 g_4\mbox{,}\quad g_2=0.57368 g_4,\quad\mbox{and}\quad g_3=0.84035 g_4,
2119: \end{equation}
2120: will minimize expression (\ref{eq:matrix-unsymmetry}) and are the desired
2121: optimum gains for the different resistor rows. Since the gain for
2122: each row of summing amplifier inputs is $g_i=-R_f/R_{S_i}$,
2123: the corresponding resistor values are given by
2124: \begin{equation}
2125: \label{eq:opt-r-vals}
2126: R_{S_1},R_{S_8}=3.7075 R_{S_4}\mbox{,}\quad R_{S_2},R_{S_7}=1.74314
2127: R_{S_4}\mbox{,}\quad R_{S_3},R_{S_6}= 1.18998 R_{S_4}\quad\mbox{and}\quad
2128: R_{S_5}=R_{S_4}
2129: \end{equation}
2130: for different weights of the input rows.
2131: With this change in the summing amplifier, the voltage representing
2132: the second moment of the signal distribution now becomes
2133: \begin{equation}
2134: \label{eq:optimize-sum-voltage}
2135: U^\Sigma=-5 R_v R_f\sum_{i,j}\left(\frac{81}{4}-i^2\right) \frac{J_{ij}}{R_{S_j}}.
2136: \end{equation}
2137: A plot of the 64 values is shown in figure \ref{fig:voltage-sums-sym}
2138: to the left. In the same figure but to the right, the
2139: relative deviation between the approximate equation
2140: (\ref{eq:optimize-sum-voltage}) and the results obtained with network
2141: simulations is shown. These discrepancies are clearly due to the fact
2142: that the current that is extracted by the analogue adder at each of
2143: the 64 inputs distorts the ideal weighting as derived in section
2144: \ref{subsection:dpc-with-2m-1D}. Furthermore, the network simulator
2145: treats all elemental devices of the circuit as real as possible, while
2146: for the derivation of equation~(\ref{eq:optimize-sum-voltage})
2147: ideal components have been supposed.
2148:
2149: \begin{figure}[t]
2150: \subfigure[][Voltage sum for the 2D-proportional resistor
2151: network with maximized symmetry.]{\label{subfig:sumvoltage-symmetrized}
2152: \psfrag{j}{\hspace*{2ex}$j$}
2153: \psfrag{i}{$i$}
2154: \psfrag{U}{\hspace*{-4ex}$\frac{U^\Sigma_{ij}}{R_vJ_{ij}}$}
2155: \includegraphics[width=0.47\textwidth]{cdr/sumvoltage-symmetrized}
2156: }\hspace*{0.5em}
2157: \subfigure[][Relative deviation in \% of the computed voltage sum from the
2158: results obtained by simulation.]{\label{subfig:sumvoltage-zymmetrized-errors}
2159: \psfrag{j}{\hspace*{2ex}$j$}
2160: \psfrag{i}{$i$}
2161: \psfrag{E}{\%}
2162: \includegraphics[width=0.47\textwidth]{cdr/sumvoltage-symmetrized-errors}
2163: }
2164: \caption[Summed voltages and deviation from simulations for the
2165: optimized case]{Summed voltages and deviation from simulations for
2166: the optimized case.}
2167: \label{fig:voltage-sums-sym}
2168: \end{figure}
2169:
2170:
2171: \subsection{Hybrid Solution}
2172:
2173: In this last case of possible implementations of charge division
2174: circuits with second moment capability, one faces a similar situation as
2175: just discussed. One once again has a circuit that is not symmetric with
2176: respect to swapping the coordinates $x$ and $y$ and once again this
2177: symmetry-breaking is not reflected in the measurement of the centroids
2178: but can be observed very well in the measurement of the second
2179: moment. It is clear from equation
2180: (\ref{eq:1D-r-chain-with-adder-total-sum}) and the fact that the four
2181: outputs $J_A,\ldots,J_D$ are connected through the downstream amplification
2182: stage (see section \ref{ch:general-preamplifier-configuration}) to
2183: virtual ground, that the voltage sum for the hybrid circuit has to be
2184: \begin{equation}
2185: \label{eq:volt-sum-2D-hybrid}
2186: U_\Sigma=-\frac{R_d}{2}\sum_{i,j} \left[g_jJ_{ij}\left(\frac{(n+1)^2}{4}-i^2\right)\kappa_\tincaps{II}(i,\epsilon)\right].
2187: \end{equation}
2188: If one supposes a summing amplifier of the second type (figure
2189: \ref{subfig:adder-with-buffer}), one can set
2190: $\kappa_\tincaps{II}(i,\epsilon) = 1$. Alternatively one can equate
2191: $\kappa_\tincaps{II}(i,\epsilon)\simeq1$ by choosing the input
2192: resistors of the summing amplifier sufficiently
2193: large. $\kappa_\tincaps{II}(i,\epsilon)$ can then be neglected for the
2194: symmetrization of equation~(\ref{eq:volt-sum-2D-hybrid}).
2195: This substantially eases the symmetrization
2196: of equation~(\ref{eq:volt-sum-2D-hybrid}). Since the rows of the hybrid
2197: circuit (figure~\ref{fig:hybrid-solution}) decouple from each other,
2198: for even $n$ one only has to set
2199: \begin{equation}
2200: \label{eq:hybrid-row-gain}
2201: g_j=\frac{4}{n
2202: \left(n+2\right)}\left(\frac{(n+1)^2}{4}-j^2\right),
2203: \end{equation}
2204: where the factor $4/(n(n+2))$ has been derived from the requirement
2205: of unity gain for both central rows, {\em e.g.}\
2206: $g_{n/2}=g_{n/2+1}=1$. For the case of odd $n$, equation
2207: (\ref{eq:hybrid-row-gain}) has to be adjusted according to this
2208: situation. With the aid of the feedback resistor $R_f$ of the summing
2209: amplifier, the global gain $G=R_f/R_{S_{n/2}}$ is introduced and the
2210: symmetrized version of the sum voltage (\ref{eq:volt-sum-2D-hybrid}) reads
2211: \begin{equation}
2212: \label{eq:volt-sum-2D-hybrid-final}
2213: U_\Sigma=-\frac{2 R_f
2214: R_d}{R_{S_{n/2}}n\left(n+2\right)}\sum_{i,j}
2215: \left[J_{ij}\left(\frac{(n+1)^2}{4}-i^2\right)\left(\frac{(n+1)^2}{4}-j^2\right)\right].
