1. Introduction
The study of rare nuclear processes (rare alpha and beta decays, neutrinoless double beta decay, or dark matter (DM) particle direct detection) is an active field of research with steady improvements utilizing more and more sensitive detectors. Experimental information on rare decays supports the understanding of nuclear structure, has applications in a variety of fields, e.g., nuclear chronometers, and is relevant as long-lived backgrounds in other rare events searches. As regards neutrinoless double beta decay, it can shed light on fundamental mechanism beyond the standard model of particle physics. Many experiments exploiting various detector techniques aiming to search for these elusive nuclear processes with half-lives longer than age of the universe (on the scale of
y) were proposed and realized within last 20 years. The highest sensitivity was achieved using “source = detector” experimental approach, where the isotope of interest is embedded into the detector material. A major advantage of the “source = detector” approach is almost 100% detection efficiency, as well being able to detect alphas and betas directly inside bulky large-mass target material. For example, leading double beta decay experiments have reached sensitivity of over
y [
1,
2]. Within this approach, the alpha decay of
180W with
116 [
3] and
151Eu with
(Eu) [
4] crystal scintillators was registered at an experimental sensitivity level of over 10
18 y. A recent and comprehensive review on rare decays and techniques of their detection is given in [
5]. However, the main disadvantage of the “source = detector” approach is the limitation to certain target elements which are suitable for detector manufacturing. Therefore, for many years, rare processes that could occur in elements such as Zr, Hf, Sn, Pt, Os, Pd, and Ru have not been explored at the highest sensitivity level due to the lack of scintillating material into which they can be embedded. Everything has changed after the first encouraging results on spectroscopic characteristics observed with
(CHC) scintillating crystals in 2015 [
6]. This crystal exhibits energy resolution of about 3% along with light yield comparable with that for NaI(Tl) standard scintillators. Consequently, this compound immediately attracted a significant attention from scientists from various fields.
Currently for high-sensitivity gamma-spectroscopy application there are needs for scintillators with targeted properties, such as a high light output (about 100,000 ph/MeV), excellent energy resolution (less than 3% at 662 keV line of 137Cs), high stopping power (Zeff larger than 70), fast scintillation decay time (less than 1 μs), good linearity down to low energies, and low cost. Thus, excellent scintillating performance in combination with the recent progress in crystals production opens up various applications, where such scintillators could be utilized along with commonly used high performance NaI(Tl), CsI(Tl), La:Ce or La:Ce scintillating crystals. However, further crystal production chain optimization aiming at general cost reduction should be addressed.
Despite that all above mentioned parameters are also required for
crystals to be used in fundamental research, another characteristic of this material turns out to be even more attractive. Indeed, the unique feature of this material is the large fraction of the embedded Hf element of about 26% in mass. It should be stressed that these crystals are compounds with stable mechanical and optical properties over time, contrary to metal-loaded liquid scintillators that typically tend to the precipitation of heavy loaded scintillators and deterioration of their optical and spectroscopic properties. Hence, such a stable compound as a
crystal with a high mass fraction of Hf, in combination with its high performance as a scintillating detector along with a relatively low internal background opens up new opportunities to search for rare nuclear processes occurring in Hf isotopes applying the “source = detector” experimental approach with an advanced sensitivity. As evidence that
crystals are very promising in search for elusive nuclear processes, the rare alpha decay of
174Hf isotope with a half-life of T
1/2 = 7 ×
y was observed with only 7 g
crystal acting as a scintillating detector (see
Section 10 below, and [
7]).
The compound crystallizes in a cubic structure with a lattice parameter a = 10.42 ± 0.01 Å (space group Fm3m) that is isostructural to a potassium platinum chloride (K2PtCl6). Hence, the belongs into a group of compounds with a general formula of A2MX6, where A = Li, Na, K, Rb, Cs; M = Hf, Zr, Ti, Pt, Sn, Te; and X = Cl, Br, or I. Accordingly, each element in the crystal structure can be substituted for an alternative element with an equal ionic charge keeping the structural type unchanged. This makes the matrix very flexible to the element of interest that can be embedded. Therefore, controlled elements substitution would allow to use CHC-family crystals as scintillating detectors in experiments aiming to register rare processes (rare alpha, beta decay and neutrinoless double beta decay) that could occur in elements such as Zr, Sn, Pt, Os, Pd, and Ru, which have not been studied with the “source = detector” experimental approach at ultimate sensitivity.
While the first part of this review is dedicated to the analysis of progress in CHC-family crystals production, including the crystal growth, raw material purification, and conditioning procedures, the second part will cover aspects related to crystal characterization. This part would focus not only on preliminary light output estimation, scintillating decay components, and energy resolution evaluations, which are indeed very important and desired characteristics for gamma-spectroscopy, but it also will include consideration of material properties such as defect structure, quenching factor for alpha particles, pulse shape discrimination ability, and internal radioactive background, which is of vital importance if the material is used in low-background experiments.
In view of the large number of articles related to this “hot” topic that have been published and may be under preparation, this review could be considered as the first attempt to systematize already existing results and achievements in CHC-family crystals growth, and to keep in one place the current status of these crystals studies, while providing hints for future investigations.
2. Cs2HfCl6 Crystal and Bromine-Containing Mixture Compounds
Despite the fact that
compound was originally discovered as a luminescent material in 1984 by [
8], it was re-invented as an attractive scintillator for gamma detection in 2015 by [
6]. The best energy resolution that was measured with a ~1 cm
3 CHC crystal sample was calculated to be 3.3% at 662 keV under irradiation by
137Cs gamma source (see
Table 1), and this was obtained with a long shaping time of 12 μs. The light yield (with photomultiplier (PMT) readout) was estimated to be comparable to that of NaI(Tl) commercial crystal. The scintillation pulse was described with model that consists of two exponential components: fast decay time is 0.3 μs (5%) and slow component is 4.4 μs (95%). Authors also measured radioluminescence spectra and assigned the broad emission band that is centered at about 400 nm to an intrinsic luminescent center based on transitions of charge transfer type of the undisturbed [
]
2− anion complex, placed in the cubic hole created by Cs
+ ions at the corners of the cube.
To grow this CHC crystal, the stoichiometric ratio of raw materials CsCl (99.998%) beads and (99.9%, trace metals basis, excluding Zr) powder was properly mixed. No additional purification step was applied. All actions with the compound preparation were performed in a glove box under argon atmosphere, with moisture and oxygen levels maintained both below 1 ppm. It should be highlighted that in all below described experiments on CHC and CHC-family crystals growth, raw materials handling was performed in glove box with controlled atmosphere and reduced moisture, unless otherwise indicated. Vertical Bridgman furnaces were used for crystal growth. In this experiment, the CHC charge was melted at 820 °C and translated from a hot zone down to the cold zone at a pulling rate of (0.5–1.0) cm/day. The temperature gradient of 5 °C/cm was established at the solid/liquid interface. Crystals used for characterization were cut from the as-grown CHC boule. The CHC density was calculated to be (3.78 ± 0.04) g/cm3, based on gravimetric measurements.
After this first demonstration of such promising scintillating characteristics of the CHC crystal, the comparative study of the luminescence and scintillation properties of
and
ZrCl
6 crystals was performed by [
9], where photoluminescence spectra, X-ray excited radioluminescence spectra, scintillation decay times, and pulse height spectra under irradiation with 662 keV gammas of
137Cs source were analyzed.
