Processing math: 100%

Cross-section measurements for the 58,60,61Ni(n, α)55,57,58Fe reactions at 8.50, 9.50 and 10.50 MeV neutron energies

  • Cross sections of the 58,60,61Ni(n, α)55,57,58Fe reactions were measured at 8.50, 9.50 and 10.50 MeV neutron energies based on the HI-13 tandem accelerator of China Institute of Atomic Energy (CIAE) with enriched 58Ni, 60Ni, and 61Ni foil samples with backings. A twin gridded ionization chamber (GIC) was used as the charged particle detector, and an EJ-309 liquid scintillator was used to obtain the neutron energy spectra. The relative and absolute neutron fluxes were determined via three highly enriched 238U3O8 samples inside the GIC. The uncertainty of the present data of the 58Ni(n, α)55Fe reaction is smaller than most existing measurements. The present data of 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reactions are the first measurement results above 8 MeV. The present experimental data could be reasonably reproduced by calculations with TALYS-1.9 by adjusting several default values of theoretical model parameters.
  • All current observations suggest that our universe contains an excess of matter over antimatter. The Planck result [1] showed that the left baryon-to-photon ratio is nB/nγ=(6.12±0.04)×1010 [2]. For theoretical discussions, it is more convenient to use the baryon-to-entropy ratio nB/s to quantify this asymmetry, which is approximately nB/s8.7×1011 as calculated from the observational result, as s7.04nγ at present. The origin of this baryon number asymmetry remains unexplained in cosmology. Conventionally, it was argued that this asymmetry was generated dynamically from an initial symmetric baryon phase at the following conditions [3]: (1) baryon number non-conserving interactions, (2) C and CP violations, (3) a departure from thermal equilibrium.

    However, if the CPT symmetry is violated, the baryon number asymmetry could be generated in thermal equilibrium [4]. For example, in Refs. [5-7], an effective interaction,

    Lin=cMμϕJμB,

    (1)

    between the dynamic dark energy (quintessence) and baryons was introduced, which considers a dimensionless coupling constant c and a cut-off mass scale M. As the universe expands, the background evolution of the scalar field ϕ breaks the Lorentz and CPT symmetries spontaneously, providing an effective chemical potential for baryons and the opposite for antibaryons:

    cMμϕJμBcM˙ϕnB=cM˙ϕ(nbnˉb),μb=cM˙ϕ=μˉb.

    (2)

    This creates a difference between the distribution functions of baryons and antibaryons in thermal equilibrium, producing an excess of baryons over antibaryons [8]:

    nB=gbT36[μbT+O(μbT)3]c˙ϕT23M,

    (3)

    where gb=2 is the internal degrees of freedom of the baryon. In terms of the entropy density s=(2π2/45)gsT3, the baryon-to-entropy ratio is written as

    nBs=15c2π2˙ϕgsMT.

    (4)

    The parameter gs denotes the total number of relativistic degrees of freedom in the universe and decreases slightly with cosmic expansion and cooling [8]. In the standard model of particle physics, gs=106.75 in the early universe during radiation domination, i.e., after reheating but well before the electroweak phase transition. Please note that to generate the baryon number asymmetry (4), there should be baryon number violating interactions in thermal equilibrium, otherwise the coupling introduced in Eq. (1) does not make sense, as LinμϕJμBϕμJμB=0 after removing a total derivative term. During this early stage, with B-violating processes in thermal equilibrium, the baryon-to-entropy ratio (4) changes as the universe expands. This occurs until the temperature of the universe cools to TD when the B-violating interactions decouple from the thermal bath, after which the baryon-to-entropy ratio remains unchanged. Therefore, the relic baryon number asymmetry observed today should be

    nBs|TD=15c2π2˙ϕgsMT|TD.

    (5)

    In such mechanisms, the scalar field ϕ in the coupling (1) is not necessarily dark energy and can be replaced by other cosmic scalar fields. In Ref. [9], the authors proposed the following interaction,

    Lin=1M2μRJμB,

    (6)

    between the curvature scalar and baryons. As a result, the baryon-to-entropy ratio is

    nBs˙RM2T.

    (7)

    As this baryon number asymmetry is generated by the derivative of the curvature, which accounts for the gravitation, this type of baryogenesis model is termed “gravitational baryogenesis.” Unfortunately, within the framework of general relativity (GR), the Einstein equation gives: R=8πGTμμ=8πG(13w)ρ, so that R and ˙R vanish in the radiation dominated epoch as w=1/3, and no baryon number asymmetry can be produced. To circumvent this problem, various methods to obtain a non-vanishing ˙R have been considered in Ref. [9], such as including significant trace anomaly effects in the radiation dominated time Tμμ0, baryon number asymmetry produced in the reheating period, during which the universe has w=0, and so on. Later, gravitational baryogenesis has been realized using other modifications, such as considering a gravity theory different from GR [10], or replacing the coupling μRJμB with μf(R)JμB [11], where f(R) is a non-linear function of the curvature scalar.

    In this study, we are interested in gravitational baryogenesis within the frameworks of teleparallel gravity (TG) [12] and symmetric teleparallel gravity (STG) [13]. These frameworks provide gravity models in non-Riemannian systems. Both TG and STG can be equivalent to GR but are formulated in flat spacetime where the curvature vanishes. In the TG model, gravity is attributed to the spacetime torsion, while in the STG model, gravity is identified with non-metricity. Within the GR equivalent TG model and the GR equivalent STG model, we will first consider the baryogenesis induced by derivative couplings to the baryon current. These couplings are similar to Eq. (6) except the curvature scalar is replaced by the torsion scalar and the non-metricity scalar, respectively. We show that in these baryogenesis models, the produced baryon-to-entropy ratios are too small to be consistent with the observed value. Subsequently, we employ gravitational leptogenesis, in which the torsion scalar and the non-metricity scalar are coupled derivatively to the current of BL. With appropriate cut-off scales, these gravitational leptogenesis models can generate the required baryon number asymmetry. We would like to point out that the gravitational baryogenesis within the framework of TG has been also studied in Refs. [14-16] in various ways. Gravitational leptogenesis from gravitational waves in inflation models was proposed in Ref. [17]. In the following sections we expand on our investigation and demonstrations. In sections II and III we start with brief introductions of TG and STG. Readers who are not familiar with these models may learn more details from reviews on these subjects, e.g., Ref. [18].

    We use the convention of most negative signatures for the metric. The spacetime tensor indices are represented in Greek letters μ,ν,σ,...=0,1,2,3, and their spatial components are denoted in Latin letters i,j,k,...=1,2,3. The tensor indices in the local Minkowski spacetime are represented in capital Latin letters A,B,C,...=0,1,2,3, and the corresponding spatial components are denoted in small Latin letters a,b,c,...=1,2,3.

    The TG theory can be considered as a constrained metric-affine theory. It is formulated in a spacetime endowed with a metric gμν and an affine connection ˆΓρμν, which are constrained by the vanishing of curvature and the metric compatibility,

    ˆRρσμνμˆΓρνσνˆΓρμσ+ˆΓρμαˆΓανσˆΓρναˆΓαμσ=0,ˆρgμν=ρgμνˆΓλρμgλνˆΓλρνgμλ=0.

    (8)

    Without curvature in this theory, gravity is described with spacetime torsion. The torsion tensor is defined as the antisymmetric part of the affine connection: Tρμν=2ˆΓρ[μν]. In terms of the language of tetrad and spin connection and the general relations: gμν=ηABeAμeBν and ˆΓμρσ=eμA(ρeAσ+ωABρeBσ), one finds that

    Tρμν=2eρA([μeAν]+ωAB[μeBν]),

    (9)

    and the spin connection under the constraints of Eq. (8) can be expressed as,

    ωABμ=(Λ1)ACμΛCB,

    (10)

    where ΛAB is an element of an arbitrary Lorentz transformation matrix which is position dependent and satisfies the relation ηABΛACΛBD=ηCD for any spacetime point.