2216: \end{equation}
2217:
2218: \begin{figure}[t]
2219: \centering
2220: \subfigure[][Voltage encoding for an $8\times8$ array of anode
2221: segments when using the hybrid solution.]{\label{subfig:hybrid-sum-voltage}
2222: \psfrag{j}{\hspace*{2ex}$j$}
2223: \psfrag{i}{$i$}
2224: \psfrag{U}{\hspace*{-5ex}$\frac{U^\Sigma_{ij}}{R_d G J_{ij}}$}
2225: \includegraphics[width=0.47\textwidth]{cdr/hybrid-sumvoltage}}
2226: \subfigure[][Relative difference in \% between theoretic model (equation \ref{eq:volt-sum-2D-hybrid-final})
2227: and simulation results for $n=8$.]{\label{subfig:hybrid-sum-voltage-errors}
2228: \psfrag{j}{\hspace*{2ex}$j$}
2229: \psfrag{i}{$i$}
2230: \psfrag{E}{\%}
2231: \includegraphics[width=0.47\textwidth]{cdr/hybrid-sumvoltage-errors}}
2232: \caption{Corrected sum voltage for the hybrid charge divider configuration.}
2233: \end{figure}
2234: The voltage values are plotted in figure
2235: \ref{subfig:hybrid-sum-voltage} for the case of $n=8$. As for the
2236: other results, expression~(\ref{eq:volt-sum-2D-hybrid-final}) for the
2237: voltage sum of the hybrid solution was verified with {\sc Spice} simulations
2238: ({\em Simulation Program with Integrated Circuits Emphasis}, see for
2239: instance Tietze and Schenk \cite{Tietze}). The small differences between the model and the
2240: simulations are shown in figure
2241: \ref{subfig:hybrid-sum-voltage-errors}. They are mainly due to the
2242: approximation made with $\kappa(i,\epsilon)=1$, that is, by the finite
2243: input impedances of the summing amplifier.
2244:
2245: \section{Anode Inhomogeneity Compensation}
2246:
2247: One point that has not been considered so far are the differences in the
2248: response of the different anode-segments. For real devices one clearly
2249: cannot expect all the detection segments to behave in exactly the same
2250: manner. The deviation in the total sensitivity $S_\mathit{tot}$,
2251: that defines the ratio of the anode current and the incident light
2252: flux can alter significantly from one segment to another, especially
2253: for large size position sensitive PMTs. $S_\mathit{tot}$ is generally
2254: an accumulation of various
2255: effects like the photocathode sensitivity $S_\tincaps{PC}$, the
2256: quantum efficiency $\mathit{QE}_\tincaps{PC}$, the photoelectron
2257: collection efficiency $\eta$, the gain $G_\tincaps{Dyn}$ of the
2258: electron multiplier system and the anode sensitivity
2259: $S_\tincaps{Anode}$ \cite{Flyckt:2002}. The variation over the
2260: detector's spatial extension of each of these factors is finally reflected
2261: in the signal distribution at the output of the detector. Due to the
2262: large number of contributions involved, the inhomogeneity of the
2263: sensitivity is mostly determined experimentally and provided by the
2264: manufacturer in the form of an anode uniformity map for each single device.
2265:
2266: As shown in chapter~\ref{ch:experiment}, the normalized nontrivial
2267: moments that are intended to be measured with the presented
2268: enhanced charge division circuits are nearly unaffected by these
2269: variations. This however does not apply to the 0th moment which
2270: represents the energy. In favor of a high intrinsic energy resolution,
2271: one is tempted to correct for this detector inhomogeneity. In systems
2272: that digitize all channels of the used photodetector independently, the
2273: way to go would be a software correction of these aberrations from
2274: their optimal behavior. In the present case, many computation steps are
2275: performed analogically and the detector-caused inhomogeneity of the
2276: signal results in a systematic error of these computations.
2277: The problem can be solved satisfyingly with additional active
2278: electronic components or, in the case of the Anger logic, with a
2279: passive compensation of the anode inhomogeneity.
2280:
2281: \subsection{Passive Compensation}
2282:
2283: The true Anger logic described in section \ref{ch:anger-approach} has
2284: the already mentioned advantage that each anode-segment has its own
2285: independent charge divider. Therefore it is possible to connect
2286: a (different) correction shunt $R^c_{ij}$ as shown in figure
2287: \ref{subfig:2D-anger-logic-with-comp} at each input of the Anger logic
2288: \mycite{Tornai}{{\em et al.}\ }{1996}. The value of $R^c_{ij}$ that is
2289: required for the compensation of the anode inhomogeneity can be easily
2290: computed from the manufacturer's anode uniformity map and the input
2291: impedance of the Anger logic at the respective position according to
2292: \begin{equation}
2293: \label{eq:input-imp-anger-logic}
2294: R_\mathit{In}=\left(\frac{1}{R_{ij}^u}+\frac{1}{R_{ij}^l}+\frac{1}{R_{ij}^d}+\frac{1}{R_{ij}^r}\right)^{-1}.
