Fast nuclide identification based on a sequential bayesian method

NUCLEAR ELECTRONICS AND INSTRUMENTATION

Fast nuclide identification based on a sequential bayesian method

Xiao-Zhe Li
Qing-Xian Zhang
He-Yi Tan
Zhi-Qiang Cheng
Liang-Quan Ge
Guo-Qiang Zeng
Wan-Chang Lai
Nuclear Science and TechniquesVol.32, No.12Article number 143Published in print Dec 2021Available online 20 Dec 2021
6500

The rapid identification of radioactive substances in public areas is crucial. However, traditional nuclide identification methods only consider information regarding the full energy peaks of the gamma-ray spectrum and require long recording times, which lead to long response times. In this paper, a novel identification method using the event mode sequence (EMS) information of target radionuclides is proposed. The EMS of a target radionuclide and natural background radiation were established as two different probabilistic models and a decision function based on Bayesian inference and sequential testing was constructed. The proposed detection scheme individually processes each photon. When a photon is detected and accepted, the corresponding posterior probability distribution parameters are estimated using Bayesian inference and the decision function is updated. Then, value of the decision function is compared to preset detection thresholds to obtain a detection result. Experiments on different target radionuclides (137Cs and 60Co) were performed. The count rates of the regions of interest (ROI) in the backgrounds between [651, 671], [1154, 1186], and [1310, 1350] keV were 5.05, 3.83, and 3.61 CPS, respectively. The experimental results demonstrate that the proposed method can identify 137Cs in 3.8 s with a full energy peak count rate of 5.05 s−1 and can identify 60Co in 4.1 s with a full energy peak count rate of 7.44 s−1. The results demonstrate that the proposed method can detect radioactive substances with low activity.

Natural radiationNuclide identificationSequential testingNuclear safety
1

Introduction

Radioactive substances are a threat to public safety when they are present in public areas. In such cases, nuclide identification is crucial for inspecting radioactive substances. Traditional radionuclide identification methods[1-2] are based on the full energy peak information of the gamma-ray spectrum and require that the particle events generated by the decay of nuclides are treated as statistics in the spectrum. Peak-finding algorithms such as the derivative peak searching method, second derivative peak identification method[3], covariance peak-search method[4], and symmetric zero-area peak searching method[5] use full energy peaks information to identify nuclides based on the energy of the gamma rays from nuclides. To obtain a low-statistic spectrum, a long spectrum recording time is required, resulting in a slow response time. To compensate for the shortcomings of traditional radionuclide identification methods, novel detection methods have been proposed, including the adaptive filtering method[6], chaos oscillator method, blind-source separation method, and stochastic resonance method[4]. The adaptive filtering method is a weak signal detection method with an optimal parameter adjustment function that was developed based on the Kalman filter. This method can automatically optimize a nuclear signal. For the Chaos oscillator method, a nuclear signal in the form of a periodic signal is superimposed with a fixed-frequency chaotic oscillator. The mixed signal is then detected and processed using phase differences. The blind-source separation method uses an analysis algorithm to obtain the best estimate of a nuclear signal. The stochastic resonance method uses the synergistic effects of noise and signals to analyze nuclear signals. Sullivan adopted wavelet transforms and the modulus maxima method[78] to achieve accurate nuclide identification using low-resolution spectra. Several artificial neural network (ANN) methods have also been introduced in previous studies[912] based on their significant success in other research fields. Liang used the K-L transform to extract gamma-ray spectrum features and trained an ANN to perform radionuclide identification. Experiments and tests demonstrated that the ANN method is effective for rapid radionuclide identification[12]. Furthermore, Yang et al. proposed a radionuclide identification method based on machine learning and pattern recognition[1318], and Chong Jie and Yi Ming used a fuzzy-logic-based algorithm to discriminate characteristic peaks and identify radionuclides[19,23,24]. Although the aforementioned gamma ray spectrum analysis methods have proven to be reliable and practical, a long recording time is required to obtain a spectrum with low statistical error. Therefore, these techniques cannot meet the requirements for real-time radionuclide inspection in public areas.

In 1945, Wald proposed the sequential probability ratio testing theory[22], which can be used to perform real-time analysis and judgment regarding whether a hypothesis is true based on recorded events. The results satisfied statistical requirements. The sequential probability ratio test combined with the Bayesian law for posterior probability distributions has been applied in many fields[23,24]. Candy proposed a sequential Bayesian radionuclide identification method in 2008. This method combines the sequential probability ratio test and Bayes rule for radiation monitoring and radionuclide identification, and establishes a radiation detection model. In 2010, Candy et al. developed SRaDS, which is a radiation monitoring system using a sequential-Bayesian-based radionuclide identification method[25]. SRaDS automatically eliminates irrelevant and unexpected photons during the detection process and improves detection efficiency. This system abandons the concept of the energy spectrum statistics used for traditional nuclide recognition. It uses a Bayesian algorithm to group each target photon, compares preset parameters for each photoelectric event sequence, and calculates the probability of the target radionuclide corresponding to the photoelectric event sequence. Finally, the statistical results of multiple single-energy ray groups are used to determine whether the target radionuclides are detected. This method can successfully detect radionuclides when the number of recorded photons is small and the measurement time is short[25,26].