    The TG model we are most concerned with is the teleparallel equivalent of general relativity (TEGR), in which the action for gravity is

    Sg=M2p2d4xeT,

    (11)

    where Mp=1/8πG is the reduced Planck mass, e=g is the determinant of the tetrad, and T is the torsion scalar, defined as

    T=TμTμ+14TαβμTαβμ+12TαβμTβαμ,

    (12)

    with Tμ=Tαμα being the torsion vector. This action is diffeomorphism invariant and identical to the Einstein-Hilbert action up to a boundary term,

    Sg=M2p2d4xg[R(e)+2μTμ],

    (13)

    where the curvature scalar R(e) is defined by the Levi-Civita connection and considered to be fully constructed from the metric, and in turn from the tetrad. The covariant derivative μ is also associated with the Levi-Civita connection. The boundary term does not affect the equation of motion, so the TEGR model is equivalent to GR. In addition, it can also be considered as a pure tetrad theory. The spin connection only contributes to the boundary term, so it represents pure gauge in the TEGR action (11), and in practice, we may fix a spin connection (as long as it satisfies Eq. (10)) that does not affect the equation of motion. The simplest choice is to use of the Weitzenböck connection, ωABν=0, which is frequently adopted in the literature. Using the Weitzenböck connection, the torsion two form is simply expressed as

    TAμν=eAρTρμν=μeAννeAμ.

    (14)

    It deserves stressing that in a general TG theory, fixing a spin connection usually means breaking the local Lorentz symmetry, but this is not the case in the TEGR model (11) as the spin connection in this model only contributes the boundary term as shown in its equivalent form, Eq. (13). One can straightforwardly prove that, when taking the Weitzenböck connection, the action (11) is unchanged under the local Lorentz transformation, eAμΛAB(x)eBμ, up to a boundary term. However, for the modified TEGR models, taking the Weitzenböck connection indeed breaks the local Lorentz symmetry; see Refs. [19-22] for recent discussions. To avoid such explicit Lorentz violation, it is better to keep the general form (10) for the spin connection and treat both the tetrad eAμ and the Lorentz matrix element ΛAB in Eq. (10) as fundamental variables.

    Within the framework of TEGR, a gravitational baryogenesis model similar to that in Ref. [9] can be constructed by considering the derivative coupling of the torsion scalar to the baryon current, LinμTJμB. This gives the baryon an effective chemical potential, which is proportional to ˙T, and the standard cosmology is unchanged, as TEGR is equivalent to GR. The key point is that T=R2μTμ differs from R by the term 2μTμ, so that T and its time derivative do not vanish in the radiation dominated epoch. This means the effective chemical potential μb for the baryon is non-vanishing, which is then expected to generate a net baryon number according to Eq. (3).

    In this section we consider the full action containing this derivative coupling,

    S=d4xg(M2p2T+1M2μTJμ)+Sm.

    (15)

    Other matter and non-gravitational interactions, including baryon number non-conserving interactions, are described by Sm. All matter is assumed to couple to the metric (or the tetrad) minimally aside from the introduced derivative coupling. As the derivative coupling depends on the torsion scalar, which accounts for gravity, this model is also classified as a modified TEGR model.

    The equations of motion follow from the variation of the action with respect to eAμ and ΛAB separately:

    (12θM2M2p)Gμν+θM2M2pTgμν2M2M2pSμνσσθ=1M2pTμν,

    (16)

    S[μν]σσθ=0,

    (17)

    where θμJμ, Gμν is the Einstein tensor, Tμν=(2/g)(δSm/δgμν) is the energy-momentum tensor of matter, and Sμρσ=(1/2)TμρσT[ρσ]μ+2gμ[ρTσ] is the so-called superpotential in TG theory and is antisymmetric under the interchange of its last two indices.

    Now, we apply these equations to the spatially-flat Friedmann-Robertson-Walker (FRW) universe with ds2=dt2a2(t)dx2 to obtain the values of T and its time derivative. Given the FRW metric, we can always choose the tetrad as eAμ=diag{1,a,a,a}. Then, the spin connection will be solved through equations (16) and (17). We first consider the case in which the current in the derivative coupling is the baryon current Jμ=JμB, so θ=˙nB+3HnB only depends on time but does not vanish during the baryogenesis process (in the radiation dominated epoch) due to the baryon number violation. With these, equations (16) and (17) require that S[μν]0=0 and Sij0δij, which in turn give the following constraints on the spin connections:

    ωabjδjaδbi=0,ωa0iδai.

    (18)

    The next step is to find the Lorentz matrix elements Λ to satisfy the above constraints according to the expression of the spin connection (10). In the homogeneous and isotropic spacetime, it is natural to consider the homogeneous Λs as the solution to the equations. Indeed, if all the Lorentz matrix elements in Eq. (10) only rely on the time, then the constraints in Eq. (18) are satisfied automatically. These considerations lead to the following result:

    T=R2μTμ=R2(0+3H)(3H1aωa0iδia)2i(1aω0a0δai)=6H2.

    (19)

    This result means that the torsion scalar does not vanish during the radiation dominated epoch, and its time derivative ˙T=12H˙H provides an effective chemical potential for baryons, μb=˙T/M2, which then induces the baryon number asymmetry:

    nBs=152π2˙TgsM2T=90π2H˙HgsM2T=180π2H3gsM2T,

    (20)

    in the last step we have employed the relation ˙H=2H2 in the radiation dominated epoch. From standard big bang cosmology, it is well known that the Hubble rate in the radiation dominated epoch is Hg1/2T2/(3Mp), where g denotes the total degrees of freedom that contributes to the radiation density, which is equal to gs in the very early universe. Hence, the baryon number asymmetry is

    nBs|TD=203π2g1/2T5DM3pM20.5×1054(TDGeV)3(TDM)2.

    (21)

    In the second step of the above equation, we have employed g=106.75 and Mp2.4×1018 GeV. The decoupling temperature of the baryon number non-conserving interaction is approximately TD100 GeV [23] and the cut-off scale M should be no less than this. Considering this, one can evaluate that the produced baryon number asymmetry is at most on the order of 1048, which is too small to be consistent with observations. This disappointing consequence has been also obtained in Ref. [14], in which the authors then turned to the modified TEGR gravity, i.e., the f(T) model, replacing the original derivative coupling with μf(T)JμB. We would like to point out that within the framework of the TEGR model there is another way to circumvent this difficulty: gravitational leptogenesis.

    In the leptogenesis scenario [24-27], there should be BL violating processes at high energy scales, and these can be realized purely by lepton number violations. In our gravitational leptogenesis model, in addition to the BL violating interactions in thermal equilibrium, we identify the current Jμ in the action (15) with JμBL. This means that the torsion scalar couples derivatively to the BL current instead of the baryon current. Through similar calculations, we obtain the thermally produced BL asymmetry:

    nBLs|TD=203π2g1/2T5DM3pM20.5×1054(TDGeV)3(TDM)2,

    (22)

    where TD is the decoupling temperature of the BL violating interaction and can be much higher than the electroweak scale. At a later time, the previously produced nBL/s asymmetry will not be washed out by the electroweak Sphaleron processes, which violate B+L but conserve BL and are in thermal equilibrium when the temperature of the universe is in the range of 100GeV<T<1012 GeV [23]. Furthermore, the electroweak Sphaleron processes will partially convert the BL asymmetry to the baryon number and lepton number asymmetries, respectively [28]:

    nBs=csnBLs,nLs=(cs1)nBLs,

    (23)

    where cs=(8Nf+4)/(22Nf+13) and Nf is the number of generations. In the standard model, Nf=3 and cs0.35, so the final baryon number asymmetry in this model is

    nBs0.18×1054(TDGeV)3(TDM)2.

    (24)

    Numerically, if TD is close to M, the current observational result, nB/s8.7×1011, requires the decoupling temperature TD to be 0.78×1015 GeV. However, TD should be lower than the inflation scale, otherwise the produced asymmetry nBL/s would be diluted during inflation. One may evaluate that the Hubble rate at TD as H(TD)g1/2T2D/(3Mp)1012 GeV, which is lower than the energy scale of the inflation process in the single field inflation models.

    Now we turn to the gravitational leptogenesis model within the framework of STG, which has not previously been discussed. The STG theory can also be considered as a constrained metric-affine theory. It is formulated in a spacetime with a metric gμν and an affine connection Γλμν, with the latter leading to zero curvature and zero torsion,

    ˆRρσμνμˆΓρνσνˆΓρμσ+ˆΓρμαˆΓανσˆΓρναˆΓαμσ=0,Tρμν=2ˆΓρ[μν]=0.