2295: \end{equation}
2296:
2297: \begin{figure}[t]
2298: \centering
2299: \begin{minipage}[c]{0.52\textwidth}
2300: \centering
2301: \subfigure[][Passive compensation method proposed by
2302: Tornai {\em et al.}\ \cite{Tornai:1996} for use with the Anger
2303: positioning logic.]{\label{subfig:2D-anger-logic-with-comp}%
2304: \psfrag{j}{$J_{ij}$}\psfrag{ru}{$R^u_{ij}$}\psfrag{rl}{$R^l_{ij}$}
2305: \psfrag{rr}{$R^r_{ij}$}\psfrag{rd}{$R^d_{ij}$}\psfrag{rc}{$R^c_{ij}$}
2306: \psfrag{jl}{$J^l$}\psfrag{jr}{$J^r$}\psfrag{ju}{$J^u$}\psfrag{jd}{$J^d$}
2307: \psfrag{jli}{$J^l_{ij}$}\psfrag{jl1}{$J^l_{ij-1}$}\psfrag{jl2}{$J^l_{ij+1}$}
2308: \psfrag{jri}{$J^r_{ij}$}\psfrag{jr1}{$J^r_{ij-1}$}\psfrag{jr2}{$J^r_{ij+1}$}
2309: \psfrag{jui}{$J^u_{ij}$}\psfrag{ju1}{$J^u_{i-1j}$}\psfrag{ju2}{$J^u_{i+1j}$}
2310: \psfrag{jdi}{$J^d_{ij}$}\psfrag{jd1}{$J^d_{i-1j}$}\psfrag{jd2}{$J^d_{i+1j}$}
2311: \includegraphics[width=0.94\textwidth]{cdr/2D-anger-node-with-compensation}}
2312: \end{minipage}\hfill
2313: \begin{minipage}[c]{0.48\textwidth}
2314: \centering
2315: \subfigure[][Simplest non-trivial case of proportional charge
2316: dividing resistor chains with intended passive compensation.]{\label{subfig:1D-r-chain-with-comp}%
2317: \psfrag{R1}{$R_1$}\psfrag{R2}{$R_2$}\psfrag{R3}{$R_3$}
2318: \psfrag{Rs1}{$R_{S_1}$}\psfrag{Rs2}{$R_{S_2}$}\psfrag{aJ}{$\alpha J$}
2319: \psfrag{bJ}{$\beta J$}\psfrag{(1-a)bJ}{\hspace*{-0.6ex}$(1\!-\!\alpha)\,\beta J$}
2320: \psfrag{(1-b)aJ}{\hspace*{-1ex}$(1\!-\!\beta)\;\alpha J$}
2321: \includegraphics[width=0.94\textwidth]{cdr/CorrectPSPMTbyDPC}}
2322: \subfigure[][Special case $0<\beta<\alpha=1$,
2323: $R_{S_1}=R_S$ and $R_{S_2}=\infty$.]{\label{subfig:1D-r-chain-with-comp-beta}%
2324: \psfrag{R1}{$R_1$}\psfrag{R2}{$R_2$}\psfrag{R3}{$R_3$}\psfrag{Jl}{$J_l$}
2325: \psfrag{Rs}{$R_{S}$}\psfrag{J}{$J$}\psfrag{bJ}{$\beta J$}
2326: \psfrag{(1-b)J}{\hspace*{-1ex}$(1-\beta)J$}\psfrag{Jr}{$J_r$}
2327: \includegraphics[width=0.94\textwidth]{cdr/CorrectPSPMTbyDPCbeta}}
2328: \end{minipage}
2329: \caption[Electronic configurations for passive anode inhomogeneity
2330: compensation]
2331: {\label{fig:multi-dim-anger-compensated}%
2332: Electronic configurations for passive compensation of the spatial
2333: sensitivity inhomogeneity of photodetectors.}
2334: \end{figure}
2335:
2336: For the proportional resistor chain and its derived positioning
2337: circuits, this compensation method cannot be used. This can be
2338: demonstrated for the simple case of a 1D-resistor chain of two inputs
2339: shown in figure \ref{subfig:1D-r-chain-with-comp}. Suppose that the
2340: two anode-segments whose signals $\alpha J$ and $\beta J$ are fed into
2341: this charge divider are of different strength, {\em e.g.}\ $\alpha\neq\beta$,
2342: and are wanted to be equalized. Without loss of generality one can
2343: further assume that $0<\beta<\alpha=1$. The passive compensation works
2344: by reducing all signals of the photodetector to the strength of the
2345: lowest one. Therefore one encounters the situation shown in figure
2346: \ref{subfig:1D-r-chain-with-comp-beta}. A first condition on the
2347: values of the resistors $R_1,R_2,R_3$ and $R_S$ is given by the fact
2348: that the computation of the first moment requires linear weights for
2349: the currents $J$ and $\beta J$. Therefore, the condition
2350: \begin{equation}
2351: \label{eq:1-codition-linear-compensated-DPC}
2352: R_2=R_3=\frac{R_1 R_S}{R_1+R_S}
2353: \end{equation}
2354: has to be fulfilled.
2355: In order to equalize both currents $J$ and $\beta J$, the fraction
2356: $(1-\beta)J$ has to be bled to ground by the resistor $R_S$ and one
2357: obtains (neglecting the trivial solution of $R_S=R_1=R_2=R_3=0$)
2358: \begin{equation}
2359: \label{eq:2-codition-linear-compensated-DPC}
2360: R_S=\frac{R_1}{3}\left(\frac{1+\beta}{1-\beta}\right).
2361: \end{equation}
2362: With this configuration, the circuit behaves as desired for the
2363: current $J$ fed into the left input of the circuit. Unfortunately,
2364: the current $\beta J$ will also be divided at the left node and
2365: the computation of the centroid gives
2366: \begin{equation}
2367: \label{eq:erroneous-centroid}
2368: \langle x\rangle\big|_{\beta\neq1} = \frac{J_r-J_l}{J_r+J_l}=\frac{5}{3}+\frac{8}{3
2369: (\beta -3)}\neq \langle x\rangle\big|_{\beta\to1}=\frac{1}{3}.
2370: \end{equation}
2371: This leads to unacceptable systematic errors in the centroid even for
2372: $\beta$ near to unity. As a consequence, one obtains for a composed
2373: signal a positioning error, which is comparable in its size to the
2374: mispositioning, that would exist if the network were not compensated
2375: at all.
2376:
2377:
2378: \subsection{Active Compensation}
2379:
2380: A second obvious possibility which works with each of the described
2381: enhanced charge dividers is to use an independent current amplifier for
2382: each of the inputs of the positioning circuit.
2383: Figure~\ref{fig:active-comp} shows this for a 1D-resistor chain. One
2384: only has to adjust the gain of each single amplifier corresponding to
2385: the anode uniformity map. Besides the universality of this circuit,
2386: another important advantage is that it does not bleed parts of the
2387: signal to ground but amplifies it. For the correct working of this
2388: charge divider circuit, the fixed gain amplifiers (FGAs) have to deliver
2389: currents at their outputs. Furthermore, a very small input impedance
2390: is needed, since photodetectors for photon-counting applications
2391: normally act as current sources. The main drawback of this method is
2392: its complicated implementation owing to power consumption and required
2393: space on the printed circuit board.
