6
Timing Measurements

The measurement of short time intervals between nuclear events is of paramount importance in many fields of nuclear science and technology. Example applications are radionuclides metrology, positron emission tomography (PET), nuclear data measurements, and nuclear physics research. We start this chapter with a discussion of the basics of pulse timing followed by a detailed description of different components of pulse timing systems. Pulse timing procedures with scintillation, semiconductor, and gaseous detectors are also separately discussed.

6.1 Introduction

Since each detected nuclear event is represented with an output detector pulse, the measurement of time intervals between the events is equal to measuring the difference between the arrival times of pulses from the corresponding detectors. The time intervals range from a few picoseconds to as large as a few microseconds. In some applications, it is only sufficient to determine if the time spacing between the nuclear events lies in a preset time window, while in many applications the distribution of the time difference between the events is required. The former is called coincidence measurement, while the latter is called time spectroscopy. Some examples of time difference measurements between detector pulses are shown in Figure 6.1. In Figure 6.1a, uncorrelated particles strike on two detectors, and thus the time difference between the arrival times of any pair of pulses will have a random value. Then, the distribution of the time difference between pulses may have a flat shape that carries no special information. When there is a correlation between the particles striking on the two detectors, the distribution of the time difference between the pulses will carry some information on the nature of correlation between the particles. Figure 6.1b shows the case when the pulses are initiated by particles that are produced at the same time, for example, two positron annihilation photons. Such events are called prompt coincidence events. In such cases, the time difference is constant and, in the absence of measurement errors, the distribution will have a delta function shape. Figure 6.1c represents the measurement of the time delay between the gamma rays that result from the formation and subsequent decay of a nuclear state. Such timing is called delayed‐coincidence measurement and is used to measure the half‐life of the nuclear state. Figure 6.1d shows an example of time‐of‐flight measurements. In this arrangement, coincidence pulses are initiated by charged particles traveling between two detectors. The time difference, that is, time of flight, will depend on the velocity of charged particles, and thus, the time spectrum can be used to obtain information such as the type of charged particles. Time‐of‐flight measurements can be also performed with photons and neutrons.

Image described by caption and surrounding text.

Figure 6.1 Some examples of time‐difference measurements between two detectors and the corresponding distributions of time differences under ideal measurement conditions (see text for details).

In many applications, the number of detectors involved in timing measurements can be more than two, and the time difference may be measured between any pair of detectors. Also in some situations, the timing measurements may be performed with respect to a reference pulse produced by a non‐detector device such as an accelerator. Nevertheless, the basic concepts of pulse timing are defined for a system involving a pair of detectors. Figure 6.2 illustrates a basic pulse timing setup, involving two detectors. The aim of the system is to produce a spectrum of the time difference between the arrival times of the pulses from the two detectors. The detector’s pulses can be delivered by a photomultiplier tube (PMT) or a preamplifier, depending on the type of detector. One of the pulses is labeled as start pulse from which the time difference of a second pulse known as stop pulse is measured. The output pulses of each detector are processed with a time pick‐off unit, also called time discriminator, which constitutes a fundamental part of any timing measurement setup and generates logic pulses whose leading edge (LE) corresponds precisely to the event of interest in its time of occurrence. Therefore, the timing of logic pulses has a very fast LE of a few hundreds of picoseconds. The start and stop logic pulses are then fed into a device that produces an output pulse with an amplitude proportional to the time interval between input start and stop pulses. This device is called time‐to‐amplitude converter (TAC). The stop pulse is normally delayed some small amount of time to ensure that the stop pulse arrives after the start pulse and the TAC operates within its linear range. When the TAC output is fed into a multichannel pulse‐height analyzer (MCA), the distribution of the amplitude of TAC output pulses recorded by the MCA will give a measure of the distribution of time intervals between the start and stop pulses, which is called the time spectrum that bears a close relation to pulse‐height spectrum. Here the horizontal scale, rather than pulse height, is the time interval length, while the vertical scale represents the number of events in each time interval.

Image described by caption and surrounding text.

Figure 6.2 A simple block diagram of a typical time spectrometer.

In an ideal timing system, that is, in the absence of time measurement errors, each time pick‐off unit produces a logic pulse that is precisely related to the arrival time of detected events, and therefore, in prompt coincidence measurements such as in Figure 6.1b, the start and stop pulses are always separated by the fixed delay value, and the TAC output will have a constant amplitude output that is therefore stored in a single channel of the MCA. However, in an actual timing system, there are always some sources of uncertainty or error in the determination of the arrival times of detector pulses, and thus even a prompt coincidence time spectrum shows a distribution rather than a delta function. If in Figure 6.2 we assume that the independent timing uncertainties in two branches are of Gaussian type, which is usually the case, then the prompt coincidence time spectrum will display a Gaussian distribution as shown in Figure 6.3. The peak that represents the true coincidence events may lie on a background of chance coincidences that are produced by uncorrelated start and stop signals. The intensity of the chance coincidences depends on the rates at which pulses are generated in either start or stop branches. The area under the peak after the subtraction of the continuum due to the chance coincidences gives the total number of detected coincidences. The accuracy of timing measurements is generally described with the concept of time resolution. The full width at half maximum (FWHM) of a prompt coincidence time distribution is called the time resolution of the system. The FWHM of the time spectrum can be written as

where FWHM1 and FWHM2 are the contributions of each branch to the time resolution of the system. For identical detectors and time pick‐off elements in both start and stop branches, the contribution of detectors in the total FWHM are equal, and thus, the time resolution of each detector can be expressed as FWHM/√2. Another figure of merit for a timing spectrum is full width at tenth maximum (FWTM), which more fairly accounts for tails sometimes observed at either side of time spectra. In general, the timing resolution must be high, that is, the timing peak must be very narrow, so that the time relationship between two closely spaced events can be measured accurately and also the overlap region with chance coincidences is minimized. Moreover, the narrow width of the peak must be maintained down to a small fraction of its maximum height to ensure minimum contamination of the true timing events with unwanted events. As it will be discussed in the rest of this chapter, the time resolution depends on the type of detector, on the type of pulse processing system used to extract the timing information from signals, and generally on the type and energy of incident particles. Figure 6.3 also shows the effect of the delay on the stop branch on the time spectrum. If there were no delay on the stop branch, the coincidence time spectrum that resulted from input pulses with exactly the same arrival time would be centered about channel zero and only half of its shape would be measured. Introducing a delay into the stop channel moves the entire time spectrum to the right by an amount equal to the delay and allows both sides of the time spectrum to be recorded. The effect of delay on the location of a coincidence peak also allows one to calibrate the time scale of the system versus channel number of MCA by using a linear relation between the peak channel numbers and corresponding delay values. More accurate time calibrations can be obtained by using time calibrator units that produce start and stop pulses of accurately known time difference by using an accurate digital clock to produce stop pulses at precisely spaced intervals after a start pulse.

Graph of time (channel number) vs. number of events displaying a coincidence time spectrum with arrows depicting delay, zero delay, FWHM, and FWTM and a shaded area labeled chance coincidences.

Figure 6.3 A coincidence time spectrum.

As it was mentioned earlier, in some applications, it is not necessary to record the time spectrum, and it is only sufficient to determine if the time difference between the events from the two detectors lies in a preset time window. One way to do this task is to use a single‐channel analyzer (SCA) as shown in Figure 6.4. In this system, the SCA selects the range of pulse amplitudes from the TAC that represents events that occur within this time window. The SCA output can be simply counted or used to gate other instruments in the system. A more simple way of selecting events lying in a time window is to use a coincidence unit. A simple coincidence unit with two inputs is shown in Figure 6.5. It produces a logic pulse output when the input logic pulses occur within the resolving time window (2τ), but it does not determine the actual time difference between two input signals or indicate which of the two signals occurred first. In counting the number of coincidence events, the resolving time must be carefully adjusted to ensure that genuinely coincident events produce an output pulse. Since detector events occur at random times, accidental coincidences can occur between two pulses that produce background in the coincidence counting. The rate of accidental or random coincidences is given by [1]

(6.2)images

where N1 is the count rate in detector number one, N2 is the count rate in detector number two, and 2τ is the resolving time. A detailed discussion of coincidence measurement errors and correction methods can be found in Ref. [1].

Image described by caption and surrounding text.

Figure 6.4 A system for selecting timing events lying in a time window by using a SCA.

Image described by caption and surrounding text.

Figure 6.5 Selection of events lying in a time window with a coincidence unit.

6.2 Time Pick‐Off Techniques

The primary function of a time pick‐off unit or a timing discriminator is to produce an output logic pulse, which is precisely related to the arrival time of the input pulse and must have extremely fast LE, but the width of the pulse is not of prime importance and may vary according to the measurement requirement. Historically, most of the commercial fast timing discriminators are designed to accept negative input pulses on a terminated 50‐Ω coaxial cable. An ideal time discriminator should produce a logic pulse at the moment that the input pulse arrives (T), as shown in Figure 6.6. However, due to the presence of electronic noise and the response time of the device, the production of the timing output at the same moment as the pulse arrival time is not possible. Therefore, practical timing discriminators should ideally produce an output logic pulse whose firing time (T1) is precisely and consistently related to the beginning of each input pulse, that is, a constant relation between T1 and T. To produce such logic pulses, different principles have been used that are discussed in the following sections.

Relation of the output logic pulse of a time discriminator displaying a square waveform (bottom) and the arrival time of the input pulse displaying a bell-shaped curve (top).

Figure 6.6 The relation of the output logic pulse of a time discriminator and the arrival time of the input pulse. Since generation of the logic pulse at time T is not possible, the logic pulse is generated at time T1 while T1 has a precise and constant relation to T.

6.2.1 Leading‐Edge Discriminator

6.2.1.1 Principles

An LE discriminator is the simplest type of time pick‐off circuits. The principle of operation and basic structure of an LE discriminator is shown in Figure 6.7. It simply produces a logic pulse when the input signal crosses a fixed threshold level, which should be above the noise level to prevent spurious triggering on noise signals. An LE discriminator can be simply built by using a fast comparator that features high switching speed with low transition time. One input of the voltage comparator receives the detector pulse, while the other input of the voltage comparator receives a reference voltage that determines the threshold level. The width of output logic pulse should be short enough to operate the coincidence circuit and is usually arranged by a mono‐stable multivibrator circuit [2–5]. An LE discriminator can be equipped with extra discriminators to select the events that lie in an energy range of interest. The energy selection capability also prevents the device from possible false triggering on the baseline noise.

Top: Graph of the principle of operation displaying curves labeled output and input pulse. Bottom: The basic structure of a leading-edge time discriminator featuring fast comparator and monostable multivibrator.

Figure 6.7 (a) Principle of operation and (b) the basic structure of a leading‐edge time discriminator.

6.2.1.2 Leading‐Edge Discriminator Errors

In principle, there are two important sources of errors in time pick‐off measurements: jitter and walk. These affect the accuracy of the relation between T1 and T, as shown in Figure 6.6. Timing errors may result from drift in the electronics that is introduced by component aging and by temperature variations. Jitter is the time uncertainty that is caused by noise and also by statistical fluctuations of the detector signals. The noise can originate from the detector, pulse processing electronics, or the discriminator itself. The effect of jitter on the time pick‐off uncertainty is shown in Figure 6.8 for an ideal LE discriminator. If the input signal is assumed to be approximately linear in the region of threshold crossing, for a noise with Gaussian probability density function and with zero mean value, the noise‐induced uncertainty in threshold‐crossing time is given with reasonable accuracy by the triangle rule as

where σv is the rms value of noise and σT is the rms value of time uncertainty. Equation 6.3 indicates that the time jitter is inversely proportional to the slope of the input signal at the threshold‐crossing time, which means that the threshold level must be set at the point with greatest slope and the pulse should not be slowed down during transit to the discriminator. It is clear that the minimum value of discriminator level is limited by the level of electronic noise. When the pulses are delivered by a charge‐sensitive preamplifier, due to the unavoidable pole created in the preamplifier, the pulses exhibit an exponentially rising region near the starting point, which is called toe [6]. In this case, the lower limit of the threshold level is limited at the region of the unavoidable toe because in this region the slope is degraded, leading to smaller slope‐to‐noise ratio. Another source of jitter results from the statistical amplitude fluctuations that cause uncertainty in the time at which the discriminator fires the output logic pulse [7]. In addition to jitters related to input pulse, there is generally a residual jitter contribution from the electronics system due to variations in the transit times of pulses [8].

Graph of time vs. amplitude displaying a curve with horizontal lines depicting discriminator level and dotted rectangles with vertical lines depicting output logic pulse.

Figure 6.8 The effect of noise‐induced jitter on the time pick‐off uncertainty.