    (25)

    With these constraints, the affine connection can be expressed generally as

    ˆΓλμν(x)=xλyβμνyβ,

    (26)

    with yβ(x) being four functions. These functions define a special coordinate system in which the affine connection vanishes. One may use the y-coordinate system to do the remaining calculations. This is a gauge choice, with this “coincident gauge” being adopted extensively in studies on STG theories in the literature. However, taking the coincident gauge will break the diffeomorphism invariance explicitly, which we try to avoid in this study. Therefore, we will keep the general form (26) for the affine connection. The gravity in the STG theory is identical to the non-metricity. As standard, the non-metricity tensor is defined as

    Qαμνˆαgμν=αgμνˆΓλαμgλνˆΓλανgμλ,

    (27)

    which measures the failure of the affine connection to be metric-compatible. The STG Equivalent of GR (STGR) model has the following action,

    Sg=M2p2d4xgQM2p2d4xg×[14QαμνQαμν12QαμνQμνα14QαQα+12˜QαQα],

    (28)

    where Q is a non-metricity scalar and the vectors Qα,˜Qα are two different traces of the non-metricity tensor, i.e., Qα=gσλQασλ,˜Qα=gσλQσαλ. In terms of the constraints (25), one can easily determine that Q=Rμ(Qμ˜Qμ), so that the STGR action (28) is equal to the Einstein-Hilbert action up to a boundary term,

    Sg=M2p2d4xg[R+μ(Qμ˜Qμ)].

    (29)

    Hence, the STGR model is equivalent to GR.

    By introducing the derivative coupling of the non-metricity scalar, the full action we consider is

    S=d4xg(M2p2Q+1M2μQJμ)+Sm.

    (30)

    The equation of motion via the variation with the metric is

    (12θM2M2p)Gμν+θM2M2pQgμν+1M2M2pPαμναθ=1M2pTμν,

    (31)

    where again we have θμJμ, which is equal to ˙n+3Hn in the FRW universe. The superpotential is defined as Pαμν=Qαμν2Q(μν)α(Qα˜Qα)gμν+gα(μQν), which is symmetric under the interchange of the last two indices. Besides the metric, the four functions yν from which the affine connection is constructed are independent variables in this model. The corresponding equation of motion is

    ˆαˆμ(gPαμνθ)=0.

    (32)

    Similarly, the aim of this section is to determine the value of Q and its time derivative by solving the above equations in the FRW universe. As the space of the FRW universe is homogeneous and isotropic, it has six spatial Killing vectors ξμ. It is reasonable to require the affine connection is also uniformly distributed in the universe, so that its Lie derivatives along the Killing vectors vanish: LξΓλμν=0. With this symmetry requirement, the non-vanished components of the affine connection have the following forms [29]:

    ˆΓ000=K1(t),ˆΓ011=ˆΓ022=ˆΓ033=K2(t),ˆΓ101=ˆΓ202=ˆΓ303=K3(t),

    (33)

    where K1(t),K2(t),K3(t) are three uniform functions. The zero curvature condition requires that

    K3(K1K3)˙K3=0,K1K2+˙K2=0,K2K3=0.

    (34)

    Now, we discuss the cases of K2=0 and K3=0 separately. In the first case, K2=0, one can obtain the non-metricity scalar and the simple form of constraint equation (32) as follows:

    Q=3(2H2+3HK3+˙K3),K3(¨θ+3H˙θ)=0.

    (35)

    As mentioned, θ=˙n+3Hn does not vanish due to the non-conservation of the corresponding quantum number. Its change with time depends on the specific particle physics model and the quantum number violation. So in general, we cannot expect a vanishing ¨θ+3H˙θ. Then, the constraint equation (32) requires that K3=0, and we obtain Q=6H2.

    In the second case K3=0, we get the non-metricity scalar and the form of constraint equation (32) as follows:

    Q=3(2H2+HK2/a2+˙K2/a2),K2(¨θ2K1˙θ+H˙θ)=0.

    (36)

    Similarly we cannot expect ¨θ2K1˙θ+H˙θ=0 in general, and we can only have K2=0. So we again obtain Q=6H2.

    In all, we obtain Q=6H2 for the model (30) within the STG theory framework. Then, we have the same premise regarding the baryon number asymmetry as that we have discussed within the TG theory framework in the previous section. If the current in the derivative coupling (1/M2)μQJμ is the baryon current JμB, the produced baryon number asymmetry is as low as 1048 and is not consistent with current observations. To solve this, one may consider the modified STGR model, such as the f(Q) model, or by changing the coupling μQJμB to μf(Q)JμB. In this study, we prefer the gravitational leptogenesis mechanism, i.e., besides assuming the existence of BL violating processes in the very early universe, the introduced derivative coupling of Q should be (1/M2)μQJμBL. Consequently, we have the BL asymmetry

    nBLs|TD0.5×1054(TDGeV)3(TDM)2,

    (37)

    and a similar order asymmetry for baryons: nB/s0.18×1054(TD/GeV)3(TD/M)2. With the decoupling temperature of BL violating interactions TDM1015 GeV, the required baryon-to-entropy ratio nB/s1010 can be obtained.

    In this study, we investigated the gravitational baryogenesis and leptogenesis models within the frameworks of teleparallel gravity (TG) and symmetric teleparallel gravity (STG). Both the TG and STG theories can include models equivalent to GR, i.e., TEGR and STGR respectively, but account for gravitational phenomena from different viewpoints. By introducing the derivative couplings of the torsion scalar (in TEGR) or the non-metricity scalar (in STGR) to baryons, the baryon number asymmetry can be produced in thermal equilibrium. In the case of baryogenesis, the produced baryon number asymmetry is too small to be consistent with observations. However, as we have shown, the leptogenesis scenario works for these cases.

    In gravitational baryogenesis models, the back reaction of the introduced derivative coupling to the spacetime can be ignored because in the radiation dominated epoch the baryon density is subdominant.

    [1] S. C. Chetal, V. Balasubramaniyan, P. Chellapandi et al., Nucl. Eng. Des. 236, 852 (2006) doi: 10.1016/j.nucengdes.2005.09.025
    [2] E. T. Cheng, A. B. Pashchenko, and J. Kopecky, Fusion Techn. 30, 1182 (1996) doi: 10.13182/FST96-A11963108
    [3] Z. Li, P. Cheng, H. Geng et al., Phys. Rev. ST Accel. Beams 16, 080101 (2013) doi: 10.1103/PhysRevSTAB.16.080101
    [4] E. M. Baum, M. C. Ernesti, H. D. Knox et al., Nuclides and Isotopes. Chart of the Nuclides, 17-th Edition, KAPL, New York, 2010
    [5] H. Jiang, Z. Cui, Y. Hu et al., Chin. Phys. C 44, 114102 (2020) doi: 10.1088/1674-1137/abadf2
    [6] A. Fessler and S. M. Qaim, Radiochimica Acta 84, 1 (1999) doi: 10.1524/ract.1999.84.1.1
    [7] T. Sanami, M. Baba, T. Kawano et al., Jour. of Nucl. Sci. and Techn. 35, 851 (1998) doi: 10.1080/18811248.1998.9733957
    [8] A. D. Majdeddin, V. Semkova, R. Doczi et al., Hungarian report to the I. N. D. C., Austria, No. 031 (1997)
    [9] Yu. M. Gledenov, M. V. Sedysheva, G. Khuukhenkhuu et al., Conf. on Nucl. Data for Sci. and Techn., Trieste, Vol. 1, p. 514 (1997)
    [10] V. V. Ketlerov, A. A. Goverdovskij, V. F. Mitrofanov et al., Vop. At. Nauki i Tekhn., Ser. Yadernye Konstanty, 1, 121 (1996)
    [11] A. A. Goverdovskiy, O. T. Grudzevich, A. V. Zelenetskiy et al., Fiz. -Energ Institut, Obninsk Reports, No. 2242 (1992)
    [12] S. M. Qaim, R. Wolfle, M. M. Rahman et al., Nucl. Sci. Eng. 88, 143 (1984) doi: 10.13182/NSE84-A28398
    [13] R. Woelfle and S. M. Qaim, Radiochimica Acta 27, 65 (1980) doi: 10.1524/ract.1980.27.2.65
    [14] N. I. Molla and S. M. Qaim, Nucl. Phys. A 283, 269 (1977) doi: 10.1016/0375-9474(77)90431-6
    [15] H. Braun and L. Nagy, Radiochimica Acta 10, 1 (1968) doi: 10.1524/ract.1968.10.12.1
    [16] T. Khromyleva, I. Bondarenko, A. Gurbich et al., Nucl. Sci. Eng. 191, 282 (2018) doi: 10.1080/00295639.2018.1463746
    [17] K. Shibata, T. Kawano, T. Nakagawa et al., Nucl. Sci. Technol. 48, 1 (2011) doi: 10.1080/18811248.2011.9711675
    [18] A. J. Koning, D. Rochman, J. Sublet et al., Nuc. Data Sheets 155, 1 (2019) doi: 10.1016/j.nds.2019.01.002
    [19] S. Kunieda, R. C. Haight, T. Kawano et al., Phys. Rev. C 85, 161 (2012)
    [20] C. Tsabaris, E. Wattecamps, G. Rollin et al., Nucl. Sci. Eng. 128, 47 (1998) doi: 10.13182/NSE128-47
    [21] R. C. Haight, F. B. Bateman, S. M. Sterbenz et al., Fusion Eng. Des. 37, 73 (1997) doi: 10.1016/S0920-3796(97)00033-1
    [22] R. C. Haight, D. W. Kneff, B. M. Oliver et al., Nucl. Sci. Eng. 124, 219 (1996) doi: 10.13182/NSE96-A28573
    [23] E. Wattecamps, in Proc. Int. Conf. Nuclear Data for Science and Technology (ed. S. M. Qaim), May 1991, Jülich, Springer-Verlag, Berlin/Heidelberg, p. 310 (1992)
    [24] S. L. Graham, M. Ahmad, S. M. Grimes et al., Nucl. Sci. Eng. 95, 60 (1987) doi: 10.13182/NSE87-A20432
    [25] D. A. Browna, M. B. Chadwickb, R. Capote et al., Nucl. Data Sheets 148, 1 (2018) doi: 10.1016/j.nds.2018.02.001
    [26] A. C. Kahler, R. E. MacFarlane, R. D. Mosteller et al., Nucl. Data Sheets 112, 2997 (2011) doi: 10.1016/j.nds.2011.11.003
    [27] A. Plompen, O. Cabellos, C. Jean et al., Eur. Phys. J. A 56, 181 (2020) doi: 10.1140/epja/s10050-020-00141-9
    [28] S. V. Zabrodskaya, A. V. Ignatyuk, V. N. Koscheev et al., Ser. Nucl. Const. 1–2, 3 (2007)
    [29] CENDL-3.2, http://www.nuclear.csdb.cn/pingjia.html
    [30] NEW VERSION OF NEUTRON EVALUATED DATA LIBRARY BROND-3.1, https://vant.ippe.ru/en/brond-3-1.html
    [31] H. Bai, Z. Wang, L. Zhang et al., Nucl. Instrum. Methods Phys. Res. A 886, 109 (2018) doi: 10.1016/j.nima.2017.12.072
    [32] A. D. Carlson et al., Nucl. Data Sheets 148, 143 (2018) doi: 10.1016/j.nds.2018.02.002
    [33] A. J. Koning, S. Hilaire, and M. C. Duijvestijn, TALYS-1.9, http://www.TALYS.eu/.
    [34] Evaluated Nuclear Data File (ENDF), https://www-nds.iaea.org/exfor/endf.htm
    [35] Experimental Nuclear Reaction Data (EXFOR), https://www.nndc.bnl.gov/exfor
    [36] A. J. Koning and J. P. Delaroche, Nucl. Phys. A 713, 231 (2003) doi: 10.1016/S0375-9474(02)01321-0
    [37] A. J. Koning, S. Hilaire, and S. Goriely, Nucl Phys. 810, 13 (2008) doi: 10.1016/j.nuclphysa.2008.06.005
    [38] C. Kalbach, Phys. Rev. C 71, 034606 (2005) doi: 10.1103/PhysRevC.71.034606
  • [1] S. C. Chetal, V. Balasubramaniyan, P. Chellapandi et al., Nucl. Eng. Des. 236, 852 (2006) doi: 10.1016/j.nucengdes.2005.09.025
    [2] E. T. Cheng, A. B. Pashchenko, and J. Kopecky, Fusion Techn. 30, 1182 (1996) doi: 10.13182/FST96-A11963108
    [3] Z. Li, P. Cheng, H. Geng et al., Phys. Rev. ST Accel. Beams 16, 080101 (2013) doi: 10.1103/PhysRevSTAB.16.080101
    [4] E. M. Baum, M. C. Ernesti, H. D. Knox et al., Nuclides and Isotopes. Chart of the Nuclides, 17-th Edition, KAPL, New York, 2010
    [5] H. Jiang, Z. Cui, Y. Hu et al., Chin. Phys. C 44, 114102 (2020) doi: 10.1088/1674-1137/abadf2
    [6] A. Fessler and S. M. Qaim, Radiochimica Acta 84, 1 (1999) doi: 10.1524/ract.1999.84.1.1
    [7] T. Sanami, M. Baba, T. Kawano et al., Jour. of Nucl. Sci. and Techn. 35, 851 (1998) doi: 10.1080/18811248.1998.9733957
    [8] A. D. Majdeddin, V. Semkova, R. Doczi et al., Hungarian report to the I. N. D. C., Austria, No. 031 (1997)
    [9] Yu. M. Gledenov, M. V. Sedysheva, G. Khuukhenkhuu et al., Conf. on Nucl. Data for Sci. and Techn., Trieste, Vol. 1, p. 514 (1997)
    [10] V. V. Ketlerov, A. A. Goverdovskij, V. F. Mitrofanov et al., Vop. At. Nauki i Tekhn., Ser. Yadernye Konstanty, 1, 121 (1996)
    [11] A. A. Goverdovskiy, O. T. Grudzevich, A. V. Zelenetskiy et al., Fiz. -Energ Institut, Obninsk Reports, No. 2242 (1992)
    [12] S. M. Qaim, R. Wolfle, M. M. Rahman et al., Nucl. Sci. Eng. 88, 143 (1984) doi: 10.13182/NSE84-A28398
    [13] R. Woelfle and S. M. Qaim, Radiochimica Acta 27, 65 (1980) doi: 10.1524/ract.1980.27.2.65
    [14] N. I. Molla and S. M. Qaim, Nucl. Phys. A 283, 269 (1977) doi: 10.1016/0375-9474(77)90431-6
    [15] H. Braun and L. Nagy, Radiochimica Acta 10, 1 (1968) doi: 10.1524/ract.1968.10.12.1
    [16] T. Khromyleva, I. Bondarenko, A. Gurbich et al., Nucl. Sci. Eng. 191, 282 (2018) doi: 10.1080/00295639.2018.1463746
    [17] K. Shibata, T. Kawano, T. Nakagawa et al., Nucl. Sci. Technol. 48, 1 (2011) doi: 10.1080/18811248.2011.9711675
    [18] A. J. Koning, D. Rochman, J. Sublet et al., Nuc. Data Sheets 155, 1 (2019) doi: 10.1016/j.nds.2019.01.002
    [19] S. Kunieda, R. C. Haight, T. Kawano et al., Phys. Rev. C 85, 161 (2012)
    [20] C. Tsabaris, E. Wattecamps, G. Rollin et al., Nucl. Sci. Eng. 128, 47 (1998) doi: 10.13182/NSE128-47
    [21] R. C. Haight, F. B. Bateman, S. M. Sterbenz et al., Fusion Eng. Des. 37, 73 (1997) doi: 10.1016/S0920-3796(97)00033-1
    [22] R. C. Haight, D. W. Kneff, B. M. Oliver et al., Nucl. Sci. Eng. 124, 219 (1996) doi: 10.13182/NSE96-A28573
    [23] E. Wattecamps, in Proc. Int. Conf. Nuclear Data for Science and Technology (ed. S. M. Qaim), May 1991, Jülich, Springer-Verlag, Berlin/Heidelberg, p. 310 (1992)
    [24] S. L. Graham, M. Ahmad, S. M. Grimes et al., Nucl. Sci. Eng. 95, 60 (1987) doi: 10.13182/NSE87-A20432
    [25] D. A. Browna, M. B. Chadwickb, R. Capote et al., Nucl. Data Sheets 148, 1 (2018) doi: 10.1016/j.nds.2018.02.001
    [26] A. C. Kahler, R. E. MacFarlane, R. D. Mosteller et al., Nucl. Data Sheets 112, 2997 (2011) doi: 10.1016/j.nds.2011.11.003
    [27] A. Plompen, O. Cabellos, C. Jean et al., Eur. Phys. J. A 56, 181 (2020) doi: 10.1140/epja/s10050-020-00141-9
    [28] S. V. Zabrodskaya, A. V. Ignatyuk, V. N. Koscheev et al., Ser. Nucl. Const. 1–2, 3 (2007)
    [29] CENDL-3.2, http://www.nuclear.csdb.cn/pingjia.html
    [30] NEW VERSION OF NEUTRON EVALUATED DATA LIBRARY BROND-3.1, https://vant.ippe.ru/en/brond-3-1.html
    [31] H. Bai, Z. Wang, L. Zhang et al., Nucl. Instrum. Methods Phys. Res. A 886, 109 (2018) doi: 10.1016/j.nima.2017.12.072
    [32] A. D. Carlson et al., Nucl. Data Sheets 148, 143 (2018) doi: 10.1016/j.nds.2018.02.002
    [33] A. J. Koning, S. Hilaire, and M. C. Duijvestijn, TALYS-1.9, http://www.TALYS.eu/.
    [34] Evaluated Nuclear Data File (ENDF), https://www-nds.iaea.org/exfor/endf.htm
    [35] Experimental Nuclear Reaction Data (EXFOR), https://www.nndc.bnl.gov/exfor
    [36] A. J. Koning and J. P. Delaroche, Nucl. Phys. A 713, 231 (2003) doi: 10.1016/S0375-9474(02)01321-0
    [37] A. J. Koning, S. Hilaire, and S. Goriely, Nucl Phys. 810, 13 (2008) doi: 10.1016/j.nuclphysa.2008.06.005
    [38] C. Kalbach, Phys. Rev. C 71, 034606 (2005) doi: 10.1103/PhysRevC.71.034606
  • 加载中

Cited by

1. Kumar, J., Paul, S., Maurya, S.K. et al. A generalized solution for anisotropic compact star model in F(Q) gravity[J]. Physics of the Dark Universe, 2025. doi: 10.1016/j.dark.2024.101764
2. Kaur, S., Maurya, S.K., Shukla, S. Anisotropic fluid solution in f (Q) gravity satisfying vanishing complexity factor[J]. Physica Scripta, 2023, 98(10): 105304. doi: 10.1088/1402-4896/acf348
3. Jasim, M.K., Maurya, S.K., Khalid Jassim, A. et al. Minimally deformed anisotropic solution generated by vanishing complexity factor condition in f(Q)-gravity theory[J]. Physica Scripta, 2023, 98(4): 045305. doi: 10.1088/1402-4896/acbfeb

Figures(8) / Tables(8)

Get Citation
Haoyu Jiang, Zengqi Cui, Yiwei Hu, Jie Liu, Haofan Bai, Jinxiang Chen, Guohui Zhang, Yu. M. Gledenov, E. Sansarbayar, G. Khuukhenkhuu, L. Krupa, I. Chuprakov, Xichao Ruan, Hanxiong Huang, Jie Ren and Qiwen Fan. Cross-section measurements for the 58,60,61Ni( n, α) 55,57,58Fe reactions at 8.50, 9.50 and 10.50 MeV neutron energies[J]. Chinese Physics C. doi: 10.1088/1674-1137/ac3412
Haoyu Jiang, Zengqi Cui, Yiwei Hu, Jie Liu, Haofan Bai, Jinxiang Chen, Guohui Zhang, Yu. M. Gledenov, E. Sansarbayar, G. Khuukhenkhuu, L. Krupa, I. Chuprakov, Xichao Ruan, Hanxiong Huang, Jie Ren and Qiwen Fan. Cross-section measurements for the 58,60,61Ni( n, α) 55,57,58Fe reactions at 8.50, 9.50 and 10.50 MeV neutron energies[J]. Chinese Physics C.  doi: 10.1088/1674-1137/ac3412 shu
Milestone
Received: 2021-09-27
Article Metric

Article Views(2019)
PDF Downloads(67)
Cited by(3)
Policy on re-use
To reuse of subscription content published by CPC, the users need to request permission from CPC, unless the content was published under an Open Access license which automatically permits that type of reuse.
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Email This Article

Title:
Email:

Cross-section measurements for the 58,60,61Ni(n, α)55,57,58Fe reactions at 8.50, 9.50 and 10.50 MeV neutron energies

    Corresponding author: Guohui Zhang, guohuizhang@pku.edu.cn
  • 1. State Key Laboratory of Nuclear Physics and Technology, Institute of Heavy Ion Physics, School of Physics, Peking University, Beijing 100871, China
  • 2. Frank Laboratory of Neutron Physics, Joint Institute for Nuclear Research, Dubna 141980, Russia
  • 3. Nuclear Research Centre, National University of Mongolia, Ulaanbaatar, Mongolia
  • 4. Flerov Laboratory of Nuclear Reactions, Joint Institute for Nuclear Research, Dubna 141980, Russia
  • 5. Institute of Experimental and Applied Physics, Czech Technical University in Prague, Horska 3a/22, Prague 2 12800, Czech Republic
  • 6. L. N. Gumilyov Eurasian National University, Nur-sultan 010000, Kazakhstan
  • 7. The Institute of Nuclear Physics, Ministry of Energy of the Republic of Kazakhstan, Almaty 050032, Kazakhsta
  • 8. China Institute of Atomic Energy, Beijing 102413, China

Abstract: Cross sections of the 58,60,61Ni(n, α)55,57,58Fe reactions were measured at 8.50, 9.50 and 10.50 MeV neutron energies based on the HI-13 tandem accelerator of China Institute of Atomic Energy (CIAE) with enriched 58Ni, 60Ni, and 61Ni foil samples with backings. A twin gridded ionization chamber (GIC) was used as the charged particle detector, and an EJ-309 liquid scintillator was used to obtain the neutron energy spectra. The relative and absolute neutron fluxes were determined via three highly enriched 238U3O8 samples inside the GIC. The uncertainty of the present data of the 58Ni(n, α)55Fe reaction is smaller than most existing measurements. The present data of 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reactions are the first measurement results above 8 MeV. The present experimental data could be reasonably reproduced by calculations with TALYS-1.9 by adjusting several default values of theoretical model parameters.

    HTML

    I.   INTRODUCTION
    • Nickel plays an important role in the field of nuclear engineering [1-3]. In the MeV neutron energy region, the radiation damage of structural materials due to the α-particles emitted from the (n, α) reactions are non-negligible. The abundances of 58Ni, 60Ni and 61Ni in natural nickel are 68.077%, 26.223% and 1.140%, respectively [4]. Therefore, accurate cross sections of the 58,60,61Ni(n, α)55,57,58Fe reactions in the MeV energy region are important to the safety design for nuclear engineering. Also, measurements of the cross sections of the 58,60,61Ni(n, α)55,57,58Fe reactions can enhance the understanding of nuclear structure and nuclear reaction mechanisms.

      For the cross sections of the 58Ni(n, α)55Fe (Q = 2.899 MeV) reaction, the residual nucleus 55Fe is radioactive so the cross section can be measured using the activation method. In the MeV neutron energy region, there are eleven existing measurement results, but most of them were obtained in the region below 8 MeV or around 14 MeV [5-15]. In the 8.50–10.50 MeV region, there are only two measurement data [6, 12] for the 58Ni(n, α)55Fe reaction, and the cross sections of Qaim [12] are higher by ~18% than those of Fessler [6], showing apparent disagreement.

      For the cross sections of the 60Ni(n, α)57Fe (Q = 1.355 MeV) and 61Ni(n, α)58Fe (Q = 3.580 MeV) reactions, measurement data are scarce because the activation method is unavailable. In the MeV energy region, for the 60Ni(n, α)57Fe reaction, there are only two measurement data [5, 16], both measured below 8 MeV. For the 61Ni(n, α)58Fe reaction, there is only one measurement datum obtained by our group [5], and this datum was measured below 6 MeV. So, there is no data for the 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reactions in the 8.50–10.50 MeV region.