2394:
2395:
2396: \begin{figure}[!t]
2397: \centering
2398: \psfrag{vga}{{\small FGA}}
2399: \psfrag{Jr}{$J_\mathit{r}$}\psfrag{Jl}{$J_\mathit{l}$}
2400: \psfrag{Ji}{$J_i$}\psfrag{a1}{anode 1}\psfrag{anp}{anode $n$}
2401: \psfrag{rnp}{}\psfrag{r1}{}
2402: \includegraphics[width=0.95\textwidth]{cdr/active-compensation}
2403: \caption[Active anode inhomogeneity compensation using current amplifiers]{\label{fig:active-comp}
2404: Anode inhomogeneity compensation using current amplifiers. Each of
2405: these amplifiers will have a different amplification according to
2406: the anode-segment that it has to correct. It is important that
2407: the input impedance of the FGA is very low. Otherwise the
2408: currents $J_i$ will generate voltages $J_i
2409: R_\mathit{In}^\tincaps{FGA}$ that are opposite in sign to the bias
2410: voltages of the photodetector and lead to a decrease or even an
2411: complete breakdown of electron collection or multiplication.}
2412: \end{figure}
2413:
2414:
2415:
2416: \section{Errors of The Center of Gravity Algorithm}
2417: \label{ch:errors-of-cog-and-cdr}
2418:
2419: At the end of the present section, the systematic and statistic errors
2420: that are introduced by the center of gravity algorithm as well as
2421: the imperfection of the photodetectors and readout electronics are
2422: discussed. Important sources for systematic errors are discretization effects
2423: and border effects. The first are a consequence of the sampling of an
2424: arbitrary signal distribution that enforcedly includes its
2425: discretization. Almost always, this destroys existing symmetries of
2426: the distribution. Border effects arise because it is impossible to realize
2427: detectors of infinite spatial extension and lead also to breaking of
2428: the symmetries of the signal distribution. On the other hand, one faces
2429: signal fluctuations that are caused by the quantization of the signal
2430: and a quantum detection efficiency which in no real case can reach 100\%.
2431: The thermal noise of the electronic components falls far short of the
2432: other mentioned error sources and it will be shown that one must not
2433: worry too much about this error.
2434:
2435: \subsection{Signal Fluctuations}
2436:
2437: Fluctuations in the signal distribution are caused by various
2438: underlying processes. The signal generation starts with the
2439: photoconversion of the $\gamma$-ray into one or more electrons within
2440: the scintillation crystal. Depending on the underlying process,
2441: {\em e.g.}\ the Compton effect, photoelectric effect or pair production, one
2442: or more primary electrons are generated. These primary electrons
2443: produce the scintillation light via the decay of the secondarily ionized
2444: atoms. Together with these random processes, inhomogeneities of the
2445: material and non-proportional energy dependency lead to important
2446: fluctuations in the total light output of the scintillator. While this
2447: affects all anode-segments by the same amount, the distribution of the
2448: finite number of these scintillation photons over the sensitive area of the
2449: detector leads to fluctuation that differs for all
2450: anode-segments. In the following step, the visible
2451: light photons will be converted independently and one by one into
2452: photoelectrons at the cathode of the detector. This is a Poisson
2453: process and probably the most important cause for segment-dependent signal
2454: fluctuations for photomultiplier tubes. Finally, the fluctuation will
2455: lead to an uncertainty $\delta_\tincaps{P}\mu_1$ in the centroid
2456: measurement $\mu_1$, which
2457: can be estimated using error propagation. For this, a set
2458: $\bar{f}^{(\!x_o\!)}=\{f^{(\!x_o\!)}_j\},\,j=[1,2,\ldots,n]$ of $n\in\mathbb{N}$ values that
2459: represent the signals of the anode-segments with center-positions
2460: $x_j$ is considered.
2461: The superscript $(x_0)$ indicates that the signals are generated by
2462: the photoconversion of a $\gamma$-photon at position $x_0$.
2463: If the errors of the anode positions are assumed to be negligible, one
2464: obtains for the error $\delta_\tincaps{P}\mu_1(\bar{f}^{(\!x_o\!)})$ in the centroid
2465: $\mu_1(\bar{f}^{(\!x_o\!)})$ the expression
2466: \begin{equation}
2467: \label{eq:stat-centroide-error}
2468: \delta_\tincaps{P}\mu_1(\bar{f}^{(\!x_o\!)})=\sqrt{\sum_j\left[\frac{\partial}{\partial
2469: f_j}\left(\frac{\sum_i x_i f_i^{(\!x_o\!)}}{\sum_i f_i^{(\!x_o\!)}}\right)\delta
2470: f_j^{(\!x_o\!)}\right]^2}=\frac{1}{\sqrt{\sum_if_i^{(\!x_o\!)}}}\sqrt{\frac{\sum_j
2471: (x_j-\mu_1(\bar{f}^{(\!x_o\!)}))^2\delta f_j^2}{\sum_i f_i^{(\!x_o\!)}}},
2472: \end{equation}
2473: where the centroid is given by
2474: \begin{equation}
2475: \label{eq:centroide-reminder}
2476: \mu_1(\bar{f}^{(\!x_o\!)})=\frac{\sum_i x_if_i^{(\!x_o\!)}}{\sum_i f_i^{(\!x_o\!)}}.
2477: \end{equation}
2478: Setting $\delta f_j\propto\sqrt{f_j}$ for the fluctuations of Poisson
2479: dominated process, equation~(\ref{eq:stat-centroide-error})
2480: reduces to
2481: \begin{equation}
2482: \label{eq:stat-centroide-error-final}
2483: \delta_\tincaps{P}\mu_1(\bar{f}^{(\!x_o\!)})\propto\frac{\sigma_\mathit{\mbox{\tiny SD}}}{\sqrt{\sum_i f_i^{(\!x_o\!)}}},
2484: \end{equation}
2485: where $\sigma_\mathit{\mbox{\tiny SD}}$ denotes the standard deviation
2486: for the set $\bar{f}^{(\!x_o\!)}$.