This method makes full use of event mode sequence (EMS) information. An EMS of the target radionuclide can be treated as a composition of EMSs emitted from several monoenergetic source components. A well-defined probability model can then be used to describe the EMS of the target radionuclide. Therefore, a sequential Bayesian detection scheme can be applied to identify radioactive nuclides in real time by incorporating energy, count rate, and emission ratio information[24,27]. Unlike the traditional full-spectrum-analysis-based method, this method can detect radionuclides under low-count conditions. However, the results of this method are affected by the background count and some simulation experiments by Xiang Qingpei verified that it may be inaccurate in cases with high background counts[25,26].

In this study, two different probabilistic models were established based on the EMSs of natural background radiation and target nuclide radiation fields. The posterior distribution parameters of the probabilistic models were then estimated using Bayes’ law. Finally, nuclide identification was performed via sequential detection. Once a photon has been detected and processed, the posterior probability distribution parameters can be estimated using Bayesian inference and a decision function can be updated and compared to preset detection thresholds to determine whether the target radionuclide is detected. In this paper, a rapid nuclide identification model and sequential detection scheme are introduced, followed by the presentation of validation experiments on target radionuclides (137Cs and 60Co).

2

Methods

2.1
Nuclide identification model based on background comparisons

The concept of monoenergetic decomposition[13] has been proposed to leverage EMS information. Suppose that the sequence of photons emitted by a radionuclide is a set of photons emitted by several separate monoenergetic sources. An Event(n;em(n),Δτm(n)), is defined as the nth gamma ray recorded by the detector. This ray is emitted by the mth monoenergetic source with a measured energy em(n) and interarrival time Δτm(n) between the current and previous event. The EMS of a target radionuclide (EMSrn) is then represented as EMSrn(N;e_,Δτ_). EMSrn(N;e_,Δτ_):=m=1Mn=1NmEvent(n;em(n),Δτm(n)) (1)

The index m represents the mth monoenergetic source of a target radionuclide and the total number of monoenergetic sources is M. The detector detects Nm events from the mth monoenergetic source during the detection period. e_ is defined as the complete set of energies composing EMSrn along with Δτ_, which is the corresponding set of interarrival times. N is the sum of the set Nm.

Considering instrument noise and experimental uncertainty, the measured energy of the gamma rays emitted by a monoenergetic source in a radiation field follows a Gaussian distribution and the interarrival times follow an exponential distribution (based on the Poisson statistics of nuclear decay[28]). The corresponding formula is presented in Eq. (2), where e¯m and σem are the expected value and standard deviation of the Gaussian distribution, respectively. λm is the expected value of the exponential distribution and αm is the emission probability of the mth monoenergetic source. P(em(n)|radionuclide)=P(em(n)|e¯m,σem2)=12πσeme(em(n)e¯m)22σem2P(Δτm(n)|radionuclide)=P(Δτm(n)|λm)=λmeλmΔτm(n)P(Event(n;em(n),Δτm(n)|radionuclide)=αm×P(em(n)|radionuclide)×P(Δτm(n)|radionuclide) (2)

When a radionuclide identification instrument is placed in a public area, the background radiation is constant. For an EMS measured in the presence of natural background radiation, the region of interest (ROI) in the spectrum consists of the Compton platform and other types of background radiation. Therefore, we modeled the measured energy of a gamma ray emitted by a monoenergetic source in an ROI uniformly and the interarrival times exponentially. P(em(n)|background)=P(em(n)|emleft,emright)=1emrightemleft,P(Δτm(n)|background)=P(Δτm(n)|λmbase)=λmbaseeλmbaseΔτm(n),P(Event(n;em(n),Δτm(n)|background)=λmbaseM(emrightemleft)eλmbaseΔτm(n). (3)

In Eq. (3), emleft and emright are the left and right energies of the ROI, respectively, and λmΔτbase is the background count rate of the ROI. The hypothesis test defined in Eq. (4) is applied to identify the nuclide through sequential detection based on the two hypotheses. H1:EMSrn(N;e_,Δτ_)~P(e_,Δτ_|radionuclide)if radionuclide is detected,  H0:EMSrn(N;e_,Δτ_)~P(e_,Δτ_|background)if radionuclide is not detected (4)

When a radionuclide is detected, P(e_, Δτ_|radionuclide) is considered as the probabilistic model of the EMS. Otherwise, the probabilistic model of the EMS is P(e_, Δτ_|background). The optimal solution to this binary decision problem is provided by the Wald sequential probability ratio test[21]. It is assumed that the Nth detected photon is emitted from the mth monoenergetic source. According to the definition in Eq. (1), there are a total of Nm photons detected from the mth monoenergetic source and the likelihood ratio Ratio(N) is obtained by applying the Newman–Pearson theorem as follows: pic pic pic (5)