      Thus, experimental data are scarce for all the 58,60,61Ni(n, α)55,57,58Fe reactions in the 8.50–10.50 MeV region. According to the JENDL-4.0 [17] and TENDL-2019 [18] libraries, the cross sections of the 58,60,61Ni(n, α)55,57,58Fe reactions are nearly equal to those of the 58,60,61Ni(n, ) reactions in the energy region below 11 MeV. So, the cross sections of the (n, ) reactions can be used as the reference for the analysis of the excitation functions of the (n, α) reactions below 11 MeV. In the 8.50–10.50 MeV energy region, there are six measurement data [19-24] for the 58Ni(n, ) reaction, four measurement data [21-24] for the 60Ni(n, ) reaction, and one measurement datum [22] for the 61Ni(n, ) reaction. Among these data, discrepancies are apparent. Furthermore, most of the experimental data for the 58,60,61Ni(n, ) reactions are characterized by large uncertainties (>10%).

      Because the existing measurement data are either scarce or show discrepancies, deviations among different evaluation libraries are noticeable. For example, in the 8.50–10.50 MeV region, the amount of variation among different evaluation libraries [17, 18, 25-30], are 10.5%, 19.9% and 23.9% for the 58Ni(n, α)55Fe, 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reactions, respectively.

      In our previous work, the cross sections of the 58,60,61Ni(n, α)55,57,58Fe reactions in the 4.50En5.50 MeV region were measured based on the 4.5 MV Van de Graaff accelerator at Peking University (PKU) [5]. In the present work, measurements were extended to the 8.50 En10.50 MeV region on the HI-13 tandem accelerator at China Institute of Atomic Energy (CIAE). Details of the experiment are described in section II, data processing and results are presented in section III, theoretical analysis is described in section IV, and the conclusions are drawn in section V.

    II.   DETAILS OF EXPERIMENTS
    • The experiments were performed using the HI-13 tandem accelerator of the China Institute of Atomic Energy (CIAE). The experimental apparatus is shown in Fig. 1, which was similar to the previous measurement [5], except that the scintillator detector was further away from the neutron source because of the strong γ-background from the source.

      Figure 1.  (color online) Schematic drawing of the experimental apparatus.

      Through the 2H(d, n)3He reaction, neutrons with kinetic energies of 8.50, 9.50, and 10.50 MeV were generated using the energetic deuteron beam from the accelerator to bombard the deuterium gas target with incident energies of 5.95, 6.91, and 7.90 MeV. The deuteron beam current was about 2.0 μA throughout the measurement. The deuterium gas target was 3.0 cm in length, 1.0 cm in diameter and 3.0 atm in pressure. It was separated from the vacuum tube of the accelerator by a molybdenum foil 10.0 μm in thickness. In addition to the main neutrons, low-energy neutrons will also be generated through the 2H(d, np)2H reaction. To correct for this interference, the neutron energy spectra were measured using an EJ-309 liquid scintillator detector. Unfolding the pulse height spectra measured with the scintillator detector [31], the neutron energy spectra were obtained as shown in Fig. 2.

      Figure 2.  (color online) Measured neutron energy spectra from the EJ-309 liquid scintillator detector.

      A twin gridded ionization chamber (GIC) with sides labelled '01 side' and '02 side' was used as the charged particle detector. The working gas of the GIC was Xe + 5.0% H2 with the pressure of 1.161 atm. The high voltages applied on the cathodes and anodes were -1500 V and 750 V, and the grid electrodes were grounded. There is a sample changer at the common cathode of the GIC with five sample positions, on which these samples were mounted: 1) back-to-back compound α-sources for the detection system calibration, 2) 238U3O8#III/Ta samples for the absolute neutron flux measurement, 3) back-to-back 58Ni#II/58Ni#I foil samples for the measurement of the 58Ni(n, α)55Fe reaction (the 58Ni#II sample was broken, so the cross section was obtained using the 58Ni#I sample), 4) back-to-back 60Ni/61Ni foil samples for the measurements of the 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reactions, and 5) Ta/Ta backings for the measurement of the background events. To monitor the relative neutron flux, the 238U3O8#I and 238U3O8#II samples were glued on each of the fission cathode electrodes of the GIC, respectively. Details of the GIC, including the measurement principle, structure and electronics, together with information of the samples, including the isotopic enrichment, thickness, diameter, unevenness and impurity content can be found in Ref. [5].

      The 58,60,61Ni(n, α)55,57,58Fe reactions were measured at En = 8.50, 9.50 and 10.50 MeV. For each neutron energy point, measurements were performed with the sample set sequentially at the five sample positions, then the GIC was rotated by 180° and measurements were performed for each sample position again, so that the forward (0°–90°) and the backward (90°–180°) cross sections of the (n, α) reaction could be measured. The experiment was conducted in October 2020, and the total neutron beam duration was ~ 30 h.

    III.   DATA PROCESSING AND RESULTS

      A.   Data processing

    • The data processing procedures are presented in Fig. 3, which were the same as the previous measurements in the 4.50–5.50 MeV range [5] except that the unfolding method for reducing the uncertainty of En was not used in the present work. The uncertainty of the En at CIAE was small enough (<1%) due to the higher energy of the incident deuteron beam, therefore the unfolding method is not necessary.

      Figure 3.  Data processing procedure.

      The forward or backward cross sections σα of the (n, α) reactions can be calculated by

      σα =σfNUNαεfρlow(1kimpurity)NNiNfεαGCf_fission,

      (1)

      where σf is the standard cross section of the 238U(n, f) reaction [32]. NU and NNi are the numbers of the 238U and 58,60,61Ni nuclei in the samples, respectively. Nα is the count of the net α events within the threshold in the anode projection spectrum after the valid-event-area determination in the cathode–anode two-dimensional spectrum and background subtraction. A typical cathode–anode two-dimensional spectrum and an anode projection spectrum are shown in Fig. 4 (a) and (b), respectively. Nf is the count of the fission events within the threshold in the fission anode projection spectrum as shown in Fig. 5. εα and εf are the detection efficiencies for the α particles from the 58,60,61Ni(n, α)55,57,58Fe reactions and fission fragments from the 238U(n, f) reaction at the specific neutron energy, respectively. ρlow is the correction coefficient of the α and fission events induced by the low-energy neutrons. kimpurity is the correction coefficient of the α events from the nickel and oxygen impurity isotopes. G is the ratio of the average neutron flux in the area of nickel sample to that of the 238U3O8#III sample area at the sample changer. Cf_fission is the ratio of the fission counts from the 238U3O8#I or 238U3O8#II sample glued on the fission cathode during the foreground measurement to that during the absolute neutron flux measurement. εα, εf, ρlow, kimpurity and G were obtained using simulation methods; whose details can be found in Ref. [5]. Values of the parameters in Eq. (1) for the three reactions are presented in Tables 13.

      ParametersValues
      En = 8.50 MeVEn = 9.50 MeVEn = 10.50 MeV
      NU1.872×1019 a, b1.872×1019 a, b1.872×1019 a, b
      NNi9.526×1019 a, b9.526×1019 a, b9.526×1019 a, b
      σf /mb1015.0 a, b1013.5 a, b1009.5 a, b
      Nα6890 a5976 b7260 a6927 b11335 a7867 b
      Nf16366 a14085 b17128 a16801 b20307 a19284 b
      ρlow1.031 a1.026 b1.097 a1.030 b1.154 a1.150 b
      εf (%)82.62 a84.23 b81.18 a83.26 b81.40 a83.25 b
      εα (%)93.97 a91.49 b87.05 a94.06 b89.05 a93.02 b
      G1.007 a1.002 b1.010 a0.999 b1.007 a1.000 b
      kimpurity0.007 a0.006 b0.021 a0.008 b0.096 a0.007 b
      a For the forward cross section. b For the backward cross section.

      Table 2.  Values of the parameters in Eq. (1) for the 60Ni(n, α)57Fe reaction.

      Figure 4.  (color online) (a) Cathode–anode two-dimensional spectrum of the foreground events (the area between the red line is the valid-event-area of α particles determined using the two-dimensional spectrum obtained from the α sources). (b) Anode projection spectrum of the foreground, background and net events together with the simulated spectrum for the measurement of the 58Ni(n, α)55Fe reaction in the forward direction at En = 10.50 MeV.