2487: This result is not unexpected. Since the standard deviation gives an
2488: idea of the dispersion of the set of variables $\{f_j^{(\!x_o\!)}\}$,
2489: equation~(\ref{eq:stat-centroide-error-final}) states that the error
2490: of the
2491: centroid is smaller the narrower the signal distribution is.
2492: Furthermore, the result also scales with the square root of the
2493: sum of all signals $\{f_j^{(\!x_o\!)}\}$, {\em e.g.}\ the total amount of photons collected
2494: by all the anode-segments. One therefore expects a better position estimate
2495: from the center of gravity algorithm the more light is released by
2496: the $\gamma$-ray and the narrower the distribution of this light is.
2497:
2498: As already mentioned, any measurement of the (characteristic) parameter
2499: of the distribution observed with the photodetector has to be sampled
2500: and therefore will be only available in the form of the finite set of
2501: values $\{f_j^{(\!x_o\!)}\}$. Normally, this sampling is done by integrating the
2502: distribution of interest piecewise over many small and equidistant
2503: intervals. Mathematically, this amounts to a convolution \mycite{Landi}{}{2002}
2504: of the real distribution $\varphi(x)$ with the interval function\footnote{The
2505: interval function is defined through $\Pi(x)=\left\{
2506: \begin{array}{ccc}
2507: 1 & & |x|< 1/2 \\ 1/2 & \mbox{for} & |x|=1/2 \\ 0 & & |x|> 1/2
2508: \end{array}\right.$.} $\Pi(x)$
2509: of width $\tau$:
2510: \begin{equation}
2511: \label{eq:sampling-convolution}
2512: f_{x_o}(x)=\int_{-\infty}^{\infty}\Pi\left(\frac{x-x'}{\tau}\right)\varphi(x'-x_0)dx'.
2513: \end{equation}
2514: The set of measured values $\{f_j^{(\!x_o\!)}\}$ is just given by the functional value of
2515: $f_{x_o}(x)$ at the positions $x=x_j$:
2516: \begin{equation}
2517: \label{eq:sampling-values}
2518: f_j^{(\!x_o\!)}=f_{x_o}(x)\big|_{x_j}.
2519: \end{equation}
2520: It is clear from equation~(\ref{eq:sampling-convolution}),
2521: that the resulting distribution $f_{x_o}(x)$ has to be wider than
2522: the original distribution $\varphi(x)$, since one always has $\tau>0$
2523: for any real photodetector. This means together with
2524: equation~(\ref{eq:stat-centroide-error-final}),
2525: that when the CoG algorithm is used together with a continuous
2526: light distribution, the pixel size of the detector used is of great
2527: importance. As an example, the projection of the normalized
2528: one-dimensional inverse square law onto the abscissa $x$,
2529: \begin{equation}
2530: \label{eq:test-func-for-stat-err-cog}
2531: \varphi(x)=\frac{d J}{\pi(x^2+d^2)},
2532: \end{equation}
2533: is used to study the influence of the detector pixel size $\tau$ on
2534: the width of the sampled distribution $f_{x_o}(x)$. For the
2535: computation one can set $x_0=0$ w.l.o.g. While the undisturbed
2536: distribution~(\ref{eq:test-func-for-stat-err-cog}) has a FWHM equal to
2537: $2d$, the sampled distribution
2538: $f_{x_o}(x)$ will have the width $\sqrt{\tau^2+4d^2}$. The results are
2539: plotted in figure~\ref{fig:sampling-broadening} for two different
2540: depths $d$ and different values of $\tau$. Obviously, this result is
2541: not directly applicable to
2542: equation~(\ref{eq:stat-centroide-error-final}) since there the standard
2543: deviation was computed. Nevertheless, the qualitative result that
2544: one gets a larger statistical error of the CoG for a
2545: coarser sampling of the distribution still holds since the standard
2546: deviations as well as the FWHMs of the distributions $f_{x_o}(x)$ and
2547: $\varphi(x)$ are strictly increasing functions.
2548:
2549: The statistical error made in the second moment is not related to the
2550: center of gravity algorithm but will be given for completeness. It can be computed
2551: equivalently to (\ref{eq:stat-centroide-error}) and yields
2552: \begin{equation}
2553: \label{eq:stat-secmom-error}
2554: \delta_\tincaps{P}\mu_2(\bar{f}^{(\!x_o\!)})=\sqrt{\sum_j\left[\frac{\partial}{\partial
2555: f_j}\left(\frac{\sum_i x^2_i f_i^{(\!x_o\!)}}{\sum_i f_i^{(\!x_o\!)}}\right)\delta
2556: f_j^{(\!x_o\!)}\right]^2}\,\,\stackrel{\delta f_j=\sqrt{f_j}}{=}\,\,\frac{1}{\sqrt{\sum_i
2557: f_i^{(\!x_o\!)}}}\sqrt{\sum_j\left[\left(x^2_j-\frac{\sum_ix_i^2f_i^{(\!x_o\!)}}{\sum_if_i^{(\!x_o\!)}}\right)^2 f_j^{(\!x_o\!)}\right]}.
2558: \end{equation}
2559: Note that $\sum_i(x_i^2f_i^{(\!x_o\!)})/(\sum_if_i^{(\!x_o\!)})$ is just the central second
2560: moment for $x_0=0$.
2561:
2562: \begin{figure}[t]
2563: \centering
2564: \subfigure[][Interaction depth
2565: $2\,\mathrm{mm}$.]{\label{subfig:sampling-broadening-2mm}
2566: \psfrag{J}{}
2567: \psfrag{x}{$x$}
2568: \includegraphics[width=0.45\textwidth]{cdr/rect-convolved-func-1}}\hspace*{0.05\textwidth}
2569: \subfigure[][Interaction depth $6\,\mathrm{mm}$.]{\label{subfig:sampling-broadening-6mm}
2570: \psfrag{J}{}
2571: \psfrag{x}{$x$}
2572: \includegraphics[width=0.45\textwidth]{cdr/rect-convolved-func-2}}
2573: \caption[Broadening of the signal distribution owing to the pixel size]{Broadening of the original distribution (solid line)
2574: introduced by a sampling with an anode-segment size of $3\,\mathrm{mm}$
2575: (dashed line) and $6\,\mathrm{mm}$ (dot-dashed line) for 2 different
2576: interaction depths. All graphs are normalized to
2577: $f_{x_o}(x)\big|_{x=0}=1$ for easier comparison of
2578: the width.}
2579: \label{fig:sampling-broadening}
2580: \end{figure}
2581:
2582:
2583: \subsection{Discretization Errors}
2584:
2585: The necessary discretization of the signal distribution leads
2586: to additional systematic errors \mycite{Landi}{}{2002}.
2587: The definition of the moments for a distribution $\varphi(x)$ was given by (see
2588: equation~(\ref{eq:func-moments}) in section \ref{ch:stat-estimates})
2589: \begin{equation}
2590: \label{eq:gen-moment-once-again}
2591: \mu_k(x)=\int_{\omega}x^k\varphi(x)dx.
2592: \end{equation}
2593: At this level and for normalized $\varphi(x)$, the method gives a
2594: perfect position measurement for $k=1$ and all other moments for
2595: $k\neq1, k\in\mathbb{N}$ whenever the integration interval $\omega$
2596: covers the support of the distribution. Unfortunately, $\varphi(x)$ is not
2597: accessible to experiments, $\varphi(x)$ is not normalized, nor is it
2598: possible to build detectors of infinite spatial extension.
2599: Each real detector performs a discrete reduction that essentially modifies the
2600: properties of the method. In its simplest form,
2601: the sampling is a set of finite disjoint integrations of the
2602: distribution. Therefore the detector readout is the set of the
2603: following $n$ numbers
2604: \begin{equation}
2605: \label{eq:sampling-values-alt}
2606: f^{(\!x_o\!)}_j=\int_{\tau(n-\half\!)}^{\tau(n+\half\!)}\!\!\!\varphi(x-x_0)\,dx,
2607: \end{equation}
2608: which is just another definition of $\{f^{(\!x_o\!)}_j\}$ equivalent
2609: to the one given with equation~(\ref{eq:sampling-convolution}). Except for the case of
2610: $\tau\to0$ and $n\to\infty$, the centroid~(\ref{eq:centroide-reminder})
2611: differs from $x_0$ almost everywhere by a systematic error due to the
2612: discretization. This error was studied in detail by G.\ Landi
2613: \cite{Landi:2002}. It was found to be
2614: \begin{equation}
2615: \label{eq:discretization-cent-error}
2616: \mu_1(\bar{f}^{(\!x_o\!)})=x_0+\frac{\tau}{\pi}\sum_{k=1}^\infty\frac{(-1)^k}{k}
2617: \sin{(2\pi kx_0/\tau)}\Phi(2\pi k/\tau),
2618: \end{equation}
2619: where $\Phi(\omega)$ is the Fourier transform of $\varphi(x)$.
2620: The $x_0$ dependency of the discretization error has the form of a
2621: Fourier Series with the amplitudes $(-1)^k\Phi(2\pi k/\tau)/k$ that
2622: scales with the sampling interval $\tau$. Evidently the discretization
2623: error completely vanishes for all $x_0=\tau n/2,\, n\in\mathbb{Z}$,
2624: that is, for all those cases when $x_0$ is located exactly over the
2625: center of one interval or exactly between two intervals. At these
2626: special points, the symmetry of the distribution will be correctly
2627: reproduced by the detector. If $\varphi(x)$ converges to the Dirac
2628: $\delta$-function, the discretization error reaches its maximum
2629: values. This is intuitively expected because the exact position of the
2630: Dirac $\delta$-function within one interval cannot be determined.
2631:
2632: One obtains the following simple form of the Fourier transform for the
2633: special signal distribution~(\ref{eq:test-func-for-stat-err-cog}) with
2634: frequencies $\omega=2\pi k/\tau$:
2635: \begin{equation}
2636: \label{eq:foureier-weigths-test-func}
2637: \Phi(2\pi k/\tau)=\frac{e^{-\frac{2 d k \pi }{\tau }} J}{\sqrt{2 \pi }}.
2638: \end{equation}
2639: With this result, one can easily find an upper limit for the
2640: discretization error (\ref{eq:discretization-cent-error}), since
2641: $e^{-2d\pi/\tau}/\sqrt{2\pi}$ majorizes\footnote{One also has to pay
2642: attention to the sine factor within the
2643: series~(\ref{eq:discretization-cent-error}). Generally it is possible
2644: that this factor cancels exactly the alternate sign
2645: $(-1)^k$ for all $k\in\mathbb{N}$ inside the sum. This happens for
2646: $x_0=\tau/2$ and $k=(2n+1)/2,\,n\in\mathbb{N}$ which is
2647: just the previously discussed case where the center of gravity
2648: algorithm does not break the symmetry of the distribution and does not
2649: contribute any error at all.} the series
2650: (\ref{eq:discretization-cent-error}) and by computing the alternating
2651: harmonic series one gets the following expression for the upper limit
2652: of the discretization error:
2653: \begin{equation}
2654: \label{eq:upp-lim-discrete-error}
2655: \delta_\tincaps{D}\mu_1(\bar{f}^{(\!x_o\!)})=\mu_1(\bar{f}^{(\!x_o\!)})-x_0=\pm\frac{e^{-\frac{2d\pi}{\tau}}\tau\ln(2)}{\sqrt{2}\pi^{3/2}}.
2656: \end{equation}
2657: From the previous section it is known that for the present example
2658: distribution $d$ is proportional to its width. Therefore
2659: equation~(\ref{eq:upp-lim-discrete-error}) states that the discretization
2660: error becomes significant for detector setups with $\tau$
2661: much larger than the width of the distribution to be measured.
2662:
2663:
2664: \subsection{Symmetry Breaking of the Current Distribution}
2665: \label{sec:symmetry-breaking}
2666:
2667: Another systematic error which is of special interest within the scope of the
2668: present work is introduced by the fact that all practicable
2669: detectors can never be of infinite size. Owing to this intrinsic
2670: property of the devices, the sampled signal distribution has a finite
2671: support. This error is well known and because of its severity it has
2672: been subjected to detailed studies by many investigation groups
2673: ({\em e.g.}\ Freifelder {\em et al.}\ \cite{Freifelder:1993}, Clancy {\em et al.}\
2674: \cite{Clancy:1997}, Siegel {\em et al.}\ \cite{Siegel:1995} and Joung et
2675: al.\ \cite{Joung:2002}) in order to avoid or combat this degradation
2676: often referred to as {\em image compression} or {\em edge effects}.