The decision function Λ(N) is obtained by taking the logarithm of Ratio(N), as shown in Equation (6). Λ(N)=Λ(N1)+lnP(Event(Nm;em(Nm),Δτm(Nm)|EMSrn(N1;e_,Δτ_),H1)lnP(Event(Nm;em(Nm),Δτm(Nm)|EMSrn(N1;e_,Δτ_),H0). (6)

With the Nth photon recorded by the detector, the posterior probabilistic distribution parameter of the EMS model under H1 is estimated using Bayesian inference. The decision function is then updated according to the resulting parameters and compared to preset detection thresholds to obtain detection results. The Wald sequential probability ratio test result is defined in Eq. (7), where T1 and T0 are the detection thresholds[22]. Λ(N)>T1 ,Accept H1,T0Λ(N)T1, Continue,Λ(N)<T0 ,Accept H0. (7)

After defining a detection method, it is necessary to infer parameters and finalize a decision function. Based on the specified probability distribution form and observed data, the posterior distribution of the mth monoenergetic source under H1 is defined in Eq. (8), where the probabilistic model is decomposed by applying Bayes’ rule[32,33] and αm(Nm) is the emission rate estimated by the Nmth photon detected from the mth monoenergetic source. P(Event(Nm;em(Nm),Δτm(Nm)|EMSrn(N1;e_,Δτ_),H1)=P(Δτm(Nm)|em(Nm),αm(Nm),EMSrn(N1;e_,Δτ_),H1)×P(em(Nm)|αm(Nm),EMSrn(N1;e_,Δτ_),H1)×P(αm(Nm)|EMSrn(N1;e_,Δτ_),H1). (8)

P(αm(Nm)|EMSrn(N1;e_,Δτ_),H1) is the probability that the detected photon is emitted by the mth monoenergetic source. Additionally, αm(Nm) represents the detection ratio for the Nmth photon. The total detected number of photons is N and there are Nm photons emitted by the mth monoenergetic source. Therefore, the following equation can be derived: P(αm(Nm)|EMSrn(N1;e_,Δτ_),H1)=NmN. (9)

The measured energy of a photon detected from the mth monoenergetic source is assumed to follow a Gaussian distribution under  H1, as shown in Eq. (10). P(em(Nm)|αm(Nm),EMSrn(N1;e_,Δτ_),H1)=12πδem(Nm)e(em(Nm)e¯m(Nm))22σem2(Nm), (10)

where e¯m(Nm) and δem(Nm) are the expected value and standard deviation of the posterior distribution, respectively.

The arrival time interval Δτ is exponentially distributed and λm(Nm) is the expected value. Therefore, the following equation can be derived: P(Δτm(Nm)|em(Nm),αm(Nm),EMSrn(N1;e_,Δτ_),H1)=λm(Nm)eλm(Nm)Δτm(Nm). (11)

Then, the probabilistic model of the EMS under hypothesis H1 is given by Eq. (12). P(Event(Nm;em(Nm),Δτm(Nm)|EMSrn(N1;e_,Δτ_),H1)=Nmλm(Nm)N2πσem(Nm)×eλm(Nm)Δτm(Nm)(em(Nm)e¯m(Nm))22σem2(Nm) (12)

The probabilistic model of the EMS under hypothesis H0 is given by Eq. (13). P(Event(Nm;em(Nm),Δτm(Nm)|EMSrn(N1;e_,Δτ_),H0)=λmbaseM(emrightemleft)eλmbaseΔτm(Nm) (13)

The decision function is obtained by substituting Eqs. (12) and (13) into Eq. (6), as shown in Eq. (14). Λ(N)=Λ(N1)+lnNmlnN+ln(λm(Nm)2πσem(Nm))λm(Nm)Δτm(Nm)(em(Nm)e¯m(Nm))22σem2(Nm)+lnM+ln(emrightemleft)ln(λmbase)+λmbaseΔτm(Nm). (14)

2.2
Implementation of the proposed method

First, a radionuclide identification instrument is installed in a public area to detect the EMS of the target radionuclide under conditions with natural background radiation for a long period (at least hundreds of seconds). The count rate λmbase of the mth monoenergetic source in the natural background radiation is then obtained by taking average counts from the ROI.

Next, a parallel detection architecture is proposed to implement the sequential detection scheme. As shown in Fig. 1(a), there are two types of detection channels that process the detected photons emitted by a monoenergetic source. The first type is designed to infer the parameters of the probabilistic model and evaluate the log-likelihood under hypothesis H1, whereas the other evaluates the log-likelihood under hypothesis H0. When a photon is detected in the first type of channel, photon discrimination is performed. If an event is accepted by one of the channels, then posterior distribution parameter estimation is performed. Finally, the log-likelihood is calculated based on the probabilistic model under hypothesis H1 with updated parameters. .Event discrimination is first performed using the second type of channel. Finally, the proposed method performs a log-likelihood evaluation based on the probabilistic model under hypothesis H0 with constant parameters. After an event is processed using the detection channels, the current decision function value is updated according to Eq. (14) and compared to the preset detection thresholds to obtain detection results according to Eq. (7).