      Figure 5.  (color onlne) Anode spectrum of the fission fragments for the absolute neutron flux measurement at En = 10.50 MeV.

      ParametersValues
      En = 8.50 MeVEn = 9.50 MeVEn = 10.50 MeV
      NU1.872×1019 a, b1.872×1019 a, b1.872×1019 a, b
      NNi1.039×1020 a b9.526×1019 a b4.883×1019 a b
      σf /mb1015.0 a b1013.5 a b1009.5 a b
      Nα4764 a4768 b4570 a6221 b9592 a8401 b
      Nf14085 a16366 b16801 a17128 b19284 a20307 b
      ρlow1.009 a1.008 b1.058 a1.048 b1.010 a1.080 b
      εf (%)84.23 a82.62 b83.26 a81.18 b83.25 a81.40 b
      εα (%)90.01 a92.81 b89.45 a93.48 b92.30 a88.02 b
      G1.012 a1.006 b1.008 a1.006 b1.004 a1.009 b
      kimpurity~ 0 a, b~ 0 a, b~ 0 a, b
      a For the forward cross section. b For the backward cross section.

      Table 1.  Values of the parameters in Eq. (1) for the 58Ni(n, α)55Fe reaction.

      ParametersValues
      En = 8.50 MeVEn = 9.50 MeVEn = 10.50 MeV
      NU1.872×1019 a, b1.872×1019 a, b1.872×1019 a, b
      NNi4.883×1019 a, b4.883×1019 a, b4.883×1019 a, b
      σf /mb1015.0 a, b1013.5 a, b1009.5 a, b
      Nα2155 a1937 b2556 a2191 b3343 a3091 b
      Nf14085 a16366 b16801 a17128 b19284 a20307 b
      ρlow1.024 a1.023 b1.031 a1.075 b1.139 a1.124 b
      εf (%)84.23 a82.62 b83.26 a81.18 b83.25 a81.40 b
      εα (%)94.88 a92.60 b91.18 a90.35 b93.02 a85.58 b
      G1.010 a0.994 b0.999 a1.002 b1.007 a1.005 b
      kimpurity0.1724 a0.1768 b0.1607 a0.1600 b0.1573 a0.1445 b
      a For the forward cross section. b For the backward cross section.

      Table 3.  Values of the parameters in Eq. (1) for the 61Ni(n, α)58Fe reaction.

    • B.   Measured results

    • Using Eq. (4), the forward and backward cross sections of the 58Ni(n, α)55Fe, 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reactions can be calculated. Total (n, α) cross section can be acquired by adding the forward cross section and backward one together. Sources of uncertainty and their magnitudes are presented in Table 4. The measured cross sections of the three reactions are shown in Tables 57 and Fig. 6 (a) – (c).

      SourceMagnitude (%)
      58Ni(n, α)55Fe60Ni(n, α)57Fe61Ni(n, α)58Fe
      NU1.0 a, b1.0 a, b1.0 a, b
      NNi1.0 a, b5.0 a, b1.0 a, b
      σf1.4 – 1.5 a, b1.4 – 1.5 a, b1.4 – 1.5 a, b
      Nα2.1 – 3.8 a, 2.7 – 6.5 b2.8 – 5.3 a, 4.6 – 8.2 b6.4 – 8.4 a, 10.2 – 13.9 b
      Nf0.7 – 0.9 a, b0.7 – 0.9 a, b0.7 – 0.9 a, b
      ρlow2.7 – 3.3 a, 2.8 – 4.6 b4.4 – 4.6 a, 4.0 – 4.6 b4.4 – 5.3 a, 4.0 – 4.8 b
      εf2.0 – 2.5 a, b2.0 – 2.5 a, b2.0 – 2.5 a, b
      εα2.6 – 3.5 a, 2.2 – 4.0 b3.0 – 6.5 a, 3.0 – 4.5 b1.7 – 2.9 a, 2.5 – 4.8 b
      G0.5 – 1.3 a, b0.5 – 1.6 a, b0.5 – 1.9 a, b
      kimpurity~ 00.1 – 4.6 a, < 0.2 b3.0 – 3.1 a, 3.1 – 3.2 b
      σα6.0 – 6.5 a, 5.6 – 9.5 b5.9 – 7.8 c7.2 – 10.7 a, 8.3 – 10.3 b7.7 – 9.6 c9.9 – 10.7 a, 12.3 – 16.0 b11.0 – 12.9 c
      En0.7 – 0.9 a, b0.7 – 0.9 a, b0.7 – 0.9 a, b
      a For the forward cross section. b For the backward cross section. c For the total (n, α) cross section.

      Table 4.  Sources of uncertainty and their magnitudes.

      En/MeVCross section/mbForward/backward ratio
      MeasurementCalculationMeasurementCalculation
      8.50 ± 0.0773.4 ± 4.376.41.22 ± 0.101.08
      9.50 ± 0.0872.6 ± 4.476.31.10 ± 0.091.17
      10.50 ± 0.0875.7 ± 5.977.01.16 ± 0.131.25

      Table 5.  Measured 58Ni(n, α)55Fe cross sections and forward/backward ratios in the laboratory reference system (results obtained from the 58Ni#I sample) compared with TALYS-1.9 [33] calculations using the adjusted input parameters.

      Figure 6.  (color online) Present (a) 58Ni(n, α)55Fe, (b) 60Ni(n, α)57Fe, and (c) 61Ni(n, α)58Fe cross sections compared with evaluations [34], TALYS-1.9 [33] calculations and existing measurements [35].

      En/MeVCross section/mbForward/backward ratio
      MeasurementCalculationMeasurementCalculation
      8.50 ± 0.0732.6 ± 2.533.71.06 ± 0.111.08
      9.50 ± 0.0833.2 ± 3.034.51.18 ± 0.161.10
      10.50 ± 0.0833.3 ± 3.234.81.27 ± 0.181.15

      Table 6.  Measured 60Ni(n, α)57Fe cross sections and forward/backward ratios in the laboratory reference system compared with TALYS-1.9 [33] calculations using the adjusted input parameters.

      En/MeVCross section/mbForward/backward ratio
      MeasurementCalculationMeasurementCalculation
      8.50 ± 0.0716.5 ± 2.018.11.18 ± 0.211.10
      9.50 ± 0.0718.0 ± 2.018.51.12 ± 0.181.16
      10.50 ± 0.0819.7 ± 2.519.6 1.09 ± 0.211.26

      Table 7.  Measured 61Ni(n, α)58Fe cross sections and forward/backward ratios in the laboratory reference system compared with TALYS-1.9 [33] calculations using the adjusted input parameters.

    • C.   Comparison of the present results with evaluations and existing measurements

    • As Fig. 6 (a) – (c) show, the present reaction cross sections were compared with the data from different evaluations [17, 18, 25-30], as well as existing measurement data of the 58,60,61Ni(n, α)55,57,58Fe reactions and the 58,60,61Ni(n, ) reactions below 11 MeV [5-16, 19-24]. The three reactions are discussed as following:

      1) For the 58Ni(n, α)55Fe reaction, compared with different evaluations, the present results are lower by 8.1%–28.8% [17, 18, 25-30]. In the 8.50En10.50 MeV region, the present results show that the change of cross section of the 58Ni(n, α)55Fe reaction with the increasing of En is quite small, and the ENDF/B-VII.1 [26] evaluation shows the similar trend.

      Compared with existing measurements, the present cross sections in the 8.50En10.50 MeV region are close to the experimental data of S. L. Graham (1987, 5.0–14.0 MeV, (n, )) [24] and E. Wattecamps (1992, 8.0–11.0 MeV, (n, )) [23] within the error range. However, the present cross sections are 42.6% and 25.9% lower than the results of S. M. Qaim (1984, 5.36–9.49 MeV) [12] and Fessler (1999, 5.36–19.4 MeV) [6]. Both Qaim [12] and Fessler [6] used the activation method, but their measurement data were characterized by large uncertainties due to the small counts and low energy of the X-ray (Kα = 5.89 keV and Kβ = 6.49 keV) of the activation product 55Fe (T1/2 = 2.7 yr). Average uncertainties of the results of Qaim [12] and Fessler [6] are 15.8 % and 11.7 %, respectively, while that of the present result is 6.6%.

