You are on page 1of 10

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO.

6, DECEMBER 2011

3411

Pulse Shape Discrimination in Impure and Mixed


Single-Crystal Organic Scintillators
Natalia Zaitseva, Andrew Glenn, Leslie Carman, Robert Hatarik, Sebastien Hamel, Michelle Faust,
Brandon Schabes, Nerine Cherepy, and Stephen Payne

AbstractNeutron/gamma pulse shape discrimination (PSD)


utilized for detection of high-energy neutrons with organic
scintillators was investigated using a model system of mixed
diphenylacetylene-stilbene single crystals of different compositions. The results of the studies, which include experimental tools
of crystal growth and characterization combined with computer
simulation, showed that the presence of impurities with lower
bandgap energies can be a major factor influencing PSD properties of organic materials. Depending on the concentration, an
impurity may suppress or increase the rate of excited triplet state
interaction leading, respectively, to a complete disappearance or
enhancement of PSD, consistent with a percolation threshold. The
results are applied to produce novel materials with controlled
decay characteristics. Single crystals with a large fraction of
delayed light and enhanced PSD have been grown for high energy
neutron detection, while crystals with suppressed delayed light
were produced for use as low-afterglow scintillators for energetic
neutron detection in time-of-flight experiments.
Index TermsDiphenylacetylene, neutron detector, organic
scintillators, pulse shape discrimination, stilbene.

I. INTRODUCTION

HE most widely used method for high-energy neutron detection in the presence of gamma radiation background
utilizes the difference in the shapes of the scintillation pulses excited by neutrons (recoil protons) and -rays in organic scintillators. Pulse shape discrimination (PSD) phenomena discovered
and demonstrated many decades ago [1] are based on the existence of two-decay component fluorescence, in which, in addition to the main component decaying exponentially (prompt
fluorescence), there is usually a slower emission that has the
same wavelength, but longer decay time (delayed emission).
According to a commonly accepted mechanism [2], [3], the fast
component results from the direct radiative de-excitation of ex, while the slow component originates
cited singlet states

Manuscript received May 13, 2011; revised August 01, 2011; accepted
September 27, 2011. Date of publication November 09, 2011; date of current
version December 14, 2011. The work was performed under the auspices of
the U.S. DOE by Lawrence Livermore National Laboratory under Contract
DE-AC52-07NA27344. Financial support provided by the U.S. Department of
Energy Office of Nonproliferation Research and Development (NA-22) and
Department of Homeland Security (DNDO).
The authors are with the Lawrence Livermore National Laboratory, Livermore, CA 94551, USA (e-mail: zaitseva1@llnl.gov; glenn22@llnl.gov;
carman1@llnl.gov hatarik1@llnl.gov; hamel2@llnl.gov; mafaust@ucdavis.
edu; brandonschabes@gmail.com; cherepy1@llnl.gov; payne3@llnl.gov).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TNS.2011.2171363

Fig. 1. (a) Basic physical processes leading to the delayed fluorescence characteristic of neutron excitation of organics with phenyl groups; (b) average waveforms indicating different levels of delayed light in neutron and gamma scintillation pulses; ratio of signal integrals, R, collected in two time gates, Tail and
Total, is used for neutron/gamma pulse separation; (c) typical PSD pattern of a
stilbene crystal obtained by digitized separation of neutron and gamma pulses;
(d) PSD profiles of experimental data used for calculation of the PSD figure of
merit (FOM).

from the collisional interaction of pairs of molecules (or exci. The process is
tons) in the lowest excited -triplet states
illustrated by Fig. 1(a). Since triplets are known to be mobile in
organic compounds, the energy migrates until two triplets collide and annihilate to form one singlet excited state :
. The lifetime of the delayed emission is deterand the rate of
collisions.
mined by the lifetime of
The short range of the energetic protons produced from neutron
collisions yields a high concentration of triplets, compared to
the longer range of the electrons from the gamma interactions,
leading to the enhanced level of delayed emission with longer
decay times in neutron-induced pulses in comparison to those
produced by the gamma excitation [Fig. 1(b)].
From this general mechanism, a high concentration of hydrogen and the presence of aromatic rings are conditions for
neutron interaction and efficient fluorescence. However, its
relevance to other specific properties of organic materials, such
as physical state, molecular and crystallographic structure, or
presence of impurities is practically unknown. The problem
relates to the limited number of organic materials historically

0018-9499/$26.00 2011 IEEE

3412

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO. 6, DECEMBER 2011

studied for their PSD properties. The earliest work on scintillation decay [4] reported slow decay component measurements
in anthracene, trans-stilbene, p-terphenyl, p-quaterphenyl,
and diphenylacetylene single crystals. Later studies [5][7]
limited to the same set of crystals established the best ability
to discriminate between different types of radiation in stilbene,
but did not confirm detection of any delayed light or PSD
properties in diphenylacetylene. Since no additional studies had
been conducted to understand the differences, stilbene received
a reputation of the best organic scintillator for fast neutron
detection, while diphenylacetylene became a classic example
cited in the literature as an aromatic crystal without PSD [1],
[3, pp. 392 and 495], respectively.
A curious discrepancy of the original results relates to the fact
that stilbene and diphenylacetylene are very similar molecules,
differing only by the double and triple bonds between the
central carbon atoms [Fig. 2(a)]. Moreover, according to the
literature, they have nearly identical crystallographic structures [10], with only small differences in the corresponding
lattice parameters [Fig. 2(b)]. They also exhibit very similar
photoluminescence spectra, with stilbene emission slightly
shifted to the region of longer wavelengths [Fig. 2(c)]. These
similarities, however, were not reflected either in PSD properties or in the scintillation efficiency, also controversially
reported for diphenylacetylene as greatly superior or inferior
to that of stilbene [11][13]. The absence of a proper understanding of the connection between PSD and the specific nature
of organic compounds made it difficult to predict the possible
presence or absence of PSD properties in broader varieties of
organic materials, limiting the choice of organic scintillators
used for efficient neutron detection.
Modern high-speed waveform digitizers [14][16] allow for
easy separation of neutron and gamma pulses [Fig. 1(c)], enabling rapid characterization of PSD properties. Organic crystals required for the tests can be obtained by different methods,
among which melt growth techniques have been so far the most
common [13]. Recent developments in low-temperature solution growth [17], [18] enabled use of alternative methods, which
allow for faster and less expensive production of a wide range of
small-scale organic single crystals. With respect to the potential
for growth to large sizes, crystals grown by solution methods
may have an advantage in producing crystals with much lower
stress than is typical for highly anisotropic crystals grown by
Bridgman techniques. In our recent work, a broad survey of
about 150 types of solution-grown fluorescent organic crystals
was conducted with a dual goal of empirical identification of
new efficient PSD materials [17] and better understanding of
the mechanisms of scintillation [19]. The results of our studies
showed that, in addition to the molecular and crystallographic
structures, which determine exchange coupling and triplet mobility in organic materials, one of the most important factors influencing PSD is the effect of impurities. The detailed consideration of this effect made with stilbene-diphenylacetylene used
as a model system explained the controversy in the properties
observed with these compounds in the past. Applied to the recent survey, the results helped to understand variations of PSD
properties measured in different types of organic crystals. In a

broader sense, the results provided better understanding of energy migration and excited state interactions in organic crystals,
opening new opportunities for design and engineering of scintillating crystals with controlled decay times and efficient PSD.
II. EXPERIMENTAL AND COMPUTATIONAL METHODS
Raw material of trans-stilbene (96%99%), diphenylacetylene (DPAC, 98%) and other solid powders used for crystal
growth were purchased from different vendors (Aldrich, Alfa
Aesar, Acros Organics and TCI America). Prior to growth, all
compounds were partially purified by recrystallization from
toluene (Aldrich, 99.8%) solutions. A simple evaporation
technique similar to that described previously [19] was used
to grow small single crystals (12 cm in size) for the initial
tests. Mixed crystals were grown from solutions containing
DPAC and stilbene in different ratios. The chemical composition of the growth solutions and crystals was measured
by gas chromatography-mass spectrometry (GC-MS) using a
Hewlett Packard (HP) 6890 GC coupled to a HP5973 mass
selective detector with the injection temperature set at 250 C.
Chromatographic separation was achieved using a DB-5MS
0.25 mm id, 0.25- m film thickness), with
column (30 m
the GC oven ramped from 40 C to 250 C over 35.5 min.
The mass selective detection was run in full-scan mode (30 to
178 and
550 m/z). DPAC and stilbene were identified by
180, respectively, and authenticated with a NIST library search.
They were quantified using pure stilbene and DPAC standards.
Photoluminescence (PL) spectra were measured using a commercial Fluoromax-2 spectrometer. The scintillation efficiency
of crystals was evaluated from gamma light yield (LY) obtained
from the position of the Compton edge in the Cs spectra relative to the Compton edge of a pure stilbene crystal. For our
studies, 500 keV gamma equivalent is defined by 50% of the
Compton edge peak.
Neutron detection properties of grown crystals were studied
using the pulse shape discrimination (PSD) technique, which
allows for the separation of the scintillation pulses produced by
neutron and gamma events via the relative increase in delayed
light for neutron stimulations [9], [10]. The measurements were
Cf source shielded with 5.1 cm of lead,
performed using a
which reduced the gamma rates to the same order of magnitude
as neutrons, to irradiate crystals coupled to an R6231-S Hamamatsu photomultiplier tube (PMT). The signals collected at the
PMT anode were recorded using a high-resolution waveform
CompuScope 14200 digitizer with a sampling rate of 200 MS/s,
for offline analysis. The ability of crystals to discriminate beCf
tween the neutrons and gamma rays emitted from the
PSD discriminator
source was evaluated using the
Fig. 1(b). The waveforms were numerically integrated over two
and a subinterval
, corresponding
time intervals:
to the total charge and the delayed component of the signal, respectively. The value of the ratio of charge
for the two time intervals indicate whether the considered event
was likely produced by a neutron (high R value) or a gamma ray
(small R value). As reported previously [19], comparative PSD
tests made with solution and melt-grown stilbene did not reveal

ZAITSEVA et al.: PULSE SHAPE DISCRIMINATION IN IMPURE AND MIXED SINGLE-CRYSTAL ORGANIC SCINTILLATORS

3413

Fig. 2. Comparison of diphenylacetylene (DPAC) and stilbene crystals: (a) and (b) Molecular and crystal structures; lattice parameters for DPAC: a = 15.488 ,
b = 5.754 , c = 12.766 , 90 ,
113.36 ,
90 ; for stilbene: a = 15.478 , b = 5.66 , c = 12.287 , 90 ,
112.03 ,
90 ; ([8] and
[9], respectively). (c) Photoluminescence spectra of pure and mixed (18.7% of stilbene) crystals; stilbene excitation spectrum shows the overlap needed for the
excitation transfer from DPAC to stilbene molecules. The side of the background square in crystal photos is 6.5 mm.

any substantial influence of growth methods, if crystals of the


same size and purity were used for the measurements.
Density functional theory (DFT) calculations [20] were used
to obtain the relaxed geometry in the first triplet excited state
and in the ground state
of both stilbene and DPAC.
The triplet excitation energy was taken as the total energy difference between the triplet and the ground states at their respective
optimized geometries [21]. The calculations were performed
with the Vienna ab initio simulation package [22] using projector augmented wave pseudopotentials to represent the ionelectron interactions [23].
III. RESULTS AND DISCUSSION
The similarity between the structures of diphenylacetylene (DPAC) and stilbene molecules explains a known fact
that both compounds form nearly identical crystallographic
structures [Fig. 2(a) and (b)]. The relatively small difference in
the lattice parameters suggests that the compounds can form
solid solutions and mixed crystals. A proof of this expectation
was provided by the frequent growth of good quality, faceted
crystals, chemical analysis of which showed mixed composition with different concentrations of DPAC and stilbene.
More detailed crystallographic analysis is required to confirm
if stilbene and DPAC form truly mixed crystals with periodic
incorporation of the different molecules in the crystallographic
lattice. However, the fact that all crystals had similar shapes and
practically identical angles between the corresponding facets
was taken as an indication of relatively homogeneous structure,
at least in crystals of a cm scale, similar to those shown in
photographs of Fig. 2.

A. Delayed Light and Pulse Shape Discrimination


Measurements made with numerous stilbene crystals grown
in our experiments always revealed excellent PSD similar to that
shown in Fig. 1(c). The situation was different for DPAC crystals, which, depending on the origin and batch of the initial material, might have large variations of PSD from a good level to its
complete absence. Combined chemical and PSD analysis performed on the crystals showed that the variation of the properties related to the presence or absence of impurities. A surprising
result obtained in analysis of DPAC crystals without PSD was
that all of them contained stilbene in the amount of 1%2%. On
the contrary, a good level of PSD was typical only for DPAC
crystals Fig. 3(a), in which no traces of stilbene were detected.
More detailed experimental studies were conducted with
about 50 DPAC crystals intentionally doped with controlled
stilbene additions. The experiments confirmed the sharp
quenching effect of the stilbene impurity on the PSD of DPAC
in the range of stilbene concentrations of about 0.110%. As
a result of the PSD loss, the neutron and gamma peaks, which
are clearly separated in pure materials, become joined in one
narrow pattern corresponding to a sharply decreased fraction
of the delayed light Fig. 3(b). We confirmed that scintillators
without clear PSD respond to both gammas and neutrons
by increasing shielding and moderator without changing the
Cf source. For example, adding 5.1 cm of
distance to the
high-density polyethylene reduced the count rate to 67% of
nominal; the addition of 5.1 cm of Pb, for a total of 10.2 cm,
reduced the count rate to 70% of nominal. Further analysis
of the mixed crystals grown at higher stilbene concentrations
showed reappearance of the delayed light resulted first in a
broadening of the single pattern [Fig. 3(c)] followed by its

3414

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO. 6, DECEMBER 2011

rapid return to neutron/gamma separation [Fig. 3(d)] and PSD


similar or even exceeding that typical for the corresponding
stilbene crystals [Fig. 3(e)(f)].
For more detailed consideration of the physical processes
giving rise to this nonlinear dependence of the delayed light
on chemical composition, the experimental results were compared in the same -equivalent calibrated energy ranges of
500550 keV, using PSD profiles typical examples of which
are shown in Fig. 4(a)(c).
Fig. 5, presents the dependence of the delayed light fraction
Cf excitation as a funccalculated for the mixed field of
tion of stilbene concentration measured in a large number of
DPAC-stilbene single crystals. The left vertical axis displays the
neutron/gamma delayed light separation [S in Fig. 1(d)] as a
function of stilbene fraction (or 1DPAC fraction). The separation corresponds to a difference between the mean delayed
, for neutrons and gammas taken as
light fraction,
a distribution for a large number, generally 40 000, of wave, is taken as the integral of the
form acquisitions. Here,
PMT baseline subtracted waveform from 65 to 1750 ns, where
, is taken as the intethe peak sample sets the 0 time.
gral from -75 to 1750 ns. The timing values were selected to
produce approximately the best neutron/gamma PSD figure of
merit (FOM) [Fig. 1(d)] for high stilbene concentrations.
Since many of the crystals were too small for accurate energy calibration, we considered comparison of separations at
different energies between samples to be the largest source of
error. For each data point, corresponding to an individual single
crystal, the waveforms were grouped into five or six subsets
corresponding to uncalibrated energy bins of roughly equal statistics, and a separation measurement was performed for each
subset. The vertical error bars of each point indicate the standard
, of the subset separation meadeviation,
surements. Several of the crystals in this study showed clearly
anomalous PSD patterns, most likely indicative of structural or
chemical inhomogeniety, however, no data from such crystals
Cf
were included in this analysis. One significant issue in
mixed field data analysis is that it is difficult to know for certain on an event by event basis, if the waveform was generated
by a neutron or a gamma, particularly for the 0.2% to 20% stilbene region, which exhibits little to no PSD. To mitigate the
effects of remaining inhomogeniety and low PSD samples, we
estimate the mean amount of delayed light for neutron stimulations by first subtracting off an estimate of the, generally more
region of
Gaussian, gamma component. The low
the distribution was fitted with a Gaussian, and the Gaussian is
distribution. The mean
subtracted from the total
of the remaining distribution is taken as a measure of the neutron delayed light, and the mean of the fitted Gaussian is taken
as a measure of the mean gamma delayed light.
The gamma/neutron separation was additionally calcuCs
lated in terms of the FOM for crystals that had a good
Compton edge and exhibited good PSD [see Fig. 4(d)]. These
FOM values for 500550-keV gamma equivalent (near the
Cs Compton edge) as measured from double Gaussian fits
of the
distributions are included in Fig. 5 with
the corresponding right vertical axis. The vertical errors are
based on uncertainties from the fitter routine for the mean

and width parameters of the double Gaussian fit propagated


through the FOM formula. The data of Fig. 5 show an absence
of neutron/gamma separation and PSD in the region of low
stilbene concentrations followed by a sudden onset of delayed
light at approximately 20% stilbene, as well as a maximum
delayed light fraction soon after the sudden increase.
Although unexpected because of stilbene acting as a PSD
quenching impurity, the initial effect of PSD disappearance in
DPAC-stilbene mixed crystals can be easily understood on the
basis of energy transfer phenomena studied previously in various mixtures of organic scintillators [2]. It is well known that in
the mixed crystals, the impurity molecules with lower energies
of the excited states may act as traps for both singlet and triplet
excitations of the host. Calculations of the triplet excitation energies made in the current work find that stilbene has a smaller
triplet excitation energy (2.4 eV) than DPAC (2.7 eV). As
a result, stilbene molecules can act as traps that prevent the
DPAC triplet energy migration and annihilation needed for the
formation of delayed light in DPAC crystals containing stilbene
in small concentrations, corresponding to no PSD regions in
Figs. 3 and 5.
In mixed DPAC-stilbene crystals containing higher stilbene
concentrations, unexpected broadening of the experimental patterns are first observed, followed by a sharp rise of the PSD
[Fig. 3(c)(d)] at a certain threshold stilbene concentration (
20 in Fig. 5). It is obvious that the sudden reappearance of
the delayed light must correspond to a restored exciton mobility
resulting from the increased concentration of stilbene. The phenomenon is similar to other mixed systems, where a sudden rise
of certain properties was studied by application of the percolation theory [24]. Percolation is a well-known phenomenon first
observed for combined mixtures of metal and plastic spheres
in which conductivity suddenly appears at a threshold volume
fraction of the conductive metal spheres. It is described as a
geometric-statistical phenomenon whereby at the percolation
threshold the probability of the existence of at least one infinite conductive cluster of metal balls becomes unity. We analogously consider the case of the existence of an infinite conductive cluster of stilbene molecules also having a percolation
threshold where the triplet excited states can reach each other to
incur the triplet-triplet annihilation and generate delayed singlet
emission.
From the basic principles of percolation, the reappearance
of the delayed light is likely due to attaining the percolation
threshold in exciton migration. We hypothesize that in pure
DPAC, even if stilbene is present at very low concentrations, less
than 1 , the average distance between stilbene traps is much
larger than the excitation density, and so the host excitation still
can migrate producing triplet recombination needed for the formation of the delayed light and PSD in the initial DPAC crystals [Figs. 3(a) and 5]. At slightly higher stilbene concentrations
(before the percolation threshold), traps capture all host triplet
excitations. The triplets, however, cannot interact because of the
still large distances between stilbene molecules, acting as triplet
exciton traps, thus leading to a quenched migration, recombination and absence of PSD [Figs. 3(b) and 5 below 20%]. When
the concentration of stilbene in the mixed crystal becomes large
enough, high trap density leads to a dramatically increased prob-

ZAITSEVA et al.: PULSE SHAPE DISCRIMINATION IN IMPURE AND MIXED SINGLE-CRYSTAL ORGANIC SCINTILLATORS

3415

Fig. 3. Energy-calibrated experimental patterns showing variation of PSD in mixed DPAC-stilbene crystals: (a) neutron/gamma discrimination properties of pure
DPAC (b) vanish at low concentrations of stilbene; (c) the delayed light reappears at higher concentrations (d-e) with further increase to an excellent separation
similar to that of (f) pure stilbene.
Cf source used for the measurements.

ability of stilbene network formation. Then, the stilbene traps


act as an excitation transport medium resulting in a sudden rise
of PSD. And finally, when stilbene network reaches infinite dimensions required for the efficient triplet migration and interaction, the scintillation pulses become more uniform, resulting in
clear peak separation, which goes through a maximum followed
by a slow drop with increasing stilbene concentrations beyond
50%.
Fig. 6 displays a similar dependence of the delayed light from
triplet excitation annihilation calculated for pure gamma excitations in the same sets of the mixed DPAC-stilbene crystals. The
data were obtained using nearly the same method as used with
the Cf source, except here the vertical axis corresponds to the
mean of the
distributions for Cs gamma stimulations. The fraction of delayed light in this case shows a clearer
50 more delayed light for a 65:35%
maximum; there is
DPAC-stilbene crystal than for pure stilbene. Fig. 6 includes a
fit based on the function for infinite cluster probability from the
percolation theory combined with dropping effective triplet confor
centration. The functional form is
and 0 otherwise. The best-fit parameters from a
minimization are
,
and
. We note that
factor. This
pure percolation theory does not include the
factor is added to account for the effect that once the concentration is high enough to form a fully connected stilbene network,
additional stilbene decreases the effective triplet concentration.
is in the range of FCC site fracThe percolation threshold,
tion (0.198) and BCC bond fraction (0.1803) percolation, while
the critical exponent, , is relatively close to the value of 0.41
expected for 3-dimensional percolation [25]. We concede that
our simplified model cannot adequately describe the full concentration range. It only attempts to model the emissions from
stilbene, so the delayed light from DPAC at the lowest stilbene

fractions is not included. Also, improvements will be needed to


better describe the 5 to threshold region, which may require
effects of nonstatistical stilbene incorporation (i.e., aggregates)
and PSD from statistically formed finite stilbene clusters. The
later dependence may be closely related to the mean cluster size
behavior described by percolation theory.
B. Scintillation Light Yield
The scintillation light yield (LY) measured in the mixed
DPAC-stilbene crystals of different composition is presented
in Fig. 7. The LY of a pure stilbene crystal grown from the
same material as in the mixtures was taken as a reference point
of a unit (corresponding to about 13 000 Ph/MeV). The large
scatter of the experimental data most likely results from the
difference in size (meaning different extent of self-absorption)
and insufficient optical quality of crystals grown by the evaporation technique. However, all measurements made with large
number of crystals, obtained in several independent sets of
growth experiments show a clearly pronounced dependence, in
which a relatively modest LY of pure DPAC crystals sharply
increases almost by an order of magnitude with very small
additions of stilbene, and then gradually decreases, with the
changing composition, to the LY of pure stilbene.
These trends observed in the LY may be explained on the
basis of classical understanding of energy transfer phenomena
influencing the emission efficiency in the mixed organic scintillators. As mentioned above, the impurity molecules with
the lower energy band gaps may trap both singlet and triplet
excitons of the host that absorbs most of the initial excitation.
Fast de-excitation of the trapped singlet states then occurs in
a regular way, producing scintillation corresponding to the
relative efficiency of the host and impurity molecules. Based on
this mechanism, the initial excitation energy of wider bandgap

3416

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO. 6, DECEMBER 2011

Fig. 4. Examples of typical PSD patterns used for evaluation of the delayed light fractions in DPAC-stilbene crystals of different composition: (a) prompt scintillation with short decay and no delayed light; (b) a pattern containing both prompt and delayed components without clear separation; (c) PSD pattern with clearly
Cs spectrum showing Compton edge typical of larger crystals.
defined separation between two components; and (d) example

Fig. 5. Change of the neutron/gamma separation and figure of merit (FOM)


near the
Cs Compton edge as a function of stilbene fraction in the mixed
Cf stimulations.
DPAC-stilbene crystals obtained from

DPAC should be transferred to stilbene present in the host


as impurity. The increased LY in the mixed DPAC-stilbene
crystals results from the singlet emission of stilbene molecules,
which are posited to have higher quantum efficiency than DPAC
molecules. The mechanism is supported by the observation
that the photoluminescence spectra of the mixed DPAC-stilbene crystals with enhanced LY resemble more the stilbene,

Fig. 6. The average amount of delayed light as a function of stilbene fraction


in mixed stilbene-DPAC crystals obtained from
Cs gamma stimulations.

rather than the DPAC spectrum. Furthermore, the spectra of


the mixed crystals exhibit a slight blue-shift in comparison to
the spectra of pure stilbene crystals [see Fig. 2(c)], reflecting,
most likely, the transformation of stilbene molecules from their
bonded state in a pure stilbene crystal to the state of single
fluorescing molecules distributed in DPAC. By analogy with
liquid scintillators, in the mixed DPAC-stilbene system, stilbene acts as a solute, producing the more efficient scintillation

ZAITSEVA et al.: PULSE SHAPE DISCRIMINATION IN IMPURE AND MIXED SINGLE-CRYSTAL ORGANIC SCINTILLATORS

3417

these studies, it could be concluded that the actual scintillation


efficiency of DPAC discussed in conflicting reports previously
[2], [11][13] is inferior to that of stilbene, because all pure
diphenylacetylene crystals, defined by chemical analysis or
presence of PSD, showed a LY (gamma, beta, and PL) almost
an order of magnitude lower than in stilbene crystals.
B. Interpretation of PSD in Newly Tested Materials

Fig. 7. Relative scintillation light yield (LY) measured in mixed DPAC-stilbene


crystals using gamma ( Cs) excitation source. LY of a pure stilbene crystal
grown from the same material as used in the mixed crystals was taken as a unit.
We estimate the one standard deviation uncertainties for determining the energy
Cs Compton edge due to counting statistics, crystal size
calibration near the
variations, and optical coupling variations to be no more than 10%.

because of decreased self-absorption in the diluted state. Due


to energy-dependent photon statistics, the higher LY efficiency
can be also reflected in the enhanced PSD produced by the
stilbene network in the mixed crystals.
IV. APPLICATIONS OF RESULTS
The DPAC-stilbene system is a rare representative of a small
group of compounds which can form mixed crystals in large
range of concentrations because of the close similarity of their
molecular and crystal structures. More experiments are required
to understand to what extend the mechanisms studied with this
system are applicable for the cases of organic crystals containing
impurities with more different structure and optical properties.
Our observations, as well as some literature reports indicate that
small additions of longer wavelength impurities do not always
lead to the quenched PSD (p-terphenyl doped by 1,4-diphenyl1,3-butadiene in [26] and [27]). Differentiation of impurities
in respect of their effects on the PSD in organic scintillators
presents an interesting subject of further studies. However, our
experience obtained with the DPAC-stilbene system already can
be helpful for better understanding of the experimental problems identified in the current or earlier works. As shown by a
few examples described below, the results also can be useful
for practical applications leading to the development and engineering of new materials.
A. Explanation of the Previous Results
The results presented here correct a previously accepted
opinion [1], [2], [9] that DPAC has a negligible delayed component of the scintillation emission and therefore is unsuitable for
PSD. Pure diphenylacetylene has PSD [Fig. 3(a)], in this respect
not being very different from stilbene or other hydrocarbon
compounds of similar structure. Our data indicate [Fig. 5] that
separation for pure DPAC is similar to
the mean
crystals with high stilbene concentration. The corresponding
FOM, however, is significantly lower. This is consistent with
DPAC having a lower light yield. Based on the results of

Studies made with the DPAC-stilbene system enable better


understanding and interpretation of the results obtained in
the PSD survey of new organic crystals. As reported previously [17], [19], a number of new materials with PSD close
or even superior to stilbene [Fig. 8(a)] were identified and
selected for further development. However, the overall results
obtained in the initial studies were difficult to interpret prior to
this work. The problem was that the majority of the measured
PSD patterns, the most typical examples of which are shown in
Fig. 8, did not exhibit any visible relation between the PSD and
molecular or crystallographic structures of the tested materials.
Now it may be understood that a complete absence of PSD or
bad PSD, similar to that shown in Fig. 8(b)(d), does not
necessary reflect intrinsic discrimination properties, but can
result from strong effects of impurities. It is a well known fact
that 95%98% purity is very typical for commercially produced
organic materials. Further purification of such materials can
be especially difficult in the case of close similarity of the
host and impurity molecules. The controversy in earlier studies
regarding delayed light in DPAC crystals likely resulted from
the fact that stilbene is typically used as the initial starting
material for DPAC synthesis [28]. Commonly introduced in
DPAC during synthesis, stilbene is very difficult to remove by
recrystallization because of its preferential incorporation into
crystals. Similar situations can be observed in other mixtures
of organic scintillating materials. Fig. 9 shows the deterioration
of the initially good PSD produced as a result of the addition
of lower-band-gap impurities in some organic crystals. As
shown by the example of the bibenzyl (BB)-stilbene crystals
[Fig. 9(a)], small concentrations of impurities may cause a
complete suppression of the delayed light and PSD not only
in DPAC-stilbene, but in other mixed crystals. In a set of very
similar molecules (bibenzyl, diphenylacetylene and trans-stilbene), bibenzyl has the shortest wavelength of emission (widest
band gap). Similar to the dependence shown in Fig. 7, small
additions of stilbene to BB crystals are also accompanied by
a noticeable increase in its scintillation LY, indicating higher
scintillation efficiency of stilbene in comparison to BB. On the
contrary, when mixed BB-DPAC crystals are grown with small
concentration of DPAC, the LY of the resulting crystals drops
significantly, supporting once more the hypothesis that DPAC
has lower scintillation efficiency than stilbene. Depending on
the kind and concentration, the impurities may lead to deterioration of PSD properties producing scintillation with a large
fraction of delayed light, but poor separation between gamma
and neutron events [Figs. 8(b) and (c) and 9(b)]. And finally,
an interesting result of these experiments is that even stilbene
discrimination properties can be spoiled by the presence of a
small amount of closely structured 1,4-diphenyl-1,3-butadiene
[Fig. 9(c)], indicating that historically known excellent PSD

3418

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO. 6, DECEMBER 2011

Fig. 8. Examples of typical PSD patterns measured in a wide range of new organic crystals grown for the survey of neutron/gamma discrimination properties:
(a) good PSD with clear separation of the prompt and delayed components; (b) intermediate PSD with less separation of broad peaks; (c) bad PSD containing
large fraction of the delayed light weakly separated from the prompt component; and (d) no PSD, with low fraction of delayed light.
Cf source used in all
measurements.

Fig. 9. Deterioration of PSD under the effect of impurities in scintillation crystals. Right column patterns correspond to pure crystals. Left column are respective patterns obtained from crystals of the same compounds grown with intentionally added lower energy bandgap impurities.
Cf source used in all measurements.

of stilbene is most likely determined by the fact that it does


not typically end up with PSD killing impurities during its
synthesis.

C. New Single Crystal Scintillators With Enhanced


PSD Performance
Identification and production of new scintillating materials
that can be used for efficient neutron/gamma discrimination is
the most important goal of the conducted studies. As shown
in Fig. 3, both DPAC and stilbene have PSD suitable for use
in neutron detectors. However, according to the main result
obtained in our study, the mixed DPAC-stilbene crystals containing stilbene in concentrations above 30%40% appear to
offer performance better than pure crystals. An optimal composition of mixed DPAC-stilbene is yet to be identified because, in
addition to the scintillation properties, the selection should take
into account the ability to grow single crystals to large sizes
required for detection devices. While in large pure stilbene
crystals, self-absorption can reduce the effective light yield, the
problem may be much less pronounced in mixed DPAC-stilbene crystals of the same size. Thus, mixed DPAC-stilbene
crystals with intermediate stilbene concentration corresponding
to the highest LY and maximum PSD may be good candidates
for efficient detector applications. It should be noted, however,
that scaling-up growth of these crystals is not a trivial task,
because prior to this work, solution growth techniques were
used mainly for production of large crystals soluble in water.
Fig. 10 shows pictures of unique, fully faceted DPAC-stilbene
crystals grown from a pure organic solvent (anisole) using the
rapid growth technique developed previously for growth of
large KDP crystals [29], [30]. The encouraging results indicate
that larger and better quality single crystals can be grown with
further development of the solution growth technique now
applied to pure organic systems. It is also important to note
that selection of such mixed systems should not be necessarily

ZAITSEVA et al.: PULSE SHAPE DISCRIMINATION IN IMPURE AND MIXED SINGLE-CRYSTAL ORGANIC SCINTILLATORS

3419

Fig. 11. Comparison of decay curves obtained with commercial plastic and
liquid scintillators and DPAC (diphenylacetyle) and BB (bibenzyl) single crystals containing stilbene added to suppress slow component of scintillation. Both
BB and DPAC have slow decay light components suppressed much stronger
than all other samples. Results were obtained using a neutron source utilizing
lead shielding and time-of-flight to suppress gamma signal in the detector by a
factor 5 to 10 to emphasize neutron response.

Fig. 10. Single crystals of (a) diphenylacetylene (DPAC), (b) stilbene, and
(c) mixed DPAC-stilbene ( 50:50) grown from organic solvents by temperature reduction method; white bars correspond to approximately 1 cm in length.
(d) Photoluminescence image showing the higher brightness of the mixed
crystal under the same UV illumination.

limited to the DPAC-stilbene combination. It can be extended


to other organic compounds that, grown from solution or melt,
are known to form mixed crystals. Examples of such combination can include aromatic compounds with similar molecules,
such as naphthalene and anthracene, biphenyl and p-terphenyl,
p-terphenyl and p-quaterphenyl, etc., differing by the number
of phenyl rings and, therefore, by the bandgap energy. These
similar compounds are known to form practically identical
crystallographic structures [10], in which previously known
effects of the scintillation efficiency changes can be combined
with the variations of PSD that can be used for the engineering
of new neutron detection materials.
D. Low-Afterglow Scintillators
Although the focus of this research is to understand the processes leading to the delayed light utilized for neutron/gamma
pulse shape discrimination, the understanding of the PSD mechanisms in mixed crystals can be used for producing scintillators with almost no delayed light. The use of organic scintillators with very low afterglow is important for ion and neutron spectrometry at high intensity pulsed laser facilities, which
require separation of temporally close signals, particularly in
cases when the delayed signals are significantly smaller than the
earlier produced signals. Neutrons with different energies can
be separated by the time-of-flight method, which uses different
velocities of neutrons traversing down a flight path. For example, separation of 14-MeV neutrons produced in a fusion reaction from the 1012-MeV neutrons scattered in the high-density compressed target [31] requires fast organic scintillators
with a minimized fraction of the delayed light not exceeding

of peak intensity 30 ns after the primary signal. Liquid


scintillators containing dissolved molecular oxygen known for
its ability to suppress the delayed light via efficient quenching
of the excited triplet states responsible for the slow scintillation
component [32] have the disadvantage that, even if the desired
short decays are reached in the liquid scintillators, the presence
of molecular oxygen leads to simultaneous suppression of not
only the delayed, but also the prompt component of fluorescence, resulting in total decrease of the scintillation efficiency
essential for high-resolution neutron spectrometry. Absence of
PSD in the mixed DPAC-stilbene crystals is also caused by suppression of the delayed light, therefore leading to a sharp decrease of the scintillation decay times. Comparative tests made
with a number of commercially available scintillators showed
that DPAC-stilbene [Fig. 3(b)] or bibenzyl-stilbene [Fig. 9(a)]
single crystals without PSD can have lower fraction of delayed
light than oxygen-quenched liquid scintillators [Fig. 11].
Another attractive feature of the single crystals demonstrated
in these DPAC-stilbene studies is that the addition of impurities
used to suppress the delayed light does not necessarily produce
any quenching effect on the prompt component of fluorescence,
but on the contrary, may lead to the enhancement of the scintillation LY. Since both suppression of the delayed light and
increase in the scintillation efficiency can be obtained at very
small additions of quenching impurities, different combinations
of host-impurity systems can be produced depending on the requirements to decay times, brightness, and wavelengths of emission.
V. CONCLUSION
Pulse shape discrimination (PSD) properties utilized for detection of high-energy neutrons were studied with a number of
new organic crystals produced by solution growth. A model
system of mixed diphenylacetylene-stilbene single crystals of
different compositions was studied to learn about the physical
phenomena of excited state migration and annihilation leading
to the formation of delayed light and PSD in organic materials.

3420

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO. 6, DECEMBER 2011

The results of the studies showed that, in addition to the molecular and crystallographic structures, presence of impurities with
lower bandgap energies may be a major factor influencing PSD
properties of organic materials. Due to trapping effects, very
small concentrations of such impurities in a host crystal can
substantially decrease the rate of triplet excited state migration,
leading to a sharp decrease in the decay times and complete
disappearance of PSD. At a higher concentration of an impurity, PSD can appear at a level exceeding that of the PSD typical for both pure host and impurity. The concentration of the
PSD reappearance is consistent with a percolation threshold resulting from the formation of a network of impurity molecules
that enables triplet energy migration. The results were applied to
the explanation of PSD properties measured in different organic
scintillators in previous works and in our recent survey of new
materials. Understanding of the energy migration processes in
complex systems resulted in growth of new organic scintillators
with controlled decay characteristics. Crystals with a large fraction of delayed light can be used for high-energy neutron detection via improved PSD, while crystals with suppressed delayed
light may be of interest for further development as low-afterglow scintillators for different applications.

REFERENCES
[1] F. D. Brooks, A scintillator counter with neutron and gamma-ray discrimination, Nucl. Instrum. Methods, vol. 4, pp. 151163, 1959.
[2] J. B. Birks, The Theory and Practice of Scintillation Counting.
London, U.K.: Pergamon, 1964.
[3] F. D. Brooks, Development of organic scintillators, Nucl. Instrum.
Methods, vol. 162, pp. 477505, 1979.
[4] B. H. Phillips and K. R. Swank, Measurements of scintillation lifetimes, Rev. Sci. Instrum., vol. 24, pp. 611616, 1953.
[5] G. T. Wright, Scintillation decay times of organic crystals, Proc.
Phys. Soc., vol. B69, pp. 358372, 1956.
[6] R. B. Owen, The decay times of organic scintillators and their application to the discrimination between particles of different specific
ionization, I.R.E. Trans. Nucl. Sci., NS-5, pp. 198201, 1958.
[7] L. M. Bollinger and G. E. Thomas, Measurement of the time dependence of scintillation intensity by a delayed, Coincidence Method,
Rev. Sci. Instrum., vol. 32, pp. 10441050, 1961.
[8] A. A. Espiritu and J. G. White, Refinement of the crystal structure
of diphenylacetylene, Zeitschrift fur Kristallographie, vol. 147, pp.
177186, 1978.
[9] A. Hoekstra, P. Meertens, and A. Vos, Refinement of the crystals
structure of trans-stilbene, Acta Cryst., vol. B31, pp. 28132817,
1975.
[10] A. I. Kitaigorodskii, Molecular Crystals and Molecules. New York:
Academic, 1973.
[11] C. F. Ravilious, J. O. Elliot, and S. H. Liebson, Gamma-scintillations
in diphenylacetylene, Phys. Rev., vol. 77, p. 851, 1950.
[12] W. S. Koski and C. O. Thomas, Scintillations in the diphenylpolyenes
and related compounds, J. Chem. Phys., vol. 19, pp. 12861290, 1951.

[13] R. C. Sangster and J. W. Irvine, Study of organic scintillators, J.


Chem. Phys., vol. 24, pp. 670715, 1956.
[14] S. D. Jastaniah and P. J. Sellin, Digital pulse-shape algorithms for
scintillation-based neutron detectors, IEEE Trans. Nucl. Sci., vol. 49,
no. 4, pp. 18241828, 2002.
[15] Y. Kaschuck and B. Esposito, Neutron/ -ray digital pulse shape discrimination with organic scintillators, Nucl. Instrum. Methods, vol.
A551, pp. 420428, 2005.
[16] N. V. Kornilov, V. A. Khriatchkov, M. Dunaev, A. B. Kagalenko, N.
N. Semenova, V. G. Demenkov, and A. J. Plompen, Neutron spectroscopy with fast waveform digitizer, Nucl. Instrum. Methods, vol.
A497, pp. 467478, 2003.
[17] G. Hull, N. Zaitseva, N. J. Cherepy, J. R. Newby, W. Stoeffl, and S.
A. Payne, New organic crystals for pulse shape discrimination, IEEE
Trans. Nucl. Sci., vol. 56, no. 3, pp. 899903, 2009.
[18] N. Zaitseva, L. Carman, A. Glenn, J. Newby, M. Faust, S. Hamel, N.
Cherepy, and S. Payne, J. Cryst. Growth, vol. 314, pp. 163170, 2011.
[19] N. Zaitseva, J. Newby, S. Hamel, L. Carman, M. Faust, V. Lordi, N. J.
Cherepy, W. Stoeffl, and S. A. Payne, Neutron detection with single
crystal organic scintillators, Proc. SPIE, vol. 7449, p. 744911-(1-10),
2009.
[20] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., vol. 77, pp. 38653868,
1996.
[21] O. Gunnarsson and B. I. Lundqvist, Exchange and correlation
in atoms, molecules, and solids by thespin-density-functional formalism, Phys. Rev., vol. B13, pp. 42744298, 1976.
[22] G. Kresse and J. Hafner, Ab initio molecular dynamics for liquid
metals, Phys. Rev. B, Condens. Matter, vol. 47, pp. 558561, 1993.
[23] P. E. Blochl, Projector augmented-wave method, Phys. Rev. B, Condens. Matter, vol. 50, pp. 1795317979, 1994.
[24] D. Stauffer and A. Aharony, Introduction to Percolation Theory.
Philadelphia, PA: Taylor & Francis, 1994.
[25] H. Scher and R. Zallen, Critical density in percolation processes, J.
Chem. Phys., vol. 53, pp. 37593761, 1970.
[26] J. Iwanowska, M. Moszynski, L. Swiderski, A. Syntfeld-Kazuch, T.
Szczesniak, P. Sibczynski, N. Z. Galunov, and N. L. Karavaeva, Composite scintillators as detectors for fast neutrons and gamma-radiation
detection in the border monitoring, in IEEE Nucl. Sci. Symp. Conf.
Rec. (NSS/MIC), 2009, pp. 14371440.
[27] N. Z. Galunov, B. V. Grinyov, N. L. Karavaeva, J. K. Kim, Y. K. Kim,
O. A. Tarasenko, and E. V. Martynenko, Development of new composite scintillation materials based on organic crystalline grains, IEEE
Trans. Nucl. Sci., vol. 56, no. 3, pp. 904910, 2009.
[28] L. I. Smith and M. M. Falkof, Organic Syntheses, Coll., vol. 3, p. 350,
1955.
[29] N. P. Zaitseva, J. J. DeYoreo, M. R. Dehaven, R. V. Vital, K. E. Montgomery, M. Richardson, and L. J. Atherton, Rapid growth of largecrystals, J. Cryst. Growth, vol. 180, pp.
scale (4055 cm)
255262, 1997.
[30] N. Zaitseva and L. Carman, Rapid growth of KDP and DKDP crystals, Progress Cryst. Growth Characterization Mater., vol. 43, pp.
1118, 2001.
[31] A. Frenje, D. T. Casey, C. K. Li, F. H. Sguin, R. D. Petrasso, V. Y.
Glebov, P. B. Radha, T. C. Sangster, D. D. Meyerhofer, S. P. Hatchett,
S. W. Haan, C. J. Cerjan, O. L. Landen, K. A. Fletcher, and R. J. Leeper,
Probing high areal-density cryogenic deuterium-tritium implosions
using downscattered neutron spectra measured by the magnetic recoil
spectrometer, Phys. Plasmas, vol. 17, p. 056311(1-9), 2010.
[32] R. Lauck, M. Brandis, B. Bromberger, V. Dangendorf, M. B. Goldberg,
I. Mor, K. Tittelmeier, and D. Vartsky, Low-afterglow, high-refractiveindex liquid scintillators for fast-neutron spectrometry and imaging applications, IEEE Trans. Nucl. Sci., vol. 56, no. 3, pp. 989993, 2009.

KH PO

You might also like