The second source of timing error is walk. In general, time‐walk is the time movement of the output logic pulse relative to the input pulse due to variations in the shape and amplitude of the input pulse. Time‐walk is a serious limitation of LE discriminators. Figure 6.9 illustrates the time‐walk of an ideal LE comparator due to input signal variations. In the upper part of the figure, two pulses of the same risetime but different amplitudes are shown. Both pulses start at the same time, but pulse 1 crosses the threshold level at time T1, while pulse 2 crosses the threshold at time T2. The difference in threshold‐crossing time causes the output logic pulse from the LE discriminator to walk along the time axis. It is clear that time‐walk increases for small amplitude pulses and thus is most pronounced for signals with amplitudes that only slightly exceed the threshold level. The lower part of Figure 6.9 shows two pulses of the same amplitude but with different risetimes. The slower pulse crosses the discriminator after the crossing time of the first pulse that results in a time‐walk along the time axis.

Image described by surrounding text.

Figure 6.9 Illustration of time‐walk in a leading‐edge discriminator due to amplitude and risetime variations of input pulses.

In a real LE discriminator, an additional contribution to the time‐walk results from the charge sensitivity of comparators [7, 9]. All physically realizable comparators are charge sensitive in a sense that even though the comparator threshold has been exceeded, a finite amount of charge has to be injected to trigger the comparator. The time‐walk due to charge sensitivity is illustrated in Figure 6.10. For pulse 1, the small amount of charge required to trigger the discriminating element moves the trigger time from T to T1, and the effective threshold level of the discriminator changes to Vth + ΔVq. These changes are related to the slope of the input signal as it passes through the threshold. The timing errors introduced by charge sensitivity are greater for signals with longer risetimes and for signals with smaller pulse amplitudes above the threshold. The characteristics of the input device are also a major factor in determining the charge sensitivity of the discriminator. By assuming that the input signal is approximately linear during the time that is required to accumulate the required charge, corresponding to area S indicated in Figure 6.10, the error in the effective sensing time is related to the slope of the input signal by [9]

(6.4)images

where v(t) is the input pulse as a function of time and T is the threshold‐crossing time of the pulse. One can see in Figure 6.10 that the time‐walk of an LE discriminator tends to decrease for pulses with shorter risetimes though the effective threshold level increases for pulses with short risetime.

Image described by surrounding text.

Figure 6.10 Schematic illustration of time‐walk due to charge sensitivity of a leading‐edge discriminator. The trigger time for pulse 1 from T moves to T1 due to the required charge determined by area S.

The discussion on the leading‐edge discriminator’s timing errors implies that when this technique is restricted to those applications that involve a very narrow dynamic range of pulse amplitudes, excellent timing results can be obtained. This is most frequently achieved by experimenting with the threshold level. Under the optimum threshold level, timing errors due to charge sensitivity and to jitter are minimized for input signals with the greatest slope at threshold‐crossing time.

6.2.1.3 Optimum Timing Filter

The timing jitter due to electronic noise may be minimized by choosing a proper noise filter. Such filter minimizes the time jitter by maximizing the slope‐to‐noise ratio according to Eq. 6.3 [10–12]. The problem of maximizing the slope‐to‐noise ratio can be treated in the same way as the problem of maximizing signal‐to‐noise ratio that was treated earlier in Chapter 4. An input signal V() with a superimposed noise of power spectrum N(ω) is processed by a filter with transfer function H(). The slope of the filter output pulse g(t) is dg(t)/dt, which in the frequency domain corresponds to a multiplication of G() with where G() is the output pulse in the frequency domain. Then the slope of the output pulse and the mean‐square value of output noise voltage are given by

(6.5)images

and

(6.6)images

The mean‐square error in the time measurement is then

By defining ρT as inverse of σT, we obtain

(6.8)images

Minimizing Eq. 6.7 is equivalent to maximizing its inverse ρT. Thus, by defining

and using the Schwarz inequality, the maximum value of the mean square of ρT is obtained as

(6.11)images

This maximum value is obtained only when

where K is a constant and images is the complex conjugate of y2(ω). By substituting Eqs. 6.9 and 6.10 into Eq. 6.12 and solving for H(), we obtain

(6.13)images

This equation represents the optimum filter for timing for an arbitrary signal and noise and also for an arbitrary measurement time. From our discussion in Chapter 4, one realizes that H() differs only by the factor from the filter optimizing the signal‐to‐noise ratio. Since a multiplication with in the frequency domain corresponds to the derivative in the time domain, the impulse response of the filter that maximizes the slope‐to‐noise ratio is equal to the derivative of the impulse response of the filter that maximizes the signal‐to‐noise ratio. However, one should note that such optimum filter is related in a unique way to the input signal shape and the noise spectrum, and thus, it may not be always realizable or practical. In practice, filters such as integrator–differentiator filters are commonly used to prepare the signal for time pick‐off (see Section 6.2.2.6). The use of such filters particularly has proved to be useful in timing with semiconductor detectors.

6.2.1.4 Extrapolated Leading‐Edge Timing (ELET)

We already discussed that the output of an LE discriminator suffers from time‐walk errors due to variations in the amplitude and shape of input pulses. Although the time‐walk can be reduced by using small threshold levels, the choice of a sufficiently low discriminating threshold for reducing the walk is again restricted by the noise level and the presence of toe region. Therefore, LE discriminators produce a poor time resolution for detectors with variable pulse shapes and amplitudes. Extrapolated leading‐edge discriminator (ELET) is a modified version of the LE discriminator that compensates for the time‐walk stemmed from the variable pulse risetimes and amplitudes. This method was first introduced by Fouan and Passerieux [13], and a variety of circuits based on the same principles have been built mainly for timing with germanium detectors [6, 14–17]. The principle of this method is shown in Figure 6.11. Two LE discriminators with different threshold levels VL and VU are used where VU is generally a multiple of VL. For pulse 2, the time‐walk with respect to pulse 1 is equal to TL2 − TL1 associated with the discriminator set at VL. Ideally we are interested to have a zero time‐walk, which corresponds to the measurement of the start time of the pulse T. This time cannot be directly marked, but if the time interval between a reference time Tref and T is intended to be measured, then the subtraction of the time TU − TL from the time TL − Tref gives the correct time interval T − Tref, if (TU − TL) = (TL − T). Here TL and TU are the crossing times of the pulse at VL and VU, respectively. In other words, in this method, the triggering times of two discriminators with different threshold levels are used to extrapolate back to a time origin. The ELET method is very effective for linearly rising pulses that do not exhibit change of slope below the upper threshold. However, the noise jitter of this method is poorer than the standard LE method because jitter is introduced by both discriminators. The ELET method has been implemented in real time by using dedicated circuits or standard nuclear instrument modules and also offline by using software extrapolation of discriminators outputs.

Timing diagram of extrapolating leading-edge time discriminator displaying 2 horizontal lines for VU and VL, 2 curves labeled 1 and 2, and vertical dashed lines with double-headed arrows labeled Tref, To, etc.

Figure 6.11 Timing diagram of extrapolating leading‐edge time discriminator.

6.2.2 Constant‐Fraction Discriminator

6.2.2.1 Principles

The problem of time‐walk in LE discriminators that results from variations in the pulse amplitudes can be avoided by using a time pick‐off circuit in which the discrimination level varies as a constant fraction of pulse amplitude. Such time discriminator was first reported by Gedcke and McDonald in 1967 [18] and is called constant‐fraction discriminator (CFD), which is in wide use today with all types of detectors due to its good timing performance and relatively simple structures. Figure 6.12a shows a basic functional diagram of a CFD. The pulse is sent into two branches. In one branch it is attenuated by the fraction f of the pulse amplitude, while in the other branch it is delayed by td. The two branches are compared and a time mark is developed when two signals are equal. This is equivalent to determining the zero‐crossing time of the difference of two signals as shown in Figure 6.12b. The formation of the difference signals or CFD pulse is an important step of proper operation of a CFD circuit. The CFD shaping steps are shown in Figure 6.13. The transfer function of such CFD shaper is given by

(6.14)images
Image described by caption and surrounding text.

Figure 6.12 Block diagrams for a CFD pulse timing.

CFD pulse shaping displaying input pulse, delayed input pulse, inverted and inverted pulse, constant-fraction pulse, and output logic pulse (top to bottom) displaying different solid curves.

Figure 6.13 CFD pulse shaping.

CFD shaping produces a bipolar pulse whose zero‐crossing point occurs at the fraction of the pulse amplitude chosen. The zero‐crossing is detected and used to produce an output logic pulse. Ideally, one can set the fraction f to minimize the jitter, setting the threshold fraction at the point of the signal shape where the slope is at maximum, as it was discussed for an LE discriminator. The minimization of time‐walk is achieved by a proper selection of shaping delay, td, depending on the characteristics of the input pulse. The role of shaping delay is explained in the following. For a theoretical linear rising input signal described with

(6.15)images

where A is a constant and t is time, the constant‐fraction (CF) zero‐crossing time tcfd can be calculated from

(6.16)images

or

This relation holds for a pulse of continuous rise. In practice, pulse has a finite amplitude, and thus the zero‐crossing time can happen before or when the attenuated input pulse has reached its maximum amplitude. In true constant‐fraction (TCF) discrimination circuits, the zero‐crossing time occurs when the attenuated pulse is at maximum. To ensure this happens, the shaping delay is selected so that

(6.18)images

where tr is the input signal risetime. In this case, the zero‐crossing time is given by [7, 9]

(6.19)images

In practice, input pulses have finite widths, and thus the shaping delay must also be made sufficiently short to force the zero‐crossing signal to occur during the time that attenuated signal is at its peak. A simple and useful test to ensure that the CFD is properly adjusting is to simultaneously view the bipolar signal, known as monitor signal, and CFD output on a fast oscilloscope that is triggered by the output of the CFD. A TCF timing fundamentally or theoretically (with perfect comparators, attenuators, and delay lines) produces no walk for signals of variable amplitude but with equal risetimes since it makes the arrival time independent of the amplitude of the input pulse. However, in many detectors, the risetime of pulses also varies that has led to another mode of pulse timing with CFDs. This mode is called amplitude and risetime compensated (ARC) timing that will be discussed in Section 6.2.2.3.

6.2.2.2 CFD Timing Errors

In a TCF the jitter is minimized by setting the CFD to provide a time pick‐off signal at the fraction of the input pulse amplitude with maximum slope‐to‐noise ratio. Similar to an LE discriminator, by assuming the CFD signal to be approximately linear in the region of zero‐crossing, the rms value of noise‐induced jitter in the CF zero‐crossing time, σT, is given by

where σv is the rms value of the noise on the CF bipolar signal and Tcf is the general zero‐crossing time. The rms value of noise on the CFD bipolar signal can be related to the noise on the input signal by making the simplifying assumptions that the CF circuit is noiseless and has an infinite bandwidth and also the noise on the input signal is time‐stationary type and has a Gaussian probability density function with zero mean value. Then, the rms value of noise on the CFD bipolar pulse can be written as [9]

(6.21)images

where σ is the rms value of the noise on the input signal, f is the CFD attenuation factor, en(t) is the input noise, φ(td) is the autocorrelation function of the input noise, and td is the CFD shaping delay. In the case of uncorrelated noise, the relation between the rms values of the noise on the CFD bipolar pulse and on the input signal is simplified to

By using Eq. 6.22 and having the slope of the pulse, one can estimate the timing jitter from Eq. 6.20. For linear rising pulses of amplitude V, the slope of the pulses is given by

Combining Eqs. 6.20, 6.22, and 6.23 leads to the noise‐induced jitter as

(6.24)images

A comparison of Eq. 6.23 with the result for a simple LE discriminator indicates that under identical input and noise conditions, the jitter is usually worse for TCF timing than that for simple LE timing by the factor (1 + f2)l/2. However, TCF timing virtually eliminates time jitter due to statistical amplitude variations of detector signals so that TCF timing often results in less overall jitter. A TCF theoretically eliminates the time‐walk due to the variation in the amplitude of pulses of fixed risetime. However, the time‐walk effect attributed to the charge sensitivity of the zero‐crossing comparator remains. Similar to a simple LE discriminator, in a CFD at the zero‐crossing point, an additional charge is required to actually trigger the comparator. This dependency leads to the variations in the charging time, and consequently, the response time of the comparator is a cause of time‐walk in a CFD. Time‐walk can be in the range of several tens to several hundred picoseconds for input signals with a 40 dB dynamic range and 1 ns risetime [19]. An analysis of CFD timing errors caused by voltage comparators can be found in Ref. [20]. Some supplementary techniques have been used to reduce the time‐walk by using automatic gain control that reduces the amplitude variation in the input signals to a comparator and, consequently, the variation in its response time [20].