      Also, the present cross sections are 19.9%, 20.2%, and 20.6% lower than the experimental data of R. C. Haight (1996, 9.85 MeV, (n, )) [22], R. C. Haight (1997, 2.0–48.5 MeV, (n, )) [21], and S. Kunieda (2012, 1.7–25.0 MeV, (n, )) [19], respectively. The three existing measurements were all carried out using the α-particle counting method. Average uncertainties of these three results are 16.3%, 8.5%, and 14.6% which are greater than that of the present results.

      2) For the 60Ni(n, α)57Fe reaction, compared with different evaluations [17, 18, 25-30], the present results are lower by 3.8%–46.4%. The present results show that the cross section of the 60Ni(n, α)57Fe reaction are nearly invariable with the increasing of En in the 8.50En10.50 MeV region, which is similar to the trend of the ENDF/B-VII.1 [26] evaluation. As shown in Fig. 6 (b), the present results in the 8.50En10.50 MeV region together with the data measured by our group in the 5.00En5.50 MeV region [5] agree well with the evaluation results of the ENDF/B-VII.1 [26] library multiplied by the coefficient of 0.54.

      Compared with existing measurements, the present cross sections in the 8.50En10.50 MeV region agree well with the measurement data of S. L. Graham (1987, 5.0–14.0 MeV, (n, )) [24] and E. Wattecamps (1992, 8.0–11.0 MeV, (n, )) [23] within the error range.

      R. C. Haight has measured the cross sections of the 60Ni(n, α)57Fe reaction in 1996 [22] and 1997 [21], yet his data in 1997 are ~20% lower than those in 1996. Compared with the data of his two measurements, the present cross sections are close to his data measured in 1997.

      3) For the 61Ni(n, α)58Fe reaction, compared with different evaluations, the present results are lower by 17.5%–60.1% [17, 18, 25-30]. In the 8.50En10.50 MeV region, the present results show that the cross section of the 61Ni(n, α)58Fe reaction grows gradually with the increasing of En, which is similar to the trend of evaluations of the CENDL-3.2 [29] and BROND-3.1 [30] libraries. The present cross sections are lower by 13.9% than the experimental data of R. C. Haight (1996, 9.85 MeV, (n, )) [22].

    IV.   THEORETICAL ANALYSIS
    • The cross sections of the 58Ni(n, α)55Fe, 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reactions were calculated using TALYS-1.9 [33]. To better agree with the present results, several input parameters of the TALYS-1.9 [33] including those of the optic model, the level density and stripping parameters were adjusted from the default values as Table 8 lists. As shown in Tables 5-7 and Fig. 6 (a) – (c), the calculated cross sections and forward/backward ratios agree well with both the present measurement results and the results measured by our group in the 4.50 En5.50 MeV region [5]. Using the adjusted parameters, the calculated cross sections of other major reaction channels, including the (n, tot), (n, el), (n, inl), (n, p) reactions and the angular distributions of elastic scattering, also agree well with the results of most evaluations [34] and existing measurements [35] (three calculations are shown in Fig. 7 as examples), by which the reliability of the adjusted parameters was verified.

      58Ni(n, α)55Fe60Ni(n, α)57Fe
      KeywordParametersKeywordParameters
      Tljadjust aa 2.00 0rvadjust ba 1.08
      Tljadjust aa 2.00 1rvadjust bp 1.57
      Tljadjust ba 1.75 2aadjust d28 60 1.22
      rvadjust ba 0.95aadjust d27 60 1.06
      rwadjust ca 1.62Tadjust e28 59 0.80
      rvadjust bp 1.05cstrip fa 0.90
      aadjust d26 55 0.8561Ni(n, α)58Fe
      Tadjust e26 55 0.92KeywordParameters
      cstrip fa 0.80rvadjust ba 1.01
      maxlevelsbin ga 8rvadjust bp 1.12
      aadjust d26 58 1.04
      Tadjust e26 58 0.93
      a Tljadjust: Multiplier to adjust the optical model potentials (OMP) transmission coefficient per l-value [33], whose parameters are “particle symbol, value and l-value”. b rvadjust: Multiplier to adjust the OMP parameter rv of Eq. (3) of Ref. [36]. whose parameters are “particle symbol and value”. c rwadjust: Multiplier to adjust the OMP parameter rw of Eq. (3) of Ref. [36], whose parameters are “particle symbol and value”. d aadjust: Multiplier to adjust the level density parameter a of Eq. (15) of Ref. [37], whose parameters are “atomic number, atomic mass number, and value”. e Tadjust: Normalization factor for the nuclear temperature T of Eq. (53) of Ref. [37], whose parameters are “atomic number, atomic mass number, and value”. f cstrip: Adjustable parameter for the stripping or pick-up process to scale the complex-particle pre-equilibrium cross section per outgoing particle [38], whose parameters are “particle symbol and value”. g maxlevelsbin: The number of included discrete levels for the nuclides resulting from binary emission that is considered in Hauser-Feshbach decay and the gamma-ray cascade [33], whose parameters are “particle symbol and value”.

      Table 8.  Adjusted input parameters for TALYS-1.9 [33].

      Figure 7.  (color online) TALYS-1.9 [33] calculations using the default and adjusted parameters for the (a) 58Ni(n, tot), (b) 60Ni(n, p)60Co, and (c) 61Ni(n, el) reactions compared with evaluations [34] and existing measurements [35].

      For En < 10.5 MeV, if the angular distributions of emitted α particles of the 58,60,61Ni(n, α)55,57,58Fe reactions are isotropic in the center-of-mass system, the forward/backward ratio in the laboratory system should be less than 1.07. However, as shown in Tables 57, the forward/backward ratios of the 58,60,61Ni(n, α)55,57,58Fe reactions are generally larger than 1.07, which reveals the non-statistical mechanisms of the 58,60,61Ni(n, α)55,57,58Fe reactions in the 8.50–10.50 MeV region. Using TALYS-1.9 [33] with the adjusted parameters, the ratio of the direct, pre-equilibrium and compound components of the (n, α) cross sections were calculated as shown in Fig. 8.

      Figure 8.  (color online) Ratios of the direct, pre-equilibrium and compound components of the (a) 58Ni(n, α)55Fe, (b) 60Ni(n, α)57Fe, and (c) 61Ni(n, α)58Fe cross sections calculated using the TALYS-1.9 [33] code with the adjusted parameters.

      As shown in Fig. 8, for the 58,60,61Ni(n, α)55,57,58Fe reactions, the compound mechanism predominates below ~ 9 MeV, and then the pre-equilibrium mechanism gradually comes into effect. The direct mechanism has little effect on the final cross sections for all three reactions below 20 MeV.

    V.   CONCLUSIONS
    • In the present work, the cross sections of the 58Ni(n, α)55Fe, 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reactions have been measured at 8.50, 9.5 and 10.50 MeV neutron energies using the HI-13 tandem accelerator, a GIC charged-particle detector, enriched nickel isotopic foil samples, 238U3O8 samples, and EJ-309 liquid scintillator detector. In the 8.50En10.50 MeV region, the present cross sections of the 58,60,61Ni(n, α)55,57,58Fe reactions change slowly with the increase of En, and they are generally ~20% lower than most existing measurement data and evaluations. By adjusting several default values of theoretical model parameters, the present experimental data could be reasonably reproduced by calculations with TALYS-1.9. Analyses show that for the three reactions the compound mechanism predominates below ~ 9 MeV, and then as the neutron energy increases the pre-equilibrium mechanism gradually comes into effect. The uncertainty of the present cross sections of the 58Ni(n, α)55Fe reaction is smaller than most of the data of existing measurements. The present cross sections of the 60Ni(n, α)57Fe and 61Ni(n, α)58Fe reaction are the first measurement results above 8.0 MeV region. Considering the difference and scarcity of existing measurements and the discrepancy among evaluations, the present results are useful in the clarification of the deviations and can provide valuable information for future evaluations.

    ACKNOWLEDGEMENTS
    • The authors are indebted to the operating crew of Beijing HI-13 Tandem Accelerator of China Institute of Atomic Energy. Prof. Zhenpeng Chen from Tsinghua University and Prof. Vlad Avrigeanu from Bucharest University are acknowledged for their helpful suggestions.

Reference (38)

目录

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return