CsCl (99.999%), HfCl4 (99.5%), and ZrCl4 (≥99.9%) powders were used as starting materials for synthesizing and (CZC) crystals, accordingly. The powders were mixed in a stoichiometric ratio and dried by heating overnight at about 470 K in vacuum. Afterwards, the prepared mixtures were sealed in quartz ampules under vacuum, and then the vertical Bridgman method was used to fabricate crystals. The temperature gradient was set to 8.7 °C/cm, while pulling rate was at 1 mm/h. The CHC (7 × 5 × 2 mm3) and CZC (5 × 4 × 3 mm3) polished samples were used for crystal characterization and comparison. The X-ray fluorescence analysis revealed that the crystal contained Zr at a 0.12 mol% level. Meanwhile, the X-ray diffraction patterns of CHC and CZC samples indicated that both crystals contained a slight CsCl crystalline phase.
The X-ray-excited radioluminescence spectra of the CHC and CZC crystals measured at 300 K both have broad emission bands with main peaks at 398 nm wavelength for CHC, and 435 nm and 510 nm for CZC, correspondingly. Authors attributed the 398 nm emission of CHC crystal to the charge-transfer luminescence of the [HfCl6]2− anion complex. The luminescence quantum yields of the and ZrCl6 crystals were estimated to be 55% and 29%, respectively.
Because scintillating pulses cannot be fitted with a single exponential model, authors have applied a double exponential assumption and fit averaged scintillating pulses with such model in the time interval up to 50 μs for CZC crystal and 200 μs for CHC crystal. The decay time constants were found to be about 2.2 and 8.4 μs for the
crystal, and 1.5 and 7.5 μs for the
ZrCl
6 crystal, respectively. It should be emphasized that authors didn’t introduce any deep physical meaning in the fitting model that contains two exponential functions, except a high quality fitting procedure. This point will be addressed later in the text (see
Section 11). The scintillation light yield was estimated in comparison to the commercial NaI(Tl) crystal taking into account emission wavelengths of each crystal and spectral sensitivity of PMT at each peak wavelength. The scintillation light yields for
and
ZrCl
6 crystals were estimated to be 27,500 and 25,100 photons/MeV, respectively.
In Ref. [
10],
(CHC) and
(CHC4B2) crystals were grown from 99.5%
with 99.999% CsCl and 99.999% CsBr. Prior to synthesis,
powder was purified by sublimation at 300–400 °C. Then, starting materials were melted and compounded in stoichiometric ratio at 900 °C for at least 48 h to produce CHC and CHC4B2, correspondingly. Furthermore, synthesized CHC compound was sublimed after compounding prior subjecting to crystal growth, while CHC4B2 was not. This additional sublimation caused the difference in the final crystal boules. Along the perimeter of the CHC4B2 boule, numerous black inclusions were observed. A secondary phase accumulated primarily at the tail of the grown crystal and throughout its core, while the rest of the bulk was clear. On the contrary, CHC boules were clean inside being milky of the surface, and secondary phase precipitates in the core and boundaries of grains with the primary phase. The CsCl was identified by a micro X-ray fluorescence spectrometry to be this secondary phase, and its formation was explained as a result of a non-stoichiometric (CsCl-rich) melt composition. This could be due to an insufficient temperature or time required for the synthesis reaction, or due to the high vapor pressure of
during compounding. This secondary phase rich in CsCl was observed in CHC4B2 boule, as well.
Clean CHC sample had a light yield and energy resolution of 30,000 ph/MeV and 3.3%, respectively, and decay components of scintillating pulses of 0.39 and 3.9 μs (see
Table 1). While, the sample of CHC4B2 crystal with a secondary phase present in the core had a light yield and energy resolution of 18,600 ph/MeV and 4.4%, correspondingly. The scintillating pulses show a faster decay with 0.38 and 2.0 μs components (see
Table 2).
Authors also investigated one very important technological parameter of these crystals in light of their wide application—hygroscopicity and moisture sensitivity. Indeed, both crystals showed minimal moisture sensitivity. CHC developed a hazy surface between days 11 and 29, while CHC4B2 developed a hazy and then opaque surface layer after 30 and 48 h, respectively. Both samples showed minimal weight changes (<1 mg) within testing period.
Authors conclude that scintillating properties of CHC and CHC4B2 crystals make these materials appropriate for most low-counting rate applications. Moreover, an additional purification, alloying, or doping may also enhance scintillation performance. The investigation and development of alternate synthesis pathways (e.g., vapor-solid synthesis or wet synthesis) are needed to generate stoichiometric material for high quality crystal growth.
The complex work done by [
11] should also be emphasized, wherein five different procedures of CHC crystal growth have been investigated in order to optimize growth conditions and scintillation properties. The authors proceeded from two approaches. The first approach was to develop a purer starting material that maintains stoichiometry better than just mixing binary halides. The second approach was to attempt to alter inclusion formation by adjusting crystal growth parameters, such as a thermal gradient, pulling rate, and crystal diameter.
Most of the CHC crystals obtained in this study were grown from the CHC charge produced following the procedure described below. An acid-saturated methanol solution was prepared by bubbling anhydrous HCl gas through 400 mL of ice-chilled methanol for 10 min. This acidic methanol solution was then diluted in a 1:10 ratio with fresh methanol to produce the reaction medium. Raw powder (99.9% grade) was added in portions to the ice-cooled round bottom flask, followed by addition of CsCl grains (99.9% grade). The white precipitate of CHC compound was formed as the CsCl was added. This mixture was refluxed for 3 h under nitrogen and then stirred overnight at room temperature. Then, CHC powder was filtered out of the solution and washed twice with fresh methanol. The final CHC powder was then heated under vacuum at 120 °C in order to remove the remaining non-reacted methanol. The charge CHC powder, synthesized in above described way was loaded into a quartz ampule, dried at 100 °C for 12 h under vacuum and then sealed. Next, this material was overheated to 870 °C, while CHC melting point is about 810 °C, to separate secondary phases. After cooling, the obtained material was extracted and grounded. At the next step, the grounded charge material was loaded into an ampoule with two quartz frits to remove the high melting point secondary phases formed in the first step. The material was then melted through the filters for 24 h at 870 °C. This led to a complete separation of the secondary phases and all non-condition material and a single phase CHC, which was melted and passed through the frit. The filtered single phase CHC compound was then subjected to growth in the same ampoule. After the completion of growth, crystal boules were cooled to room temperature over 72 h.
Only one sample (crystal C) was grown from a CHC charge synthesized from raw CsCl and powders in a stoichiometric ratio and then purified through hydrochlorinating. For this process, a portion of the pre-synthesized CHC compound was heated up to the melting point of 810 °C for 3.5 h under a flow of HCl gas.
Five crystals were grown in these studies and designated as crystals A to E. Crystal A (∅22 mm) was grown with a thermal gradient of 21 °C/cm and a pulling rate of 1 mm/h. Crystal B (∅22 mm) was grown with a higher thermal gradient of 34 °C/cm but the same pulling rate of 1 mm/h. Crystal C was grown in the same conditions as crystal B, except instead of using melt filtering to remove secondary phases, the material was hydrochlorinated. Crystal D (∅13 mm) was grown with a thermal gradient of 34 °C/cm and a pulling rate of 0.5 mm/h. While the crystal E (∅13 mm) was grown with a thermal gradient of 34 °C/cm and a higher pulling rate of 1 mm/h.
While X-Ray diffraction (XRD) measurements did not show difference in pattern for all analyzed crystals, the XRD spectrum of the black material shows the presence of hafnium oxide. Authors suggest that the hafnium oxide could be formed due to residual methanol or water from the synthesis reacting with the at high temperature. As an indirect confirmation of this hypothesis could be the observation of black material presented at the top of the ampules is amorphous suggesting that it could be a carbon coating, created during the dissociation of methanol. The measurements taken with energy dispersive X-ray spectrometer (EDXS) were performed for each crystal on both the dark and the light regions, and demonstrate that the black regions are mainly composed of the CsCl phase.
Analyzing the set of CHC crystal samples obtained in growth runs with various gradient and pulling rate, variation of ampule diameter and other parameters authors noticed following effects: (1) an increase in thermal gradient resulted in less cloudiness or fewer inclusions in the bulk of the crystal; (2) a decrease in an ampule diameter results in a significant increase in the number of inclusions in the center of the crystal; (3) a decrease in the growth rate results in an effect similar to that of increasing the thermal gradient (the crystal sample is clearer near the core). As one possible explanation for the improvement in crystals quality that were grown at a higher thermal gradient or slower growth rate, authors suggest considering the constitutional supercooling.
In fact, the crystal sample C, that exhibits the best scintillating performance with an energy resolution (FWHM) of 4.0% at 662 keV gamma line and a light yield of (36,000 ± 500) ph/MeV (see
Table 1), was grown in 22 mm diameter quartz ampule in thermal gradient 34 °C/cm and 0.5 mm/h pulling rate. Meanwhile, all obtained here CHC crystals exhibit superior proportionality with deviation from 1 less than 3% down to energy of 5 keV for gamma quanta, in comparison to SrI(Eu) scintillator with 6% or NaI(Tl) crystals with 15% deviation, respectively.
The radioluminescence emission spectra from the whole range of CHC crystals consist of a single peak with a maximum between 400 and 410 nm, and no other significant differences between spectra for crystals with different quality, beside just slight width variation of the emission peak. From this observation, the authors made the important conclusion that secondary phase of CsCl that could be present in CHC crystal is not producing a significant secondary emission.
The average scintillation pulses were fit with model contains two exponential components: 4.4 us (75%) and 1.0 us (25%). Meanwhile, for the purest crystal (hydrochlorinated material, sample C), these two components correspond to 5.8 us (86%) and 1.4 us (14%). This effect of scintillating kinetics elongation the authors attributed to its higher purity. Moreover, similarity of scintillating kinetics for CHC with different concentration of black inclusions within the crystals suggests that such inclusions are not affecting the mechanism of scintillation, while resulting in a stronger light scattering and absorption in the bulk. This hypothesis was confirmed by dedicated measurements of light absorption within CHC samples. The best performing CHC crystals offer the highest optical transmission in the region of 350–450 nm, which corresponded to the CHC peak emission, while crystals with the poorest performance exhibit the lowest transmission in that region.
To summarize, the authors suggested that the usage of the pre-synthesized CHC powder with a subsequent hydrochlorination and growth with a high thermal gradient maintains melt stoichiometry, and results in a nearly single phase CHC crystal with a high scintillating performance.
In 2019, the addition of bromine into the
original lattice, resulting crystal structure and its scintillation properties were investigated in detail [
16]. The main goal of that work was to provide improvements of CHC decay time of scintillating pulses, as the long decay time of 4.37 μs in initial compound limits the materials performance in a wide range of applications.
To produce starting materials were used 99.998% pure CsCl 99.999% pure CsBr, and 99.8% powder (trace metals basis, exclusive of Zr). Before synthesis, the powder was treated to three-fold static sublimation (with 50 g load each time) to reduce the concentration of low-vapor pressure contaminants. Each purification stage was performed at 400 °C for 12 h. After the third stage, the fraction of purified material was obtained nearly 80% of the initially loaded mass. The stoichiometric ratio of the purified starting materials was sealed in a quartz ampoule and maintained at 800 °C for 24 h to ensure the homogeneity of the CHC4B2 charge. The crystal was grown with a pulling rate of (0.5–1) cm/day and temperature gradient of 5 °C/cm at the solid/liquid interface. At the end of the crystal growth, the temperature of the furnace was uniformly cooled to room temperature in 72 h. Despite dedicated efforts and technological features, some precipitates were still observed in the central part of the crystalline boule.
Pulse height spectra were acquired with 10 μs shaping time under irradiation by
137Cs gamma source, and demonstrated an energy resolution of 4.5% at 662 keV peak. The light yield was estimated to be 37,000 ph/MeV by comparison to a standard commercially available NaI(Tl) crystal (see
Table 2). The light yield non-proportionality for CHC4B2 crystal was observed to be better (less than 1% deviation from 1) than for CHC crystal in the energy range of (20–400) keV. In addition to improved non-proportionality of scintillating response in CHC4B2, one would also expect better energy resolution than in CHC crystals due to a smaller band gap (3.7 eV), and consequently, larger number of photons at the same deposited energy. Moreover, the radioluminescence spectrum of CHC4B2 crystal, centered at 420 nm, makes it better spectrally matched with a SiPM device. Therefore, one could expect further energy resolution improvement with
crystals coupled to such type of light detector.
The scintillation decay time response consists of two components: fast is 0.18 μs (8%) and a slow of 1.78 μs (92%), that is 2.5 times faster compared to previously reported for such type of crystal in [
10]. This improvement of scintillation decay time authors attributed to reduction of CsCl secondary phase concentration, as well as to a quantum effect caused by the replacement of Cl
− ion with the heavier Br
- ions and breaking the symmetry the Hf
4− activator ion. Shorter scintillation decay times are typically expected for heavier alkali halides. Authors are expecting to achieve further improvements of light yield and energy resolution through additional purification of raw materials.
Further progress in
and
crystals production was shown in [
12]. As starting materials, CsCl (99.999%) and
(99.9%) for CHC, along with CsCl, CsBr (99.995%),
and HfBr
4 (99.9%) for CHC4B2,
were utilized. In this experiment, initial HfCl
4 powder of low
chemical purity underwent an improved purification by a one-fold sublimation. There
200 g of HfCl
4 powder sealed in a quartz ampoule under high vacuum was placed into a single-zone horizontal furnace with temperature set to 220 °C. The sublimation process lasted for 72 h. For crystal growth stoichiometric amounts of corresponding starting materials loaded in quartz ampule and sealed under high vacuum were placed into a two-zone vertical furnace. The top furnace’s zone temperature was set a few degrees above melting point of CHC or CHC4B2, while the bottom furnace’s zone a few degrees below the corresponding melting point. The pulling rate was set at (3–4) mm/hour. The temperature gradient for CHC growth run was about 10 °C/cm, while for CHC4B2 was 12 °C/cm. When the growth cycle was completed both zones of furnace were cooled down to room temperature at the rate of (100–150) °C/day. As described, transparent, clear, and crack-free 25 mm diameter
and
boules were obtained, and samples with dimensions of ∅23 × 30 mm
3 (CHC) and ∅23 × 26 mm
3 (CHC4B2) were cut for crystal characterization.
The response of CHC and CHC4B2 samples to gamma irradiation was measured by placing a crystal in a mineral oil within a quartz cup lined with Teflon tape as a diffuse reflector. The
137Cs spectrum collected with a ∅23 × 30 mm
3 CHC sample showing an energy resolution of 3.5% (FWHM) at 662 keV, while the light yield was measured to be 23,000 ph/MeV (see
Table 1). The light yield was evaluated by comparing the 662 keV full energy peak position for CHC with that of NaI:Tl taking into account the quantum efficiency of the PMT at 410 nm (emission peak of CHC) and at 415 nm (emission peak of NaI:Tl). The slightly lower light yield (LY) value for this sample, than previously reported in [
10], authors attributed to not optimal light collection in those experimental conditions. The CHC4B2 sample demonstrates an energy resolution of 3.7% (FWHM) at 662 keV. The light yield, calculated with the same procedure as with CHC crystal, is 20,000 ph/MeV (see
Table 2). It was suggested by the authors that better crystal quality contributes to a better energy resolution.
The averaged scintillating signals of CHC crystal were fitted with sum of two exponential functions, and decay time constants determined to be 0.25 μs (7%) and 3.8 μs (93%). In case of CHC4B2 crystal, the characteristic decay times were found to be 0.33 μs (10%) and 1.8 μs (90%). While the authors noticed that substituting some Cl- ions with Br- ions appears to halve the primary decay time from 3.8 μs to 1.8 μs, there are still no clear theoretical explanations of this phenomenon.
Next results of this group of researchers were published in [
14], where the effect of raw materials purity on crystals scintillating properties was investigated. As starting materials were used CsCl (99.999%), CsBr (99.99%), and
(99.9%), respectively. To grow
crystal, as-purchased raw materials in stoichiometric ratio were mixed. The first CHC crystal, CHC-1, was produced from a stoichiometric mixture of as-purchased raw materials with no additional purification. Meanwhile, the second CHC crystal, CHC-2, was produced from
powder that underwent an improved purification by a one-fold sublimation, as described in [
12]. After materials loading, each ampoule was subjected to low temperature dehydration under high vacuum to remove moisture that might be introduced during loading and subsequent steps. Then, ampules were sealed and subsequently placed in a two-zone vertical furnace for crystals growth.
The response of CHC and CHC4B2 samples to gamma irradiation was measured with the same approach, as described in [
12], where studied crystal was placed into a mineral oil. The
137Cs energy spectrum collected with a ∅10 × 15 mm
3 CHC-1 sample, showing an energy resolution of 2.8% (FWHM) at 662 keV, and light yield was measured to be 23,000 ph/MeV. Energy resolution of 3.2% and light yield of about 27,000 ph/MeV were obtained for CHC-2 sample with slightly smaller dimensions, ∅8 × 8 mm
3 (see
Table 1). The CHC4B2 crystal sample demonstrates an energy resolution of 5.1% (FWHM) at 662 keV and light yield of 21,000 ph/MeV (see
Table 2, correspondingly). The light yield value for all studied crystals was evaluated by comparing the 662 keV full energy peak position for CHC/CHC4B2 with that of a commercial ∅25 × 25 mm
3 NaI:Tl crystal.
The averaged scintillating signals of CHC crystals were fitted with sum of two exponential functions, and decay time constants determined to be 0.25 μs (13%) and 3.3 μs (87%) for CHC-1, and 0.26 μs (12%) and 3.4 μs (88%) for the CHC-2 sample, correspondingly. In the case of CHC4B2 crystal, the characteristic decay times were found to be 0.19 μs (11%) and 2.9 μs (89%). Hence, one could notice that while for CHC crystals of different quality the kinematics of scintillation pulse is very stable, CHC4B2 crystals still demonstrate noticeable variation in the evaluated decay components for different crystal samples (see, for comparison, values reported in [
12]).
The most recent progress in
(CHC) and CHC-family crystals growth has been made at Fisk University was summarized in [
13]. Since the vendors could only provide Hf-halide compounds up to 99.9% purity level, impurities contained in these starting materials contributed much to inclusion and growth problems of the CHC and CHC-family compounds. Thus, a purification stage is required. The main difference in a purification process by an improved single-fold sublimation applied here to that as described in [
12,
14], was in a decrease of a loaded
or Hf
powder for a single purification run, from 200 g to 150 g. Then, for each growth, a stoichiometric mixture starting materials were loaded into quartz growth ampoules that were pre-cleaned and baked at 700 °C. Loaded ampoules were attached at the dehydration-sealing station, where the mixed materials were then dehydrated at 90 °C for a few hours, and then sealed under high vacuum. Crystal growth was performed in a two-zone vertical furnace, following thermal conditions and pulling rates as described in [
12] (see above).
At the beginning, 6.3% of energy resolution (FWHM) was measured for 7 × 7 × (4–5) mm3 sample cut from the CHC crystal boule grown from unpurified raw materials. With samples retrieved from the CHC boule grown from purified precursors the improved energy resolution of (2.8–3.2)% at 662 keV was achieved. The results for samples from unpurified and purified by single-fold sublimation raw materials suggested that the purity level of the starting materials affected the CHC crystal quality and performance more than the influence of growth parameter alteration. Therefore, the next growth was performed with raw materials that were additionally purified, specifically sublimed-purified (no details of this specific sublimation process listed in the cited article). The clear and transparent boule shows that both tuning growth parameters and employing purified starting materials improve the crystal quality.
At the last step, 25 mm diameter CHC crystal from additionally purified raw materials was grown, and the energy resolution of 3.5% (FWHM) at 662 keV was measured with a large ∅23 × 30 mm
3 CHC sample (see
Figure 1 and
Table 1), which is comparable to that earlier measured for crystals of smaller volume. A primary decay time between 3.5 and 4 μs is typical for CHC crystals and was obtained for both small and large volume crystals.
The same procedures, including the purification of starting materials, were followed also to grow CHC4B2 crystal. The energy resolution of 3.7% (FWHM) at 662 keV was measured for 25 mm diameter CHC4B2 crystal grown with purified starting materials (see
Table 2). The primary decay time measured for this CHC4B2 sample is 1.8 μs.
Authors also perform the first studying of a permanent packaging of CHC crystals. An aluminum packaging with a quartz window was used as encapsulation. Prior to packaging, each sample was tested in a typical set-up with a mineral oil (see, for example [
12]). The energy resolution of 3.6% (FWHM) at 662 keV was measured with bare crystal, while after packaging this value slightly degrades to 3.9% (FWHM) at 662 keV.
Authors concluded that similar results, achieved both with small and large volume CHC crystals, indicate that crystal quality almost reached the highest level, and there is no internal scattering, absorbing centers, or impurities that reduce light output and prevent efficient light collection. Thus, the applied purification technique and growth conditions make it possible to produce large volume CHC crystals with the same performance as small size crystals.
While a number of articles are dedicated to
crystal production and characterization, authors in [
17] attempt to produce the
(CHC3B3) crystal and investigate the effect of further constitution of Cl to Br ions. In order to synthesize
compound the
(99.9%) and Hf
(98.5%) powder, along with CsCl (99.999%) and CsBr (99.995%) beads were mixed in a stoichiometric ratio and sealed into a quartz ampule under vacuum. Then, the ampule was placed into a vertical furnace and pulled down with a speed of 1 mm/h during the crystal growth.
An obtained the CHC3B3 crystalline boule with rough dimensions ∅15 × 15 mm
3 had a yellowish color, contained many cracks and its surface was opaque, but the crystal was well transparent inside. Several crystal pieces were successfully fabricated from the obtained crystal boule for the optical measurements. Single crystal X-ray diffraction indicated that the CHC3B3 has a cubic Fm3m structure and the lattice constant was determined to be 10.67 Å. By substituting Br for Cl, the emission peak of CHC3B3 crystal was shifted when compared to CHC (415 nm) or
HfCl
4Br
2 (around 420 nm) towards longer wavelength region of 450 nm. The spectra of CHC3B3 also had a broad shoulder around 530 nm. The scintillation decay was fitted by a single exponential function and decay constant was estimated to be (2.2 ± 0.3) μs. Such results might suggest that halogen substitution causes the emission wavelength shift for CHC with increasing the atomic number of substituted halogen atoms, i.e., Br and I, while the scintillation decay slowed down with high content of Br in the Cl sub-lattice, in comparison to CHC4B2 crystal with decay time of about 1.8 μs. No estimation on relative light yield was made with those samples (see
Table 2).
The synthesis of a single-phase stoichiometric material, as could be seen from above discussed researches, is an uneasy task due to the high volatility and hygroscopicity, and kinetics of the CsCl and reaction in solid phase. CHC crystal growth suffers from CsCl-phase precipitation, which leads to inclusions in the first-to-freeze sections, compound decomposition, and non-stoichiometry. Thus, the successful growth of high-quality CHC crystal requires complete synthesis of this compound and removal of oxygen-containing impurities before growth process. Therefore, besides an extensive characterization of scintillating properties of CHC crystal, it is vital to characterize the behavior of as a chemical compound under different thermal treatments. The former will help to improve the technological procedure and thermal regimes of charge synthesis and crystal growth.
Authors in [
18] prepared a high-quality single-phase undoped
crystal and described material stability and thermal behavior by the simultaneous non-isothermal differential scanning calorimetry and thermogravimetry (DSC–TG) analysis. In addition, the influence of various atmospheres (nitrogen and vacuum) and conditions (enclosed system in a sealed quartz ampule) on the CHC compound stability were investigated. This study contributes to the improvement and optimization of the thermal conditions for CHC crystal growth from the melt by the Bridgman method. First, the congruent melting point was determined to occur at ca. (821–822) °C. The temperature position of the effect was the same even at 10-cycle measurement, pointing to the reversibility of the observed effect and its stability. Similarly, no undercooling of
compound was determined as the solidification occurred at practically similar temperature of ca. (822–824) °C. This indicates that the
solidification occurs at a similar temperature as its melting. Such a result is not common in halides as high undercooling was observed, e.g., in PbCl
2 [
19] and CsCaI
3 [
20]. In this study, it was also shown that the CHC compound decomposition occurred at 300 °C under vacuum, while under N
2 atmosphere CHC stay without significant mass losses until the eutectics at 587 °C. Such high temperature instability could limit the
crystal application as a scintillator since CHC crystal luminescence and scintillating properties are strongly correlated with its chemical composition and stoichiometry.
It is necessary to mention research reported in [
21], where the question of doping of CHC crystal by Te
4+ ions was addressed. In this study the photoluminescence (PL) and scintillation properties of 1.0, 3.0, 5.0, and 10 mol% Te
4+-doped
crystals were investigated along with photoacoustic (PA) spectroscopy, which enables direct monitoring of the non-radiative de-excitation processes that take place in the sample after absorption of irradiation energy. There,
crystal plays the role of host material, since the (HfCl
6)
2− octahedra have a perfect octahedral symmetry and do not share anions, i.e., the tetravalent metal ions, such as Zr
4+, Te
4+, Sn
4+ Pt
4+, Os
4+, etc., can occupy the center of an undistorted (HfCl
6)
2− octahedron. For this reason, CHC-family crystals appeared to be ideal as model compounds for investigating the photoluminescence properties of tetravalent metal ions.
In this study unusual cleaning procedure was applied to quartz ampules. Quartz ampoules were cleaned in a strong alkali solution to remove organic impurities and then were thoroughly rinsed with ultrapure water. As a second step, ampoules were annealed at 1000 °C for 24 h. Afterwards, the stoichiometric mixture of CsCl (99.99%), (99.9%), and TeCl4 (99.9%) raw materials without any additional purification were loaded into the clean ampoules, dried under vacuum at 300 °C for 24 h, and then sealed. Crystals were grown in two-zone vertical furnace at the temperature gradient of 13 °C/cm and at the pulling rate of (24−48) mm/d. Crystal samples for whole range of doping concentrations have a yellowish color and visible cracks induced during the crystal growth and cooling processes. Samples with dimensions 2.0 × 2.0 × 1.0 mm3 were sliced and polished to be used in further measurements.
The radioluminescence spectra exhibited a yellow emission band peaking at 575 nm. The scintillation intensity for the 3.0 mol% Te4+-doped crystal was the highest among studied crystals. The decay time of scintillation pulses is in agreement with a three-exponential function, with primary, secondary, and tertiary decay time values of 0.02 μs (2%), 0.63 μs (18%), and 2.34 μs (80%), respectively. The value of the dominant two decay components is consistent with that of the corresponding PL values, which are due to transitions of Te4+ ions. The origin of the primary fast component was not clearly understood, but it could be due to the quenching of the intrinsic emission of the CHC host with an efficient energy transfer to Te4+ ions. Scintillating pulse decay components decreased with increasing of Te4+ dopant concentration.
The pulse height spectra under irradiation by
137Cs gamma source were recorded with a 6 μs shaping time. The scintillation light yield of the samples was calculated by comparing the determined 662 keV gamma line photopeak position in the spectrum with that of a CsI:Tl (LY = 54,000 ph/MeV, wavelength of the emission maximum (λ
em) = 550 nm) under the same experimental conditions. The scintillation light yield of the Te
4+-doped crystals was estimated to be approximately 11,700 (Te, 1.0 mol%), 13,100 (Te, 3.0 mol%), 9600 (Te, 5.0 mol%), and 9000 (Te, 10 mol%) ph/MeV. Energy resolution was not quoted for these measurements (see
Table 3).
The measurement of photoacoustic (PA) spectrum was performed with a custom setup, where a xenon-lamp was used as the excitation source and a monochromator combined with a mechanical chopper plays the role of the optical system. The excitation light was modulated by the mechanical chopper at 20 Hz frequency. The acoustic signal was detected with the sample being placed in a customized PA cell fitted with an electret microphone. The output signal from the microphone was amplified by a preamplifier and then fed into a lock-in-amplifier with a reference signal input from the chopper. Finally, the signal from the lock-in-amplifier was converted into a digital form and transmitted for analysis by computer through a digital multi-meter. The spectra were normalized for changes in the excitation light intensity using carbon-black as reference. As a result, an inverse correlation between scintillation light yield and PA intensity with different Te-concentrations was confirmed experimentally. This study is interesting in the field of scintillating materials investigation, where the main focus will be on studying the energy transfer mechanism between (HfCl6)2− complex ion and dopant ions using thermal measurements of the photoluminescence spectrum and decay curves.
Another attempt to grow doped CHC crystals was described in [
22]. In this study, the luminescence and scintillation properties of Tl- and Ce-doped CHC crystals was investigated. Stock powders of CsCl (99.999%),
(99.9%), TlCl (99.9%), and
× 17H
2O (99.9%) compounds were used for the synthesis of corresponding charges. The mixtures in stoichiometric ratio were dried by heating over 150 °C for 12 h under vacuum. Then, mixtures were sealed in quartz ampules and crystals were grown at a temperature gradient of 8.7 °C/cm and a pulling rate of 1.0 mm/h in the vertical furnace. As-grown CHC:Tl and CHC:Ce boules were cut and polished to sizes of 3.5 × 2.5 × 1 and 5 × 3 × 1 mm
3, respectively.
It should be noted that actual Tl and Ce concentrations measured by mass-spectrometry technique in the grown crystals, 0.08 mol% and 0.15 mol%, were significantly lower than those in the corresponding melt, 0.5 mol% and 0.2 mol%, respectively. However, no clear idea has been proposed to explain such unexpectedly low Tl-concentration in the final crystal, the lower Ce-concentration was explained by differences in the ionic radii and charges of the impurity (Ce3+) and host (Cs+) ions.
RL bands for Tl- and Ce-doped CHC crystals were observed at 405 and 430 nm, respectively. The 405 nm band can be attributed to the same origin as undoped CHC. While the 430 nm band authors attributed to charge-transfer transitions in [HfCl6]2− or complex ion luminescence from [ZrCl6]2− impurities, no extrinsic luminescence originating from the dopants in either the Tl- or Ce-doped CHC crystals was observed by RL spectroscopy. The authors explained this by a low concentration of dopant ions.
Scintillation decay time constants were obtained using the double exponential assumption. Decay time constants were found to be about 0.6 μs (24%) and 4.1 μs (76%) for CHC:Tl crystal, and 1.2 μs (33%) and 6.1 μs (67%) for CHC:Ce sample. The light yield was estimated in comparison to that of Gd
2SiO
5:Ce (GSO:Ce), using the photopeak position of 662 keV gammas. Assuming that the light yield of GSO:Ce crystal is about 10,000 ph/MeV, the light yield for CHC:Tl and CHC:Ce crystals was estimated to be 23,700 and 15,700 ph/MeV, respectively (see
Table 3). The authors explain the lower light yield for doped crystals compared to undoped CHC, by the low energy transport efficiency from the host to dopant ions.
4. Tl-Containing Crystals
Significant and impressive progress in CHC crystals growth technology improvement and its high scintillating performance triggered further studies of compounds from the same family, but which would better satisfy requirements for scintillating materials in gamma-spectroscopy in terms of a high stopping power, i.e., high Zeff.
In 2018, two independent groups of authors, reported about success in Tl
2HfCl
6 (THC) and Tl
2ZrCl
6 (TZC) crystals growth and very encouraging results after first testing of its scintillating characteristics. First group [
25] came to these compounds as a result of sequence studies of crystals such as Rb
2HfCl
6 and
Hf
that demonstrated high scintillating light yield and good energy resolution without any dopant. Following this route and trying to further profit from high
Z of Tl-ions, authors focused on Tl
2HfCl
6 and
crystals, which having high Z
eff results in excellent photoelectric conversion efficiency for X-rays and gamma quanta. Crystals were grown from raw materials of commercial grade (TlCl—99.9%,
—99.9%, and
—99.99%) without any additional purification step. Small samples with dimensions about 2 × 1 × 1 mm
3 were used for crystal characterization. Both crystals showed broad radioluminescence spectra with emission band peak at 450–470 nm. Scintillating signals were fitted by sum of two exponential decay components with characteristic times of 0.29 μs and 6.34 μs for THC, and 0.70 μs and 2.36 μs for TZC crystal. Scintillating light yield was estimated in comparison to commercial NaI(Tl) crystal. It was approximately estimated to be 24,200 and 50,800 ph/MeV for THC and TZC crystals, respectively. Energy resolution acquired at the 662 keV gamma line was 17.7% for THC and 5.6% for TZC, correspondingly (see
Table 5 below).
The second group [
26], initially focused only on the production of TZC crystals. Commercial raw materials TlCl (99.999%) and ZrCl
4 (99.9%) without any purification step mixed in stoichiometric ratio were used for synthesis of the TZC compound. A transparent crack-free sample with dimensions 8 × 8 × 1.5 mm
3 was used for crystal characterization. A broad band in radioluminescence spectrum in the range 350–700 nm with a peak at 468 nm was observed. The luminescence in TZC crystal was assigned to the charge transfer in the [ZrCl
6]
2− anion complex or self-trapped exiton (STE) emission. From the pulse height spectrum acquired under irradiation by
137Cs gamma source, 4.3% of energy resolution was achieved. The light yield was estimated to be (47,000 ± 4,700) ph/MeV, comparing to
:Ce (LYSO) crystal (see
Table 5). This high value of the measured light yield was explained by authors reminding that Tl
+ ion has a higher electronegativity than the alkali ion [
27], thus band gap could decrease, which leads to increase of LY for the same energy deposition.
A bit later, results on the excellent pulse shape discrimination (PSD) achieved with the same TZC crystal sample (8 × 8 × 1.5 mm
3) under irradiation by 662 keV gamma quanta of
137Cs and 5.48 MeV external alpha particles from
241Am sources were published in [
32]. The energy resolution was calculated to be 4.3% for 662 keV gammas (see
Table 5) and 6.5% for 5.48 MeV alphas, correspondingly. Slightly worsened energy resolution for alpha particles was explained by roughness of the crystal surface and possible light leakage through the hole for alpha particle irradiation in Teflon tape wrapping crystal. The quenching factor (QF) was estimated to be 0.29. By more careful fitting of the scintillating pulses with a model that contains three exponential components, authors achieved excellent particle discrimination. For example, with optimum filter method the factor of merit (FOM) parameter was calculated to be 3.5, and it is higher than for CLYC crystal (FOM > 3). The greater FOM value indicates that alpha particles can be clearly separated from gammas in the energy range from 500 to 2000 keVee.
The best energy resolution of 4.0% FWHM at 662 keV was obtained with a 5 mm thick THC sample in [
28]. This significant improvement in the energy resolution in comparison to earlier published values (17.7%) was assigned to a better crystal quality due to the modified method of growth. As well as the separation of impurities by the sublimation resulted in a better quality crystal, leading to a notable improvement in the scintillation characteristics. The light yield of the THC crystal was estimated by comparing its 662 keV photopeak position with that of a calibrated LYSO:Ce scintillating crystal, and it was calculated to be (32,000 ± 3,200) ph/MeV (see
Table 5). Scintillating pulses under gamma quanta excitation consisted of three decay constants 0.36 μs, 1.04 μs and 14.9 μs with relative amplitudes 18.9%, 61.2% and 19.8%, respectively. At the same time, pulses under irradiation by alpha particles were described by fitting with four components—0.09 μs (8.5%), 0.46 μs (28.9%), 1.04 μs (44.9%) and 11.2 μs (18%). The quenching factor was estimated to be 0.34 for 5.4 MeV alpha particles. This crystal also exhibited a great pulse shape discrimination ability for alpha particle and gamma quanta with FOM of 2.6.
It should be noticed that detection efficiency of the THC crystal to a full absorption peak was significantly higher in comparison to CHC due to the presence of heavy Tl+ ions. For example, the detection efficiency of 3 cm thick THC crystal at the energy of 3 MeV was 280%, 220%, and 170% better than those for the NaI(Tl), CHC and La crystals, respectively.
Moreover, very recently, the PSD ability was demonstrated for TZC crystals not only for gamma quanta and alpha particles, but also for neutrons and protons [
29]. Owing the presence of
35Cl in its chemical formula (Tl
2ZrCl
6), fast neutrons can be detected though (n,p) and (n,alpha) nuclear reactions. Three different clusters of signals were detected under irradiation of the TZC crystal by AmBe source, and ascribed to registration of gammas, alpha particles, and protons. Thus, this material is a potential candidate for both high energy neutrons and gamma detection and spectroscopy.
A group of authors [
31] implemented one alternative way to grow THC from raw materials with purity grade TlCl (99.99%) and ZrCl
4 (99.9%), using a low melting point halide flux (with melting point less than 350 °C) to control a high vapor pressure of
component. They succeeded in obtaining the THC crystal sample with volume of about 1 cm
3 and to characterize it. The light yield of this THC crystal was estimated in comparison to
(BGO) scintillating crystal, and calculated to be about 20,000 ph/MeV. The energy resolution at that energy was about 5% (see
Table 5).
To date, the largest volume of high-quality THC and TZC crystals was grown by group [
30]. This group is well-known because of the re-invention of CHC crystal in 2015 [
6]. Since then, all efforts have been focused into production of large volume high-quality CHC-family crystals with a highest scintillating performance. In this study, the THC and TZC crystal boules were grown from raw materials with purity grade TlCl (99.99%),
(99.9%) and ZrCl
4 (99.9%), and following purification and growth procedures as described in [
13] for production of the high quality CHC crystal. At the end, transparent and crack-free crystal boules with outer diameter of 16 mm and about 80 mm long were obtained. For crystal characterization, samples with ∅16 × 16 mm
3 rough dimensions were cut by diamond saw and polished up to 4000 grit Al
2O
3 polishing pad. This group typically utilizes unusual method to measure scintillating light yield and energy resolution of crystals under gamma source irradiation. Here, crystal samples were placed into a quartz cup filled with a mineral oil and wrapped by Teflon tape on lateral side, while top was closed by Teflon sheet. In that configuration, one could expect to achieve light output enhancing due to minimization of refraction index mismatch between THC/TZC crystals and mineral oil, and later, this extracted scintillation light is more effectively redirected to the PMT window by diffuse reflector (Teflon tape and sheet) with further homogenization. Both these parameters lead to an improvement of the relative light yield and energy resolution detected with a crystal in such configuration.
The energy resolution of 3.7% was measured for the THC crystal sample under irradiation by 662 keV line of
137Cs gamma source, while energy resolution of 3.4% was achieved with the TZC crystal. It should be noticed that these values of energy resolution are the best among all quoted up to date for those scintillating materials. The relative light yield of 27,000 ph/MeV for THC and 35,000 ph/MeV for TZC was measured, respectively (see
Table 5). Primary decay times improve from about 4 μs for CHC to about 1 μs for THC and 2 μs for TZC, which makes these compounds more preferable in gamma-spectroscopy applications.
Brief information about recently produced THC and TZC crystals by different groups are listed in
Table 5, including the chemical purity of used raw materials and additional conditions (like a low temperature flux or additional sublimation of raw materials), dimensions of crystal samples used for its characterization, relative light yield, and energy resolution measured under irradiation by 662 keV gamma line.
All authors noticed that further improvements of the light yield and energy resolution may be pursued through additional purification of raw materials and crystal growth process modification. An obvious path is the further purification of Hf and Zr-halides, since market available purity grade is no more than 99.9%, which is less than required for a high-quality crystal production. One should aim for chemical purity of about 99.999%, and only in that case the highest quality crystals are expected to be produced in furnaces with optimized thermal gradient and growth conditions. For example, previously produced CHC crystals from additionally purified powder by multi-fold sublimation exhibit better scintillating properties than one produced from as-received raw materials, as was described above in CHC section.
THC and TZC crystals, where Cs+ ions are substituted by Tl+ ions, possessed higher density, effective atomic number (Zeff) and faster scintillation kinetics, and all these parameters are vital for enhancing an X-ray/gamma-ray detection efficiency. Therefore, recent measurements and characterization mostly focused on its preliminary light yield evaluation, energy resolution determination and scintillating pulses decay times estimation. From achieved results, it is clear that these materials are very promising for homeland security, nuclear proliferation, gamma spectroscopy and other applications, like fast neutrons detection via reactions on 35Cl isotope. For the moment, there are no results of measurements aiming to evaluate its internal radioactive contamination (internal background) that will help to understand applicability of such crystals in fundamental research looking for rare nuclear processes or other low-background applications. Typically, scintillating materials should contain radioactive nuclides on the activity level of about 1 mBq/kg, or less. Such a demanding radiopurity level can only be achieved by careful selection of raw materials from suppliers that provide the lowest initially radioactive contamination, as well as through the use of highly effective purification and synthesis technologies.
11. Discussion
As one could see from the presented results, within last five years, a large number of articles were published on CHC and CHC-family crystals production, technology variation and its optimization, as well as on characterization of their scintillating performance. Additionally, studies of the crystal defect structure and luminescence centers nature, influence of the variety of doping elements, chemical impurities level, internal radioactive background, and the first confident registration of rare decays that occurs in Hf isotopes have been performed using CHC crystals.
Obviously, CHC and its related compounds are very promising as materials for gamma-spectroscopy, and therefore many groups are attempting to contribute in technology development and ultimate scintillating performance achievement. Despite the fact that one could see a quite random situation with a chemical purity grade of utilized raw materials, mainly for Hf and Zr-based compounds (99–99.9%, no more), as was already mentioned above, and also noticed by some authors, only the highest purity raw materials in combination with highly effective purification methods will allow to obtain the high-quality large volume and crack free CHC crystals. Therefore, detailed studies on raw materials purification methods and conditioning are strongly desired. Moreover, to be able to do more quantitative analysis, and consequently, have a clearer view of the crystal quality driving parameters, efforts made in CHC growth technology development should be accompanied by raw material characterization by mass-spectrometry at each step of the technological sequence. This could include but is not limited to full characterization of as-received raw materials, materials after purification, the leftover after purification, as-grown CHC crystal, and impurities distribution along the crystal growth axis. Knowledge of present impurities and their concentrations will help to identify and properly assign defects related to undesired impurities or to main luminescent centers, which are responsible for a high scintillating performance of these materials. To date, such mass-spectrometry measurements of initial raw materials, materials after purification and the final CHC crystal were done only once in [
40].
One of the main questions that every researcher is interested to answer regarding CHC and CHC-family crystals is—what is the absolute light yield of these crystals and how to reach its maximum? However, currently, no measurements have been performed to determine the absolute light yield. One of the most confident ways to measure it—is to use the light integration sphere. While, values that are presented in all above listed articles, where the peak position under irradiation of known energy gamma-line is comparing to that measured with scintillator of well-defined light yield is the relative light yield measurements, or even to be more correct—the relative light output measurements. Moreover, to correctly estimate the relative light output, one should take into account many experimental parameters that are varying from one crystal to another. For example, light collection efficiency parameter, which is showing how much light is escaped from the crystal and reached PMT window. This strongly depends on the crystal size, its shape and surface treatment, type of the reflector placed around the crystal surface, transmittance of the crystal material to the emitted light, the difference in refractive index of the studied crystal—optical grease—light detector window system, light loses in the optical grease, etc. Without knowing this correction parameter, one cannot compare crystals of different sizes and shapes, since this will introduce a significant error into the calculated value of the relative light output. Another important parameter is the matching spectrum of emitted light and spectral sensitivity of the used light detector. In the best case, it should take into account not only the spectral sensitivity of the light detector at the peaking wavelength, but should be deconvolution of the entire crystal emission spectrum and the entire sensitivity spectrum of the light detector. One important but often omitted experimental parameter is the integration time of the certain set-up and characteristic scintillating time of the standard crystal used for comparison. Since, typically the relative light output is compared to that of the NaI(Tl) scintillating crystal, which has decay time of about 0.25 μs, within the time window of the integration (shaping) time (6–10) μs, its emitted light will be fully collected, while for the CHC crystal, with decay time of about 4 μs, one would need to utilize longer integration time to collect the whole light, otherwise light output will be underestimated. As an alternative solution, one should use a standard scintillator with comparable decay characteristics of scintillating pulses. Therefore, a wide spread of the light output values from different groups, is not only related to the quality of the obtained CHC crystals, but could partially be caused by the above mentioned experimental parameters.
One should also say few words about measurements of the energy resolution under irradiation of high energy gamma sources, like 137Cs. While for a small size samples (less than cm3), one could easily observe second peak at lower energy, which is corresponded to escaped X-rays of Hf, Tl, Cs or Zr elements from the bulk, with larger samples—the main and escape peaks are merged, thus spreading this cumulative gamma-peak. Hence, there is a logical question: should one describe a gamma-peak as a sum of two Gaussians, or as a single peak at full-absorption energy peak? Depending on your choice, one will get a slightly different FWHM value. Consequently, the energy resolution achieved with your device could be affected by a fitting procedure and not fully represent the quality of the crystal. Thus, this point should be carefully addressed during crystal characterization measurements.
Many questions could occur to careful readers concerning determination of the characteristic decay time components of scintillation pulses. For the sake of quantitative representation, the scintillation decay curves are often approximated by the sum of several exponential functions: where is background, and are amplitudes and decay time constants, respectively. The scintillation decay curves of CHC samples were fitted using from one to four exponential functions that ensures the best quality of the fit at certain experimental conditions, mainly pulse acquisition time (typically a few tens of μs). However, it should be highlighted that the slow decay time components are essential for the analysis of the final part (tail) of the decay curves, since they would significantly contribute to amplitude of the scintillating pulses, when acquiring in a wide time interval, and consequently, will affect the light output estimation. Moreover, the proper description of scintillating pulse decay components, their actual number and relative amplitudes could be achieved only as a result of a dedicated complex studies involving measurements with TSL and EPR methods, detail photoluminescence analysis under different excitation energies and low-temperature measurements in an optical cryostat, where scintillating signals are recorded over few hundreds of μs, to ensure the full collection of the emitted light. Otherwise, values of the exponential components as a result of the fitting scintillation pulses could be considered only as a preliminary estimation, or as an instrumental characteristic time, demonstrating the fact how kinetics of the light emission in a certain crystal is matching with a shaping tract of the spectrometric unit. Therefore, one could naturally expect a slight variation of decay time values for the same crystal type presented by various groups as a reflection of different chemical purity of raw materials and consequently uncontrolled impurities presence into the crystal (that could lead to slow components contribution enhancement), and by different time intervals of scintillation pulses recording. Further detailed studies on kinetics of scintillation pulses for whole members of CHC-family crystals are strongly desired.
Most characterization measurements with CHC-family crystals were performed under irradiation by gamma quanta, mainly, from a 137Cs gamma source. This is a useful and prompt way how to estimate the general detector performance for gamma-spectroscopy. While only few groups are turning their attention to characterization of the scintillating response of these crystals to alpha particles. More information is needed on energy resolution, quenching factor and pulse shape of scintillating pulses as a function of alpha particles energy. Since alpha particles interact very locally with detector material (in the range of a few to ten μm), while gamma quanta are involving bulk volume through interaction, one could gain a very important information on the difference in excitation energy transfer to luminescence centers at different ionization density, which occur through alpha and gamma events. In addition, the saturation effect at luminescence centers could be studied. Moreover, due to redistribution of the incident energy between decay components of the scintillating pulse for alphas with respect to that for gammas events, it will be possible to study the effect of the involvement of various luminescence centers into the scintillating pulse formation.
As previously mentioned above, with CHC crystals the quenching factor for alpha particles was measured to be about 0.35–0.40 in the energy region of (4–8) MeV, which is a factor of two larger than for any oxide crystals. Based on some phenomenological model, this material could also have an intensive scintillating pulse for nuclear recoil events. However, this assumption should be proved in dedicated measurements with CHC crystal under neutron irradiation. CHC-family crystals could be promising target materials for the direct DM search experiments due to high light output, excellent energy resolution, low energy threshold, expected high quenching factor for nuclear recoils and an ultimate flexibility of crystalline matrix for element substitution. The latter is very important, when considering different models of DM particles with a wide spread of expected masses, hence having an adjustable target material one could match the expected energy of the nuclear recoil, based on kinematics of the DM interaction with regular matter, to energy region with optimal detector performance. As an additional advantage, CHC-family crystals have no directionality of scintillating response due to their cubic crystalline structure. Thus, the energy resolution for alpha particles and nuclear recoils should not depend on the direction of the interaction with detector, in contrast for example to CdWO
4 [
3] crystals. It leads to better background discrimination and, as a consequence, to better experimental sensitivity. In light of non-fundamental applications, knowledge of pulse shape and quenching factor will help to design dual-channel detectors, i.e., gamma-neutrons or gamma-alphas spectrometers, which could be useful in homeland security applications.
To provide a fast turn-over with a new type of crystal that related to the CHC-family and facilitate progress in that direction, one should utilize an approach that would allow for a fast test-growth under set of desired changeable conditions (various raw material purity grade, alternative purification procedures, doping with numerous different activators, different thermal conditions and pulling rate, etc.). Recently, such method using the micro-pulling-down apparatus, so-called “miniaturized-vertical-Bridgman” (mVB), was described in detail [
41]. The structural and optical quality of the
crystal grown within the mVB approach was confirmed to be comparable to the quality of
crystals grown by the standard vertical Bridgman method. Hence, authors claim the introduction of a time- and cost-effective method for the single crystal growth that is suitable for a fast screening of A
2MX
6 family compounds.
CHC-family crystals are promising not only as scintillating detectors, but also could have a wide usage in optoelectronic applications. For example, recently the first colloidal synthesis of vacancy-ordered lead-free perovskite
ZrCl
6 nanocrystals was reported in [
42]. The unique vacancy-ordered structure of CZC results in a strongly localized charge-carriers with the formation of self-trapped excitons (STEs). CZC nanocrystals exhibit a high photoluminescence quantum yield of 60.4% at room temperature, with a broadband photoluminescence emission peaked at 446 nm. The emission color can be easily tuned from blue to green by synthesizing the
ZrBr
xCl
(6-x) mixed-halide nanocrystals. In addition, the CZC nanocrystals exhibit a high thermal stability up to 650 °C and long-term air stability for over 6 months. These results suggest that
ZrCl
6 nanocrystals, and more generally A
2MX
6 family compounds, are promising for optoelectronic applications, and this first study could stimulate future research in the design of new environmentally friendly (Pb-free) nanocrystals.