Fig. 1
Detection architecture
pic

A two-stage structure was designed to perform photon discrimination, as shown in Fig. 1(b). The energy discriminator is used to determine which channel a photon should be processed in, after which a detection rate discriminator uses the interarrival time for verification[13]. The energy discriminator performs a confidence interval test, as shown in Eq. (15), where n is the particle ordinal number, kem is the confidence coefficient of the energy, and emt and σmt are the mean and standard deviation of the energy, respectively, which can be obtained experimentally. emt kemσmt em(n) emt+ kemσmt (15)

The interarrival time discriminator also performs a confidence interval test, as shown in Eq. (16). Δτmbase kγ¯σΔτbaseΔτ¯m  Δτmbase+ kγ¯σΔτbase, σΔτbase= Δτbase/n, (16) where Δτ¯m is the average interarrival time and kγ¯ is the confidence coefficient. Δτmbase is the mean interarrival time under conditions with natural background radiation and  σΔτbase is the width of the confidence interval.

The posterior probability distribution parameter estimation of the EMS under hypothesis H1 is achieved using Bayesian inference. Because energy is modeled as a Gaussian distribution and the interarrival time is modeled as an exponential distribution (both belonging to the exponential family of distributions), it is convenient to find conjugate prior distributions and infer the posterior distribution in analytical form. Following research on Bayesian inferencing[28,29], the Gaussian–Gamma distribution was applied to model the prior distribution of e¯m and (σem2)1 because the precision parameter precem is defined as (σem2)1. The prior distributions τem(0) and e¯m(0) are defined in Equation (17), where a0, b0, u0, and v0 are hyperparameters[30]. For a weak prior, u0 is set to emt, a0 and b0 are set to one and 100, respectively, and λ0 is set to one. precem(0)~Gamma(a0,b0),e¯m(0)~N(u0,(v0precem(0))1),P(e¯m(0),precem(0))=N(e¯m(0)|u0,(v0precem(0))1)Gamma(precem(0)|a0,b0) (17)

Equation (18) models the prior distribution λm(0) of the expected value of the interarrival λm as Gamma. c0 and d0 are also hyperparameters, where c0 is set to one and d0 is set to λmbase. λm(0)~Gamma(c0,d0)P(λm(0))=Gamma(λm(0)|c0,d0) (18)

Up to the nth arrival from the mth monoenergetic source, the posterior distributions of e¯m(n), τem(n), and λm(n) are updated as shown in Equation (19), where un, vn, an, bn, cn, and dn are the hyperparameters of the posterior distributions[31]. pic (19)

The hyperparameters can be inferred in analytical form[31] via conjugacy, as shown in Equation (20). emmean(n)=i=1nem(i)nλmmean(n)=i=1nλm(i)nun=λ0u0+i=1nem(i)λ0+nvn=v0+nan=a0+n/2bn=b0+12i=1n(em(i)emmean(n))2+λ0n(emmean(n)u0)2λncn=c0+n/2dn=d0+12i=1nλm(i) (20)

The expectation of the posterior distribution can be used to obtain the parameters of the EMS model under H1. e¯m(n)=unσem(n)=1τem(n)=bnanλm(n)=cndn (21)

Then, the decision function is updated according to the parameters and the detection result is obtained by comparing the decision function to the preset detection thresholds. If the decision function value exceeds the threshold T1, a target radionuclide exists in the detection environment. If the decision function value exceeds the threshold T0, then there is no target radionuclide in the detection environment. Otherwise, the detection process continues until one of the thresholds is exceeded.

3

Experiments

3.1
Background experiment

In this experiment, the spectrum of the background was recorded for a long duration. The count rate of the ROI was stable and the relative statistical error was small. The target radioactive sources were added and the spectrum was collected to form a stable spectrum. The energy distribution model of the characteristic peaks in the ROI was then analyzed.

For a target radionuclide, the count rate of the ROI in the background can be obtained through a long-term detection experiment[34]. 137Cs and 60Co were selected as target radionuclides. The radioactivity of the 137Cs source was 8700 Bq and the radioactivity of the 60Co source was 1500 Bq. A LaBr3 (Ce) detector of size Ф1.5×1.5 inches was used in this experiment. The EX-03 multi-channel analyzer was developed by the Chengdu University of Technology to obtain the EMS. The parameters for Eqs. (15) and (16) are listed in Table 1. The experiment was performed in a laboratory environment to obtain the count rate of the ROI in the background and the detection time was set to 250 s. The threshold used in this study was obtained experimentally and the results of the experiments demonstrated that a threshold of 3.98 is acceptable.

Table 1
Preset detection parameters
Detection parameter Value
e1tof the 1st component of 137Cs 661 keV
σ1t of the 1st component of 137Cs 10 keV
e1t of the 1st component of 60Co 1170 keV
σ1t of the 1st component of 60Co 16 keV
e2t of the 2nd component of 60Co 1330 keV
σ2t of the 2nd component of 60Co 20 keV
kem,kγ¯  1.0
T0  −3.98
T1  +3.98
Show more

The spectrum and EMS detected in the background are presented in Fig. 2. The ROIs in the spectrum are labeled with red, blue, and green regions. The ROI of 137Cs ranges from 651 to 671 keV. The first ROI of 60Co ranges from 1154 to 1186 keV and the second ROI of 60Co ranges from 1310 to 1350 keV. It is clear that the counts in the ROI approximately follow a uniform distribution.

Fig. 2
(Color online) Detected EMS and spectra in different radiation fields. (a) Description of the test bench; (b) Spectrum obtained under background conditions with a detection interval of 250 s; (c) 137Cs radiation energy spectrum; (d) 60Co radiation energy spectrum
pic

Figure 2(a) presents a description of the test bench. Figure 2(b) presents the energy spectrum and EMS time-domain scatter diagram for natural background radiation. Figure 2(c) presents the energy spectrum and EMS time-domain scatter diagram of the target nuclide 137Cs and natural background radiation. Figure 2(d) presents the energy spectrum and EMS time-domain scatter diagram of the target nuclide 60CO and natural background radiation.

Mathematically, the Chi-square goodness-of-fit test was performed to validate that the energy detected in the ROI under natural background radiation conforms to a uniform distribution[31]. The relevant formula is presented in Eq. (22), where LeftChannel and RightChannel are the edges of the selected ROI, Spectrumi is the count of the ith channel in the spectrum, and T is obtained by taking the average of the counts in the ROI. Chisquare=i=LeftChannelRightChannel(SpectrumiT)2T. (22)

For the ROI of 137Cs, the Chi-square test result was 26.32. For the first ROI of 60Co ([1155, 1185] keV), the result was 35.85 and for the second ROI of 60Co ([1310, 1350] keV), the result was 47.10. The calculated chi-square values were compared to the 95% confidence threshold values, which were 31.41, 43.77, and 55.76, respectively. All of the calculated values are less than the corresponding threshold values, indicating that the energy in the ROI measured under the background conditions was uniformly distributed.

The count rate of the ROI under natural radiation background, denoted as λΔτbase, was also calculated. The average interarrival time Δτbase was obtained by taking the reciprocal of the count rate. The average count rates in the background for [651, 671] keV, [1154, 1186] keV and [1310, 1350] keV were 5.05, 3.83 and 3.61 per second, respectively. The corresponding average interarrival times were 0.1980 s, 0.2611 s, and 0.2770 s, respectively.

The rapid nuclide identification method was used to detect target radionuclides under background conditions. The detection results for 137Cs under background conditions are presented in Fig. 3(a). The detection function gradually decreases along with events that are processed until the detection function exceeds the low threshold, indicating that the background is detected in 5 s. The detection results for 60Co under background conditions are presented in Fig. 3(b). This detection function also gradually decreases with the events that are processed until the detection function exceeds the low threshold, indicating that 60Co is absent in the background.

Fig 3.
137Cs and 60Co Detection results under background radiation conditions
pic
3.2
Radionuclide detection experiment

The 137Cs source was placed 35 cm in front of the LaBr3 detector to verify the proposed method under a radiation field and detection was performed for 100 s. The detected EMS and spectrum are presented in Fig. 2 (b). A fitted Gaussian function (red line) is placed above the ROI, and it is clear that the energy detected in the ROI under a radiation field approximately follows a normal distribution. In addition to this graphical depiction, the Chi-square test was performed to validate that the energy detected in the ROI under a radiation field approximately follows a normal distribution. The relevant formula is presented in Eq. (23). Chisquare= i=LeftChannelRightChannel(SpectrumiIntegralValuei)2IntegralValuei, (23) where LeftChannel and RightChannel are the edges of the selected ROI, Spectrumi is the count of the ith channel in the spectrum, and IntegralValue 

is obtained using Eq. (24). IntegralValuei=TotalCount ×eleftierighti12πσte(eet)2(δt)2, (24) where TotalCount is the total count in the ROI, and elefti and erighti are the left and right energies of the ith bin in the spectrum, respectively. Additionally, et and σt are preset distribution parameters from Table 1.

The chi-square test result for the ROI [651,671] was 26.32. When comparing the calculated Chi-square value to the value of the 95% confidence threshold of 31.41, the calculated value is less than the threshold, indicating that the energy spectrum in the ROI under the 137Cs radiation field and long-term detection conforms to a normal distribution.

A rapid nuclide identification method was used to detect 137Cs in the corresponding radiation field. The detection function is presented in Fig. 4(a), where the decision function gradually increases with events that are processed and eventually exceeds the high threshold, indicating that the 137Cs source is detected in 3.8 s, where the count rate of the ROI is 5.05 s−1.

Fig. 4
(Color online) 137Cs and 60Co decision results in 137Cs and 60Co spectra. (a) Decision function and spectrum obtained when 137Cs is identified. (b) Decision function and spectrum obtained when 60Co is identified.
pic

There were only a few counts in the ROI before 137Cs that were identified and the counts in the ROI were not sufficient to form a peak. According to Eqs. (23) and (24), the Chi-square test was performed to determine whether the spectrum in the ROI follows a Gaussian distribution. The test value was 41.76, which is greater than the threshold value of 31.41, indicating that traditional full-spectrum-analysis-based radionuclide identification methods cannot identify 137Cs at low values.

The 60Co source was placed 20 cm in front of the LaBr3 detector and detection was performed for 100 s. The EMS and spectrum detected are presented in Fig. 2(c), where two fitted Gaussian functions (red and blue lines) are placed in the ROI. The energy detected in the ROI under the 60Co radiation field approximately follows a Gaussian distribution. By using Eqs. (23) and (24), the Chi-square test was also performed for validation.

The Chi-square test results for ROI [1155, 1185] and [1310, 1350] keV were 35.85, and 47.10, respectively. The calculated Chi-square values were compared to the values of the 95% confidence threshold, which were 43.77 and 55.76, respectively. All calculated values were less than the threshold values, indicating that the energy in the ROI detected in the spectrum under radiation and long-term detection conforms to a Gaussian distribution.

The rapid nuclide identification method was used to detect 60Co in the corresponding radiation field. The detection results are presented in Fig. 4(b), the decision function gradually increased with events that are processed and eventually exceeds the high threshold, indicating that there is a 60Cs source in the radiation field. The proposed method can detect 60Cs in 4.1 s, where the count rate of the corresponding ROI is 7.44 s−1. Only a few counts in the ROI before 60Co were detected. To determine if the counts in the ROI were insufficient to form a peak, according to Eqs. (23) and (24), a Chi-square test was performed to demonstrate that the spectrum in the ROI did not follow a Gaussian distribution. The test values of the ROIs of [1155, 1185] keV and [1310, 1350] keV were 48.59 and 60.34, respectively, and the 95% confidence threshold values were 43.77 and 55.76, respectively. It is clear that the calculated values are greater than the detection thresholds. Therefore, traditional full-spectrum-analysis-based radionuclide identification methods cannot identify 60Co with such a low counts.

3.3
Detection performance experiments

Experiments were conducted while varying the distance between the source and detector to evaluate the performance of the proposed method. The step length was 5 cm and the distance ranged from 5 to 100 cm. For each distance, the detection experiments were repeated 20 times to analyze the statistics of the identification time. The average detection time and variance for each test group were calculated and plotted. The results for the identification time required to cover the distance between the source and detector are presented in Fig. 5.

Fig. 5
Relationship between detection time and detection distance. Identification times required to cover the distances to the 137Cs source (left) and 60Co source (right).
pic

As shown in the figure above, when the detection distance is set to 5 cm, the 137Cs detection time is 0.2 s. As the distance increases, the detection time also increases. When the distance reaches 95 cm, the average detection time increases to 6.19 s. Regarding the two types of sources with low activity in the experiment, when the detection distance reaches 90 cm, the detection times of the target nuclide 137Cs and target nuclide 60CO reach 7 s and 12 s, respectively. For the measurement of 60Co, the average detection time is within 2 s when the detection distance is within 25 cm. As the detection distance increases, the detection time also increases and the relative distance reaches 60 cm. The average detection time is 6 s, which meets the requirements for rapid identification.

4

Conclusions and prospects

Based on the Bayesian sequential detection method, this paper introduced a rapid nuclide identification method for radionuclide inspection in public areas. Unlike other traditional full-spectrum-analysis-based methods, in this study, based on the EMS measurement of background radiation and the target radiation field, two EMS sequences were established, a priori models were estimated based on effective particle events, and an a posteriori model was obtained. The target nuclides were identified via sequential detection of the posteriori models.

To prove that the efficiency of the proposed background-comparison-based radionuclide identification method can meet the requirements of rapid identification under different measurement conditions, several groups of experiments were conducted. The relative distance between the detector and two radioactive sources was controlled and multiple measurements were performed at distances of 0 to 100 cm to obtain detection time curves for the radioactive sources. The average detection time was 6 s for 60Co (with an activity of 1500 Bq) at a distance of 60 cm from the detector. The average detection time was 7 s for 137Cs (with an activity of 8700 Bq) at a distance of 90 cm from the detector. These results demonstrate that rapid nuclide identification can be achieved using the proposed background-comparison-based radionuclide identification method.

To improve the adaptability of the proposed method to complex environments, further research should focus on the rapid detection and identification of radionuclides in multiple-nuclide mixtures under conditions with complex motion.

References
[1] G. F. Knoll, Radiation Detection and Measurement, 3rd ed. Hoboken, NJ: Wiley, 2000.
[2] G. Gilmore and J. Hemingway, Practical Gamma-Ray Spectrometry. Hoboken, NJ: Wiley, 2003.
[3] M.A. Mariscotti,

A method for automatic identification of peaks in the presence of background and its application to spectrum analysis

. Nucl. Instrum. Methods 50, 309-320 (1967). doi: 10.1016/0029-554X(67)90058-4
Baidu ScholarGoogle Scholar
[4] J.M. Zhang,

Methods and experiments for weak nuclear signal detection and nuclide identification. Doctoral dissertation

, University of Science and Technology of China (In Chinese)
Baidu ScholarGoogle Scholar
[5] R. Shi, X.G. Tuo, H.L. Li et al.,

Unfolding analysis of LaBr3:Ce gamma spectrum with a detector response matrix constructing algorithm based on energy resolution calibration

. Nucl. Sci. Tech. 29, 1 (2018). doi: 10.1007/s41365-017-0340-6
Baidu ScholarGoogle Scholar
[6] Y.L. Song, F.Q. Zhou, Y. Li et al.,

Methods for obtaining characteristic γ-ray net peak count from interlaced overlap peak in HPGe γ-ray spectrometer system

. Nucl. Sci. Tech. 30, 11 (2019). doi: 10.1007/s41365-018-0525-7
Baidu ScholarGoogle Scholar
[7] C.J. Sullivan, S.E. Garner, K.B. Butterfield,

Wavelet analysis of gamma-ray spectral

. IEEE Nuclear Science Symposium Conference Record Nuclear Science 2004, 281-286 (2004). doi: 10.1109/NSSMIC.2004.1462198
Baidu ScholarGoogle Scholar
[8] C.J. Sullivan, M.E. Martinez, S.E. Garner,

Wavelet analysis of sodium iodide spectral

. IEEE T. Nucl. Sci. 5(33): 2916 (2006). doi: 10.1109/TNS.2006.881909
Baidu ScholarGoogle Scholar
[9] M. Kamuda, J. Stinnett, C.J. Sullivan,

Automated isotope identification algorithm using artificial neural networks

. IEEE T. Nucl. Sci. 64, 1858-1864 (2017).doi: 10.1109/TNS.2017.2693152
Baidu ScholarGoogle Scholar
[10] J.-P. He, X.-B. Tang, P. Gong et al.,

Spectrometry analysis based on approximation coefficients and deep belief networks

. Nucl. Sci. Tech. 29, 69 (2018). doi: 10.1007/s41365-018-0402-4
Baidu ScholarGoogle Scholar
[11] J. He, X. Tang, P. Gong et al.,

Rapid radionuclide identification algorithm based on the discrete cosine transform and BP neural network

. Ann. Nucl. Energy 112, 1-8 (2018). doi: 10.1016/j.anucene.2017.09.032
Baidu ScholarGoogle Scholar
[12] L. Chen, Y. Wei,

Nuclear identification algorithm based on K-L transform and neural networks

. Nucl. Instrum. Meth. Physics Research Sect. A 598(2): 450-453 (2009). doi: 10.1016/j.nima.2008.09.035.
Baidu ScholarGoogle Scholar
[13] C. Bobin, O. Bichler, V. Lourenço et al.,

Real-time radionuclide identification in γ-emitter mixtures based on spiking neural network

. Appl. Radiat. Isot. 109, 405-409 (2016). doi: 10.1016/j.apradiso.2015.12.029
Baidu ScholarGoogle Scholar
[14] E.G. Androulakaki, M. Kokkoris, C. Tsabaris et al.,

In situγ-ray spectrometry in the marine environment using full spectrum analysis for natural radionuclides

. Appl. Radiat. Isot. 114, 76-86 (2016). doi: 10.1016/j.apradiso.2016.05.008
Baidu ScholarGoogle Scholar
[15] C. Tsabaris, E.G. Androulakaki, S. Alexakis et al.,

An in-situ gamma-ray spectrometer for the deep ocean

. Appl. Radiat. Isot. 142, 120-127 (2018). doi: 10.1016/j.apradiso.2018.08.024.
Baidu ScholarGoogle Scholar
[16] D.K. Fagan, S.M. Robinson, R.C. Runkle,

Statistical methods applied to gamma-ray spectroscopy algorithms in nuclear security missions

. Appl. Radiat. Isot. 70, 2428-2439 (2012). doi: 10.1016/j.apradiso.2012.06.016.
Baidu ScholarGoogle Scholar
[17] C.J. Sullivan, J. Stinnett,

Validation of a Bayesian-based isotope identification algorithm

. Nucl. Instrum. Meth. Phys. Res. Sect. A 784, 298-305 (2015). doi: 10.1016/j.nima.2014.11.113.
Baidu ScholarGoogle Scholar
[18] Y.G. Huo, H.Z. An, Y.L. Wu et al.,

Fast radioactive nuclide recognition method study based on pattern recognition

. Nucl. Electron. Detection Technol. 51, 53 (2014). doi: 10.1.1.637.5584.
Baidu ScholarGoogle Scholar
[19] C.J. Wang, D.M. Bao, C. Song et al.,

Investigation on the fuzzy recognition mechanism for γ-ray fingerprints of nuclear materials

. Acta Physica Sinica 57(9), 5361-5365 (2008). doi: 10.7498/aps.57.5361
Baidu ScholarGoogle Scholar
[20] E. Rohée, R. Coulon, F. Carrel et al.,

Benchmark of the non-parametric Bayesian deconvolution method implemented in the SINBAD code for X/γ rays spectra processing

. Nucl. Instrum. Meth. Physics Research Sect. A 836, 91-97 (2016). doi: 10.1016/j.nima.2016.08.025
Baidu ScholarGoogle Scholar
[21] E. Barat, T. Dautremer, T. Montagu et al.,

Nonparametric Bayesian inference in nuclear spectrometry

. 2007 IEEE Nuclear Science Symposium Conference Record., 2007, pp. 880-887. doi: 10.1109/NSSMIC.2007.4436469.
Baidu ScholarGoogle Scholar
[22] Y.M. Wang, Y.Z. Wei,

Fuzzy logic based nuclide identification for γ ray spectra

. Journal of Tsinghua University (NATURAL SCIENCE EDITION) 52, 1736-1740 (2012). (in Chinese)
Baidu ScholarGoogle Scholar
[23] A. Wald, Sequential Analysis. New York: Wiley, 1947, (Reprint Dover Publications, 1973). doi: 10.1038/2121008b0.
[24] J. V. Candy, E. Breitfeller, B.L. Guidry et al.,

Physics-based detection of radioactive contraband: a sequential bayesian approach

. IEEE T. Nucl. Sci. 56, 3694-3711 (2009). doi: 10.1109/TNS.2009.2034374
Baidu ScholarGoogle Scholar
[25] Y. Chao, L.Q. Xu, L.Q. Hu et al.,

Gaussian process tomography based on Bayesian data analysis for soft x-ray and AXUV diagnostics on EAST

. Chinese Physics B 29, 095201 (2020). doi: 10.1088/1674-1056/aba2e4
Baidu ScholarGoogle Scholar
[26] Z.H. Qian, Z.H. Ding, M.Z. Ai et al.,

Bayesian Optimization for wavefront sensing and error correction

. Chinese Physics Letters 38, 064202 (2021). doi: 10.1088/0256-307X/38/6/064202
Baidu ScholarGoogle Scholar
[27] Q.P. Xiang, D.F. Tian, J.Y. Zhu et al.,

Numerical study on the sequential Bayesian approach for radioactive materials detection

. Nucl. Instrum. Meth. Phys. Research Sect. A 697, 107-113 (2013). doi: 10.1016/j.nima.2012.09.031.
Baidu ScholarGoogle Scholar
[28] J. V. Candy,

Model-based signal processing

. J. Acoustical Society of America 119(5), 2553 (2006). doi: 10.1121/1.2184129
Baidu ScholarGoogle Scholar
[29] J. V. Candy, Bayesian Signal Processing: Classical, Modern and Particle Filtering Method. Wiley, 2009. ISBN:0470180943, 9780470180945.
[30] Q.P. Xiang, D.F. Tian, F.H. Hao et al.,

Off-line experiments on radionuclide detection based on the sequential Bayesian approach

. Nucl. Instrum. Meth. Physics Research Sect. A 729, 212-219 (2013). doi: 10.1016/j.nima.2013.07.046.
Baidu ScholarGoogle Scholar
[31] C.M. Bishop, Pattern Recognition and Machine Learning (Information Science and Statistics). Springer-Verlag New York, Inc. 2006 ISBN:0387310738.
[32] J. Ling, G.J. Liu, J.L. Li et al.,

Fault prediction method for nuclear power machinery based on Bayesian PPCA recurrent neural network model

. Nucl. Sci. Tech.  31, 75 (2020). doi: 10.1007/s41365-020-00792-9
Baidu ScholarGoogle Scholar
[33] H. Torkey, A.S. Saber, M.K. Shaat et al.,

Bayesian belief-based model for reliability improvement of the digital reactor protection system

. Nucl. Sci. Tech.  31, 101 (2020). doi: 10.1007/s41365-020-00814-6
Baidu ScholarGoogle Scholar
[34] Q.X. Zhang, J.K. Zhao, Y. Gu et al.,

Evaluation of terrestrial radiation in the Chengdu Plain using 1/250,000-scale geochemical prospecting data

Nucl. Sci. Tech. 28, 150 (2017). doi: 10.1007/s41365-017-0293-9
Baidu ScholarGoogle Scholar