6.2.2.3 Amplitude and Risetime Compensated (ARC) Timing

The true CFD method described so far is most effective when used with input signals having not only a wide range of amplitudes but also a narrow range of risetimes and pulse widths like outputs of scintillator detectors. However, the output pulses of many detectors such as semiconductors are subject to risetime variations that make the TCF method unsuitable. A CFD can be made insensitive to both pulse‐height and risetime variations by operating it in ARC timing mode, which is achieved by setting the CFD shaping delay so that the time of zero‐crossing occurs before the attenuated input signal has reached its maximum amplitude. Figure 6.14 illustrates the signal formation in an ideal CFD for ARC timing with linear input signals of different risetimes and amplitudes. For ideal linear input signals that begin at time zero, the zero‐crossing time TARC from Eq. 6.17 is given by

(6.25)images
Signal waveforms in ARC timing mode CFD for ideal linear rising input pulses displaying input pulses, attenuated and inverted, delayed input, and CFD pulses (top to bottom) with an arrow labeled time pick-off.

Figure 6.14 Signal waveforms in ARC timing mode CFD for ideal linear rising input pulses.

When the risetime of input pulses is variable, the criterion for td to ensure ARC timing is

(6.26)images

where tr(min) is the minimum expected risetime for any input signal [7, 9]. Opposite to TCF method, in ARC timing, the fraction of the input pulse amplitude at which the time pick‐off signal is generated is not constant. The effective triggering fraction for each input pulse is related to the attenuation fraction by the input signal risetime. For linear input signals, the effective ARC timing triggering fraction is given by [9]

(6.27)images

This value is always less than f, while in TCF is always equal to f. One should note that the immunity to risetime variations described for ideal linear rising input pulses does not hold for a general case because in practice deviations from linear rise may occur, introducing timing errors.

For linear rising input signals, the jitter due to electronic noise in ARC timing can be obtained by having the pulse slope and noise. By assuming a linear input pulse of amplitude V and risetime tr, the slope of the CF signal at zero‐crossing point is given by

By combining Eqs. 6.3, 6.22, and 6.28, we have

(6.29)images

A comparison of this equation with that of TCF indicates that for identical input signal and noise conditions and for the same attenuation fraction, the jitter effect in ARC timing is worse than that for TCF timing due to the smaller slope of the ARC zero‐crossing signal. However, the ARC timing remains beneficial when the time‐walk due to risetime variations dominates the timing performance.

6.2.2.4 Practical CFD Circuits

The design of fast timing circuits requires a high level of designer skill as such circuits are very susceptible to the effects of stray parameters, ringing, and spurious signals. The interaction of active devices with the circuit layout, stray inductance, and capacitances and the presence of interference signals all place real limits on circuit performance. The principal functions that are generally performed in a CFD include attenuation, delay, inversion, summing, arming a zero‐crossing detector, detection of the zero‐crossing of the CFD bipolar pulse, and development of an output logic pulse. Most commercial CFDs have also provisions for adjusting the threshold level, the shaping delay, and the walk. The fraction is generally set at 0.2 or 0.3 that can be adjustable.

The initial CFD circuits were built around a tunnel diode, but nowadays high‐speed monolithic comparators are commonly used. The simplest CFD circuit based on monolithic comparators is shown in Figure 6.15a [7, 9, 21]. The circuit is basically the same as that of Figure 6.12a, but it is equipped with an arming circuit, comprised of an LE discriminator to inhibit the output timing pulse unless the input signal exceeds the arming threshold. This prevents the sensitive zero‐crossing device from triggering the baseline noise events. The lowest threshold setting is generally a few millivolt and is determined by the characteristics of the LE comparator. The CFD bipolar pulse is formed actively in the input differential stage of the lower comparator (CF). A monitor signal can be taken at the output of the CF comparator. A one‐shot circuit is placed in series with the output circuit to prevent multiple output signals from being generated from one input. A disadvantage of this circuit is that large time‐walk can result for small and slow rising input pulses, which is explained as follows: neglecting the propagation delays of the active circuit components, the input signal is applied to the CF circuitry and to the LE discriminator simultaneously. The LE discriminator produces an output pulse that crosses the logic threshold of the zero‐crossing AND gate at time tLE, while the CFD bipolar signal crosses the logic threshold of the zero‐crossing gate at time tcf. If tLE < tcf, then the time of occurrence of the logic pulse at the output of the zero‐crossing AND gate is tcf, but for input pulses with amplitudes near the LE discriminator threshold and for input signals with exceptionally long risetimes, the time at which the input signal actually crosses the LE discriminator threshold can come after tcf. Consequently, the output logic pulse will be based on tLE and major timing errors due to LE time‐walk can occur. To improve this situation, in the circuit shown in Figure 6.15b, an additional LE discriminator is used. The lower‐level leading‐edge (LLLE) discriminator arms the zero‐crossing AND gate G1, while the upper‐level leading‐edge (ULLE) discriminator determines if the input signal is of sufficient amplitude to produce an output pulse from the instrument. The lower‐level threshold is internally set at one‐half the upper‐level threshold. Consequently, signals that satisfy the upper‐level threshold requirement exceed the lower‐level threshold by a factor of two. This dual comparator arrangement greatly reduces LE time‐walk for pulses that just slightly exceed the upper threshold level [9, 21]. Figure 6.15c shows the block diagram of a CFD in which the CFD bipolar pulse is formed passively by using a differential transformer of wide bandwidth [9]. The arming and zero‐crossing detector circuits are the same as in Figure 6.15a. The monitor output is also picked‐off at the input to the CF comparator. There are several other arrangements and designs for CFDs. In some of the designs, more LE comparators are added to allow the selection of an energy range of interest [9, 22]. Such circuits are called differential CFDs and produce an output timing pulse when the input signal exceeds the lower‐level threshold and does not exceed the upper‐level threshold. Another facility provided in some designs is the possibility of rejecting slow risetime events [23]. Since slow risetime events are inevitably accompanied with some timing errors, the rejection of such events helps to maintain the good time resolution achievable with faster pulses of semiconductor detectors that exhibit large risetime variations. However, this is achievable at the cost of some efficiency. To achieve a good time resolution, it is also necessary to stabilize the zero‐crossing point that may be disturbed due to parameters such as input signal dc level, ac interferences, input pulse shape, and high input frequencies. Special stabilization mechanisms have been used to minimize the time‐walk that results from zero‐crossing point fluctuations [24–26]. The walk error of a CFD is generally used to specify its performance, since it indicates the change in the timing moment as a function of the amplitude of the input timing pulse. The walk error of the CFDs can be measured by detecting the change in the peak location of a prompt coincidence time spectrum as a function of the amplitude of one of the input pulses.

Image described by caption and surrounding text.

Figure 6.15 The block diagram of some commercial CFD designs by ORTEC [7, 9].

6.2.2.5 Monolithic Time Discriminator Circuits

There are applications such as particle physics experiments and medical imaging in which the large number of detector channels necessitates the use of integrated front‐end circuits with accurate timing capability. To provide such timing signals, both LE and CFD methods have been implemented in ASIC systems based on complementary metal–oxide–semiconductor (CMOS) technology. Implementation of LE method has been rather straightforward by implementing fast comparators [27], while that of CFD method requires monolithic‐compatible solutions to replace the shaping delay that in discrete circuits is simply implemented by using a coaxial cable of a particular length [28–30]. Some proposed solutions for a monolithic‐compatible CFD pulse shaping are shown in Figure 6.16 [30]. Figure 6.16a shows a simple CFD pulse shaping method based on a CR differentiator [31]. Since a derivative of a signal with peak exhibits a sign change, a CR differentiator will produce a bipolar signal while no fraction circuit is required. However, in this method the slope through the zero‐crossing and overdrive is dependent on the trailing edge of the input pulse, which can result in more timing jitters and possibly time‐walks [31]. Figure 6.16b shows the lumped‐element RC shaping network [32], which generates a delay signal with a single RC low‐pass filter. The disadvantages of this method are signal amplitude and slope degradation due to the low‐pass filtration. Figure 6.16c shows the distributed RC delay‐line shaping network [33]. In this method, the delay signal for shaping is realized with a serpentine layer of polysilicon over a second grounded polysilicon plate. The polysilicon line is a lossy transmission line with delay and dispersion. Figure 6.16d shows the Nowlin method [34], which utilizes the CR differentiation with a fraction circuit. The attenuated signal is subtracted from the differentiated signal to produce the CFD bipolar pulse. This shaping circuit can be also optimized to produce ARC timing measurement. There are also other non‐delay‐line CFD configurations based on multiple pole low‐pass filters that are discussed in Ref. [29]. However, one should note that none of these methods can achieve the same performance as the ideal delay‐line method.

Image described by caption and surrounding text.

Figure 6.16 Some CFD shaping methods for integrated time pick‐off systems [30–33].

Implementation of integrated CFD circuits faces other limitations such as offset variations that can be high in CMOS circuits due to process parameter variations. The presence of offset voltage at the input of the zero‐crossing discriminator can destroy the amplitude invariability, creating residual time‐walk. Therefore, static or dynamic offset compensation schemes are often required to cancel dc offsets. In an integrated CFD circuit developed for use with scintillators in PET systems, the same principles as that in Figure 6.15a are used, but a five‐order low‐pass filter is used to produce the required delay, and both the CF and arming circuits are equipped with continuous‐time baseline restorer circuits to cancel dc errors due to circuit offset and event rate effects [35]. Examples of ASIC CFD systems based on a high‐pass CR circuit for generating a CFD bipolar pulse can be found in Refs. [36, 37].

6.2.3 Timing Filter Amplifiers

As it was discussed in section, the slope‐to‐noise ratio of a pulse can be improved by using a proper pulse filter before feeding it into a time pick‐off circuit. This generally happens when signals are delivered by a charge‐sensitive preamplifier where pulses are considerably contaminated with electronic noise. The output pulses of a charge‐sensitive preamplifier are generally in the range of several tens of millivolt, and some amplification is also required before feeding the pulses into a time discriminator. In practice, filtration and amplification of preamplifier pulses for timing measurements is performed by using a timing filter amplifier (TFA) [38]. Figure 6.17 shows a simplified circuit diagram of a TFA. It consists of integrator and differentiator sections that are implemented by means of RC and CR circuits. The time constants of the integrator and differentiator are adjusted to reduce the high and low frequency noise, respectively. The differentiation also reduces the duration of preamplifier pulses. The integrator time constant should be optimized in order to not impair the slope‐to‐noise ratio by unnecessarily slowing down the pulse. The required gains can be provided with suitable feedback resistors on the fast operational amplifiers.

Image described by caption and surrounding text.

Figure 6.17 Block diagram of a timing filter amplifier.

6.2.4 Timing Single‐Channel Analyzers

The LE and CFD timing methods are the two common methods of time pick‐off in high accuracy timing measurements. However, in some applications, only a crude approximation of a pulse arrival time is sufficient. This generally happens when the timing and energy information need to be related together (see, e.g., Figure 6.33). A timing single‐channel analyzer (TSCAs) is a signal channel pulse‐height analyzer that, in addition to pulse amplitude discrimination, provides information on the arrival times of pulses. The time pick‐off may be performed in different ways. Figure 6.18 illustrates three methods of time pick‐off from slow pulses. In Figure 6.18a and b, the TSCA utilizes the zero‐crossing time of a bipolar pulse from a pulse shaping amplifier to derive time pick‐off, while the peak amplitude of the pulse is used for energy discrimination. The production of bipolar pulse can be based on a doubly differentiated or double delay‐line‐shaped preamplifier pulse. A double delay‐line‐shaped signal performs no integration on the pulses and provides better timing resolution. Integration significantly increases the risetimes of the pulses, resulting in a large timing jitter. Figure 6.18c illustrates the trailing‐edge CF technique that can be incorporated in the TSCAs. The advantage is that it can be used with either unipolar or bipolar pulses to derive a time pick‐off pulse after the peak time of the pulse delivered by a shaping amplifier. In this technique, the linear input signal is stretched and attenuated and then used as the reference level for a timing comparator. The time pick‐off signal is generated when the trailing edge of the linear input signal crosses back through the fraction reference level. As it is seen in Figure 6.18, all these time pick‐off methods remove the amplitude‐dependent time‐walk, but the time pick‐off signal is dependent on the shape of the input signal and the slope‐to‐noise ratio is poor compared with LE and CFD methods.

Image described by caption and surrounding text.

Figure 6.18 Methods of time pick‐off from a slow pulse. (a) Zero‐crossing of double‐differentiated pulse. (b) Double delay‐line method. (c) Trailing‐edge constant fraction discrimination.