2677: Actually, this error is a simple consequence of applying the CoG
2678: algorithm with an integration interval smaller than the support for
2679: the signal distribution. It can be observed for all moments whose
2680: distribution is (necessarily)
2681: truncated by the detector design. Being one of the most important of
2682: all errors considered here, it often forces the designer of the
2683: detector to restrict the usable area to a central region and render
2684: inoperative a large peripheral area. There are many recipes to
2685: extract the best results from each detector technology and geometry,
2686: {\em e.g.}\ scintillator or semiconductor, pixels or continuous. The error
2687: discussed in this section depends crucially on the detector
2688: configuration. Here, only the special configuration of a scintillation
2689: detector with continuous crystal is considered.
2690: Even in this case, the response performance can be varied by different
2691: approaches in order to optimize total light yield and other important
2692: parameters. Some well-known design enhancements, for example
2693: diffusive reflective coating, are, however, excluded for an optimal performance of
2694: the depth of interaction measurement. Since reflection on the crystal
2695: edges would destroy the correlation between the distribution width and
2696: the depth of interaction (not the second moment), all crystal
2697: surfaces that are not read out by a photodetector have to be coated
2698: anti-reflective.
2699:
2700:
2701: First, the 0th moment is considered since it is required for the
2702: energy measurement and for the normalization of the center of gravity
2703: algorithm according to equation~(\ref{eq:func-centroid}). The error
2704: that is made by truncating the integration becomes clear from
2705: splitting the integral into the integration over the support $L$ of
2706: the detector and two residual parts $]\!-\infty,\frac{-L}{2}]$ and
2707: $[\frac{L}{2},\infty\;[$ to the left and to the right respectively:
2708: \begin{equation}
2709: \label{eq:sym-error-energy}
2710: \mu_0^{(\!x_0\!)}=J\int_{-\frac{L}{2}}^{\frac{L}{2}}\varphi(x-x_0)\;dx=
2711: J-J\int_{\frac{L}{2}}^{\infty}\left\{\varphi(x-x_0)+\varphi(-x-x_0)\right\}dx .
2712: \end{equation}
2713: Note that the signal distribution will have its symmetry center
2714: at $x_0$ and not $x=0$. Therefore, except for the case $x_0=0$, the
2715: integration of $\varphi(x-x_0)$ over the intervals $]\!-\infty,\frac{-L}{2}]$ and
2716: $[\frac{L}{2},\infty\;[$ give different results. The generalization of equation
2717: (\ref{eq:sym-error-energy}) to any $k$-th moment is straightforward and
2718: given by
2719: \begin{equation}
2720: \label{eq:sym-error-k-mom}
2721: \mu_k^{(\!x_0\!)}=J\int_{-\frac{L}{2}}^{\frac{L}{2}}x^k\varphi(x-x_0)\;dx=
2722: J\mu_k-J\int_{\frac{L}{2}}^{\infty}x^k\left\{\varphi(x-x_0)+(-1)^k\varphi(-x-x_0)\right\}dx,
2723: \end{equation}
2724: where the $\mu_k$ are the normalized distribution moments according to
2725: its definition~(\ref{eq:func-moments}), and the normalization of the
2726: moments $\mu_k^{(\!x_0\!)}$ has to be done with
2727: $\mu_0^{(\!x_0\!)}$. From equations~(\ref{eq:sym-error-energy}) and
2728: (\ref{eq:sym-error-k-mom}) it becomes clear that the introduced error
2729: depends on the signal distribution $\varphi(x)$ in a non-trivial way.
2730: Nevertheless, some general properties can be extracted from them.
2731: Consider the case of a distribution that is symmetric around the
2732: impact point $x_0$, {\em i.e.}\ $\varphi(x-x_0)=\varphi(x_0-x)$. Actually,
2733: this is a property of many signal distributions of interest in the
2734: field of experimental high energy physics and/or when using
2735: scintillation detectors. Result (\ref{eq:sym-error-k-mom}) splits then
2736: into two cases depending on whether $k$ is odd or even. The error in the
2737: odd moments vanishes completely for $x_0=0$.
2738: Contrariwise, the error in the even moments will never vanish for
2739: the class of non-trivial distributions with $\varphi(x)\neq0,\,x\in[L/2,\infty\;[$.
2740: Obviously this is the symmetry behavior of the CoG
2741: algorithm and the error that distinguishes the real impact position
2742: $x_0$ from the centroid $\mu_1^{(\!x_0\!)}$ comes along with the
2743: breaking of the symmetry of the distribution by the integration over
2744: the finite interval.
2745:
2746: \begin{figure}[t]
2747: \centering
2748: \subfigure[][Behavior of positioning for different depth of
2749: interactions ($d=0,2,4,\ldots,10\,\mathrm{mm}$) if the center of gravity
2750: algorithm is used. $L$ was set to $20\,\mathrm{cm}$.]{\label{subfig:theo-sample-centroid}
2751: \psfrag{x}{\hspace*{-8ex}true position [mm]}
2752: \psfrag{y}{\hspace*{-6ex}centroid [mm]}
2753: \includegraphics[width=0.45\textwidth]{cdr/theo-sample-cent}}\hspace*{0.05\textwidth}
2754: \subfigure[][Broadening of the point spread function due to the
2755: nonlinearity in the position response of the truncated center of
2756: gravity algorithm. The identity $x=y$ is plotted as dashed line.]{\label{subfig:inversion-broadening}
2757: \psfrag{x}{\hspace*{-9ex}measured centroid [mm]}
2758: \psfrag{y}{\hspace*{-9ex}inverted position [mm]}
2759: \includegraphics[width=0.45\textwidth]{cdr/inversion-broadening}}
2760: \caption[Behavior of positioning of the center of gravity
2761: algorithm]{Behavior of positioning of the center of gravity
2762: algorithm and consequences for the linearization.}
2763: \label{fig:theo-sample-centroid}
2764: \end{figure}
2765:
2766: For the test distribution (\ref{eq:test-func-for-stat-err-cog}) the
2767: centroid actually measured is given by:
2768: \begin{equation}
2769: \label{eq:moments-for-sample-dist}
2770: \mu_1^{(\!x_0\!)}=x+\frac{d \log \left(\frac{4 d^2+(L-2 x)^2}{4 d^2+(L+2 x)^2}\right)}{2 \left(\cot ^{-1}\left(\frac{2 d}{L-2 x}\right)+\cot
2771: ^{-1}\left(\frac{2 d}{L+2 x}\right)\right)}.
2772: \end{equation}
2773: The other finite support moments $\mu_k^{(\!x_0\!)}$ can be obtained similarly by straightforward
2774: application of the definition (\ref{eq:sym-error-k-mom}). Result
2775: (\ref{eq:moments-for-sample-dist}) is plotted for various impact
2776: depths $d$ and a typical detector size $L$ in figure
2777: \ref{subfig:theo-sample-centroid}. From equation
2778: (\ref{eq:moments-for-sample-dist}) and its graphical representation in
2779: figure \ref{subfig:theo-sample-centroid}, it is seen that the truncation
2780: of the signal distribution at the detector's limits compresses the
2781: identity $y=x$ at the detector's edges. The closer the impact position
2782: is to the edge, the stronger the compression. The degree of
2783: compression varies also with the depth of interaction. This compression not
2784: only results in a characteristic mapping of the impact position, but
2785: also leads to a resolution loss at the detector's edges. Since it will
2786: be attempting to find the inverse mapping of the dependences in figure
2787: \ref{subfig:theo-sample-centroid}, the point spread function of the
2788: detector will be broadened as shown in figure
2789: \ref{subfig:inversion-broadening}. A detector response that
2790: compresses the images at its edges is corrected by the application of an
2791: expanding reverse mapping. However, not only will the center-points of the point
2792: spread function be dispersed, but also the point
2793: spread function itself. Together with the depth dependence of the
2794: detector's positioning characteristic, the distortion at peripheral
2795: zones of the sensitive area are subjected to a significant resolution
2796: degradation. The different image compressions for different impact
2797: depths easily lead to superposition of impact positions that makes the
2798: bare center of gravity algorithm useless except for a small region at
2799: the center, especially for applications that require thick
2800: scintillators.
2801:
2802: \subsection{Electronic Noise}
2803:
2804: Finally, the influence of thermal noise generated by the passive
2805: and active electronic devices is considered. In general, the
2806: spectral density of the noise power $\hat{P}$ at the frequency $f$ is given by
2807: $\hat{P}=k_BTf$, where $k_B$ denotes the Boltzmann constant and
2808: $T$ the absolute temperature of the device (Tietze and Schenk \cite{Tietze}). Since the
2809: power is also given by $P=U^2/R$ and by virtue of Ohms law, the
2810: spectral density of the noise voltage and current can be derived to yield
2811: \begin{equation}
2812: \label{eq:noise-voltage}
2813: \hat{U}=\sqrt{4Rk_BTf}\quad\mbox{ and }\quad\hat{I}=\sqrt{4R^{-1}k_BTf},
2814: \end{equation}
2815: where the factor 4 arises from averaging the $\sin^2(t)$ function over a
2816: full period. At room temperature, $4k_BT$ amounts to $1.6\cdot10^{-20}\,\mathrm{Ws}$
2817: and therefore the noise voltage of a resistor is given by
2818: $\hat{U}_R\simeq0.13\,\mathrm{nV}\times\sqrt{Rf}$. Clearly the noise voltages of
2819: any two resistors are not correlated at all and the $n\times n$
2820: noise voltages of the summing amplifier have to be added
2821: quadratically, while one has to take into account that they will be
2822: amplified along with the desired signal voltage.
2823: \begin{equation}
2824: \label{eq:sum-noise-voltage}
2825: \hat{U}_\mathit{noise}^\Sigma\simeq\sqrt{\hat{U}^2_\mathit{noise_{OP}}+\hat{U}^2_\mathit{R_C}+4k_BTf\sum_{i,j}^{n,n}g^2_{ij}R_{ij}}
2826: \end{equation}
2827: Here, $\hat{U}_\mathit{noise_{OP}}$ is the noise contribution to the output
2828: signal of the operational amplifier itself and
2829: $\hat{U}^2_\mathit{R_C}$ the contribution caused by the compensation
2830: resistor $R_c$ (refer to figure \ref{fig:analogue-adder} in section
2831: \ref{sec:sim-measurement-of-sec-mom}). Using the network configuration
2832: of the two-dimensional proportional charge divider (figure
2833: \ref{fig:bidim-schem}) together with the optimum values for the
2834: summation weights computed in
2835: section~\ref{sec:proportional-resistor-chains}, one obtains a typical
2836: total noise voltage of $U_\mathit{noise}^\Sigma\approx160\,\mathrm{\mu V}$.
2837: The spectral noise $\hat{U}_\mathit{noise_{OP}}$ of the
2838: operational amplifier was assumed to be
2839: of $10\,\mathrm{nV\sqrt{Hz}}$, which is a typical value for
2840: modern operational amplifier. Therefore, the strongest contribution in the sum
2841: (\ref{eq:sum-noise-voltage}) arises, with $\approx140\,\mathrm{\mu V}$, from the
2842: operational amplifier itself. The reason for the rather low contribution
2843: from the summation network ($\approx25\,\mathrm{\mu V}$) lies in the posterior
2844: attenuation of the signal. Actually, the presented summation amplifier,
2845: with a mean gain of factor $\approx0.003$, is more likely to be called
2846: a summation attenuator. This very small gain is needed in order
2847: to avoid saturations of the operational amplifier but also effectively
2848: suppresses the thermal noise of the resistor network.
2849:
2850: Typical signal amplitudes observed with our experimental
2851: setup described in chapter \ref{ch:experiment} are of $50\,\mathrm{mV}$
2852: and more. Thus, the electronic noise of the summing amplifier can
2853: be neglected compared with the other error sources discussed in this
2854: section. The same holds for the currents $J_A$, $J_B$, $J_C$ and $J_D$.
2855:
2856: \chapterbib
2857:
2858: %%% Local Variables:
2859: %%% mode: latex
2860: %%% TeX-master: "~/documents/tex/thesis/thesis"
2861: %%% End:
2862: