Introductory and historical view of nuclear clustering
Clustering is a general phenomenon that appears at every hierarchical layer of the matter universe, including the largest star systems [1] and the smallest hadron systems [2]. In nuclear systems, clustering commonly occurs in light nuclei, at the surface of heavy nuclei (e.g., α-particle formation and decay), and in weakly-bound or excited nuclei [3-6], as illustrated in Fig.1. Clustering in many-body systems results from local few-body correlations becoming stronger than overall central interactions. Hence, clustering leads to fundamental changes in the motion regulation and main structural degree-of-freedom [7].
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F001.jpg)
Before the discovery of the neutron in 1932, the α-particle was considered to be one of the basic constituents of the atomic nucleus, as given by the then well-known α-decay phenomenon [8, 9]. Following the recognition of the proton-neutron composition in the nucleus and the initialization of the independent-particle concept for nuclear structure in the 1930s [10], the α-cluster model continued to be one of the main approaches for describing nuclear structure [10, 11]. The microscopic basis for nuclear clustering, including the full antisymmetrization among all protons or neutrons, was originally introduced through the famous resonating group method (RGM) [12], and this was followed by its variants, including the generator coordinate method (GCM) [13] and orthogonal condition model (OCM) [14]. Notably, the explicit delineation of the coordinates collecting the center of mass (c.m.) of the clusters (groups) allows for the description of the collective cluster motion controlled by the effective interaction among them [15-17]. The importance of the nuclear clustering was highlighted by the discovery of the 7.65 MeV (0+2) state in 12C, known as the Hoyle state, which exhibits a characteristic 3α cluster configuration and plays a crucial role in the synthesis of elements across carbon in stars [18-20].
Since the 1960s, studies on nuclear clustering were revived along with the occurrence of some exotic excitation spectra that were inconsistent with single-particle picture of nuclei [21]. During this time, the wave-packet representation of a nuclear state, known as the Brink wave function (w.f.), was also proposed [22]. The Brink w.f. is particularly appropriate for describing clustering phenomena, considering its favorable model space, compared with that of the orbit-presentation, which is suitable for describing independent particle motion in a centralized mean field. In 1968, the Ikeda diagram was introduced to indicating the favorable conditions for cluster formation in the vicinity of the corresponding cluster separation threshold [23, 3]. At the highest order of clustering, the Ikeda diagram gives rise to a pure α composition for α-conjugate nuclei. If all α-particles move in relative s-waves, then the Bose–Einstein condensation (BEC) can be adopted to characterize the nuclear system, as represented by the Tohsaki—Horiuchi—Schuck–Röpke (THSR) w.f. [24, 25]. The Hoyle state is an example of such a BEC-like state [26, 19, 15]. However, despite recent, remarkable advancements in theoretical predictions [27], the experimental determination of BEC-like states is still limited. The latest report on the experimental observation of Hoyle-like resonances in 16O [28] is encouraging regarding the further exploration of condensation states in heavier α-conjugate nuclei. Notably, the Ikeda diagram was extended to systems with excessive valence neutrons or protons, in addition to cluster cores, which gave rise to more abundant threshold arrangements associated with the formation of the molecular structures [3, 29, 30]. These topics will be further addressed in the following.
Over the past three decades, nuclear physics has entered into a new era of exploring the largely expanded nuclear chart away from the β-stability line [3, 4, 31]. Several theoretical approaches were further developed to investigate various aspects of nuclear clustering in unstable nuclei, including the antisymmetrized molecular dynamics (AMD) and fermion molecular dynamics (FMD) models [32-35] aswell as the container model extended from the THSR w.f. [27, 36-38]. These models employ the Brink-type wave-packet representation and start from the nucleon degree of freedom with full antisymmetrization to cover the major cluster features from the microscopic basis. Another category of models assumes cluster formation a priori and concentrates on the role of valence nucleons, which may form various molecular bonds, such as the molecular orbital (MO) model [39, 3] and the generalized two-center cluster model (GTCM) [40]. Efforts were also made to apply single-particle-type or collective-vibration-type frameworks to account for clustering phenomena [41-43]. Thus far, the experimentally identified molecular structures in unstable nuclei are still quite limited and concentrated on light nuclei, such as neutron-rich beryllium and carbon isotopes [28, 44-50], as illustrated in Fig. 1. Important progress has been made in recent years regarding linear chain molecular states in 14, 16C [49, 50]. These results imply the possible existence of extremely exotic structures, such as ring configurations in heavy, neutron-rich nuclei [51]. Another interesting phenomenon is the possible emergence of the compact two-neutron cluster (dineutron, 2n) in low-density environments, which may largely affect the structures of neutron-rich nuclei and even the inner crust layers of neutron stars [52-58]. The latest finding regarding the BEC-type 4He + 2n + 2n structure in 8He(0+2) [59, 60] is encouraging for the future exploration of the rich clustering configurations in heavy, neutron-rich nuclei (Fig.1). Recently, Tanaka et al. measured α clusters at the surfaces of neutron-rich Sn isotopes using the quasi-free (p, pα) knockout reaction and examined their isotopic dependence [61]. Their results provide direct experimental evidence for the existence of α clusters at the surfaces of heavy nuclei and thus provide a natural explanation for the source of α decay. This work also reveals an interesting interplay between the formation of α clusters and development of the neutron skin, which will further impact the constraints on one key parameter of the nuclear equation of state, namely, the density-dependence parameter of the symmetry energy [62-64].
Along with the perfection of the model and experimental findings, a number of physics concepts have been proposed and justified, including the threshold effect [23], strong nucleon correlations in a low-density environment that potentially lead to the appearance of BEC states [15, 16], wave-packet description of a nuclear system with configuration mixing [59, 65, 66], and orthogonality of the quantum states driven by the valence neutron occupancy [40, 65, 66].
We outline the major theoretical models that have been frequently applied in the literature by emphasizing their conceptual and logical connections in Sect. 2. The major experimental methods and observations are selectively described in Sect. 3. Further perspectives are given in Sect. 4.
Theoretical descriptions of nuclear clustering
Over the past six decades, the description of the cluster structure in nuclei has progressed along with the observations of various clustering phenomena. Here, we outline the major models that have been widely used in the field and are still active in many applications of current research on nuclear clustering.
Resonating group method
The RGM was proposed by Wheeler in 1937 [12]. At that time, the nucleon- and cluster-based models were competing to account for the observed properties of nuclei. Within the RGM approach, nucleons in a nucleus are configured into many groups, with the number of groups and numbers of protons and neutrons in each group being defined in a selective way. The spin of each group is constrained such that all spins must sum to the total spin of the nucleus. In addition, different configurations can be added together to present the complete w.f. [12]. The RGM is distinguished by its definition of the c.m. of each group and the relative coordinates between these c.m. positions (inter-cluster coordinates), upon which the interactions between the groups (cluster interactions) and the w.f. for the group–c.m. distribution (cluster w.f.) can be defined. The real cluster w.f. (probability distribution of the groups) can then be obtained according to the standard energy minimizing method, assuming the known Hamiltonian, which is also configured in line with the group definitions. In general, RGM focuses on addressing the cluster interactions and the cluster w.f. while treating the internal interaction and structure of each cluster (group) as a known basis. Hence, the selection of the group configurations should not be arbitrary. Instead, the selection should be performed in accordance with the real structure possibilities, as reflected by the existing observations and solid predictions. Otherwise, the results from the calculation might be different from reality due to the limited knowledge of the interactions (Hamiltonian) and unavoidable truncation of the configuration possibilities (state space). The RGM provides a flexible means by which to describe a complex many-body system that can suitably mix single-particle-type and cluster-type structures for use with nuclei that are far from the β-stability line. However, their method of use is still an unsolved challenge associated with experimental hints, theoretical background knowledge, and physical understanding [15, 59]. In addition, the full anti-symmetrization among all neutrons or protons of a system would cause large difficulties in real calculations. This constraint could partially be relaxed via the selection of a simple group configuration and the application of numerical methods, such as the generator coordinate method.
Here, we use a simple example to illustrate the RGM application [67, 15]. A system with A nucleons can be divided into two clusters with A1 and A2 nucleons. Then, the corresponding RGM w.f. can be expressed as:
Generator coordinate method
The GCM was originally proposed by Griffin and Wheeler to address the collective motion of the nuclear system [13]. This method aims to address overall coordinates, such as the deformation parameter of a nucleus, at several discrete values to generate the basis of the structure w.f.. These basis w.f. are then weighted and summed (integrated) to represent the structure w.f. for the system. The unknown weight factors, as a function of the generator coordinate, are determined according to the variational principle under the operation of the appropriate Hamiltonian. Hence, the complex calculations using the continuous variations of the dynamic coordinates can be simplified and extended to relatively heavier systems.
The GCM has naturally been combined with the RGM by using the inter-cluster coordinates as the generator coordinates [71, 67]. In the case of 9-11Be nuclei, the analysis using GCM reveals a decrease of the α-cluster contents from 9Be to 11Be. In addition, the results of this analysis predict a strong mixing of the 6He + 6He and 4He + 8He configurations in 12Be nuclei. Recently, the combination of RGM and GCM has become the fashion for studying the structural configurations of nuclei far from the β -stability line. An exemplar is presented in Ref. [59], which demonstrates a case of 2n clustering8. The w.f. of the 0+ states (8He) is given in terms of the RGM + GCM:
This overlap is equivalent to the projection of the eigenstate w.f. onto the RGM clustering basis. The BEC character of the 8He(0+2) state obtained from this RGM + GCM calculation is displayed in Fig. 2. This work demonstrates the flexibility and powerful ability of using the RGM combined with the GCM.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F002.jpg)
Notably, when extending the inter-cluster coordinate to a large range, the GCM can be applied to scattering problems. The first treatment of the scattering problem using GCM was reported in 1970 by Horiuchi [71]. Since then, significant progress has been achieved in this direction [15, 72-75].
Orthogonality condition method
The calculations in the RGM + GCM approach can be simplified further using the orthogonality condition method (OCM), which was initially proposed by Saito in 1968 [14]. The idea is to exclude the states for the relative motion of the clusters that are occupied by the internal structure of the clusters themselves, namely
OCM was proven to be highly effective in describing the complex nuclear cluster structure and scattering. In the 1970s, Matsuse and Kamimura conducted quantitative computations using OCM and indicated that the first
Molecular orbit model and generalized two-center cluster model
The molecular viewpoint of nuclear structure was introduced by Wheeler in 1937 [82], the same year the release of the RGM [12]. This is natural, as the interactions between groups (clusters) are quite similar to molecular bonds in atomic physics. The molecular picture supports the excessive binding energy of the α-conjugate nuclei, which is roughly proportional to the number of bonds between α-particles [11]. This concept was further developed and applied to light nuclei to account for the energy-level spectra, which could be classified into the molecular rotational bands together with the proposal of the “threshold rule (Ikeda diagram)” [23, 21, 83, 84].
The MO model was proposed for nuclei with excessive nucleons in addition to cluster cores, particularly unstable neutron-rich or proton-rich nuclei [85, 29, 30, 39, 3, 86]. The basic idea is to exploit the multicenter potential, corresponding to the cluster cores, within which the valence nucleons move around one core or multiple cores, similar to the valence electrons playing roles in ionic or covalent bonding between atoms, respectively. The MO states may occur at energies below the cluster-separation threshold and create bound (lower-lying) molecular structures [3, 86]. Here, we illustrate a simple case of a two-center system, such as that for 9,10Be. The molecular w.f. can be formed by the linear combination of nuclear orbitals (LCNO):
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F003.jpg)
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F004.jpg)
Okabe and Abe reproduced the properties of the low-lying levels of 9Be [29, 30] using the MO model. Itagaki et al. predicted the exotic behaviors in excited states of 16C in which two neutrons occupy a σ orbit while another two neutrons occupy a π orbit, thus offering new insights into neutron-rich carbon isotopes [87].
Since the MO model is limited in its descriptions of the low-lying states of light neutron-rich nuclei, an extended molecular-type model, namely, the GTCM, was introduced by Ito et al. [40, 88, 89]. This method combines the covalent configuration, similar to the LCNO, with the ionic configuration and solves the equation of state as a function of the distance between the two centers (which is used as a generator coordinate) in a unified way. The model can then be applied to predict highly excited resonant states in which the two correlated centers may become two quasi-independent clusters, with each valence particle being combined into one as an ionic system. These resonant states then undergo decays, according to the cluster partial widths. Taking 12Be (α+α+4n) as an example, the basis function within the GTCM approach can be expressed in the form [40, 89]:
The total w.f. is obtained by taking a superposition over S, 𝒎, and K as:
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F005.jpg)
Antisymmetrized molecular dynamics and fermion molecular dynamics
In the 1960s, the Brink w.f. was proposed to describe the nuclear structure using the wave-packet concept rather than the orbit configurations [22]. Basically, each nucleon or a group of nucleons is presented by a Gaussian function centered at a position R. Many Rs values can become the generator coordinates, and various wave-packet configurations can be combined and fixed, according to the weights determined by the GCM equation. The wave-packet basis is in favor of describing the clustering system within a limited state space, whereas the orbit configurations are better suited to account for the single-particle behaviors in a centralized mean field. Brink-type w.f.s have since been accepted and extended by many cluster-oriented models, such as AMD and THSR [32, 25].
In the AMD model, each nucleon is described by a Gaussian wave packet, which can be regarded as the extreme of the Brink w.f.. The model originated from the quantum molecular dynamics (QMD) approach used to simulate the heavy-ion reactions, but was implemented by applying the full antisymmetrization among all nucleons to account for the real nuclear structure [91, 32, 33]. FMD is similar to AMD, but FMD incorporates a variable width parameter for the Gaussian-type wave packet [92]. Both AMD and FMD consistently suggest that the second 0+ state of 12C (the Hoyle state) has a prominent gas-like 3α configuration with an extended size [92-94].
The AMD w.f. for a nucleus is expressed as a Slater determinant composed of single-nucleon Gaussian wave functions:
The Hamiltonian of the AMD can be given by:
Over the past thirty year, AMD has been developed to provide not only the energy eigenvalues and nuclear radii but also other observables, such as the cluster decay widths, magnetic dipole and electric quadrupole moments, and electromagnetic transition probabilities. Recently, AMD has successfully reproduced experimental properties of various nuclei, including Be, C, O, and Ne isotopes [33], and it has become a powerful tool for investigating the cluster and molecular structures in atomic nuclei [17]. Here, we illustrate typical examples that are closely related to the latest experimental findings.
In 1998, Kanada-En’yo used the AMD model to reproduce all energy levels below 15 MeV in 12C together with the related E2 transition rates and β-decay strengths [97]. Subsequently, the AMD model was extended to investigate the excited states of 10Be using the variation after projection technique [98]. Remarkably, without assuming preformed 2α clusters, AMD calculations unveiled a notable 2α + 2n cluster structure in many intrinsic states of 10Be [98]. To extract the behaviors of the valence nucleons, the single-particle Hamiltonian under the AMD framework can be defined in the following form [99]:
Later, Baba et al. extended the AMD calculation to 16C, 14C, and 14O nuclei [65, 66, 100, 101]. They identified 3-cluster structures with various valence-neutron configurations, as shown in Fig. 6. Panels (a) and (b) show the ground-state rotational bands of 16C [66], which exhibit approximately spherical structures. Panels (c) and (d) display asymmetric triangular rotational band structures, with
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F006.jpg)
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F007.jpg)
Various exotic clustering configurations have been predicted [65, 100, 99] and observed [102, 103] for 14C, including triangular, π-bond, and σ-bond linear chain structures. Baba et al. analyzed the associated decay patterns. Their results indicated that the σ-bond linear chain structure primarily decays into the ~6 MeV states of 10Be, whereas other resonant states mainly decay into the ground and first excited states (3.4 MeV, 2+), as illustrated in Fig. 8. These selective decay paths provide an important means by which to identify the existence of the σ-bond linear chain configuration, which was effectively used in later experiments conducted by Li et al. [48] and Han et al. [50].
A theoretical comparison was made between the mirror nuclei 14C and 14O, regarding their molecular structures [100]. Fig. 9 shows the excitation energies of the 0+ states in 14C and 14O, corresponding to the ground, triangle, and linear chain configurations. The shift between a pair of states, mostly caused by Coulomb interactions, also known as the Thomas–Ehrman Shift (TES) [104-106], is related to the expansion (in terms of size and deformation) of the corresponding systems. The measurement of the TES for 14C and 14O, in comparison to the calculated values, would provide new evidence for the existence of the intriguing linear chain structure in exotic nuclei.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F009.jpg)
THSR wave function and container model
According to the Ikeda diagram [23], the α conjugate nuclei may be transformed into a multi-α clustering system at the vicinity of the corresponding separation energy. Owing to the small relative energy (decay energy) in this situation, the relative motion of these α-clusters could all be in s-wave. These results have stimulated the idea of α-condensation in nuclei, as proposed by Röpke et al. in 1998 [24]. Subsequently, the THSR w.f. was developed to investigate the 3α condensation in 12C [25]. This w.f. can be regarded as a variant of the Brink w.f.. Whereas the Brink w.f. allows the subsystems (wave packets) to remain at random positions inside the nucleus, the THSR w.f. requires all subsystems (each with a spin of 0) to move in s-wave around the center of mass of the entire system [25], similar to the case of BEC [25, 26, 15, 17].
As an example, the THSR w.f. for a 3α system can be expressed as:
The THSR w.f. has been widely adopted as a standard approach to evaluate the intrinsic components of BEC-like structures in any quantum state. For instance, the Hoyle state of 12C obtained via the general RGM/GCM possesses an almost 100% overlap with an extended single THSR w.f. [108, 107, 15]
A container model based on the THSR w.f. was recently proposed [15, 38]. Unlike traditional cluster models, such as the Brink wave packet, the container model focuses on the dynamic evolution of containers to capture cluster motion. The container size evolves from smaller to larger values as the excitation energies increase. The cluster motion inside the mean-field potential is characterized by using the size parameter under the condition of full antisymmetrization among all nucleons. This approach has been studied and documented by Funaki et al. [15, 38]. The w.f., which is generally nonlocalized, can result in localized density distributions for clusters due to the Pauli blocking effect [37, 15]. With its relatively simple picture and parameter settings, the container model offers a relatively simple conceptual framework and parameter settings, thereby showing promise as a powerful alternative to traditional methods, such as the RGM and Brink-GCM, for describing cluster dynamics [15, 16]. As an example, Fig. 10 schematically illustrates the α + 8Be cluster structure in 12C and the α + 12C cluster structure in 16O under the “container” framework. In this model, 2α or 3α clusters are confined in a “container” characterized by the parameter β1, while the remaining α cluster is confined in a larger “container” characterized by β2 (where β2 is related to the parameter B, as defined in Eq. 18.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F010.jpg)
In 2023, Zhou et al. reported the microscopic five-body calculations for 20Ne [27], thus revealing a 0+ state located approximately 3 MeV above the 5α threshold, which could correspond to the state observed in recent experiments [111]. This state exhibited the
α-cluster formation and decay from heavy nuclei
α clustering in heavy nuclei is essentially connected to α decay, which is generally described as the quantum tunneling of the preformed α particles via a semi-classical model initiated by George Gamow in the 1920s [112, 113, 9]. Various theoretical approaches have been developed to predict the half-life of α decay [114-125]. However, the preformation probability of the α-cluster within a heavy nucleus still cannot be reproduced from the shell-type calculations [126]. These results are most likely attributable to the difficulties associated with accounting for the four-nucleon correlation in a complex multi-nucleon environment. Generally, this probability can only be obtained empirically by using the experimental decay probability and subtracting the calculated barrier penetrating factor [119, 127-130].
Recently, a novel microscopic approach for describing the α formation was proposed by Röpke et al. [131, 132] and developed by Xu et al. [124, 133]. The model aimed to create the quartetting w.f. (QWFA) for the four valence nucleons upon the remaining core of the nucleus. The introduction of the c.m. coordinates (new degree-of-freedom) for the quartet system (RGM–Brink-type) should have captured the main feature of the cluster-formation dynamics [121, 124, 131-134]. In QWFA, the w.f. of four valence nucleons is divided into two components [133]: the c.m. motion relative to the core and the intrinsic motion of the four nucleons relative to their c.m. [124]:
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F011.jpg)
Two different regions for R<Rsep and R>Rsep are considered to describe the α-decay process [135, 136, 121]. When R < Rsep, the intrinsic w.f. of the quartet represents four independent nucleons in quasi-particle states within the shell model framework. When R > Rsep, the nature of the intrinsic w.f. undergoes a transformation, signifying the emergence of a bound state that exhibits characteristics similar to those of an α-particle [124, 119]. Considering that the α-like state can only exist at densities lower than the critical value ρc(Rc)=0.02917 fm-3 and dissolves at higher densities, the α-cluster preformation probability Pα can be calculated by integrating the bound-state w.f. Φ(r) from the critical radius Rc to infinity [121]:
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F012.jpg)
Experimental studies of nuclear clustering
Over the past thirty years, experimental studies of nuclear clustering have largely progressed along with the availability of radioactive ion beams(RIBs) for light, neutron-rich nuclei [49, 4, 31, 50, 60]. The experimental methods and techniques have also advanced to meet the requirements for clearly identifying nuclear clustering phenomena under the strong influence (mixing) of the shell-like effects. The experimental measurements generally started with the inclusive missing mass (MM) method in the early stages and gradually evolved into more and more exclusive methods incorporating various coincident detection techniques. The cluster decay and cluster knockout measurements are of special importance in the selective probing of the cluster structures from high-density single-particle states. These measurements, particularly those in inverse kinematics with the RIBs, normally require multi-particle detection in a small solid angle. Hence, the state-of-the-art position-sensitive (pixelated) detection systems have also advanced over the past two decades [28, 48-50].
Missing mass method and cluster production cross-sections
Experimental studies of cluster structure in light nuclei often start with inclusive measurements that typically require simple detection systems. The method relies on the assumption of a two-body system in both the initial and final channel of the reaction, i.e., A(a, b)B or a + A→ b + B. With the properties of the beam (a) and target (A) particles being known and by measuring the energy and emission angle (position) of the projectile nucleus (b), the effective mass of the assumed recoil nucleus, denoted by mB*, can be deduced according to the conservation of energy and momentum in the reaction process. This process is equivalent to the deduction of the effective reaction Q-value described in the textbook [137]. The difference between this deduced effective mass (mB*) and its g.s. mass (mB) is called the MM of the reaction, or the excitation energy of the nucleus B. Typical examples for the production of the excited states in 14C can be found in Refs. [138, 3]. In these works, various reaction mechanisms are employed to populate the excited states in 14C, including the two-neutron-transfer reactions 12C(14N, 12N), 12C(15N, 13N), and 12C(16O, 14O), two-proton-transfer reaction 16O(15N, 17F), and multi-nucleon-transfer reaction 9Be(7Li, d). The two-nucleon-transfer reactions are mostly in favor of populating the shell-like states, whereas the multi-nucleon transfer reaction can strongly populate the cluster-type states, owing to the associated multi-particle, multi-hole configurations. Different reaction Q-values, corresponding to different momentum mismatches, in the direct reactions may introduce different momentum transfers into the final particles which, in turn, may be placed into different spin-value ranges. Based on these carefully designed reaction tools and combined with theoretical model calculations, the shell-like states can be classified into various deformation schemes. The remaining states, particularly those that are significantly populated only in the multi-nucleon transfer reaction, could be grouped into cluster-based molecular rotational bands. Fig. 13 shows the states in 14C populated from the multi-nucleon-transfer reaction 9Be(7Li, d). The cluster-like states are grouped into two molecular rotational bands, corresponding to the parity-inversion-doublets [138, 3]. Notably, this double-band configuration at high excitation energies resulting from the reflection-asymmetric configuration, such as 10Be + α in 14C, constitutes strong evidence for the molecular (cluster) structure formation in atomic nuclei [21]. The energy gap between the two bands can be estimated according to the asymmetric structure configuration. Similar exemplars for the study of cluster structures in nuclei via inclusive MM measurement can also be found in works such as Ref. [139] for 20O (See Fig. 14).
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F013.jpg)
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F014.jpg)
The inclusive measurements may also incorporate the experimental determination of the cluster production cross-section in various reaction processes. An early typical study of this kind was reported in Ref. [140], where the He-cluster breakup and neutron removal cross-sections for 10-12, 14Be were reported. It is interesting to see the persistence of the He-cluster structure within the neutron-rich Be isotopes up to the dripline nucleus 14Be, as was demonstrated by comparing the measured He-cluster breakup cross-section with the phase-space estimation. In an inverse kinematics experiment, using A and a to denote the projectile and the target, respectively, the aforementioned MM measurement of the recoil nucleus b can be combined with the selected cluster production related to the breakup of the nucleus B in order to probe the cluster production cross-section as a function of the excitation energy. A representative example can be found in Ref. [141], in which the MM measurement of the triton particle from the 19F(3He, t)19Ne reaction was made in coincidence with the proton or α decayed from the possible resonant states of 19Ne. As such, the α-cluster production is observed as a function of the 19Ne-excitation energies. This provides crucial information about the cluster-formation probability along the excitation axis, comprising both direct breakup and resonance-decay two-step processes, as depicted in Fig. 15.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F015.jpg)
Invariant mass method
The invariant mass (IM) method is dedicated to the resonant states, often at higher excitation energies that decay into fragments. As described above, the MM measurement is inclusive and sensitive to all kinds of states characterized by such as single-particle motion, collective motion, and cluster structures. Particularly, at high excitation energies the spectrum is generally dominated by a continuum associated with a high-level density. To improve the sensitivity to the clustering states and avoid as much of the overwhelmingly strong continuum background as possible, it is essential to measure the cluster-decay process [49]. The latter can be used to reconstruct the excitation energy of the reaction-produced cluster-like mother nucleus via the IM method, to determine the spins of the resonances based on the model-independent angular correlation analysis, extract the spectroscopic factor of the cluster structure in an observed resonant state, deduce the strength of a typical cluster excitation mode (e.g. the monopole-transition strength), and connect the unknown structures of the mother nucleus to the known structures of the detected daughter fragments via selective decay paths [46, 142, 49, 50]. Here, we highlight some useful quantities for the measurements and data analyses.
Kinematics
The IM method lies in a two-step process, namely, the resonance excitation followed by the cluster decay, as expressed by a(A, B*→c+d)b in inverse kinematics and depicted in Fig. 16. Using relativistic expressions, the sum of the squares of the four-dimensional momentum (Pμ) is a Lorentz invariant quantity in different inertial reference frames. For a resonance-decay process, the IM method can be expressed as:
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F016.jpg)
The relative energy of the decay system, such as that shown in Fig. 16, is defined by:
Similar to the description in the MM method, the Q-value of the reaction–decay process, as illustrated in Fig. 16 and corresponding to the mass deficit of the initial two-body channel and the final three-body channel, can also be deduced from the measured energy–momentum of the particles involved. If such a measured Q-value is smaller than the g.s. Q-value, then excitation can be expected in some of the final nuclei.
Useful observations
In the analysis of reaction–decay processes, as depicted in Fig. 16, some quantities (observables) are useful and often employed to clarify the true process or structural effects, such as the energy–momentum (E–P) plot, monopole transition strength, angular correlation analysis, cluster spectroscopic factor. A brief outline of these quantities is given below.
E–P plot for target component selection
In a real experiment, some composite materials may have to be used for targets, such as (CH2)n and (CD2)n. In such cases, the accurate differentiation of the recoil target nucleus is crucial to obtain the clear Q-value spectrum and the excitation-energy spectra [143, 144]. The E–P plot is an effective method for selecting the target component based on the measured quantities [145, 50]. Defining x and y from the measured quantities:
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F017.jpg)
Monopole transition strength
As described in Section 2.D for GTCM, the enhanced monopole-transition strength between two nearby 0+ states is an imprint of the cluster formation [90]. Experimentally, the extraction of the monopole-transition matrix element from either the electromagnetic process [M(E0)] or nuclear process [M(IS,0)] is an important aspect of a cluster structure study. These matrix elements are defined in the following forms [147, 148]:
In 2014, an experimental method by which to extract the monopole transition strength using inelastic scattering data was established [46]. Fig. 18 shows the experimental differential cross-sections for the 10.3 MeV (0+3) state of 12Be. The cross-sections were reproduced by the DWBA calculations with a normalization factor of a0 = 0.034(10), which is associated with the EWSR fraction (6727.9 fm4MeV2). The deduced monopole transition matrix element (M(IS)) is 7.0±1.0 fm2, which is comparable to those of typical cluster states [154, 155, 90].
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F018.jpg)
Angular correlation analysis
Spin is always an important quantity in nuclear structure characterization. In the case of a cluster or molecular structure, spin-parity determination is a precondition to addressing the excitation and decay modes as well as allocating the members in a rotational band. In early inclusive MM experiments, the spin of a produced recoil nucleus associated with transferred orbital angular momentum [137] is typically determined by measuring the angular distribution of the probe. However, this approach is valid only for direct reactions (scattering) to low lying states, and it is questionable when applied to highly excited resonant states. Fortunately, the resonance-decay provides another independent and effective means by which to determine the spin of the mother nucleus, regardless of its production mechanisms [44, 46, 49, 102, 103, 142, 153, 156-158].
By assuming a two-step process for the resonance production and decay, as depicted in Fig. 16 for the resonant nucleus B*, the conservation of the total angular momentum adheres to:
According to the relation between the primary reaction and sequential decay processes, angular correlations can be divided into in-plane and out-of-plane correlations. The correlation angles are defined in Fig. 1 of Ref. [159]. In-plane correlation refers to the alignment between the decay plane and the reaction plane, where the azimuthal angle χ is constrained around 0 or 180. The double-differential cross-section only depends on the scattering angle θ* and decay angle ψ. In some experiments [142, 157-166], owing to the geometric arrangement of the detectors, both the reaction and decay planes closely align with the detector plane, thereby satisfying the conditions for in-plane correlation. When the spins of the initial and final-state particles are zero, the events of in-plane correlation exhibit a linear “ridge” structure in the θ*–ψ plot, which is commonly employed in spin analyses [158-160, 162-164]. This linear “ridge” structure is described by the following function:
The reason for the appearance of the linear “ridge” structure in the in-plane correlation is that when the spins of all the initial and final particles are zero, the spin JB* of B* can be determined by the change of orbital angular momentum (JB*=li-lf). Moreover, there is a strong coupling between JB* and the final-state orbital angular momentum lf in their orientations (mB*=-mf). Rae and Marsh also provided rigorous derivations of this phenomenon from a quantum mechanical perspective [160, 167]. Hence, substituting Eq. 33 for the transition amplitude of compound nucleus formation into Eq. 32 for the differential cross-section:
In the case of out-of-plane correlation, the angular difference between the decay plane and the reaction plane is significant, indicating a large deviation of the χ angle from 0 or 180. For instance, a wider range of θ* and χ angles could be covered by employing detectors with larger sensitive areas or placing them closer to the target [159]. Another example comes from the use of a zero-degree detector that covers θ*~0° as well as the entire ranges of θ* and χ angles [49, 46, 168, 153]. In the case of spin-0 for all initial and final particles, the “ridge” structure on the θ*–ψ plot deviates from linearity [159], making it unsuitable for spin analysis. To enhance the structure in cases of out-of-plane correlation, additional constraints (projections) are typically needed. There exist two experimental cases that may greatly simplify the analysis of the angular correlation [159]. The first involves the zero-degree measurements (θ*~0°), where the motion of the compound nucleus B* and relative motion between B* and b are almost aligned along the z-axis, resulting in mf = 0 and mB*=m0=0 (where ψ is not necessarily equal to 0). Therefore, the differential cross-section distribution follows the square of a Legendre polynomial and is solely correlated with ψ in the form:
Cluster spectroscopic factor
In a nuclear state, the cluster structure might be mixed with other configurations, such as shell-like contents. Therefore, it is important to quantify the probability of having the cluster structure in an energy eigenstate. This has been realized in the literature by using the cluster spectroscopic factor within the R–matrix framework [169]. First the cluster-decay partial width, which is the product of the spectrum-extracted (BW form) total width of the resonance and the fraction of the IM-reconstructed cluster cross-section over the total MM cross section, should be determined experimentally. Within the single-channel R–matrix approach, this measured partial cluster width, denoted by Γα(E), can be related to the reduced width
Experimental investigations of cluster structures
Over the past three decades, there have been a growing number of experiments studying nuclear cluster structures by utilizing the IM method. The detection systems were mostly composed of position-sensitive silicon detectors with advanced data acquisition systems [172-174] to measure multiple particles in coincidence and with high resolution. These experiments encompass a wide range of reaction types, including the inelastic scattering, transfer reaction, and resonant scattering. The important criteria for identifying the cluster (molecular) structure in a nucleus include the extremely large moment of inertia deduced according to the energy-spin systematics for a rotational band, relatively large cluster decay width or cluster spectroscopic factor, enhanced characteristic (e.g., isoscalar monopole) transition strength, and selective decay paths. All these criteria depend on the cluster decay measurement and spin determination. Because a single experiment may not be able to obtain all these quantities, complementary works are generally required. We outline a few selected experimental works here to demonstrate the major progress made regarding IM measurements.
Studies on cluster structure in beryllium isotopes
The neutron-rich Be isotopes possess an inherent two-center feature, considering the unambiguous 2α structure of 8Be. The clustering configurations and the molecular rotational bands in 10Be have been intensively been studied since the 1990s [3, 44, 45]. Here, we focus on 12Be to demonstrate the recent experimental progress.
In 1999, Freer et al. conducted an inelastic breakup experiment for 12Be at 31.5 MeV/u on a (CH2)n target [175]. They utilized silicon telescopes to measure helium fragments and reconstruct 12Be resonant states via the 4He + 8He and 6He + 6He decay channels. They successfully observed a few cluster states above 13.2 MeV and with spins J exceeding 2. Fig. 19 (c) and (d) display the IM spectra for these decay channels. Subsequently, several experiments employed the IM method to further investigate the highly excited cluster states in 12Be. However, despite all efforts, the identification of the band head state, which should be closer to the cluster separation threshold, remained elusive throughout the following decade [176].
Significant progress was made by Yang et al. in 2014 through a novel inelastic scattering experiment of 12Be at 29 MeV/u on a carbon target [46, 153]. A major difference compared with the previous experiment was the use of a zero-degree telescope composed of high-performance silicon-strip detectors that substantially improved detection efficiency at energies close to the threshold, as displayed in Fig. 19(b) (dotted line). This experiment enabled identification of the band head (10.3 MeV) of the 4He + 8He molecular rotational band. Furthermore, the spin-parity of this state was firmly assigned as 0+, based on the angular correlation analysis, associated isoscalar monopole transition strength, and extracted cluster spectroscopic factor. All these criteria consistently support the dominant 4He + 8He cluster structure in this prominent and near-threshold resonance [46, 153, 168].
Studies on cluster structures in carbon isotopes
Owing to the possibility of 3α cores, more abundant cluster structures can be expected for carbon isotopes, including the triangle and linear chain configurations [3, 41, 65, 66, 99, 177-179]. Compared with prior work on beryllium isotopes with 2α cores, experimental investigations of carbon systems are more challenging. Here, the decay fragments, such as 10,12Be, may possess bound excited states. Hence, the conversion from the measured relative energy spectrum to the physically meaningful excitation energy spectrum necessarily depends on the specific states of the decaying fragments, corresponding to different Q-values (or total energies in the final channel) [48-50]. A series of breakup reaction experiments (Fig. 16) that were dedicated to the simultaneous reconstruction of the relative energy and reaction Q-value have been performed to search for the exotic cluster structures in carbon isotopes [48, 158, 180-185]. These measurements generated consistent resonant energies but little information on the spin-parities, primarily due to the limited resolutions or statistics [183, 186]. Traditional resonant scattering experiments using the thick-target (4He gas) technique played an important role in the study of Be isotopes but were not able to measure the reaction Q-values and hence may have overlapped the excitation-energy spectra for different threshold energies [102, 103, 47]. Therefore, the high resolution Q-value spectrum is crucial when studying the cluster structure in carbon or heavier isotopes, especially when determining the correct application of the decay path analysis [49, 156].
The linear chain structure was initially proposed for the 0+2 state of 12C (the Hoyle state) [51]. Previous studies have revealed the instability of the chain structure in 12C [177, 99]. Hence, it is crucial to search for linear chain molecular bands in 14, 16C. One major step forward was made by Liu et al. in their work on the inelastic scattering of 16C at 23.5 MeV/nucleon on (CD2)n targets [49, 156]. As illustrated in Fig. 16, in previous IM experiments, it is typical to measure only two of the three final particles (particles b, c, and d) [184, 157]. The energy of the remaining particle is deduced according to the conservation of momentum by considering the beam energy. Because the PF-type RIB used to have a large energy spread, the obtained Q-value spectrum did not have a sufficient energy resolution [184]. In novel new inelastic scattering experiment for 16C conducted by Liu et al. [49, 156], all three final particles were coincidentally detected using high-resolution silicon detectors. Hence, the beam energy could be deduced event-by-event from the final particles, based on the momentum conservation. This way, the obtained Q-value spectrum resolution does not depend on the RIB quality and instead depends only on the detector performance. In addition, a state-of-the-art analysis method was developed to recover events with two nearby fragments hitting adjacent strips of the detector. The zero-degree telescope composed of multilayer silicon-strip detectors played an essential role in this experimental use of RIB in inverse kinematics. The high resolution Q-value spectra have enabled the clear selection of the decay paths from the 16C resonances to various states of the final fragment 10Be (+ 6He) or 12Be (+ 4He). Fig. 20 shows the observed relative decay strengths for different decay paths. The excellent agreement between the observed and predicted characteristic decay patterns gives, for the first time, strong evidence for the existence of the linear chain configuration in 16C. In addition, the experimental determination of the spin-0 for the 16.5 MeV band head provides further support for the chain-type molecular rotational band. These assigned band members correspond to the
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F020.jpg)
Subsequently, a similar inelastic scattering experiment was also conducted for 14C at 23.0 MeV/nucleon on (CH2)n targets [50]. Proton targets were chosen here, at the expense of lower c.m. energy, to provide a better energy resolution and avoid the target breakup effect as significantly occur for deuteron target. The complex experimental setup, centered around the zero-degree telescope composed of multilayer-silicon-strip detectors, is schematically shown in Fig. 21. Again, three-fold coincidence measurements were realized for all final particles (4He and 10Be decay fragments plus the recoil proton), providing an almost background-free and high-resolution Q-value spectrum as well as well-established resonant states in 14C. One state at 13.9 MeV was firmly identified for the first time. The contribution of this work is its determination the spin-parities of the 13.9 MeV, 14.9 MeV, and 17.3 MeV states based on the clear angular correlation analyses and differential cross-section analyses (Fig. 22). These states agree excellently with the members of the predicted linear chain molecular rotational band with the π2-bond configuration.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F021.jpg)
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F022.jpg)
There are also several experimental works devoted to the study of cluster structures for other carbon isotopes, such as 11C [187], 13C [79], and its mirror nucleus 13N [188]. Abundant cluster (molecular) configurations have also been unveiled in these nuclei.
Studies on BEC-like structures
As indicated in section II.F, the concept of BEC-like states in nuclei was initially raised and accepted in the literature for the
It is natural to anticipate more BEC-analog structures in other nuclei, such as 16O, 20Ne, and 24Mg. However, experimental studies on these heavier systems face more challenges due to the requirement of simultaneous detection and identification of more bosonic particles in the final state. Recently, the 0+6 state at 15.1 MeV in 16O, located immediately above the 4α separation threshold at 14.4 MeV, has attracted significant attention, owing to its predicted BEC-analog structure [38, 192, 193]. Previous experiments [194-196] relied on only partial α-PID, resulting in high background levels in the Ex spectra. This background contamination has hindered the possible 4α resonant states [194-197].
In a recent inelastic excitation decay experiment for 16O [28], four α-particles in the final channel were fully detected and clearly identified. By selecting events with 3α forming the Hoyle state of 12C and the remaining one as a valence particle, the reconstructed IM spectrum of 16O exhibits four narrow resonances located immediately above the α +
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F023.jpg)
Some pioneering experiments have also been performed to probe multi-α, BEC-like states in heavier systems [199], including 20Ne [200, 111], 24Mg [201-203], 28Si [204, 205]. The measurement of α-cluster formation and decay from heavy nuclei have also been advanced in recent years [206-209].
A new form of BEC in the nuclear system is related to the 2n cluster in low-density environments. After the discovery of the two-neutron (2 n) halo in the mid-1980s, the concept of neutron halo was proposed, based on the idea of valence-neutron clustering in neutron-rich nuclei, such as 11Li [52, 210]. It has been gradually realized that 2n-formation can be enhanced at the surface of neutron-rich nuclei or in the inner crust layers of neutron stars [53-58]. Moreover, 2n is defined as a spatially compact entity with a size of 2~5 fm, comprising a spin-singlet neutron pair in an internal s-orbital that behaves like a boson. Therefore, in very neutron-rich nuclei, numerous excessive neutrons at the surface of the nucleus might form a 2n-BEC-analog state. Experimentally, it is interesting to investigate the 8He system, which might be the lightest one composed of multi-2n in the BEC configuration [59]. Recently, an inelastic scattering experiment was performed using an intense 8He beam at 82.3 MeV/nucleon [211]. The decay products 6He+2n were detected with high precision (Fig. 24). One important contribution of this work is the clarification of the excitation-decay mechanism via the application of a novel angular correlation analysis method that enabled the significant reduction of background events resulting from direct breakup processes [60]. The emergence of a prominent peak at a relative energy of 4.54 (6) MeV was then observed with a high significance level, as displayed in Fig. 25 (a). Combining these results with the spin determination, monopole-transition analysis, and 2n-correlation analysis, this resonant state can be classified as the
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F024.jpg)
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F025.jpg)
Resonant scattering and thick target method
The resonant scattering method can be regarded as a variant of the IM method. It typically utilizes a thick gas target in which the beam particle loses energy as a function of the track depth. When the center-of-mass energy of the beam–target system coincides with the corresponding resonant energy of the compound nucleus, the reaction cross-section may suddenly increase. Then, the unstable compound nucleus decays back to the projectile and target nuclei. To satisfy the conditions of the inverse kinematics, the beam particle should be heavier than the target. This allows the light fragments to penetrate a sufficient depth into the gas to reach the particle detectors [19, 202]. The excitation energy of the compound nucleus can then be deduced by measuring the energy of the fragments. In these experiments, silicon detectors are commonly placed in a gas-filled target chamber. Fig. 26 illustrates a typical setup for a resonant scattering experiment.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F026.jpg)
Using resonant scattering setups, several experiments have been conducted to investigate nuclear cluster structures, including the
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F027.jpg)
The limited resolution of the reaction position within the thick target and interference from inelastic scattering can introduce large uncertainties in the determination of the resonant states. The situation may become more serious when the final fragments have bound excited states, as the determination of the reaction position might be mixed with the determination of the excited states of the final particle [49, 50]. This problem can be resolved by employing the active target time projection chamber (AT-TPC) which may independently recognize the reaction position without relying on particle energy detection [214-216]. These results are due to the tracking capability of the TPC in addition to the energy measurements. Fig. 28 shows a schematic for a typical AT-TPC setup.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F028.jpg)
The resonant scattering experiment conducted by Fritsch et al. in 2016 on the 10Be + 4He system [103] is a notable example. The trajectories of the initial and final state particles were recorded in an AT-TPC, as depicted in Fig. 29(a). The angular correlation between the two final particles can be used to distinguish the ground and excited 2+ states of the 10Be fragment, as shown in Fig. 29(b). Of course, it is still difficult to see the decay to the ~6 MeV states in 10Be and, thus, insufficient to sort out the decay paths related to various molecular configurations, as demonstrated by experimental works using advanced silicon-strip detectors [49, 50]. The choice of the active gas as both the target material and detection medium presents another limitation.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F029.jpg)
New detection techniques have been introduced by combining the precise tracking TPC (MicroMegas type) with surrounding Si-CsI(Tl) telescope detectors, as shown in Fig. 30 [217]. This combined detector system enhances the capabilities of resonant scattering experiments by providing an improved tracking accuracy, a better energy resolution, and an excellent PID capability. A series of works have reported the use of this equipment, including the level scheme in 9C [218], α clustering in 18Ne [219], and various cluster configurations in 13N [188].
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F030.jpg)
Despite the numerous advantages of using AT-TPC in resonant scattering experiments, such as its high efficiency, tracking capability, and low detection threshold, the choice of the working gas (typically 4He) remains a significant limitation. One solution might be to incorporate a thin solid target inside the TPC, as has already been realized in previous fission measurements [220], as depicted in Fig. 31. Here, the TPC plays the role of a detector, not the active target. The primary advantages would come from having a lower detection threshold and larger detection solid angle, compared with those of silicon detectors. These factors are crucial, especially for low-energy experiments.
Using knockout reactions to investigate cluster composition in nuclear ground states
According to gRDF calculations, α clusters can be formed at the low-density surface regions of heavy nuclei, but, in this case, there is a close interplay between the α-cluster formation and development of the neutron skin[62]. This interplay leads to a reduction in the neutron-skin thickness, compared with the theoretical predictions, without considering the existence of α clusters, which will further impact the experimental constraint on the density-dependence parameter L of the symmetry energy. However, the formation of α clusters is gradually hindered as the neutron skin becomes thicker. This hindrance manifests itself as a monotonic decrease in the strength of the α-cluster formation along the isotopic chain and thus can be directly examined experimentally. Hence, it is crucial to carry out new experiments to find direct experimental evidence for the existence of preformed α particles (clusters) at the surfaces of heavy nuclei and further examine the systematics of α-clustering strength along an isotopic chain into a neutron-rich region accompanied by a steady increase of the neutron skin.
When unbound, α clusters can spontaneously come out of the mother nucleus to take advantage of the quantum tunneling effect (penetration of the Coulomb barrier). This phenomenon lies at the heart of α decay from heavy nuclei and experimental studies of α-cluster structures in the excited states of light nuclei. However, α clusters are generally bound in the ground states of nuclei, particularly in neutron-rich nuclei with well-developed neutron skins, and cannot be emitted spontaneously. Therefore, to probe α clustering in these nuclei, one needs a new reaction probe that takes the α clusters out of the mother nucleus with minimized impact from the residual.
The quasi-free (p, pα) reaction is the method of choice. Fig. 32(a) shows a schematic of the (p, pα) reaction in normal kinematics. Generally, a high-energy (several hundred MeV) proton beam is sent onto the target of interest. At the instant of the reaction, a certain amount of kinetic energy is transferred from the incident proton to the preformed α cluster, which gets knocked out of the target nucleus. When the transferred energy is significantly large, compared with the α separation energy, the (p, pα) reaction can be considered as quasi-free. Thus, the α cluster is knocked out without being disturbed by the residual, which can simply be considered as a spectator. Hence, the corresponding reaction cross-section provides a good measure of the strength of the preformed α cluster in the mother nucleus. Moreover, similar quasi-free nucleon knockout reactions, such as (e, ep), (p, 2p), and (p, pn) have been well-established as experimental probes to investigate the single-particle structures of nuclei [221, 222]. Furthermore, the (p, pα) reaction was widely used in the 1970s and 1980s to study the cluster structures in light stable nuclei, such as 9Be and 12C [223-225]. However, further extending such (p, pα) experiments into a heavier mass region was largely constrained by the then available experimental equipment and theoretical tools with which to connect the experimental results to the internal cluster structure. During the past decade, important advances have been made in both experimental techniques and theoretical tools, making the (p, pα) reaction a sensitive probe for clustering in the ground states of nuclei (e.g., Refs. [226, 227]).
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F032.jpg)
In 2021, Tanaka et al. performed a quasi-free (p, pα) experiment at the Research Center for Nuclear Physics (RCNP) of Osaka University to measure the α-clustering strength and isotopic dependence of 112,116,120,124Sn [61]. As illustrated in Fig. 33, the experiment was performed with the 392 MeV proton beam at the WS beam line. The scattered protons and α particles after the (p, pα) reaction were detected in coincidence using the Grand Raiden and LAS spectrometers. The experimental setup was designed according to the kinematics of the proton scattering off a preformed α particle and optimized to achieve the detection of low-energy α particles (down to ~50 MeV) and a high signal-to-noise ratio. For all four tin isotopes, the MM spectrum [Fig. 34 (top panel)] shows a clear peak located at the known α-separation energy, which is simply determined by the mass, as expected for the quasi-free knockout of the preformed α clusters. Thus, these results provide direct evidence for the formation of α clusters in these tin isotopes. In addition, the observed momentum distribution of the α particles further reaffirms that the formation of the α particle indeed occurs in the low-density surface region of heavy nuclei, as predicted by theoretical calculations.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F033.jpg)
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F034.jpg)
The MM spectrum is fitted using a combination of the Gaussian for the ground-state peak and the simulated line shape of the continuum background. For each tin isotope, the (p, pα) cross section σp, pα is then deduced from the integral of the ground-state peak. As shown in the bottom panel of Fig. 34, σp, pα gradually decreases as the mass number increases, with an approximately twofold decrease from 112 Sn to 124Sn. The observed isotopic systematics of σp, pα are well-reproduced by theoretical calculations that consider the radial density distributions of the α clusters of the gRDF prediction and reaction mechanism. Further analytical results confirmed that the observed decline in σp, pα was predominantly caused by the decrease in the α-clustering strength, whereas the effect of the reaction mechanism was minor. Thus, these results support the tight interplay between the surface α-clustering and Δrnp in heavy nuclei and thereby lead to a reduction of Δrnp, in comparison to theoretical calculations that do not consider the α-clustering effect as predicted by the gRDF calculations [62]. A linear correlation between Δrnp and the slope parameter L has been predicted via mean-field model calculations [228] and is generally used to constrain L. Many projects worldwide are ongoing to measure the neutron skin thicknesses of heavy nuclei, such as 208Pb, with sufficient precision. These results suggest the necessity of considering the effects of nuclear clustering when constraining EOS parameters from the neutron skin thickness [63].
The recent increase in the availability of secondary beams of radioactive isotopes provides new opportunities to investigate α cluster structures in the ground states of unstable nuclei [17, 229, 230]. Recently, Li et al. conducted the (p, pα) reaction in inverse kinematics (Fig. 32(b)) on the neutron-rich unstable nucleus 10Be [231]. A well-developed α-2 n-α molecular-like cluster structure, that is, a dumbbell-like arrangement of two α cores surrounded by two valence neutrons occupying the π orbit, was predicted for the ground state of 10Be by microscopic cluster models, THSR [226], and AMD [232]. The experiment was performed at the Radioactive Isotope Beam Factory (RIBF) of the RIKEN Nishina Center, and the experimental setup is illustrated in Fig. 35. From the measured angle and energy of the recoil proton and α particle, the excitation energy of the residual 6He was reconstructed. This reconstruction is equivalent to the α-particle separation energy spectrum usually used in (p, pα) reactions in normal kinematics, such as [61]. In the experiment, the two reaction channels leading to different final states of the residual, namely, 6He10Be(p, pα)6He(g.s.), 10Be(p, pα)6He(e.x.), were observed. Fig. 36 shows that the experimental triple differential cross section (TDX) for the 10Be(p, pα)6He(g.s.) channel is nicely reproduced by the distorted-wave impulse approximation (DWIA) calculations that incorporate the microscopic α-cluster w.f. provided by THSR and AMD calculations. This experiment provides strong evidence for the molecular-like cluster structure of the ground state of 10Be, as predicted by THSR and AMD.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F035.jpg)
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F036.jpg)
Summary and perspectives
The nucleus is a typical quantum many-body system controlled mainly by short-range nuclear forces. This system, having neither a concrete center nor a confinement boundary, may exhibit large structural flexibility, including the overall shape changes and formation of the internal substructure, such as clusters. When moving toward the driplines or getting excited toward some cluster separation thresholds, small or even negative binding energy, corresponding to a relatively weak overall interaction, favors the local correlation and cluster formation [23, 3, 4, 233, 234]. Furthermore, the antisymmetrization among nucleons (Pauli blocking effect) tends to stabilize the cluster separation around the threshold energy, and the orthogonality between the quantum states drives the system to more intriguing exotic structure configurations. Cluster structures also play important roles in nuclear astrophysics processes, such as the synthesis of elements via fusion reactions, which are sensitive to cluster resonant states that are close to the threshold [20, 19]. Attention was also attracted to the inclusion of nuclear structure effects, such as deformation and clustering, into heavy-ion collision processes at intermediate and relativistic energies [235, 236].
The idea of RGM, introduced by Wheeler [12], provides a reasonable working base to address the flexibility of the nuclear structure. It is particularly useful to describe the mixing of the single-particle and clustering configurations by applying different representations to different groups of nucleons, as already realized in many instances (e.g. Ref. [59]). The main difficulties here involves the complexity of real calculations, although some simplification methods have been developed for certain cases, such as combining RGM with GCM or OCM. Moreover, RGM + GCM or RGM + OCM calculations can be applied to scattering (reaction) process as long as the c.m. coordinates are extended to large values ([237] and the references cited therein). The reaction theory is generally much less developed, compared with the structure theory, when multi-nucleon (cluster) correlation effects are encountered. Considering the increasing interest in applying multi-nucleon transfer (MNT) reactions to produce heavy neutron-rich nuclei and subsequently executing fission or neutron-evaporation processes [238-240], the RGM-type reaction models may need to be developed via the invention of appropriate approximation methods. Owing to the importance of antisymmetrization in forming the cluster structure and shaping the reaction processes, it is challenging to maintain this antisymmetry throughout the MNT calculations. The application of AMD-type approaches to heavy systems might present a promising solution. Because the local nucleon correlation is sensitive to the specific nuclear force, the cluster structure can serve as a testing ground for ab initio-type nuclear forces rooted in quantum chromodynamics (QCD) [17, 241]. Considering the complexity of the anticipated problems, artificial intelligence (AI) and machine learning should be of great help when implementing the theoretical solutions.
This review demonstrates the advantages of using the Brink-type wave-packet presentation to describe clustering phenomena. The orbital (eigenstate of the angular momentum) presentation favors the description of the independent particle motion in a centralized average potential, making it easier to reproduce the shell-like structure, including the magic numbers. However, it is difficult to address the strong multi-nucleon correlations unless an infinitely large model space is used. In contrast, the wave packet presentation may directly capture the local correlation instead of the orbital behavior. When moving to heavy neutron-rich systems, as can be expected with the future RIB facilities, the structure may reasonably become a shell-like core plus an expanded clustering surface, as depicted by the inset image in the bottom-right corner of Fig. 1. The flexible combination of both presentation modes under the RGM framework will likely be a key point in future theoretical approaches.
In addition to the investigation of the cluster structure around the separation thresholds via the aforementioned MM or IM methods, cluster structures in the nuclear ground states have recently become another focus of research. Leveraging the operating and upcoming RIB facilities worldwide, such as RIBF (Japan), FRIB (US), GSI/FAIR (Germany), HIAF (China) and RAON (South Korea), cluster knockout experiments, such as (p, pα), will be applied to more neutron-rich nuclei to probe cluster structures in their ground states. Theoretical calculations predicted that the formation of clusters could be favored in the ground state of light, neutron-rich nuclei, such as beryllium and carbon isotopes [17, 229, 230]. Neutron-rich nuclei with novel mixed states of clusters and neutrons could also serve as unique settings for investigations regarding the properties of neutron-rich matter, particularly when considering inhomogeneity due to the formation of clusters. Dedicated detector systems, such as TOGAXSI at RIBF of the RIKEN Nishna Center, are currently under development for the (p, pα) reaction in inverse kinematics on unstable nuclei [242]. It would also be interesting to investigate the formation of other light clusters, such as deuterons, tritons, and 3He, by using similar quasi-free knockout reactions, which are predicted to behave similarly to α clusters [63, 243-245].
Experimentally, combining the knockout reaction method at high energies and the excitation decay IM method at lower energies would still be necessary to probe different aspects of nuclear clustering. The former is applicable to the ground and low-lying states of the nucleus, whereas the latter is able to touch many more thresholds where new, exotic cluster configurations may emerge. Because neutron coupling and multi-neutron emissions will be essential in future experiments for investigating heavy, neutron-rich systems, relevant detection systems should be developed to simultaneously record multi-fragments and multi-neutrons (Fig. 37), although these tasks will be challenging.
-202412-小份文件打包/1001-8042-35-12-007/alternativeImage/1001-8042-35-12-007-F037.jpg)
The galactic center massive black hole and nuclear star cluster
. Rev. Mod. Phys. 82, 3121-3195 (2010). https://doi.org/10.1103/RevModPhys.82.3121The hidden-charm pentaquark and tetraquark states
. Phys. Rep. 639, 1-121 (2016). https://doi.org/10.1016/j.physrep.2016.05.004Nuclear clusters and nuclear molecules
. Phys. Rep. 432, 43-113 (2006). https://doi.org/10.1016/j.physrep.2006.07.001Physics of exotic nuclei
. Nat. Rev. Phys. (2024)The nubase2020 evaluation of nuclear physics properties *
. Chin. Phys. C 45,Nuclidic mass formula on a spherical basis with an improved even-odd term
. Prog. Theor. Phys. 113, 305 (2005).History of cluster structure in nuclei
. J. Phys. Conf. Ser. 111,Nuclear physics a. stationary states of nuclei
. Rev. Mod. Phys. 8, 82-229 (1936).The alpha-particle model of the nucleus
. Phys. Rev. 54, 681-692 (1938). https://doi.org/10.1103/PhysRev.54.681On the mathematical descrition of light nuclei by the method of resonating group structure
. Phys. Rev. 52, 1107-1122 (1937). https://doi.org/10.1103/PhysRev.52.1107Collective motiona in nuclei by the method of generator coodinates
. Phys. Rev. 108, 311-327 (1957). https://doi.org/10.1103/PhysRev.108.311Effect of Pauli principle in scattering of two clusters
. Prog. Theo. Phys. 40, 893-894 (1968). https://doi.org/10.1143/PTP.40.893Cluster models from RGM to alpha condensation and beyond
. Prog. Part. Nucl. Phys. 82, 78-132 (2015). https://doi.org/10.1016/j.ppnp.2015.01.001Colloquium: Status of α-particle condensate structure of the hoyle state
. Rev. Mod. Phys. 89,Microscopic clustering in light nuclei
. Rev. Mod. Phys. 90,On nuclear reactions occuring in very hot stars.i. the synthesis of elements from carbon to nickel
. Astrophys. J. Suppl. S. 1, 121-146 (1954). https://doi.org/10.1086/190005The Hoyle state in 12C
. Prog. Part. Nucl. Phys. 78, 1-23 (2014). https://doi.org/10.1016/j.ppnp.2014.06.001Nuclear clusters in astrophysics
. Nucl. Phys. A 834, 647c-650c (2010). https://doi.org/10.1016/j.nuclphysa.2010.01.113A molecule-like structure in atomic nuclei of 16O* and 20Ne*
. Prog. Theor. Phys. 40, 277 (1968).Alpha cluster model
.The systematic structure-change into the molecule-like structures in the self-conjugate 4n nuclei
. Prog. Theor. Phys. Suppl. E 68, 464-475 (1968). https://doi.org/10.1143/PTPS.E68.464Four-particle condensate in strongly coupled fermion systems
. Phys. Rev. Lett. 80, 3177-3180 (1998). https://doi.org/10.1103/PhysRevLett.80.3177Alpha cluster condensation in 12C and 16O
. Phys. Rev. Lett. 87,Concepts of nuclear α-particle condensation
. Phys. Rev. C 80,The 5α condensate state in 20Ne
. Nat. Commun. 14, 8206 (2023). https://doi.org/10.1038/s41467-023-43816-9New evidence of the hoyle-like structure in 16O
. Sci. Bull. 68, 1119-1126 (2023). https://doi.org/10.1016/j.scib.2023.04.031The Structure of 9Be Nucleus by a Molecular Model. I)
. Prog. Theor. Phys. 57, 866-881 (1977). https://doi.org/10.1143/PTP.57.866The Structure of 9Be by a Molecular Model. II
. Prog. Theor. Phys. 61, 1049-1064 (1979). https://doi.org/10.1143/PTP.61.1049Nuclear clustering in light neutron-rich nuclei
. Nucl. Sci. Tech. 29, 1-16 (2018). https://doi.org/10.1007/s41365-018-0522-xStructure of Li and Be isotopes studied with antisymmetrized molecular dynamics
. Phys. Rev. C 52, 628 (1995). https://doi.org/10.1103/PhysRevC.52.628-646Antisymmetrized molecular dynamics and its applications to cluster phenomena
. Prog. Theor. Exp. Phys 2012,Fermionic molecular dynamics
. Nucl. Phys. A 515, 147-172 (1990). https://doi.org/10.1016/0375-9474(90)90328-JContact representation of short-range correlation in light nuclei studied by the high-momentum antisymmetrized molecular dynamics
. Phys. Rev. C 99,The container picture with two-alpha correlation for the ground state of 12C
. Prog. Theor. Exp. Phys. 2014, 101D01-101D01 (2014). https://doi.org/10.1093/ptep/ptu127Nonlocalized clustering: A new concept in nuclear cluster structure physics
. Phys. Rev. Lett. 110,“Container” evolution for cluster structures in 16O
. Phys. Rev. C 97,Molecular orbital structures in 10Be
. Phys. Rev. C 61,Coexistence of covalent superdeformation and molecular resonances in an unbound region of 12Be
. Phys. Rev. Lett. 100,Rod-shaped nuclei at extreme spin and isospin
. Phys. Rev. Lett. 115,α-clustering in atomic nuclei from first principles with statistical learning and the hoyle state character
. Nat. Commun. 13, 2234 (2022). https://doi.org/10.1038/s41467-022-29582-0Giant dipole resonance as a fingerprint of α clustering configurations in 12C and 16O
. Phys. Rev. Lett. 113,α:2n:α Molecular Band in 10Be
. Phys. Rev. Lett. 96,Determination of the cluster-decay branching ratio from a near-threshold molecular state in 10Be
. Phys. Rev. C 101,Observation of enhanced monopole strength and clustering in 12Be
. Phys. Rev. Lett. 112,Experimental investigation of a linear-chain structure in the nucleus 14C
. Phys. Lett. B 766, 11-16 (2017). https://doi.org/10.1016/j.physletb.2016.12.050Selective decay from a candidate of the σ-bond linear-chain state in 14C
. Phys. Rev. C 95,Positive-parity linear-chain molecular band in 16C
. Phys. Rev. Lett. 124,Nuclear linear-chain structure arises in carbon-14
. Commun. Phys. 6, 220 (2023). https://doi.org/10.1038/s42005-023-01342-6Interpretation of some of the excited states of 4n self-conjugate nuclei
. Phys. Rev. 101, 254 (1956). https://doi.org/10.1103/PhysRev.101.254The neutron halo of extremely neutron-rich nuclei
. Europhysics Lett. 4, 409 (1987). https://doi.org/10.1209/0295-5075/4/4/005Bound state properties of borromean halo nuclei: 6He and 11Li
. Phys. Rep. 231, 151-199 (1993). https://doi.org/10.1016/0370-1573(93)90141-YSpatial structure of neutron cooper pair in low density uniform matter
. Phys. Rev. C 73,Coexistence of BCS- and BEC-like pair structures in halo nuclei
. Phys. Rev. Lett. 99,Exploring the manifestation and nature of a dineutron in two-neutron emission using a dynamical dineutron model
. Phys. Rev. C 97,Fermion pair dynamics in open quantum systems
. Phys. Rev. Lett. 126,Surface localization of the dineutron in 11Li
. Phys. Rev. Lett. 125,Dineutron formation and breaking in 8He
. Phys. Rev. C 88,Observation of the exotic 02+ cluster state in 8He
. Phys. Rev. Lett. 131,Formation of clusters in dilute neutron-rich matter
. Science 371, 260-264 (2021). https://doi.org/10.1126/science.abe4688Neutron skin thickness of heavy nuclei with α-particle correlations and the slope of the nuclear symmetry energy
. Phys. Rev. C 89,Equations of state for supernovae and compact stars
. Rev. Mod. Phys. 89,Clustering in nuclear systems: from finite nuclei to neutron stars
. Science Bulletin 66, 2054-2056 (2021). https://doi.org/10.1016/j.scib.2021.07.009Three-body decay of linear-chain states in 14C
. Phys. Rev. C 95,Characteristic α and 6He decays of linear-chain structures in 16C
. Phys. Rev. C 97,Chapter II. Theory of Resonating Group Method and Generator Coordinate Method, and Orthogonality Condition Model
. Prog. Theor. Phys. Suppl. 62, 11-89 (1977). https://doi.org/10.1143/PTPS.62.11Scattering of alpha particles by oxygen. i. bombarding energy range 5.8 to 10.0 MeV
. Phys. Rev. 160, 782-790 (1967). https://doi.org/10.1103/PhysRev.160.782Resonating-group method for nuclear many-body problems
. Phys. Rep. 47, 167-223 (1978).Low-energy α+6He elastic scattering with the resonating-group method
. Phys. Rev. C 59, 817-825 (1999). https://doi.org/10.1103/PhysRevC.59.817Generator coordinate treatment of composite particle reaction and molecule-like structures
. Prog. Theor. Phys. 43, 375-389 (1970). https://doi.org/10.1143/PTP.43.375Resonances and scattering in microscopic cluster models with the complex-scaled generator coordinate method
. Phys. Rev. C 107,Microscopic cluster model analysis of 14O+p elastic scattering
. Phys. Rev. C 72,Microscopic analysis of 10,11Be elastic scattering on protons and nuclei, and breakup processes of 11Be within the 10Be+n cluster model
. Phys. Rev. C 91,2α+t cluster structure in 11B
. Phys. Rev. C 98,A study of α-widths of 20ne based on 16O and α-cluster model
. Prog. Theor. Phys. 49, 1765-1767 (1973). https://doi.org/10.1143/PTP.49.1765Many-cluster problem by the orthogonality condition model: general discussion and 12C problem
. Prog. Theor. Phys. 53, 447-460 (1975).α-cluster structures and monopole excitations in 13C
. Phys. Rev. C 92,Enhanced monopole transition strength from the cluster-decay of 13C
. Sci. China-Phys. Mech. Astron 62, 1-7 (2019). https://doi.org/10.1007/s11433-018-9258-7Cluster structures of the 0+ states of Λ13C Cwith the 3α+Λ four-body orthogonality condition model
. Phys. Rev. C 107,α+α+t cluster structures and 12C(02+)-analog states in 11B
. Phys. Rev. C 82,Molecular viewpoints in nuclear structure
. Phys. Rev. 52, 1083 (1937). https://doi.org/10.1103/PhysRev.52.1083On the Stability of α-Cluster Structures in 8Be and 12C Nuclei
. Prog. Theor. Phys. 46, 352-354 (1971). https://doi.org/10.1143/PTP.46.352Recent developments in nuclear cluster physics
. Prog. Theor. Phys. Suppl. 192, 1-238 (2012). https://doi.org/10.1143/PTPS.192.1A molecular-obital model of the atomic nuclei
. Prog. Theor. Phys. 49, 800-824 (1973). https://doi.org/10.1143/PTP.49.800Two-center molecular states in 9B, 9Be, 10Be, and 10B
. Z. Phys. A 354, 37-43 (1996). https://doi.org/10.1007/s002180050010Molecular-orbital structure in neutron-rich C isotopes
. Phys. Rev. C 64,Application of the generalized two-center cluster model to 10Be
. Phys. Lett. B 588, 43-48 (2004). https://doi.org/10.1016/j.physletb.2004.01.090Unified studies of chemical bonding structures and resonant scattering in light neutron-excess systems, 10,12Be
. Rep. Prog. Phys. 77,Imprint of adiabatic structures in monopole excitations of 12Be
. Phys. Rev. C 83,Fragment formation studied with antisymmetrized version of molecular dynamics with two-nucleon collisions
. Phys. Rev. Lett. 68, 2898-2900 (1992). https://doi.org/10.1103/PhysRevLett.68.2898Structure of the hoyle state in 12C
. Phys. Rev. Lett. 98,Cluster structures within fermionic molecular dynamics
. Nucl. Phys. A 738, 357-361 (2004). https://doi.org/10.1016/j.nuclphysa.2004.04.061Dineutron structure in 8He
. Phys. Rev. C 76,Nucleus-nucleus interaction and nuclear saturation property: Microscopic study of 16O+16O interaction by new effective nuclear force
. Prog. Theor. Phys. 64, 1608-1626 (1980). https://doi.org/10.1143/PTP.64.1608Hartree-fock-bogolyubov calculations with the D1 effective interaction on spherical nuclei
. Phys. Rev. C 21, 1568-1593 (1980). https://doi.org/10.1103/PhysRevC.21.1568Variation after angular momentum projection for the study of excited states based on antisymmetrized molecular dynamics
. Phys. Rev. Lett. 81, 5291-5293 (1998). https://doi.org/10.1103/PhysRevLett.81.5291Structure of excited states of 10Be studied with antisymmetrized molecular dynamics
. Phys. Rev. C 60,Structure and decay pattern of the linear-chain state in 14C
. Phys. Rev. C 94,Coulomb shift in the mirror pair of 14C and 14O as a signature of the linear-chain structure
. Phys. Rev. C 99,Pure σ-bond linear chain in 16C
. Phys. Rev. C 102,Resonances in 14C observed in the 4He(10Be,α)10Be reaction
. Phys. Rev. C 90,One-dimensionality in atomic nuclei: A candidate for linear-chain α clustering in 14C
. Phys. Rev. C 93,On the displacement of corresponding energy levels of 13C and 13N
. Phys. Rev. 81, 412-416 (1951). https://doi.org/10.1103/PhysRev.81.412An analysis of the energy levels of the mirror nuclei, 13C and 13N
. Phys. Rev. 88, 1109-1125 (1952). https://doi.org/10.1103/PhysRev.88.1109Structure and decay mechanism of the low-lying states in 9Be and 9B
. Phys. Rev. C 108,Analysis of previous microscopic calculations for the second 0+ state in 12C in terms of 3-α particle bose-condensed state
. Phys. Rev. C 67,Description of 8Be as deformed gas-like two-alpha-particle states
. Prog. Theor. Phys. 108, 297-322 (2002). https://doi.org/10.1143/PTP.108.297Nonlocalized cluster dynamics and nuclear molecular structure
. Phys. Rev. C 89,New concept for the ground-state band in 20Ne within a microscopic cluster model
. Phys. Rev. C 86,Candidates for the 5α condensed state in 20Ne
. Phys. Lett. B 819,Zur quantentheorie des atomkernes
. Zeitschrift für Physik 51, 204-212 (1928). https://doi.org/10.1007/BF01343196Mass defect curve and nuclear constitution
. Proc. Roy. Soc. A 126, 632 (1930). https://doi.org/10.1098/rspa.1930.0032Surface-clustering and -decays of 212Po
. Nuclear Physics A 323, 45-60 (1979). https://doi.org/10.1016/0375-9474(79)90415-9Absolute alpha decay width of 212Po in a combined shell and cluster model
. Phys. Rev. Lett. 69, 37-40 (1992). https://doi.org/10.1103/PhysRevLett.69.37Abrupt changes in α-decay systematics as a manifestation of collective nuclear modes
. Phys. Rev. C 81,Photon-neutron correlations and microscopic description of alpha decay
. Nuclear Physics A 549, 407-419 (1992). https://doi.org/10.1016/0375-9474(92)90087-Zα decay in the complex-energy shell model
. Phys. Rev. C 86,Global calculation of α-decay half-lives with a deformed density-dependent cluster model
. Phys. Rev. C 74,Exact estimate of the α-decay rate and semiclassical approach in deformed nuclei
. Phys. Rev. C 92,α-cluster formation and decay: The role of shell structure
. Phys. Rev. C 104,α-cluster formation in heavy-emitters within a multistep model
. Physics Letters B 813,New insight into α clustering of heavy nuclei via their α decay
. Physics Letters B 777, 298-302 (2018). https://doi.org/10.1016/j.physletb.2017.12.046α-decay width of 212Po from a quartetting wave function approach
. Phys. Rev. C 93,Improved empirical formula for α-decay half-lives
. Phys. Rev. C 101,Alpha-particle formation and decay rates from skyrme-HFB wave functions
. Phys. Scr. 89,Model-independent analysis on the regular behavior of α preformation probability in heavy nuclei
. Phys. Rev. C 108,Analytic formula for estimating the α-particle preformation factor
. Phys. Rev. C 102,α decay and cluster radioactivity within the redefined preformed cluster method
. Phys. Rev. C 102,Preformation factor for α particles in isotopes near N=Z
. Phys. Rev. C 89,Nuclear clusters bound to doubly magic nuclei: The case of 212Po
. Phys. Rev. C 90,Alpha-like clustering in 20Ne from a quartetting wave function approach
. J. Low Temp. Phys. 189, 383-409 (2017). https://doi.org/10.1007/s10909-017-1796-9α-cluster formation and decay in the quartetting wave function approach
. Phys. Rev. C 95,α decay to a doubly magic core in the quartetting wave function approach
. Phys. Rev. C 101,Decay width and the shift of a quasistationary state
. Phys. Rev. Lett. 59, 262-265 (1987). https://doi.org/10.1103/PhysRevLett.59.262Novel approach to tunneling problems
. Phys. Rev. A 38, 1747-1759 (1988). https://doi.org/10.1103/PhysRevA.38.1747Search for cluster structure of excited states in 14C
. Eur. Phys. J. A 21, 193-215 (2004). https://doi.org/10.1140/epja/i2003-10188-9Structures in 20O from the 14C(7Li, p) reaction at 44 MeV
. Eur. Phys. J. A 47, 44 (2011). https://doi.org/10.1140/epja/i2011-11044-1Helium clustering in neutron-rich be isotopes
. Phys. Lett. B 580, 129-136 (2004). https://doi.org/10.1016/j.physletb.2003.11.035Sub-threshold states in 19Ne relevant to 18F(p,α)15O
. Phys. Rev. C 103,Spin determination by in-plane angular correlation analysis in various coordinate systems
. Chin. Phys. C 43,Helium breakup states in 10Be and 12Be
. Phys. Rev. C 63,New measurement of 4He and proton decays from resonant states in 19Ne
. Chinese Physics C 47,A procedure for the analysis of the data of a three body nuclear reaction
. Nucl. Instrum. Meth. A 295, 373-376 (1990). https://doi.org/10.1016/0168-9002(90)90715-IStudy on the π-bond linear-chain molecular band in 14C
. Ph.D. thesis,Isoscalar monopole excitations in 16O: α-cluster states at low energy and mean-field-type states at higher energy
. Phys. Rev. C 85,Isospin and macroscopic models for the excitation of giant resonances and other collective states
. Nucl. Phys. A 472, 215-236 (1987). https://doi.org/10.1016/0375-9474(87)90208-9Folded potential analysis of the excitation of giant resonances by heavy ions
. Phys. Rev. C 52, 1554-1564 (1995). https://doi.org/10.1103/PhysRevC.52.1554Missing monopole strength in 58Ni and uncertainties in the analysis of α-particle scattering
. Phys. Rev. C 55, 285-297 (1997). https://doi.org/10.1103/PhysRevC.55.285Compressional-mode giant resonances in deformed nuclei
. Phys. Lett. B 549, 58-63 (2002). https://doi.org/10.1016/S0370-2693(02)02892-7Incompressibility of nuclear matter from the giant monopole resonance
. Phys. Rev. Lett. 82, 691-694 (1999). https://doi.org/10.1103/PhysRevLett.82.691Helium-helium clustering states in 12Be
. Phys. Rev. C 91,Energy levels of light nuclei A = 16-17
. Nucl. Phys. A 460, 1-110 (1986). https://doi.org/10.1016/0375-9474(86)90038-2Monopole excitation to cluster states
. Prog. Theor. Phys. 120, 1139-1167 (2008). https://doi.org/10.1143/PTP.120.1139Observation of the π2σ2-bond linear-chain molecular structure in 16C
. Phys. Rev. C 105,Investigation of the 14C+α molecular configuration in 18O by means of transfer and sequential decay reaction
. Phys. Rev. C 99,α decay of excited states in 14C
. Phys. Rev. C 75,The analysis of angular correlations in breakup reactions: the effect of coordinate geometries
. Nucl. Instrum. Meth. A 383, 463-472 (1996). https://doi.org/10.1016/S0168-9002(96)00866-2Reaction mechanism and spin information in the sequential breakup of 18O at 82 MeV
. Nucl. Phys. A 420, 320-350 (1984). https://doi.org/10.1016/0375-9474(84)90445-7Decay angular correlations and spectroscopy for 10Be*→4He+6He
. Phys. Rev. C 64,Angular correlations for α-particle decay in the reaction 12[ 12C,12C(31−)→8Be(01+)+α ]12C* at Ec.m.=32.5 MeV
. Phys. Rev. C 53, 2879-2884 (1996). https://doi.org/10.1103/PhysRevC.53.2879Excitation of 24Mg states through the interaction of 85 MeV 16O ions with 12C and 24Mg targets
. Phys. Rev. C 44, 111-118 (1991). https://doi.org/10.1103/PhysRevC.44.111New experimental investigation of the structure of 10Be and 16C by means of intermediate-energy sequential breakup
. Phys. Rev. C 93,Observation of 6He+t cluster states in 9Li
. Phys. Rev. C 103,Angular correlation measurements for the α+6He decay of 10Be
. Phys. Rev. C 70,Angular correlations in heavy ion reactions
. Phys. Lett. B 153, 21-24 (1985). https://doi.org/10.1016/0370-2693(85)91433-9Determination of the cluster spectroscopic factor of the 10.3 MeV state in 12Be
. Sci. China-Phys. Mech. Astron 000, P.1613-1617 (2014). https://doi.org/10.1007/s11433-014-5555-5The r-matrix theory
. Rep. Prog. Phys. 73,Proton decay of excited states in 12N and 13O and the astrophysical 11C(p,γ)12N reaction rate
. Phys. Rev. C 87,R-matrix theory of nuclear reactions
. Rev. Mod. Phys. 30, 257-353 (1958). https://doi.org/10.1103/RevModPhys.30.257Digital signal acquisition system for complex nuclear reaction experiments
. Nucl. Sci. Tech. 35, 12 (2024). https://doi.org/10.1007/s41365-024-01366-9Control and data acquisition system for collinear laser spectroscopy experiments
. Nucl. Sci. Tech. 34, 38 (2023). https://doi.org/10.1007/s41365-023-01197-0Two annular CsI(Tl) detector arrays for the charged particle telescopes
. Nucl. Sci. Tech 34, 159 (2023). https://doi.org/10.1007/s41365-023-01319-8Exotic molecular states in 12Be
. Phys. Rev. Lett. 82, 1383 (1999). https://doi.org/10.1103/PhysRevLett.82.1383Particle decay of 12Be excited states
. Phys. Rev. C 76,Cluster structures of excited states in 14C
. Phys. Rev. C 82,Structure of low-lying states of 12C and Λ13C in a beyond-mean-field approach
. Phys. Rev. C 109,Dipole oscillation modes in light α-clustering nuclei
. Phys. Rev. C 94,4He decay of excited states in 14C
. Phys. Rev. C 68,The 6He scattering and reactions on 12C and cluster states of 14C
. Nucl. Phys. A 730, 285-298 (2004). https://doi.org/10.1016/j.nuclphysa.2003.11.011Alpha-decay of excited states in 13C and 14C
. Nucl. Phys. A 765, 263-276 (2006). https://doi.org/10.1016/j.nuclphysa.2005.11.004Measurement of α and neutron decay widths of excited states of 14C
. Phys. Rev. C 78,Investigation of the near-threshold cluster resonance in 14C*
. Chin. Phys. C 42,Cluster decay of the high-lying excited states in 14C*
. Chin. Phys. C 40,Novel evidence for the σ-bond linear-chain molecular structure in 14C
. Chin. Phys. C 45,Cluster structure of 11C investigated with a breakup reaction
. Phys. Rev. C 107,Cluster structure of 3α+p states in 13N
. Phys. Rev. C 109,Missing monopole strength of the hoyle state in the inelastic α+12C scattering
. Phys. Lett. B 660, 331-338 (2008). https://doi.org/10.1016/j.physletb.2007.12.059Determination of nuclear radii for unstable states in 12C with diffraction inelastic scattering
. Phys. Rev. C 80,Cluster states and monopole transitions in 16O
. Phys. Rev. C 89,Characterization of the proposed 4-α cluster state candidate in 16O
. Phys. Rev. C 95,8Be and α decay of 16O
. Phys Rev C 51, 1682-1692 (1995). https://doi.org/10.1103/PhysRevC.51.16828Be+8Be decay of excited states in 16O
. Phys Rev C 70,8Be + 8Be and 12C+ α breakup states in 16O populated via the 13C(4He,4α)n reaction
. Phys Rev C 94,Search for the hoyle analogue state in 16O
. Eur. Phys. J. A 57, 286 (2021). https://doi.org/10.1140/epja/s10050-021-00592-8α-particle condensate states in 16O
. Phys. Lett. B 684, 127-131 (2010). https://doi.org/10.1016/j.physletb.2009.12.066Unexpectedly enhanced α-particle preformation in 48Ti probed by the (p,pα) reaction
. Phys. Rev. C 103,Yield measurements for resonances above the multi-α threshold in 20Ne
. Phys. Rev. C 87,Search for alpha inelastic condensed state in 24Mg
. J. Phys. Conf. Ser. 436,Searching for states analogous to the 12C hoyle state in heavier nuclei using the thick target inverse kinematics technique
. Phys. Rev. C 98,Search for the 6α condensed state in 24Mg using the 12C + 12C scattering
. Phys. Lett. B 848,Examination of evidence for resonances at high excitation energy in the 7α disassembly of 28Si
. Phys. Rev. C 99,Experimental investigation of α condensation in light nuclei
. Phys. Rev. C 100,New isotope 220Np: Probing the robustness of the N = 126 shell closure in neptunium
. Phys. Rev. Lett. 122,Short-lived α-emitting isotope 222Np and the stability of the N = 126 magic shel
. Phys. Rev. Lett. 125,New α-emitting isotope 214U and abnormal enhancement of α-particle clustering in lightest uranium isotopes
. Phys. Rev. Lett. 126,Discovery of new isotopes 160Os and 156W: Revealing enhanced stability of the N = 82 shell closure on the neutron-deficient side
. Phys. Rev. Lett. 132,Recent experimental progress in nuclear halo structure studies
. Prog. Part. Nucl. Phys. 68, 215-313 (2013). https://doi.org/10.1016/j.ppnp.2012.07.001Recoil proton tagged knockout reaction for 8He
. Phys. Lett. B 707, 46-51 (2012). https://doi.org/10.1016/j.physletb.2011.12.009Evidence for 15O+α resonance structures in 19Ne via direct measurement
. Phys. Rev. C 96,α-cluster structure of 18O
. Phys. Rev. C 90,Prototype AT-TPC: Toward a new generation active target time projection chamber for radioactive beam experiments
. Nucl. Instrum. Meth. A 691, 39-54 (2012). https://doi.org/10.1016/j.nima.2012.06.050Performance of a small AT-TPC prototype
. Nucl. Sci. Tech. 29, 97 (2018). https://doi.org/10.1007/s41365-018-0437-6Performance of the AT-TPC based on two-dimensional readout strips
. Nucl. Sci. Tech. 32, 85 (2021). https://doi.org/10.1007/s41365-021-00919-6Texas active target (texat) detector for experiments with rare isotope beams
. Nucl. Instrum. Meth. A 957,Structure of 9C through proton resonance scattering with the texas active target detector
. Phys. Rev. C 100,α-cluster structure of 18Ne
. Phys. Rev. C 106,A time projection chamber for high accuracy and precision fission cross-section measurements
. Nucl. Instrum. Meth. A 759, 50-64 (2014). https://doi.org/10.1016/j.nima.2014.05.057Proton-induced knockout reactions with polarized and unpolarized beams
. Prog. Part. Nucl. Phys. 96, 32-87 (2017). https://doi.org/10.1016/j.ppnp.2017.06.002Quenching of single-particle strength from direct reactions with stable and rare-isotope beams
. Prog. Part. Nucl. Phys. 118,Absolute spectroscopic factors from the (p,pα) reaction at 100 MeV on 1p-shell nuclei
. Phys. Rev. C 15, 69-83 (1977). https://doi.org/10.1103/PhysRevC.15.69Alpha-clustering systematics from the quasifree (p,pα) knockout reaction
. Phys. Rev. C 23, 576-579 (1981). https://doi.org/10.1103/PhysRevC.23.576Alpha-particle spectroscopic strengths using the (p,pα) reaction at 101.5 MeV
. Phys. Rev. C 29, 1273-1288 (1984). https://doi.org/10.1103/PhysRevC.29.1273Manifestation of α clustering in 10Be via α-knockout reaction
. Phys. Rev. C 97,Quantitative description of the 20Ne(p,pα)16O reaction as a means of probing the surface α amplitude
. Phys. Rev. C 100,Nuclear equation of state from ground and collective excited state properties of nuclei
. Prog. Part. Nucl. Phys. 101, 96-176 (2018). https://doi.org/10.1016/j.ppnp.2018.04.001Antisymmetrized molecular dynamics studies for exotic clustering phenomena in neutron-rich nuclei
. Eur. Phys. J. A 52, 1273-1288 (2016). https://doi.org/10.1140/epja/i2016-16373-9Coexistence of cluster and shell-model aspects in nuclear systems
. Front. Phys. 13,Validation of the 10Be ground-state molecular structure using 10Be(p,pα)6He triple differential reaction cross-section measurements
. Phys. Rev. Lett. 131,Proton radii of Be, B, and C isotopes
. Phys. Rev. C 91,Progress and perspective of the research on exotic structures of unstable nuclei
. Nucl. Tech 46, 189-194 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080020Advancements in nuclear physics research using beijing HI-13 tandem accelerator
. Nucl. Tech 46, 51-62 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.080004Low density in-medium effects on light clusters from heavy-ion data
. Phys. Rev. Lett. 125,Signatures of α clustering in light nuclei from relativistic nuclear collisions
. Phys. Rev. Lett. 112,Imprints of α clustering in the density profiles of 12C and 16O
. Phys. Rev. C 107,Pathway for the production of neutron-rich isotopes around the N = 126 shell closure
. Phys. Rev. Lett. 115,Mass correlation between light and heavy reaction products in multinucleon transfer 197Au + 130Te collisions
. Phys. Rev. C 97,Detailed study of multinucleon transfer features in the 136Xe + 238U reaction
. Phys. Rev. C 109,Unbound spectra of neutron-rich oxygen isotopes predicted by the gamow shell model
. Phys. Rev. C 103,Designing togaxsi: Telescope for inverse-kinematics cluster-knockout reactions
. Nucl. Instrum. Meth. B 542, 4-6 (2023). https://doi.org/10.1016/j.nimb.2023.05.054Composition and thermodynamics of nuclear matter with light clusters
. Phys. Rev. C 81,Low density nuclear matter with light clusters in a generalized nonlinear relativistic mean-field model
. Phys. Rev. C 95,Nuclear clustering-manifestations of non-uniformity in nuclei
. Phil. Trans. R. Soc. A. 382,Dedicated to Professor Wenqing Shen in honour of his 80th birthday