6.3 Time Interval Measuring Devices

6.3.1 Time‐to‐Amplitude Converters (TACs)

A TAC measures the time interval between pulses to its start and stop inputs and generates an analog output whose amplitude is proportional to the time interval. TAC circuits have been long used for high precision timing measurements (<100 ps) when the time interval is shorter than 100 ns [39]. The heart of a TAC unit is the analog part that converts the time interval into a pulse amplitude. Figure 6.19 illustrates the principle of conversion. It basically includes two switches, a precision current source and a converter capacitor. The start and stop switches are typically designed to accept the fast logic signals from timing discriminators. The switches are initially closed. On the arrival of the LE of the start pulses, switch S1 opens and the converter capacitor begins to charge at a rate set by the constant‐current source. On the arrival of the stop pulse, switch S2 opens and prevents any further charging of the capacitor. Because the charging current I is constant, the voltage developed on the capacitor is given by

where Tstart is the time start pulse received, Tstop is the time stop pulse received, and C is the capacitance of the converter capacitor. Equation 6.30 indicates that the output voltage is proportional to the time interval between the start and stop pulses. This voltage pulse is buffered and amplified to a suitable level and passed through a circuit that results in a rectangular output pulse with a width of a few microseconds and amplitude proportional to the time interval between the input pulses as TAC output. After this time, all the switches return to the closed condition, and the capacitor is discharged to the ground potential to be ready for the next pair of start and stop pulses. The TAC output pulse is generally processed by an ADC or an MCA to produce a time spectrum. The accuracy of the measurement primarily depends on the precision of current source and capacitor and thus TAC can deliver an exceptionally fine time resolution (~10 ps). This means that the measured time resolution is practically controlled by the timing errors originated from the start and stop time discriminators. A proper operation of TAC however requires appropriate logic structure, which is described in Refs. [40–43].

Basic configuration of a simple TAC displaying 2 curves along a dashed line labeled stop and start, 2 switches labeled S1 and S2 with connecting line to the capacitor, and a triangle labeled Buffer.

Figure 6.19 The basic configuration of a simple TAC.

The important properties of a TAC are linearity, pulse handling capacity, resolution, and stability. The linearity of a TAC can be measured with two pulse generators and an MCA. One pulse generator feeds pulses to the start channel of the TAC at a constant rate, while the other pulse generator feeds randomly distributed pulses to the stop channel of the unit. The random pulse generator can be produced by using a PMT with a radiation source or a noise generator. The TAC output pulses are then analyzed with the pulse‐height analyzer. In the ideal case, the output of pulse‐height analyzer should be uniform within the statistics, but in practice deviations are observed due to nonlinearities. Both differential and integral nonlinearities can be calculated from the distribution [40]. The differential nonlinearity (DNL) of a channel is calculated as the relative deviation of the channel width from the average channel width, while the INL is the largest deviation of any single measurement that results from the calibration line fitted to the measurement points [40]. In general, the linearity of a TAC circuit is limited when the start and stop arrival times are nearly equal. In this case, S1 and S2 are trying to switch at nearly the same time, and the current onto C is not stable for this short time duration, leading to nonlinearities. This problem is minimized by transferring the time intervals of interest to the most linear part of the time range of the TAC by using proper delay values. The TAC circuits have been also implemented as monolithic devices. In a topology proposed by Tanaka et al. [44], to overcome the nonlinearity at the low end‐of‐conversion range, two conventional TAC circuits like the one shown in Figure 6.19 are used, while a third clock signal produces the start and stop signals required by the TACs. The first and second TACs each produce an output that is proportional to the difference of the arrival times of the input start and input stop pulses, respectively, and the clock signals. The nonlinearity is minimized by taking the difference of the two TAC signals by a differential amplifier as the final TAC output. However, the stability of these circuits depends upon the matching and temperature tracking of the current sources and capacitors of the individual TACs, and also a differential amplifier with its associated errors is required. Figure 6.20 shows the block diagram of a modified version of this TAC circuit that avoids the problems due to capacitor mismatch and requires less circuitry [45].

Block diagram of a differential current-mode TAC displaying 2 curves labeled stop and start, 2 switches labeled S1 and S2, and other circuit parts labeled VDD, Reset, C, Vout, etc.

Figure 6.20 Block diagram of a differential current‐mode TAC [45].

A high pulse handling capacity is very important in many applications such as nuclear physics experiments where the time intervals are usually randomly distributed. Commercial TACs usually have a rather poor pulse handling capacity, and the dead time of a TAC for every start pulse is typically ~1 µs, which can cause significant losses at high counting rates. In particular, if a start pulse is not followed by a stop pulse, the TAC will spend a lot of time responding to this pulse that will cause excessive dead time in the TAC without producing useful data. Such situations may be experienced when a very high start pulse rate is accompanied with low stop pulse rate, and thus, to minimize the dead time, the higher pulse rate should be fed into the stop input even if the physics of experiment indicates the opposite order of start and stop events.

6.3.2 Time‐to‐Digital Converter (TDC)

To measure a time interval in digital form, one obvious method is to digitize the output of TAC by using an analog‐to‐digital converter (ADC). This approach yields a good precision but it is slow and expensive if several channels should be processed. Instead, a time‐to‐digital converter (TDC) can directly convert the time intervals into a digital output. TDCs were originally developed for particle physics experiments and have found applications in nuclear physics and radiation imaging systems. There are many methods for time‐to‐digital conversion that are summarized in a book [46] and several reviews [47, 48]. The simplest TDC is based on counting clock pulses from a stable reference oscillator during the time interval. The count of clock pulses starts with the start pulse, and at the arrival of the stop pulse, the counter stops, yielding a number proportional to the time interval between the pulses. This method has excellent linearity and the dynamic range can be large, depending on the bit capacity of the counter, but time resolution is limited to ±1 clock cycle, for example, even with a clock frequency of a few gigahertz, a resolution of ~1 ns can be achieved. To get better resolutions, the use of interpolation methods is necessary. An interpolation method is illustrated in Figure 6.21. The time interval between the rising edges of start and stop signals can be divided into three parts: T1, T2, and T12, as shown in Figure 6.21. The relation between the time intervals is written as

(6.31)images
Timing diagram of a counter-based TDC and interpolation method displaying square waves with double-headed arrows for CLK, start, stop, fractional period T1 and T2, and time measured by the counter (top to bottom).

Figure 6.21 Timing diagram of a counter‐based TDC and interpolation method.

The major time interval T12 is synchronous with the precise reference clock and is digitized simply by the main counter, while interpolation techniques are used to measure the fractional periods T1 and T2. In a well‐known interpolation method first proposed by Nutt [49], the time phase between the start signal and the oscillator is measured by switching a charging current into a capacitor at the time of the start signal. At the time of the oscillator first pulse, the charging current is switched off and only a small discharging current is fed to the capacitor. This current, which is a known fraction of the charging current, discharges the capacitor to zero. As the capacitor voltage goes beyond zero, an end‐of‐conversion pulse is generated. The interval between the time when the charging current is switched off and the end‐of‐conversion pulse represents a known expansion of the time between the oscillator first pulse and the start pulse that lets T1 be determined. A similar procedure is used to measure the time between the stop pulse and the following oscillator pulse, T2. The time interval between the start pulse and the first oscillator pulse is added to the main time interval, and the time interval between the stop pulse and the following oscillator pulse is subtracted from the main time interval. A very good time resolution of subpicosecond has been reported with such interpolation technique [50].

Another TDC method is the vernier technique whose basic principle is illustrated in Figure 6.22 [47]. In this approach, two oscillators of slightly different time periods (T1 and T2) are used to measure the time difference T between start and stop pulses. The slow oscillator with period T1 begins oscillating when the start pulse arrives, while the fast oscillator with period T2 begins oscillating when the stop pulse arrives. The pulses from the fast and slow oscillators are counted by a coarse and a fine counters, respectively. Since T2 < T1 and as stop pulse arrives later, eventually the fast oscillator will coincide or lead the slow oscillator, which is detected by a phase detector. The time interval T can be then calculated as

(6.32)images

where n1 and n2 are, respectively, the number of counts in coarse and fine counter at coincidence and ΔT is the difference in the periods of the oscillators.

Image described by caption and surrounding text.

Figure 6.22 Timing diagram of a TDC with vernier principle.

TDC units have been widely used in data acquisition systems of modern PET imaging systems [28]. In such systems each detector channel may be equipped with a TDC where a digital time stamp of the event is produced by measuring the time interval between a reference clock signal and a trigger signal issued by an ASIC CFD [51, 52]. The complexity and large number of channels in a typical PET system necessitates the use of multichannel TDCs with low power consumption and subnanosecond precision. Such systems are built based on integrated CMOS or field programmable gate array (FPGA) technologies. Deep‐submicron CMOS technologies with high density and low mass‐production costs are very promising in realizing TDCs with fine resolutions. The FPGAs also offer a simple implementation and verification platform and low prototype manufacturing costs. The TDC methods applicable to FPGA are summarized in Ref. [53]. The quality of a TDC is judged by similar parameters as an ADC such as resolution, robustness, integral and DNLs, dead time, and other parameters that will be discussed in Chapter 9. Resolution is the smallest time interval that can be digitized in a TDC. Dynamic range refers to the maximum measurable time interval using the TDC circuit. Robustness is specified over a range of variations in process, voltage, and temperature. DNL is the difference between a real least significant bit (LSB) and an ideal LSB. INL is the difference between a real transition point and an ideal transition point. Dead time is defined as the minimum time delay between two successive digitization operations.

6.3.3 Coincidence Units

As it was already mentioned, in many experiments, it is only required to determine if input pulses represent coincident events. This is equal to applying an AND logic operation to the input logic pulses. An AND coincidence gate can be constructed with several inputs. Figure 6.23 illustrates a two‐input AND gate and a simple coincidence circuit based on diodes. The AND gate generates a logic 1 output only when logic 1 pulses are present on all inputs, which means the output is generated only for the time during which the input A and B pulses overlap. This can be achieved, for example, with the simple circuit shown in Figure 6.23, where the bias of the diodes necessitates that all the diodes simultaneously receive input pulses to change the output voltage status. One can see that for logic input pulses the resolving time (±τ) is 2τ = δ1 + δ2 where δ1 and δ2 denote the width of input pulses that are normally equal for standard pulses. In nuclear instrument modules, coincidence modules are sometimes classified as slow‐coincidence and fast‐coincidence circuits. In slow‐coincidence method, the width of the input pulses is directly used in a time‐overlap evaluation, but in fast‐coincidence method, pulses of standard width are used for each input, and any overlap of the standardized pulses is detected. This approach allows adjusting the width of coincidence window by feeding the standard pulses into an internal resolving time network, where pulse widths are varied to satisfy the 2τ resolving time according to the experiment requirement. The outputs of the resolving time networks are then fed into an AND network, and when portions of the reshaped input pulses overlap each other, the AND circuit recognizes a coincidence event within the resolving time. There are other methods for detecting coincidences such as additive type in which logic input pulses are added together, and if they overlap, the resulting pulse will exceed a threshold level, leading to an output pulse [54]. For detecting truly coincident pulses, the delays through the electronics producing the pulses must be the same for both detectors, and the width of each pulse must be equal to the maximum timing uncertainty for its respective detector. If the pulse width is too narrow or the delays are not quite matched, some of the truly correlated pulses will not overlap, leading to a loss of coincidence detection efficiency. On the other hand, if the logic pulses are too wide, uncorrelated events will have a higher probability of generating an output due to accidental overlap.

Image described by caption and surrounding text.

Figure 6.23 An AND gate representing an overlap coincidence unit with two inputs (left) and a simple AND gate circuit based on diodes and resistors (right). If no pulses arrive at inputs, then the diodes will be conducting and the output voltage will be maintained at zero. If a pulse of amplitude V arrives at input channels, the diodes in the other channel will continue to conduct and the output of the system will remain at zero. But if all input logic pulses arrive simultaneously, no diode will conduct and a logic pulse of amplitude V will now be generated at the output of the circuit, indicating that a coincidence has been detected.

In some applications it is important to identify the events that are not accompanied with other events, for example, to suppress unwanted background events. In such situations anticoincidence circuits are used. As it was discussed in Chapter 2, the logic of a two‐input anticoincidence circuit is designed so that an output logic pulse is generated only when a signal is received at one input and none simultaneously at another input (see Chapter 2). Such logic function can be obtained by combining AND, OR, and NAND gates as shown in Figure 6.24.

Image described by caption and surrounding text.

Figure 6.24 An anticoincidence circuit.

6.3.4 Multichannel Scaler

A TAC or TDC can measure only a single time interval for each start pulse. This limits the data acquisition rate in the measurement of long time intervals, particularly when the arrival time of the events from a repetitive trigger signal needs to be measured, for example, a pulsed source of radiation. In such situations, when time intervals greater than 10 µs should be recorded, a multichannel scaler (MCS) is an advantageous option. In an MCS, after a trigger or common start signal, the events are counted as a function of their arrival times on the counting input of the MCS. The result is a spectrum of the number of events versus the time after excitation. An MCS after each repetition of a trigger signal scans through its channels, spending a fixed amount of time on each, which is called the dwell time. For each channel, it adds the number present at its input to the total for that channel. After a certain number of scans, the measurement is completed and the data are read out to produce the time spectrum. In principle, an MCS unit can be considered a clock generator, addition logic and memory, bookkeeping logic to keep track of the number of scans, and readout memory. MCS units are generally built as add‐ons to MCAs, but separate instruments have been also developed [55]. The performance of an MCS in measuring shorter time ranges is limited by the intrinsic time resolution set by the dwell time that has been reduced to 50 ns in some systems. Moreover, the period between start pulses must be longer than the time interval being measured plus any end‐of‐scan dead time. For the measurement of time intervals less than 1 µs a TAC is the better option, while for time ranges from 1 to 10 µs, other factors will generally determine whether the TAC or the MCS is more appropriate.

6.3.5 Delay Elements and Signal Transmission

In performing timing measurements, it is often required to delay analog or logic pulses for purposes such as aligning fast timing channels to operate coincidence circuits, calibrate the timing equipment, and operate CFDs, TACs, and so on. The common solution to signal delay is to use coaxial cables. As it was discussed in Chapter 2, coaxial cables such as RG‐58A/U 50 Ω cables produce signal delays of 5.05 ns/m, and thus certain length of cables can be used to produce the desired amount of delay. The use of coaxial cables for pulse delay is inexpensive and familiar and can have reasonably low attenuation and acceptable pulse risetimes for resistive terminations. Coaxial cables are also used to interconnect timing circuits. In using long cables, one should consider the undesirable effects that may arise from the delay, attenuation, waveform alterations, and reflection of pulses, as it was discussed in Chapter 2.

In some applications, the use of coaxial cables for delay purposes is problematic due to the fact that long length of cables is bulky and difficult to handle and store in large quantities. An alternative option is to use lumped‐constant delay lines (LC delay) [56]. An LC delay line consists of a series of low‐pass reactive filters composed of discrete series inductors and shunt capacitors as shown in Figure 6.25. Such transmission lines are characterized by the time delay, that is, the time required for a pulse to propagate from one end to the other, and by a characteristic impedance Z, which is analogous to the characteristic impedance of a coaxial cable and is equal to the resistance required to terminate the line without causing pulse reflection. By using the image impedance techniques, one can easily calculate the impedance of the line to be [57]

where ωc is the cutoff frequency given by ωc = 2(LC)−1/2. For pulses with frequency contents well below the cutoff frequency, the image impedance is nearly constant, and a finite series of sections may be terminated with a resistor of value (L/C)1/2. For such inputs, the attenuation and distortion in pulse shapes is insignificant and it can be shown that the delay of the structure is approximately

(6.34)images

where n is the number of LC sections. If the input pulse contains appreciable components above the cutoff frequency, the nonlinearities in the phase and the attenuation of the line can severely distort the output pulse. The highest frequency component, ωh, of a pulse may be approximated by the relation

(6.35)images

where tr is the risetime of the pulse estimated from the 10 to 90% of its LE. Thus, the fastest pulse that can be transferred without distortion can be estimated by setting ωc = ωh. Pulse delay can be also produced by using distributed delay lines that can be considered as a continuous limit of a LC network. A coaxial cable is one example of a distributed delay line. For a continuous limit, the inductance and capacitance of each element tends to zero so that in Eq. 6.33, Z → (L/C)1/2 and ωc → ∞. This implies that the phase is a linear function of frequency so that the delay is independent of frequency. As discussed in Chapter 2, the attenuation of these lines at high frequencies is predominantly due to the resistive losses in the conductors and shows some frequency dependence due to the skin effect. Dispersive and absorptive losses can be significant at very high frequencies but are generally negligible for frequencies less than 300 MHz. An example of commercially available distributed delay lines is the Spiradel delay line, which consists of an inductor up to several tens of centimeters long, combined with a capacitance that is distributed uniformly along its length. Size reduction is accomplished by winding the inductor on a flat thin strip and reducing the coaxial sheath to a multiple flat strip conductor, placed parallel to the strip inductor. These long lengths are then wound into a spiral and encapsulated. They have a resistive characteristic impedance that is independent of frequency and so they can be terminated in resistive loads. Besides passive delay generators, active circuits have been used to delay logic pulses. In a circuits descried in Ref. [58], a long delay cable is simulated by making a signal go through a short cable a certain number of times.

A lumped delay line displaying from "In" (left) to 8 capacitors labeled C and 12 inductors labeled L with 2 curves in between and then to "Out" (right).

Figure 6.25 A schematic diagram of a lumped delay line.

6.3.6 Non‐detector Trigger Signals

In many experiments it is required to measure the time interval between a detector pulse and a reference time that is not produced by a detector. For example, in nuclear physics time‐of‐flight experiments with pulse beam accelerators, the beam arrival time is used to provide the zero time reference. A method for producing a fast start signal from a charged particle beam is to use a cylindrical capacitor. As the beam pulse passes through a cylinder, a current pulse is induced in the wire connected to the capacitor that can be amplified and applied to a time discriminator to define the beam arrival time [59]. In many types of experiments using particle beams provided by cyclotrons, accurate time measurements relative to the arrival time of the bunched beam on target is to measure the arrival time of reaction products relative to the radio frequency (RF) of the cyclotron. The RF signal conveniently yields a timing signal for every individual beam bunch [60]. One of the most common ways to determine the energy of a thermal neutron is to measure its time of flight over a measured distance. In the measurements of the velocity spectrum of neutrons from a reactor by timing the flight of individual neutrons over a distance, a start signal can be produced by chopping the neutron beam [61]. A neutron beam chopper is a disk of neutron‐absorbing material that rotates about an axis above and parallel to the neutron beam line, as illustrated in Figure 6.26. The disk carries radial slots around the periphery of the disk assembly, and thus every time the neutron beam crosses the slit, a burst of neutrons is admitted to the flight path. The reference time for the origin of each of the neutron bursts can be produced by the periodic interruption of a light beam passing parallel to the neutron beam line between a tiny photodiode and photodetector. The light signal is used to set the time origin of a MCS that is fed by the pulses from a neutron detector.

A neutron time-of-flight spectrometer with neutron beam chopper displaying parts labeled Start signal, Stop signal, Photodetector, Photodiode, Neutron beam, Neutron detector, and Flight path.

Figure 6.26 Schematic diagram of a neutron time‐of‐flight spectrometer with neutron beam chopper.

6.4 Timing Performance of Different Detectors

It was already discussed that the timing performance of a detector is evaluated by the time resolution determined from a prompt time spectrum. In principle, the time resolution of a detector can be measured by using the setup shown in Figure 6.27. The setup can contain two identical detectors of unknown time resolution or a detector of unknown time resolution that is placed against a detector of known time resolution. In either case, Eq. 6.1 can be used to determine the contribution of the detector(s) to the overall time resolution. As illustrated in Figure 6.25, the overall time resolution is governed by three factors: (1) fluctuations in the interaction times of the radiations with the detectors, (2) time variations in the response of the detectors to the radiations, and (3) time variations associated with the pulse timing electronic system. The first group of timing errors may stem from variations in the source‐to‐detectors flight times of the radiations. Such timing errors can be minimized by a proper design of the measurement system, leaving a basic limitation of the obtainable time resolution. The errors due to electronic system may result from time pick‐off units, electronic noise, cables, and time interval‐measuring devices such as TACs and TDCs. The contribution due to the detectors can result from the fluctuations in the shape, risetime, and amplitude of the pulses and is obviously dependent on the type of the detector. In the following sections, we will separately discuss the origin of such timing uncertainties in scintillator, semiconductor, and gaseous detectors.

Basic setup for the measurement of the time resolution of a detector (1 and 2) displaying boxes connected by arrows labeled Time pick-off, Delay, and TAC + MCA or TDC.

Figure 6.27 A basic setup for the measurement of the time resolution of a detector(s).

6.4.1 Timing with Scintillator Detectors

Scintillators are coupled to photodetectors such as PMTs, photodiodes, avalanche photodiodes (APDs), and silicon photomultipliers (SiPMs). The time resolution of a scintillator detector results from both the scintillator and photodetector. We first consider the contribution from the scintillator, and the effects of different types of photodetectors will be separately discussed in the following sections. The concept of intrinsic time resolution of a scintillator is illustrated in Figure 6.28. The intrinsic time resolution is defined as the timing uncertainty associated with scintillation photons generated in the scintillator that reach to an ideal photodetector followed by an ideal time discriminator. Under such conditions, the time resolution originates from the statistical fluctuations of the registration times of individual scintillation photons that are converted to photoelectrons in the photodetector at the early stage of the signal generation where the ultimate timing information lies. The intrinsic time resolution is the theoretical limit of timing resolution that a practical scintillator detector can achieve. In 1950, Post and Schiff [62] first discussed the limitations on time resolution that arise from the statistics of photon detection. They assumed that, following the excitation of the scintillator by an incident radiation, the photodetector converts the collected scintillation photons to photoelectrons without time spread and that the resulting output pulse is fed into an ideal discriminator that triggers when a definite number of photoelectrons have been collected. From Poisson statistics, they showed that the probability that the Nth photoelectron occurs between t and t + dt can be described with

where f(t) is the average or expected number of photoelectrons emitted between the initial excitation at t = 0 and time t. For a scintillator producing on the average R scintillation photons per excitation and with zero risetime and an exponential decay time constant of τ, one can write f(t) = R(1 − et/τ), and Eq. 6.36 becomes

Intrinsic time resolution of a scintillator with a shaded part of a rectangle indicated by an arrow labeled Ideal photodetector with connecting line to a triangle labeled Ideal discriminator and Threshold level.

Figure 6.28 The intrinsic time resolution of a scintillator.

If the time discriminator is triggered when its input pulse reaches the voltage corresponding to the Nth photoelectron, then the distribution of triggering time should be given by Eq. 6.37. It can be shown that for various values of R, the best time resolution is obtained for N = 1 though in practice this may be inhibited by the effect of electronic noise [63]. For N = 1, the time distribution is nearly a pure exponential with effective time resolution of τ/R, which means that a good time resolution is directly related to the light output and decay time constant of the scintillator, and the time resolution improves by increasing the energy deposition in the detector. This timing model for slow scintillators such as NaI(Tl) with coincidence timing resolution in the range of 2–4 ns or larger leads to very accurate results because the statistics of photoelectron production dominates other effects involving the time resolution. However, for fast scintillators with subnanosecond time resolutions, for example, LSO and LaBr3 : Ce, an accurate modeling of timing should take into account a large number of detector and photodetector properties. One of these parameters is the scintillator risetime as the light pulse always has a nonzero risetime due to a complicated luminescence process that leads to the scintillation [64]. The risetime can play a determinant role in the timing resolution as it affects the photoelectron density in the early stage of the signal. A long risetime reduces the rate of the detected photons in the initial phase of the light pulse, which has the same effect on the timing resolution as a reduced light output. The effect of finite risetime has been included in a bi‐exponential timing model, leading to better accuracy in the calculation of the time resolution of fast scintillators [65]. In addition to the scintillator intrinsic risetime, as illustrated in Figure 6.29, the fluctuations in the transit times of scintillation photons from the emission point to the photodetector can affect the observed risetimes [66]. This effect becomes important when good time resolutions, for example, better than 200 ps, are required and is controlled by parameters such as scintillator’s geometry, reflective coating, and polish. In particular, by increasing the size of the scintillator, the risetime increases, and also the self‐absorption and reemission processes can increase the decay time of the output pulses that all result in larger time jitter.

Illustration of variations in light collection times on the time response of a scintillator displaying a rectangle with arrows labeled Reflector and Scintillator and a shaded area for photodetector (bottom).

Figure 6.29 An illustration of variations in light collection times on the time response of a scintillator.

6.4.1.1 Scintillators Coupled to PMTs

The shape of output pulses from a scintillator coupled to a PMT is affected by the time response of the PMT, and the effect becomes particularly important with fast scintillators. The contribution of a PMT to time resolution is mainly associated with the following [67, 68]:

  1. The spread in the transit time of electrons traveling from photocathode to the anode. The amount of transit time spread generally depends upon photomultiplier geometric characteristics and its operating conditions.
  2. The numbers of photoelectrons released from the photocathode, which is a function of the photocathode sensitivity and efficiency of photoelectrons collection.
  3. The spread in the gain of the electron multiplier that translates to fluctuations in the amplitude of the output pulses.

In addition to these parameters, the frequency bandwidth of the PMT and its output circuitry may affect the timing performance. Figure 6.30 illustrates the time response characteristics of a typical PMT. The electron transit time is defined as the average time difference between the arrival time of a δ‐function light pulse and that of the corresponding output current pulse. The transit time spread is defined as the FWHM of the probability distribution of the fluctuations in the transit times. The larger the spread, the less time resolution the PMT has. The risetime is conventionally defined as the time the anode pulse takes to rise from 10 to 90% of its final value in response to a δ‐function input pulse and in fact indicates the frequency bandwidth limitation of the PMT. In addition to this intrinsic bandwidth limit, the output circuitry can place a frequency limit on the output pulses. The dependence of the frequency response on the external circuitry was already described with Eq. 3.79 and is a function of total capacitance and resistance at the PMT output. The total capacitance represents the combined capacitance of the tube and the stray capacitance of the cables and is generally very small, but the choice of load resistor needs some care. A large load resistor reduces the bandwidth but increases the output voltage. Increasing the load resistance also makes the potential drop between the last dynode and the anode small. The consequence of this is the buildup of space charge at the last dynode and poor collection of charges, leading to nonlinearity in the response. Therefore, one must choose an optimized load resistor value based on the requirements of the particular application.

Graph illustrating time response of a PMT to a δ-function incident light pulse, displaying a bell-shaped curve labeled FWHM and lines representing for output current, input light, transit time, time, etc.

Figure 6.30 The time response of a PMT to a δ‐function incident light pulse.

In the previous section, we discussed the intrinsic time resolution of a scintillator where it was assumed that there was no transit time spread in the photodetector. When the photodetector is a PMT, the effect of PMT time spread can be included by replacing f′(t) in Eq. 6.36 with the following convolution [69]:

(6.38)images

where I(t) is the probability density function of arrival times of photons at the photocathode produced by an event at t = 0 and Ppmt(t) is the probability density function of the transit time spread of electrons in the PMT. This modified version of timing theory has been successfully applied to predict the time resolution of scintillator detectors [70–72] with the result that the PMT time spread can be considerable even for slow scintillators. However, with modern fast PMTs used with slow scintillators such as Nal(Tl), the PMT probably has a negligible effect on the final time resolution. But for fast scintillators, achieving minimum time resolution requires a PMT with minimum time spread.

6.4.1.2 Effect of Pulse Processing System

The time resolution contribution coming from the timing electronics circuits is generally very small and may be neglected in many applications. But in fast timing measurements in the range of tens of picoseconds, the intrinsic time resolution of the electronics may be considerable [73]. The time fluctuations may arise from time pick‐off circuits, TACs and ADCs or TDCs, signal dividers, fan‐in/fan‐out circuits, and so on. Therefore, for fast scintillators, it is crucial to reduce the electronics modules as much as possible. The most critical part of a timing setup is the time pick‐off circuit. The popular timing methods with PMTs are the LE and CF timing. The LE method produces very good results in measurements with the narrow dynamic range of input pulse amplitudes [74]. To achieve the optimum time resolution, one needs to set the threshold level of the discriminator at maximum slope‐to‐noise ratio that generally corresponds to between 5 and 20% of pulse amplitudes. The choice of a threshold levels reduces time‐walk, but the risk of noise‐induced triggers increases. Although the time‐walk error from an LE discriminator can be also corrected by using the information of the pulse amplitudes offline, in most of the situations, the CFD method is preferred because it cancels the time‐walk online and thus more suitable for large dynamic range measurements. However, in the measurements with fast scintillators such as LaBr3(Ce), the time‐walk due to the charge sensitivity of the comparator may still become quite noticeable. This time‐walk is dependent on the amplitude of the pulses, that is, the energy of input pulses, and thus it can be measured as a function of energy. The effect of charge sensitivity time‐walk on the calibration of time spectra and its correlation to PMT operation voltages has been studied in regard to its application in nuclear structure fast timing measurements [75].

A problem encountered in timing with slow scintillators is multiple triggering [26]. This is associated with the fact that as a result of the slow decay time of scintillators such as NaI(Tl), the last portion of each anode pulse consists of individual single‐photon pulses. Therefore, the discriminator will trigger once on the LE of the anode pulse and then multiple times at the end of the anode pulse, which can seriously affect the measurement. Multi‐triggering is also a problem with fast scintillators such as BaF2 that have a slow component as well. To remedy the multi‐triggering problem, a non‐extending dead time is deliberately imposed to the CFD that makes it insensitive to the following individual single‐photon pulses. The length of the dead‐time duration should be long enough compared with the pulse duration and can be ~1 µs for NaI(Tl) detectors.

6.4.1.3 Fast–Slow Measurements

In many radiation detections systems, a simultaneous extraction of both timing and energy information is required. This also allows the measurement of the time resolution as a function of the energy range of input signals. In such applications, the pulse processing arrangement includes a fast and a slow pulse processing channels that extract, respectively, the timing and energy information [22]. Figure 6.31 shows two arrangements for fast–slow measurements with scintillators coupled to PMTs. In the top panel of the figure, an integral mode CFD is used as the time pick‐off device in each branch, and the slow channel is composed of a preamplifier, a shaping amplifier, and a SCA to select pulses in the energy range for which timing information is desired. The shaping amplifier is generally a delay‐line amplifier due to insignificant effect of electronic noise. If two detected events fall within the selected energy ranges, and if they are coincident within the time widow adjusted on the coincidence unit, then precise timing information related to these events is strobed from the TAC and the resulting signal is stored in an MCA. This system can provide both excellent energy selection and timing characteristics but at high count rates since the TAC must handle all start–stop pairs regardless of their energy or coincidence, a count rate limitation is imposed by the TAC, and also the heating effects in the active circuitry may become serious. In the arrangement shown in the lower panel of Figure 6.31, different CFDs are used that generate the timing information and determine the energy range of interest simultaneously. In this case, a CFD output is generated if two detected events fall within the selected energy ranges, and if they are coincident within the resolving time of the coincidence unit, then TAC is gated on to accept the precise timing information. This reduces the events rate of the TAC as it must only handle start–stop signals for events that satisfy the energy and time restrictions. One should note that in this approach the energy discrimination is generally performed based on the amplitude of PMT current pulses as CFD does not integrate the pulses, yet still the energy selection accuracy is satisfactory for most applications.

Image described by caption and surrounding text.

Figure 6.31 Fast–slow measurements with scintillator detectors.

6.4.1.4 Scintillators Coupled to Photodiodes and Avalanche Photodiodes

Scintillators coupled to photodiodes have been used in some timing applications mainly due to advantages such as compactness, insensitivity to magnetic fields, and high quantum efficiency. Figure 6.32 shows an arrangement for timing measurement with photodiodes. The signal from photodiode is generally read out by using a charge‐sensitive preamplifier [76]. As it was discussed in the previous chapters, the preamplifier output signal is contaminated with the electronic noise that is determined by the diode leakage current, capacitance, and input transistor of the preamplifier. The time jitter due to the electronic noise imposes a severe limitation on the performance of a time discriminator, and thus a TFA is used to improve the slope‐to‐noise ratio by performing independent integration and differentiation on the preamplifier signal. The output of the filter is then fed into a time discriminator. The output of the charge‐sensitive preamplifier can be also simultaneously processed in a slow channel to extract the energy information. In spite of the noise reduction act of TFA, the timing performance of photodiodes is still insufficient for many applications. For example, the time resolution of a combination of CsI(Tl) and photodiode at 60Co energies was reported to be in the range of above 100 ns [77] though better time resolutions may be achieved for very large amounts of energy depositions in the scintillator [77, 78].

Block diagram of a timing setup with scintillators coupled to photodiodes and APDs, from curve arrows labeled Scintillation photons at the photodiode and bias to TFA (triangle) and to CF or LE discriminator (box).

Figure 6.32 Block diagram of a timing setup with scintillators coupled to photodiodes and APDs.

In the case of APDs due to their internal gain, the effect of electronic noise is significantly reduced. Modern APDs, in addition to high gain (>200), exhibit small leakage current, low excess noise factor, and high quantum efficiency that make them an attractive option for use in applications involving pulse timing. The principle of deriving a time pick‐off signal from APDs is the same as that from photodiodes as shown in Figure 6.32 [79–82]. If an LE discriminator is used for time pick‐off, then the contribution of electronic noise to the time resolution can be estimated from Eq. 6.3. The slope of the pulse is determined by the charge collection time in the diode, the bandwidth or risetime of the preamplifier attached to the diode, the decay time constant of the scintillator, and the TFA shaping time constants. The charge collection time in the APDs is generally very fast; in the order of a few nanoseconds and by using a preamplifier with large bandwidth, for example, 100 MHz, one can also preserve the pulse risetime. If the absence of preamplifier and shaping effects, for the fully integrated pulse from a scintillator, dV/dt in Eq. 6.3 is proportional to the number of electron–hole pairs Ne−h and inversely proportional to the decay time of the scintillator τ [83]:

(6.39)images

The rms value of noise at the input of the time discriminator is the combination of noise due to the electronic noise from APD‐preamplifier and photon noise due to the light collection process [84]. The noise behavior of an APD‐preamplifier system in most respects is similar to that of semiconductor detectors. The electronic noise rms value can be calculated from the electronic noise power spectral density, referring to the input of the APD and the frequency response of the entire preamplifier–shaper system. In regard to photon noise, however, since the current due to scintillation light, Iphoton(t), decays exponentially with time, the noise is not stationary. Therefore, the mean‐square noise due to the photon flux should be calculated in the time domain from Campbell’s theorem:

(6.40)images

where w(t) is the weighting function or the time domain response of the system, F is the excess noise factor, and q is the electric charge. The total value of root‐mean‐square noise voltage at the discriminator is then

(6.41)images

where ve is the rms value of electronic noise. By having dv/dt and ve, the time jitter can be estimated from Eq. 6.3. The effect of electronic noise can be considerably reduced by cooling the device, thereby improving the time resolution [85]. In practice, the timing performance may be also affected with the nonuniformity and fluctuations in APDs gain and statistical error of the primary electron–hole pairs. The results of timing with LSO crystals coupled to APD show that subnanosecond timing is possible [83] for gamma rays though it is still inferior to that obtainable with PMTs. This is probably because PMTs provide an almost noiseless gain of around 106, while an APD provides only a gain of around 102, and thus with APDs, many photoelectrons must be collected before sufficient charge has been collected to trigger a time discriminator. Nevertheless, the previously mentioned advantages of APD over PMTs have led to the use of such systems for PET imaging applications for which several ASIC signal processing systems have been developed to address several thousands of signal channels [36, 86].

6.4.1.5 Scintillators Coupled to SiPM

SiPMs have fast response and an amplification factor similar to PMTs (around 106) that make them very attractive for use in timing applications. Studies have shown that the contribution of SiPM to the time resolution of an SiPM‐based scintillator detector is mainly determined by its photon detection efficiency (PDE), single‐photon timing jitter, noise (mainly dark counts and optical cross talk), time transit delay, and single photoelectron gain [87–89]. A circuit of time pick‐off from a SiPM generally involves the use of a fast amplifier followed by an LE discriminator. The properties of the amplifier that convert the SiPM current pulse to a voltage pulse play a very important role in the timing performance [90, 91]. A dedicated, high‐bandwidth, low input impedance front‐end electronics with low noise is required to deliver the SiPM pulses to the discriminator with maximum slope. In particular, the combination of SiPM output capacitance with a large input impedance of the amplifier can form an RC circuit with long decay time constant that consequently slows down the pulses. Fast amplifiers suitable for SiPMs have been built based on monolithic components [91], ASIC systems [92], and commercially available RF amplifiers [90]. Excellent results have been reported for LE timing of SiPM coupled to fast scintillators. For example, Schaart et al. [91] reported 100 ps time resolution for 3 × 3 × 5 mm3 LaBr3(Ce) crystals coupled to Hamamatsu SiPMs at 511 keV gamma‐ray energy. However, these results have been achieved by using waveform digitizers and performing LE discrimination in digital domain. This has been attributed to the accuracy of baseline determination in digital domain that enables setting lower triggering thresholds. The recently developed digital SiPM employs a different approach of time pick‐off with that of analog SiPM. In such devices, each cell has its own readout circuit that is connected to an on‐chip TDC via a configurable trigger network to start the TDC at the first photon or higher photon levels. Time resolutions equivalent or better than analog SiPMs have been reported for digital SiPMs [93, 94].

6.4.2 Timing with Semiconductor Detectors

The timing performance of semiconductor detectors depends on both the detector’s pulse characteristics and preamplifier properties. The role of preamplifier is to preserve the timing information reflected in the risetime of the pulse. In principle, it is possible to exploit either the current pulse or the charge pulse of a detector for timing purposes. A current pulse is intrinsically fast because its risetime is only determined by the bandwidth of the system. However, the electronic noise can completely dominate the timing measurement, and thus it is only useful when the noise effect is insignificant, that is, when the energy deposition in the detector is large and charge collection time in the detector is very fast. The charge pulse on the other hand is produced by integrating the current pulse on the detector or preamplifier capacitance, and thus the risetime is equal to the width of the current pulse. However, the slope‐to‐noise ratio can be significantly higher than that of current pulses, and thus charge pulses are often used for timing purposes. In the choice of a charge‐sensitive preamplifier for timing measurements, its risetime needs to be only fast compared with the charge collection time inside the detector because if the charge collection time exceeds the risetime of the preamplifier, then increasing the bandwidth will only increase the electronic noise. A detector can be also simultaneously operated with both charge‐ and current‐sensitive preamplifiers as it will be discussed for silicon detectors.

6.4.2.1 Timing with Germanium Detectors

Timing measurement with germanium gamma‐ray detectors has been of interest particularly for nuclear physics experiments. With germanium detectors, the best time resolution can be achieved by deriving the timing signal from the output of a fast charge‐sensitive preamplifier. In principle, the timing performance of a germanium detector is limited by both time jitter and walk. The jitter results from the electronic noise at the output of the preamplifier, while the time‐walk stems from the fact that the shape of the charge pulses strongly varies with the interaction location of gamma rays. The time‐walk error normally overwhelms the noise jitter effect, and thus the timing technique must be focused on overcoming the time‐walk error. There have been several methods to minimize the effect of time‐walk including extrapolated LE method [13–17], ARC timing [6, 95], and more recently digital pulse timing techniques. Among the analog methods, the ARC timing method is the most widely used one. A system for the measurement of the time resolution of germanium detectors using ARC timing is shown in Figure 6.33 [21, 96]. The germanium detector is placed against a fast scintillator detector such as BaF2 or LaBr3(Ce). Since the time resolution of fast scintillators is much less than that of the germanium detector, the measured time resolution essentially reflects the time resolution of germanium detector. To derive a time pick‐off signal, the preamplifier output is presented to a TFA that performs integration and differentiation on the pulses to reduce the duration of pulses and optimize the slope‐to‐noise ratio. The TFA also provides the required gain before the timing discriminator. The TFA output is then used for time pick‐off with a CFD that is operated in ARC timing mode by setting the delay to ~30% of the minimum risetime of TFA output pulses. As it was previously discussed, ARC timing generates a timing marker independent of amplitude and risetime, provided that each pulse has a constant slope throughout its LE. However, real pulses from germanium detectors deviate from linear rise with slope changes before the time of zero‐crossing, and thus ARC timing cannot completely remove the time‐walk. Due to this limitation, the time resolution of large volume HPGe detectors is generally limited to about 8–10 ns.

Setup for time resolution measurement of a germanium detector displaying 11 blocks with connecting arrows labeled Scintillator, CFD, HPGe, TAC, MCA, TFA, Delay, etc.

Figure 6.33 A setup for time resolution measurement of a germanium detector.

6.4.2.2 Timing with Silicon and Diamond Detectors

Silicon detectors are widely used for timing measurements of charged particles, and depending on the experimental conditions, time resolutions below a few hundreds of picoseconds are quite common. The selection of suitable preamplifier depends on the delivery of a large signal with small risetime (or equivalently large slew rate dV/dt), and thus both charge‐ and current‐sensitive preamplifiers have been used, depending on the situations [97, 98]. If a charge‐sensitive preamplifier is used, the output is processed with a TFA followed by a CFD. The time resolution will basically depend on the charge collection time, noise, and uniformity of signals shapes. The minimum risetime of charge pulses is equal to the collection time of the electrons and holes in the detector sensitive region, which is determined by the distribution of the generated electron–hole pairs and of the electric field. The minimization of detector thickness reduces the charge collection time, but it increases the capacitance, thereby increasing noise. The effect of capacitance is particularly acute when very thin transmission silicon detectors are used in ΔE–E and time‐of‐flight measurements. An approach to reduce capacitance can be to segment the detector’s sensitive area while each segment is connected to its own readout electronics. The risetime of the pulses may be also affected with the plasma effect discussed in Chapter 1. A current‐sensitive preamplifier can be used when the energy deposition is large, and thus the slope‐to‐noise ratio is inherently high [99–101]. With current‐sensitive preamplifiers, the signals can be directly processed with a CFD. Although a current‐sensitive preamplifier can potentially lead to better time resolutions, the energy resolution will be inferior to that obtainable with charge‐sensitive preamplifiers, while in most experiments in nuclear physics, both good time and energy resolution are simultaneously required. In such cases, charge‐ and current‐sensitive preamplifiers can be simultaneously employed. The current pulse is used to pick off the time signal, while the charge pulse is used for energy measurements. There have been several methods for coupling a detector to both charge‐ and current‐sensitive preamplifiers [97, 102–104]. A common method of coupling a detector to two preamplifiers is shown in Figure 6.34 where a contact of the diode is grounded not directly but via the input of a fast low noise current‐sensitive preamplifier.

Arrangement for simultaneous derive of charge and current pulses from a silicon detector to energy measurement and to timing measurement.

Figure 6.34 Arrangement for simultaneous derive of charge and current pulses from a silicon detector.

Diamond detectors have attracted considerable interest for use in timing measurements of charged particles. The high mobility of charge carriers in conjunction with very high breakdown electric fields and a low dielectric constant provides detector signals of fast risetimes suitable for timing measurements. Moreover, the wide bandgap of diamond produces very low leakage current noise. Diamond detectors are generally used with current‐sensitive preamplifiers in timing measurements, and the main challenge is to preserve the slope of fast current pulses. In this regard, the bandwidth of the preamplifier and the total capacitance at the detector–preamplifier connection play a very important role [105, 106]. With the simplifications that the output current pulse of the preamplifier has a linear rising edge, the slope of the pulse can be estimated as dv/dt = 0.8V/tr where V is the amplitude of the output pulse and tr is the preamplifier risetime (10–90% amplitude level). For a single pole preamplifier, the bandwidth (BW) is related to risetime with BW = 0.35/tr. Then, the slope of the pulse can be written as dv/dt = 2.28Q·BW/C where Q is the total charge deposited in the detector and C is the total input capacitance. By assuming that the noise results only from the bias resistor and the preamplifier is noiseless, the time jitter from Eq. 6.3 can be obtained as

(6.42)images

where the noise rms value has been estimated from Eq. 2.75. This simple relation reflects the importance of the reduction of the total capacitance including the diamond capacitance, the preamplifier input capacitance and parasitic capacitances, and the role of the bandwidth of the system and the amount of charge deposition. This discussion can be extended to a more accurate definition of the noise contribution of the preamplifier that can be found in Ref. [105]. Many studies have reported time resolutions in the range of sub 100 ps for diamond detectors [107].

6.4.2.3 Timing with Compound Semiconductor Detectors

Timing with semiconductor detectors such as CdTe, CdZnTe, TlBr, and HgI2 has been of interest for use in PET imaging systems. Compound semiconductors are available in different structures such as planar, coplanar, strip, and pixelated. For planar detectors, the conventional ARC timing method as discussed for germanium detectors is commonly used, resulting in time resolutions of 10 ns for 1 mm thick CdTe detectors at 511 keV energy [108]. The poor timing performance stems from the significant variations in the shape of pulses that are due to the large difference in the mobility of electrons and slow drifting holes and electronic noise mainly due to the leakage current of the detectors. Improvement in the timing performance has been achieved by cooling the detectors. For compound semiconductor detectors employing single polarity charge‐sensing technique, in spite of improvement in the energy resolution, the timing performance is generally poor due to the limit imposed by the pulse generation mechanisms. For example, in coplanar detectors, the output pulse is obtained by subtracting the induced pulses on collecting electrode and non‐collecting electrodes, eliminating the hole component of the pulse, which is the major cause of the degradation of energy resolution. However, the subtracted pulse appears with a delay with respect to the time of the gamma‐ray interaction in the detector because electrons generated by the gamma‐ray interaction need to travel from their creation point to the vicinity of the anode grids. Therefore, the variation in the interaction locations sets a basic limit on the timing performance of the detectors. Similarly, in pixelated detectors, the electrons need to travel from the interaction location to the vicinity of the anode to give rise to charge pulses, and thus, in spite of pulses of relatively fast risetime, the time resolution remains inherently limited, larger than 10 ns [109, 110].

6.4.3 Timing with Gaseous Detectors

Gaseous detectors have been mainly used in timing measurements involving charged particles. The timing performance largely depends on the detector operating region, electrodes geometry, gas pressure, applied voltage, electronic noise, and amount of energy deposition. In the following we briefly describe the timing performance of different gaseous detectors. Frisch grid ionization chambers are used for timing of heavily ionizing particles such as fission fragments. Both charge‐ and current‐sensitive preamplifiers have been used in these applications, and subnanosecond time resolutions have been reported [111, 112]. Single‐wire proportional counters filled with BF3 or He3 gases are used for timing measurements of slow neutrons. The time resolution is inherently limited by the geometrical distribution of primary electrons throughout the counter volume as they must drift to the vicinity of anode wire to participate in the gas amplification process [113]. The drift paths have a distribution from zero up to the maximum, given by the inside radius of the cathode. Moreover, the output signal in such counters is induced by the slow‐moving positive ions, and thus the limited slope of the pulses combined with the effect of electronic noise limits the accuracy of time pick‐off. In spite of these limitations, the time resolutions are sufficient for slow neutron time‐of‐flight measurements where time intervals in the range of microseconds are measured. MWPCs operating at normal pressures also have a poor timing performance, that is, time resolution on the order of several tens of nanoseconds, due to the long drift time of electrons released in the sensitive volume toward the sense wire, where amplification occurs. This drift time is dependent upon the applied field, electrode spacing, wire pitch, gas pressure, and type of gas [114, 115]. Micropattern gaseous detectors have much faster charge collection times than traditional wire chambers, and thus subnanosecond time resolutions can be achieved by using the electron component of current pulses [116, 117]. Among these parameters the gas pressure plays a very important role so that the time resolution of a very slow detector operating at normal pressures (time resolutions above 100 ns) may be improved to subnanoseconds by operating at low gas pressures (below 20 torr) [118–120]. Low pressure gaseous detectors have been generally used in parallel‐plate and MWPC geometries and are common for fast timing measurement of heavily ionizing particles. At low gas pressures, the reduced electric field (E/P) becomes sufficiently large so that the gas amplification process takes place throughout the detector’s sensitive volume, and thus the effect of variations in the drift times of electrons to the amplification region is eliminated. In such detectors the electron component of the current pulses is sufficiently large to be measured with a proper preamplifier; for example, a design can be found in Ref. [121]. The risetime of the electron component of the current pulse can be estimated because, due to the exponential behavior of gas amplification, the bulk of electrons are produced at the last electron mean free path, and thus the risetime of the electron component of the pulse is approximately given by

(6.43)images

where ve is the drift velocity of electrons and α is the first Townsend coefficient. For a typical electron drift velocity of 5 cm/µs and Townsend coefficient of 100 cm−1, a risetime of 2 ns is achieved, leading to time resolutions as good as 100–200 ps.

References

  1. 1 G. F. Knoll, Radiation Detection and Measurement, Fourth Edition, John Wiley & Sons, Inc., New York, 2010.
  2. 2 E. Sattler, Nucl. Instrum. Methods 64 (1968) 221–224.
  3. 3 A. Borghesi, et al., Nucl. Instrum. Methods 137 (1976) 605–607.
  4. 4 H. Murayama, Nucl. Instrum. Methods 177 (1980) 433–440.
  5. 5 N. Y. Kim, et al., Nucl. Instrum. Methods A 563 (2006) 104–107.
  6. 6 Z. H. Cho and R. L. Chase, Nucl. Instrum. Methods 98 (1972) 335–347.
  7. 7 T. J. Paulus, IEEE Trans. Nucl. Sci. 32 (1985) 1242–1249.
  8. 8 H. Spieler, IEEE Trans. Nucl. Sci. 29 (1982) 1142.
  9. 9 ORTEC, “Principles and Applications of Timing Spectroscopy,” Application note, AN42, 1982.
  10. 10 T. D. Doughlass and C. Williams, Nucl. Instrum. Methods 66 (1968) 181–192.
  11. 11 E. Gatti, et al., Nucl. Instrum. Methods A 457 (2001) 347–355.
  12. 12 W. Blum, W. Riegler, and L. Rolandi, Particle Detection with Drift Chambers, Springer‐Verlag, Berlin, Heidelberg, 2008.
  13. 13 J. P. Fouan and J. P. Passerieux, Nucl. Instrum. Methods 62 (1968) 327–329.
  14. 14 B. Gottschalk, Nucl. Instrum. Methods 190 (1981) 67–70.
  15. 15 S. Bose, et al., Nucl. Instrum. Methods A 295 (1990) 219–223.
  16. 16 S. Y. Kim, et al., Nucl. Instrum. Methods A 414 (1998) 372–385.
  17. 17 A. R. Frolov, et al., Nucl. Instrum. Methods A 356 (1995) 447–451.
  18. 18 D. A. Gedde and W. J. McDonald, Nucl. Instrum. Methods 55 (1967) 377–380.
  19. 19 H. Lim, IEEE Trans. Nucl. Sci. 61 (2014) 2351–2356.
  20. 20 D. M. Binkley and M. E. Casey, IEEE Trans. Nucl. Sci. 35 (1988) 226–230.
  21. 21 M. O. Bedwell and T. J. Paulus, IEEE Trans. Nucl. Sci. 23 (1976) 234–243.
  22. 22 M. O. Bedwell and T. J. Paulus, IEEE Trans. Nucl. Sci. 26 (1979) 422–427.
  23. 23 W. J. McDonald and D. C. S. White, Nucl. Instrum. Methods 119 (1974) 527–532.
  24. 24 J. Bialkowski, et al., Nucl. Instrum. Methods A 228 (1984) 110–117.
  25. 25 J. Bialkowski and W. Schoeps, Nucl. Instrum. Methods A 270 (1988) 481–486.
  26. 26 J. Bialkowski, et al., Nucl. Instrum. Methods A 281 (1989) 657–659.
  27. 27 T. O. Tumer, et al., IEEE Nuclear Science Symposium Conference Record, Vol. 1, October 29–November 1, 2006, pp. 384–388.
  28. 28 K. Iniewski (Editor). Electronics for Radiation Detection, CRC Press, Hoboken, 2010.
  29. 29 D. M. Binkley, IEEE Trans. Nucl. Sci. 41 (1994) 1169–1175.
  30. 30 R. G. Jackson, IEEE Trans. Nucl. Sci. 44 (1997) 303–307.
  31. 31 M. L. Simpson, et al., IEEE Trans. Nucl. Sci. 42 (1995) 762–766.
  32. 32 B. T. Turko and R. C. Smith, Conference Record of 1991 IEEE Nuclear Science Symposium and Medical Imaging Conference, November 2–9, 1991, pp. 711–715.
  33. 33 M. L. Simpson, et al., IEEE Trans. Nucl. Sci. 43 (1996) 1695–1699.
  34. 34 C. H. Nowlin, Rev. Sci. Instrum. 63 (1992) 2322–2326.
  35. 35 D. M. Binkley, et al., IEEE Trans. Nucl. Sci. 49 (2002) 1130–1140.
  36. 36 V. C. Spanoudaki, et al., Nucl. Instrum. Methods A 564 (2006) 451–462.
  37. 37 M. Tanaka, et al., IEEE Trans. Nucl. Sci. 39 (1992) 1321–1325.
  38. 38 L. Karlsson, Nucl. Instrum. Methods 107 (1973) 375–380.
  39. 39 R. G. Roddick and F. J. Lynch, IEEE Trans. Nucl. Sci. 11 (1964) 399–405.
  40. 40 J. Kostamovaara and R. Myllyla, Nucl. Instrum. Methods A 239 (1985) 568–578.
  41. 41 M. Bertolaccini and S. Cova, Nucl. Instrum. Methods 121 (1974) 547–556.
  42. 42 I. J. Taylor and T. H. Becker, Nucl. Instrum. Methods 99 (1972) 387–395.
  43. 43 W. M. Henebry and A. Rasiel, IEEE Trans. Nucl. Sci. 13 (1966) 64–68.
  44. 44 M. Tanaka, et al., IEEE Trans. Nucl. Sci. 38 (1991) 301–305.
  45. 45 J. M. Rochelle and M. L. Simpson, IEEE Conference on Nuclear Science Symposium and Medical Imaging, October 25–31, 1992, pp. 468–470.
  46. 46 S. Henzler, Time‐to‐Digital Converters, Springer Series in Advanced Microelectronics, Vol. 29, Springer‐Verlag, New York, 2010.
  47. 47 D. I. Porat, IEEE Trans. Nucl. Sci. 20 (1973) 36–51.
  48. 48 J. Kalisx, Metrologia 41 (2004) 17–32.
  49. 49 R. Nutt, Rev. Sci. Instrum. 39 (1968) 1342–1345.
  50. 50 D. Vyhlidal and M. Cech, IEEE Trans. Instrum. Meas. 65 (2016) 328–335.
  51. 51 D. P. McElroy, et al., Phys. Med. Biol. 50 (2005) 3323.
  52. 52 A. Di Francesco, et al., Nucl. Instrum. Methods A 824 (2016) 194–195.
  53. 53 A. M. Amiri, et al., IEEE Trans. Instrum. Meas. 58, 3 (2009) 530–540.
  54. 54 P. W. Nicholson, Nuclear Electronics, John Wiley & Sons, Ltd, London, 1974.
  55. 55 D. M. Murphy, Nucl. Instrum. Methods A 236 (1985) 349–353.
  56. 56 J. P. Sandoval, et al., Nucl. Instrum. Methods 216 (1983) 171–176.
  57. 57 D. M. Pozar, Microwave Engineering, Second Edition, John Wiley & Sons, Inc., New York, 1998.
  58. 58 J. G. Marques and C. Cruz, Nucl. Instrum. Methods A 745 (2014) 50–56.
  59. 59 D. A. Gedcke and W. J. McDonald, Nucl. Instrum. Methods 56 (1967) 148–150.
  60. 60 M. Steinacher and I. Sick, Nucl. Instrum. Methods A 394 (1997) 305–310.
  61. 61 J. R. D. Copley, Nucl. Instrum. Methods A 273 (1988) 67–76.
  62. 62 R. F. Post and L. I. Schiff, Phys. Rev. 80 (1950) 1113–1113.
  63. 63 F. J. Lynch, IEEE Trans. Nucl. Sci. 13 (1966) 140–147.
  64. 64 P. Lecoq, et al., IEEE Trans. Nucl. Sci. 57 (2010) 2411–2416.
  65. 65 Y. Shao, Phys. Med. Biol. 52 (2007) 1103–1117.
  66. 66 W. Choong, Phys. Med. Biol. 54 (2009) 6495–6513.
  67. 67 M. Moszynski, Nucl. Instrum. Methods A 337 (1993) 154–164.
  68. 68 M. Moszynski, et al., Nucl. Instrum. Methods A 567 (2006) 31–35.
  69. 69 F. J. Lynch, IEEE Trans. Nucl. Sci. 22 (1975) 58–64.
  70. 70 E. Gatti and V. Svelto, Nucl. Instrum. Methods 4 (1959) 189–201.
  71. 71 L. G. Hyman, Rev. Sci. Instrum. 36 (1965) 193–197.
  72. 72 D. M. Binkley, IEEE Trans. Nucl. Sci. 41 (1994) 386–393.
  73. 73 J. W. Zhao, Nucl. Instrum. Methods A 823 (2016) 41–46.
  74. 74 V. Vedia, et al., Nucl. Instrum. Methods A 795 (2015) 144–150.
  75. 75 J.‐M. Regis, et al., Nucl. Instrum. Methods A 684 (2012) 36–45.
  76. 76 M. Suffert, Nucl. Instrum. Methods A 322 (1992) 523–525.
  77. 77 Y. K. Gupta, et al., Nucl. Instrum. Methods A 629 (2011) 149–153.
  78. 78 D. V. Dementyev, et al., Nucl. Instrum. Methods A 440 (2000) 151–171.
  79. 79 R. Lecomte, et al., Nucl. Instrum. Methods A 278 (1989) 585–597.
  80. 80 C. Carrier and R. Lecomte, Nucl. Instrum. Methods A 299 (1990) 115–118.
  81. 81 A. Q. R. Baron and S. L. Ruby, Nucl. Instrum. Methods A 343 (1994) 517–526.
  82. 82 J. A. Hauger, et al., Nucl. Instrum. Methods A 337 (1994) 362–369.
  83. 83 M. Moszynski, et al., Nucl. Instrum. Methods A 485 (2002) 504–521.
  84. 84 M. E. Casey, et al., Nucl. Instrum. Methods A 504 (2003) 143–148.
  85. 85 K. Inoue and S. Kishimoto, Nucl. Instrum. Methods A 806 (2016) 420–424.
  86. 86 D. M. Binkley, et al., IEEE Trans. Nucl. Sci. 47 (2000) 810–817.
  87. 87 S. Gundacker, et al., J. Instrum. 8 (2013) 1–28.
  88. 88 S. E. Drenzo, et al., Phys. Med. Biol. 59 (2014) 3261–3286.
  89. 89 S. Dolinski, et al., Nucl. Instrum. Methods A 801(2015) 11–20.
  90. 90 J. Y. Yeom, et al., Phys. Med. Biol. 58 (2013) 1207–1220.
  91. 91 D. R. Schaart, et al., Phys. Med. Biol. 55 (2010) N170–N189.
  92. 92 T. Ambe, et al., Nucl. Instrum. Methods A 771 (2015) 66–73.
  93. 93 G. Gundacker, et al., Nucl. Instrum. Methods A 787 (2015) 6–11.
  94. 94 D. R. Schaart, et al., Nucl. Instrum. Methods A 809 (2016) 31–52.
  95. 95 R. L. Chase, Rev. Sci. Instrum. 39 (1968) 1318.
  96. 96 T. J. Paulus, et al., IEEE Trans. Nucl. Sci. 28 (1981) 544–548.
  97. 97 H. Spieler, IEEE Trans. Nucl. Sci. 29 (1982) 1142–1158.
  98. 98 N. Cartiglia, et al., Nucl. Instrum. Methods A 796 (2015) 141–148.
  99. 99 N. Taccetti, et al., Nucl. Instrum. Methods A 496 (2003) 481–495.
  100. 100 P. Volkov, Nucl. Instrum. Methods A 451 (2000) 456–461.
  101. 101 V. Eremin, et al., Nucl. Instrum. Methods A 796 (2015) 158–164.
  102. 102 M. Moszynski and B. Bengtson, Nucl. Instrum. Methods 91 (1971) 73–77.
  103. 103 F. Vogler, et al., Nucl. Instrum. Methods 101 (1972) 167–170.
  104. 104 R. T. deSouza, et al., Nucl. Instrum. Methods A 632 (2011) 133–136.
  105. 105 M. Ciobanu, et al., IEEE Trans. Nucl. Sci. 58 (2011) 2073–2083.
  106. 106 M. Nakhostin, Nucl. Instrum. Methods A 703 (2013) 199–203.
  107. 107 M. O. Frégeau, et al., Nucl. Instrum. Methods A 791 (2015) 58–64.
  108. 108 K. Giboni, et al., Nucl. Instrum. Methods A 450 (2000) 307–312.
  109. 109 A. Groll, et al., IEEE Trans. Nucl. Sci., Vol. 1 (2017) 3–14.
  110. 110 P. Vaska, et al., IEEE Nuclear Science Symposium Conference Record, Vol. 5, October 23–29, 2005, pp. 2799–2802.
  111. 111 H. Rosler, et al., Nucl. Instrum. Methods 99 (1972) 477–486.
  112. 112 A. Chalupka, Nucl. Instrum. Methods 164 (1979) 105–112.
  113. 113 O. K. Harling, Nucl. Instrum. Methods 34 (1965) 141–145.
  114. 114 M. Breidenbachi, et al., Nucl. Instrum. Methods 108 (1973) 23–28.
  115. 115 L. Gruber, et al., Nucl. Instrum. Methods A 632 (2011) 69–74.
  116. 116 F. Sauli, Nucl. Instrum. Methods A 805 (2016) 2–24.
  117. 117 G. Charpak, Nucl. Instrum. Methods 478 (2002) 26–36.
  118. 118 A. Breskin, et al., IEEE Trans. Nucl. Sci. 27 (1980) 133–138.
  119. 119 A. Breskin and N. Zwang, Nucl. Instrum. Methods 144 (1977) 609–611.
  120. 120 K. Assamagan, et al., Nucl. Instrum. Methods A 426 (1999) 405–419.
  121. 121 U. Lynen, et al., Nucl. Instrum. Methods 162 (1979) 657–675.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset