This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Bottomonia in quark-antiquark confining potential

Ritu Garg, K.K Vishwakarma , Alka Upadhyay [email protected]
(School of Physics and Materials Science, Thapar Institute of Engineering and Technology, Patiala-147004, INDIA)
Abstract

In this paper, we comprehensively explore bottomonia mass spectra and their decay properties by solving the non-relativistic Schrodinger wave equation numerically with approximate quark-antiquark potential form. We also incorporate spin-dependent terms - spin-spin, spin-orbit, and tensor terms to remove mass degeneracy and to obtain excited states (nS,nP,nD,nF,n=1,2,3,4,5nS,nP,nD,nF,n=1,2,3,4,5) mass spectra. By using Van Royen - Weisskopf formula, we investigate leptonic decay constants, di-leptonic, di-gamma, tri-gamma, di-gluon decay widths and also incorporate first-order radiative corrections. We also computed radiative transition widths, which give a better insight into the non-perturbative aspects of QCD. The present results for mass spectroscopy and decay properties are in tune with available experimental values and other theoretical predictions. Our results may provide better insight to upcoming experimental information in the near future.

1 Introduction

Different experimental facilities like LHCb, COMPASS, BESIII, STAR, CMS, ATLAS, BABAR, etc., are continuously trying to produce enormous data in the field of heavy-flavor hadrons. The theoretical studies are also heavily invested in the field to interpret the experimental data and predict the properties of hadrons. The heavy quark hadrons like charmonium and bottomonium have also indeed received remarkable experimental growth. Through the efforts of various experimental collaborations and facilities, significant progress has been made in establishing bottomonium spectra. In this paper, we are focussing on the spectra and properties of bottomonium. The states Υ(1S)\Upsilon(1S), Υ(2S)\Upsilon(2S) and Υ(3S)\Upsilon(3S) were observed in 1977 by the E288 collaboration at the Fermi National Accelerator Laboratory (FNAL) [1, 2] and notated as Υ,Υ,Υ′′\Upsilon,\Upsilon^{{}^{\prime}},\Upsilon^{{}^{\prime\prime}} states respectively. After the early detection of systems composing of bb¯b\bar{b}, many of the properties and new states were discovered as given in PDG[3]. During the year 1982, the first excited state χbJ(2P)\chi_{bJ}(2P) were identified in E1 transition from Υ′′\Upsilon^{{}^{\prime\prime}} state [4, 5]. Subsequently, in 1983, an exciting breakthrough came with the discovery of χbJ(2P)\chi_{bJ}(2P) with J = 0, 1, 2 in the E1 transition from Υ\Upsilon^{{}^{\prime}} state [6, 7]. In the following year, 1984, Υ(4S),Υ(5S),Υ(6S)\Upsilon(4S),\Upsilon(5S),\Upsilon(6S) are announced in e+ee^{+}e^{-} cross section above BB¯B\bar{B} threshold [8, 9]. These findings opened the search window to investigate higher excited states of bottomonia spectra. After several experimental attempts, in 2008, the BABAR collaboration [10] announced the observation of ηb(1S)\eta_{b}(1S), spin partner of Υ(1S)\Upsilon(1S). In 2010, BABAR collaboration observed Υ(13DJ)\Upsilon(1^{3}D_{J}) with J=2J=2 in decay chain Υ(3S)γγΥ(13DJ)γγΥ(1S)\Upsilon(3S)\rightarrow\gamma\gamma\Upsilon(1^{3}D_{J})\rightarrow\gamma\gamma\Upsilon(1S) [11]. In 2011, the BABAR collaboration announced hb(1P)h_{b}(1P) observed via decay process Υ(3S)π0hb(1P)\Upsilon(3S)\rightarrow\pi^{0}h_{b}(1P) [12]. In 2011, Belle Collaboration reported observation of states hb(1P)h_{b}(1P) and hb(2P)h_{b}(2P) in e+ehb(nP)π+πe^{+}e^{-}\rightarrow h_{b}(nP)\pi^{+}\pi^{-} process [13]. In the same year, the ATLAS group reconstructed χb(nP)\chi_{b}(nP) via radiative decay χb(nP)Υ(1S)γ\chi_{b}(nP)\rightarrow\Upsilon(1S)\gamma and χb(nP)Υ(2S)γ\chi_{b}(nP)\rightarrow\Upsilon(2S)\gamma. A new state χb(3P)\chi_{b}(3P) is identified in these decay modes [14]. This state χb(3P)\chi_{b}(3P) was also reconfirmed by D0 collaborations [15] and the LHCb group [16]. The latest CMS experiment shows resolved peaks of χb1(3P)\chi_{b1}(3P) and χb2(3P)\chi_{b2}(3P) with mass difference of 10.60±0.64(stat)±0.17(syst)10.60\pm 0.64(stat)\pm 0.17(syst) MeVMeV using 80 fb1fb^{-1} in pp collisions at a center-of-mass energy of 13 TeVTeV [17]. In 2012, Belle announced the first evidence of the existence of state ηb(2S)\eta_{b}(2S), pseudoscalar partner of Υ(2S)\Upsilon(2S) [18]. Some charged candidates like Zb(10610)Z_{b}(10610) and Zb(10650)Z_{b}(10650) are also announced in π±Υ(nS)\pi^{\pm}\Upsilon(nS)(n = 1,2,3) and π±hb(mP)\pi^{\pm}h_{b}(mP) (m=1,2m=1,2) mass spectra by Belle experiment [3]. But, their nature is still enigmatic. These states have been thoroughly investigated and analyzed in theoretical frameworks and interpreted these states as molecular states, exotic states, and tetraquarks states [19, 20, 21]. Recently, there has been a resurgence of interest in the exploration and investigation of states Υ(10860)\Upsilon(10860) and Υ(11020)\Upsilon(11020), which was first announced in 1985 by CUSB experiment [9]. A recent study in 2019 by Belle investigated Υ(10860)\Upsilon(10860) state in process e+eΥ(nS)e^{+}e^{-}\rightarrow\Upsilon(nS) (n=1,2,3n=1,2,3) with high significance and measured masses are 10752.7±5.910752.7\pm 5.9 MeVMeV,and decay widths are 35.517.6+17.635.5^{+17.6}_{-17.6} MeVMeV respectively [22]. Now this vector state is mentioned as Υ(10753)\Upsilon(10753) state in PDG. Authors in Ref.[23], investigated Υ(10753)\Upsilon(10753) using relativistic flux tube and give assignment of 33D13^{3}D_{1} bb¯b\bar{b} state. In Ref.[24], with QCD sum rules, Wang et al. interpreted this state as a hidden bottom tetraquark state. Theoretically, the prevailing consensus posits that Υ(10860)\Upsilon(10860) and Υ(11020)\Upsilon(11020) states correspond to the S-wave vector bb¯b\bar{b} states denoted as Υ(5S)\Upsilon(5S) and Υ(6S)\Upsilon(6S), respectively [25, 26]. Authors in Ref.[27, 28] explicated these states (Υ(10860)\Upsilon(10860) and Υ(11020)\Upsilon(11020)) with instanton-induced potential obtained from the instanton liquid model for QCD vacuum and elucidated Υ(10860)\Upsilon(10860) as an admixture of 53S163D15^{3}S_{1}-6^{3}D_{1} states and Υ(11020)\Upsilon(11020) as an admixture of 53S153D15^{3}S_{1}-5^{3}D_{1} states respectively. Also state Υ(11020)\Upsilon(11020) studied in Ref.[27, 28] and suggested as an admixture of 63S153D16^{3}S_{1}-5^{3}D_{1} states. Very recently, in March 2023, the BelleII detector examined the decay process e+eωχbJ(1P)e^{+}e^{-}\rightarrow\omega_{\chi_{bJ}}(1P) (J=0,1,2J=0,1,2) and concluded states Υ(10860)\Upsilon(10860) and Υ(10753)\Upsilon(10753) may have different internal structures [29], which is considered same in previous studies [22]. All these recent observations of numerous states of heavy mesons developed an interest in theoreticians to explore them in all aspects. Different QCD potential models have studied all the above-observed states. It is believed that QCD potential models are the most successful phenomenological approach in the non-perturbative region. There are various kinds of potentials in literature like Cornell potential [30, 31, 32], Martin potential [33, 34], Logarithmic potential [35], Richardson potential [36], and Song and Lin potential [37], which successfully explore quarkonium and their properties. Another potential is used to study charmonium spectroscopy [38] and its decay properties. In the same context, we explore bottomonium spectra and their decay properties using the same non-relativistic potential model from Ref. [38]. Similar potentials have been harnessed with n=2n=2 to study the dynamics of light hadrons in the framework of the Bethe- Salpeter equation under Covariant Instantaneous Ansatz (CIA) [39, 40, 41]. It can extensively explore many processes like digamma decays, radiative decays, decay constants, and the structure of hadrons. This framework can also be explored for various energy scales to have better insights. Using quark - antiquark confining potential form [38], we calculated masses and decay properties of bottomonia - leptonic decays, gamma decays, gluons decays, and decay constants, which are very useful in revealing the non-perturbative domain of QCD.

The paper is summarized as follows: Section 2 gives a brief description of phenomenological quark-antiquark confining potential and extraction of potential parameters from Chi-square fitting. Section 3 gives a description of the decay rates of quarkonium. Section 4 presents the numerical analysis where we predict the mass spectra of bottomonia for nS,nP,nD,nF,n=1,2,3,4,5nS,nP,nD,nF,n=1,2,3,4,5 and their decay properties. Section 5 gives the conclusions of the paper.

2 Framework

Various theoretical approaches describe the spectrum of quarkonium states (charmonium, bottomonium, beauty charmed mesons). Still, the phenomenological potential model is one of the most popular and reliable approaches. The famous form of potential model exploited in heavy quarkonium spectroscopy is coulomb plus linear potential. The Coulomb potential component arises from the one-gluon exchange (Lorentz vector exchange) interaction between the quark and antiquark, which is analogous to the electromagnetic interaction between charged particles. On the other hand, the linear potential arises from the strong force (usually associated with Lorentz scalar exchange) that acts between the quark and the antiquark. This potential increases linearly with the separation between the quark and the antiquark. As discussed in the previous section, many other forms of confining potentials are used in literature. We have adopted the following potential form to study the spectroscopy of bottomonium bound states [38].

V(r)=VV+VS=4αs3r+Ar2(1+4Brp)12V0\displaystyle V(r)=V_{V}+V_{S}=\frac{-4\alpha_{s}}{3r}+\frac{Ar^{2}}{(1+4Br^{p})^{\frac{1}{2}}}-V_{0} (1)

where VVV_{V} is the vector part of the potential (Coulomb), VSV_{S} is the scalar part of the potential (confining), αs\alpha_{s} is the running coupling constant. AA and BB are potential parameters estimated with chi-square fitting of low-lying bottomonium states. Since for charmonium, as mentioned in Ref.[38], the results are not consistent for p=2p=2, we have also worked for p=1p=1 and B=1B=1 GeVpGeV^{p}. The value of the running coupling constant can be obtained by the expression :

αs(μ2)=4π(1123nf)(lnμ2Λ2)\displaystyle\alpha_{s}(\mu^{2})=\frac{4\pi}{\left(11-\frac{2}{3}n_{f}\right)\left(ln\frac{\mu^{2}}{\Lambda^{2}}\right)} (2)

Where Λ\Lambda is the QCD scale taken as 0.12 GeVGeV, nfn_{f} is the number of active flavors, number of flavors lighter than scale μ\mu (mqμm_{q}\ll\mu). μ\mu is the renormalization scale equal to 2mQmQ¯mQ+mQ¯\frac{2m_{Q}m_{\bar{Q}}}{m_{Q}+m_{\bar{Q}}}, for bottomonium nf=4n_{f}=4, as flavors uu, dd, ss, and cc are considered light. A similar potential is used to study the ground states of light-flavoured and heavy-light mesons [42, 43, 44, 40, 41, 45, 46, 47, 48, 49]. The confining term in the potential Eq. (1) is supposed to behave as linear confinement (rr) for the heavy quark (cc, bb) and hadronic form of the (r2r^{2}) for the light quarks. V0V_{0} is a state-dependent constant potential. In adopted potential in eq 1, we added spin-dependent terms - spin-spin, spin-orbit, and tensor to describe the spectrum’s splitting structure and account for each state’s different quantum numbers. The spin-dependent part VSDV_{SD} [50, 51] is given by:

VSD=VSS[S(S+1)32]\displaystyle V_{SD}=V_{SS}\left[S(S+1)-\frac{3}{2}\right] +VLS[12(J(J+1)S(S+1)L(L+1))]\displaystyle+V_{LS}\left[\frac{1}{2}(J(J+1)-S(S+1)-L(L+1))\right]
+VT[S(S+1)3(S1.r)(S2.r)r2]\displaystyle+V_{T}\left[S(S+1)-\frac{3(S_{1}.r)(S_{2}.r)}{r^{2}}\right] (3)

Where coefficients VSSV_{SS}, VLSV_{LS}, and VTV_{T} depend on derivatives of vector VVV_{V} and scalar VSV_{S} contributions from the adopted potential in eq 1. The expressions for coefficients are the following [50, 52]:

VSSij(r)\displaystyle V_{SS}^{ij}(r) =13MiMj2VV=16παs9MiMjδ3(r)\displaystyle=\frac{1}{3M_{i}M_{j}}\nabla^{2}V_{V}=\frac{16\pi\alpha_{s}}{9M_{i}M_{j}}\delta^{3}(r) (4)
VLSij(r)\displaystyle V_{LS}^{ij}(r) =12MiMjr[3dVVdrdVSdr]\displaystyle=\frac{1}{2M_{i}M_{j}r}\left[3\frac{dV_{V}}{dr}-\frac{dV_{S}}{dr}\right] (5)
VTij(r)\displaystyle V_{T}^{ij}(r) =16MiMj[d2VVdr21rdVVdr]\displaystyle=\frac{1}{6M_{i}M_{j}}\left[\frac{d^{2}V_{V}}{dr^{2}}-\frac{1}{r}\frac{dV_{V}}{dr}\right] (6)

Where MiM_{i}, MjM_{j} are quark masses. All these spin-dependent interactions are corrections in total mass, treated as first-order perturbation corrections in heavy quark-bound states. As VLSV_{LS}, VSSV_{SS}, and VTV_{T} expressions are proportional to 1/m21/m^{2} (Mi=Mj=mM_{i}=M_{j}=m for bottomonium), they justify their treatment as first-order perturbative corrections. The spin-orbit term containing VLSV_{LS} and tensor term containing VTV_{T} describe the fine structure of the state, whereas the spin-spin term containing VSSV_{SS} describes hyperfine splittings. Mass is the prime property to study the spectroscopy of hadrons. To calculate spectra of bottomonium, we estimated the parameters V0V_{0}, bottom quark mbm_{b}, and parameter AA appearing in potential in eq (1) with Chi-square fitting procedure. We considered these parameters as free parameters in the fitting procedure for ground-state bottomonium meson in the following range:

0<A<10<A<1 (GeV3)(GeV^{3}); 0<V0<0.50<V_{0}<0.5 (GeV)(GeV); 4<mb<54<m_{b}<5 (GeV)(GeV)

The results for parameters are: mbm_{b} = 4.72 GeVGeV, V0V_{0} = 0.10 GeVGeV, AA = 0.20 GeV2GeV^{2}. The running coupling constant is αs=0.2053\alpha_{s}=0.2053. The state dependent potential strength V0V_{0} is given by relation:

V0(n+1,l)=V0+b(n1)+cl\displaystyle V_{0}(n+1,l)=V_{0}+b(n-1)+cl (7)

Where nn and ll are the radial and angular quantum numbers of the states. Also, bb and cc are unknown parameters estimated by fitting the experimental masses of 2S2S, 1P1P states of bottomonium. We find bb = -0.105 GeVGeV, cc = 0.0007 GeVGeV. Using all these parameters and adopted potential form, we computed mass spectra of nSnS, nPnP, nDnD, and nFnF bottomonium states listed in Table 1, Table 2, Table 3, and Table 4.

3 Decay rates of heavy quarkonia

Apart from the masses of bottomonium meson states, precise predictions of decay rates are crucial properties of any successful model. Recently there are several investigations on various phenomena encompassing strong, radiative, and leptonic decays of heavy quarkonium. Such investigations directly probe the hadron structure and shed light on some aspects of the quark-gluon structure. Leptonic decay constants serve as expedient probes for the short-distance structure of hadron and act as tools for studying quark dynamics in this domain. The extracted model parameters and radial wave functions are used here to compute the di-leptonic, two-photon, and two-gluon annihilation rates. Since these rates are related to the wave- function, they provide a better insight into quark-antiquark dynamics within mesons.

3.1 Leptonic decay constants

The study of leptonic decay constants of heavy quarkonia is crucial properties for understanding weak decays. It also provides information about CKM (Cabibbo-Kabayashi-Maskawa) matrix elements. The expression for leptonic decay constants of pseudoscalar and vector mesons is given by [53]:

fP/V2=3|RnsP/V(0)|2πMnsP/VC¯2(αs)\displaystyle f^{2}_{P/V}=\frac{3|R_{nsP/V(0)}|^{2}}{\pi M_{nsP/V}}\bar{C}^{2}(\alpha_{s}) (8)

Here the QCD correction factor C¯2(αs)\bar{C}^{2}(\alpha_{s}) [54, 55] is given by:

C¯2(αs)=1αsπ(δP,VmQmQ¯mQ+mQ¯lnmQmQ¯)\displaystyle\bar{C}^{2}(\alpha_{s})=1-\frac{\alpha_{s}}{\pi}\left(\delta_{P,V}-\frac{m_{Q}-m_{\bar{Q}}}{m_{Q}+m_{\bar{Q}}}ln\frac{m_{Q}}{m_{\bar{Q}}}\right) (9)

with δP=2\delta_{P}=2 and δV=8/3\delta_{V}=8/3 in case of Bottomonium mesons. The second term in equation 9 will disappear in the case of bottomonium. Using these expressions, we computed leptonic decay constants and listed them in Table 5 and Table 6.

3.2 Electromagnetic Transition widths

The study of electromagnetic transition can be understood in terms of electric and magnetic multipole expansion, and their investigation provides better sight into the non-perturbative regime. The selection rule for E1E1 transitions is ΔL=0\Delta L=0 and ΔS=±1\Delta S=\pm 1. For M1M1 transitions, ΔL=±1\Delta L=\pm 1 , and ΔS=0\Delta S=0 are selection rules. To check the authenticity of calculated masses and chi-square fit parameters, we computed E1E1 and M1M1 transitions, and the formulas for these transitions [56, 57, 58, 59] are given below:

Γ(n2S+1LiJin2S+1LfJf+γ)\displaystyle\Gamma(n^{2S+1}L_{i}J_{i}\rightarrow n^{\prime 2S+1}L_{f}J_{f}+\gamma) =4αeeQ2ω33(2Jf+1)SifE1|MifE1|2\displaystyle=\frac{4\alpha_{e}\left<e_{Q}\right>^{2}\omega^{3}}{3}(2J_{f}+1)S_{if}^{E1}|M_{if}^{E1}|^{2} (10)
Γ(n3S1n1S0+γ)\displaystyle\Gamma(n^{3}S_{1}\rightarrow n^{\prime 1}S_{0}+\gamma) =αeμ2ω33(2Jf+1)|MifM1|2\displaystyle=\frac{\alpha_{e}\mu^{2}\omega^{3}}{3}(2J_{f}+1)|M_{if}^{M1}|^{2} (11)

where, eQ\left<e_{Q}\right> is mean charge of QQ¯Q\bar{Q} system, μ\mu is magnetic dipole moment and ω\omega is photon energy and expression for them are following :

eQ=|mQ¯eQeQ¯mQmQ+mQ¯|\displaystyle\left<e_{Q}\right>=\left|\frac{m_{\bar{Q}}e_{Q}-e_{\bar{Q}}m_{Q}}{m_{Q}+m_{\bar{Q}}}\right| (12)
μ=eQmQeQ¯mQ¯\displaystyle\mu=\frac{e_{Q}}{m_{Q}}-\frac{e_{\bar{Q}}}{m_{\bar{Q}}} (13)

and

ω=Mi2Mf22Mi\displaystyle\omega=\frac{M_{i}^{2}-M_{f}^{2}}{2M_{i}} (14)

The symmetrical factor SifE1S_{if}^{E1} is also given in the following way:

SifE1=max(Li,Lf)(Ji1JfLfSLi)2\displaystyle S_{if}^{E1}=max(L_{i},L_{f})\begin{pmatrix}J_{i}&1&J_{f}\\ L_{f}&S&L_{i}\end{pmatrix}^{2} (15)

The matrix element |Mif||M_{if}| for E1 and M1 transition can be written as

|MifE1|=3ωf|ωr2j0(ωr2)j1(ωr2)|i\displaystyle|M_{if}^{E1}|=\frac{3}{\omega}\left<f\left|\frac{\omega r}{2}j_{0}\left(\frac{\omega r}{2}\right)-j_{1}\left(\frac{\omega r}{2}\right)\right|i\right> (16)

and

|MifM1|=f|j0(ωr2)|i\displaystyle|M_{if}^{M1}|=\left<f\left|j_{0}\left(\frac{\omega r}{2}\right)\right|i\right> (17)

The computed values of E1E1 and M1M1 transitions are collected in Table 7 and Table 8, respectively.

3.3 Leptonic decay widths

Single virtual photons (QQ¯l+lQ\bar{Q}\rightarrow l^{+}l^{-}) are used to disintegrate quarkonium into leptons (eμ,τe^{-}\mu^{-},\tau^{-}). If the quarkonium state has the same quantum number as the photon, it decays to the lepton pair, i.e., JPC=1J^{PC}=1^{--}. Using Van Royen Weisskopf formula, the leptonic decay width of n3S1n^{3}S_{1} and n3D1n^{3}D_{1} states of bottomonium, including first-order radiative QCD corrections [53, 60], is given below:

Γ(n3S1e+e)=4eQ4α2|RnS(0)|2MnS2(116αs3π)\displaystyle\Gamma(n^{3}S_{1}\rightarrow e^{+}e^{-})=\frac{4e_{Q}^{4}\alpha^{2}|R_{nS}(0)|^{2}}{M_{nS}^{2}}\left(1-\frac{16\alpha_{s}}{3\pi}\right) (18)
Γ(n3D1e+e)=25eQ2α2|RnD′′(0)|22mQ4MnD2(116αs3π)\displaystyle\Gamma(n^{3}D_{1}\rightarrow e^{+}e^{-})=\frac{25e_{Q}^{2}\alpha^{2}|R_{nD}^{{}^{\prime\prime}}(0)|^{2}}{2m_{Q}^{4}M_{nD}^{2}}\left(1-\frac{16\alpha_{s}}{3\pi}\right) (19)

where eQe_{Q} is charge on quark, MnSM_{nS}, MnDM_{nD} is masses of decaying corresponding states, α=1/137\alpha=1/137 is fine structure constant that describes the strength of electromagnetic force, and RnS(0)R_{nS}(0), RnD′′(0)R_{nD}^{{}^{\prime\prime}}(0) are normalized reduced wave-function at origin for SS waves and second order derivative of normalized reduced wave function for DD waves respectively. Bracketed terms in expressions 18, 19 are the lowest order QCD corrections. The computed values of leptonic decay widths are mentioned in Table 9.

3.4 Digamma and trigamma decay widths

A decay into two photons is forbidden to J=1J=1 states by the Yang theorem [61, 62]. For other resonances, the conversation of charge parity requires the SS wave states to be in a spin-singlet state, whereas PP wave states of being in the spin-triplet state. The formula for digamma decay widths of S01{}^{1}S_{0} and PJ3(J=0,2){}^{3}P_{J}(J=0,2) is given below [53, 60]:

Γ(n1S0γγ)\displaystyle\Gamma(n^{1}S_{0}\rightarrow\gamma\gamma) =3eQ2α2|RnS(0)|2mQ2(13.4αsπ)\displaystyle=\frac{3e^{2}_{Q}\alpha^{2}|R_{nS}(0)|^{2}}{m_{Q}^{2}}\left(1-\frac{3.4\alpha_{s}}{\pi}\right) (20)
Γ(n3P0γγ)\displaystyle\Gamma(n^{3}P_{0}\rightarrow\gamma\gamma) =27eQ4α2|RnP(0)|2MP032mQ5[1+(π23289)(αsπ)]\displaystyle=\frac{27e^{4}_{Q}\alpha^{2}|R^{{}^{\prime}}_{nP}(0)|^{2}M_{{}^{3}P_{0}}}{2m_{Q}^{5}}\left[1+\left(\frac{\pi^{2}}{3}-\frac{28}{9}\right)\left(\frac{\alpha_{s}}{\pi}\right)\right] (21)
Γ(n3P2γγ)\displaystyle\Gamma(n^{3}P_{2}\rightarrow\gamma\gamma) =41527eQ4α2|RnP(0)|2MP232mQ5[1+(163)(αsπ)]\displaystyle=\frac{4}{15}\frac{27e^{4}_{Q}\alpha^{2}|R^{{}^{\prime}}_{nP}(0)|^{2}M_{{}^{3}P_{2}}}{2m_{Q}^{5}}\left[1+\left(\frac{-16}{3}\right)\left(\frac{\alpha_{s}}{\pi}\right)\right] (22)
Γ(n3S1γγγ)\displaystyle\Gamma(n^{3}S_{1}\rightarrow\gamma\gamma\gamma) =4(π29)eQ6α3|RnS(0)|2mQ2(112.6αsπ)\displaystyle=\frac{4(\pi^{2}-9)e^{6}_{Q}\alpha^{3}|R_{nS}(0)|^{2}}{m_{Q}^{2}}\left(1-\frac{12.6\alpha_{s}}{\pi}\right) (23)

RnlR^{{}^{\prime}}_{nl} is the first derivative of the normalized reduced wave function at the origin in PP wave formulas. eQe_{Q} is charge on quark. α=1/137\alpha=1/137 is a fine structure constant. Terms in brackets are the next to leading-order QCD radiative corrections. The calculated digamma, tri-gamma decay widths are collected in Table 10 and Table 11.

3.5 Digluon decay widths

At short range, two gluon decay widths are sensitive for quarkonia and its derivatives near the origin. The digluon decay width gives information about the total decay width of hadronic decay below the QQ¯Q\bar{Q} threshold. Employing VRW (Van Royen Weisskopf) method [53, 60] and including QCD radiative corrections [63, 64, 65], expressions for digluon decay widths are:

Γ(n1S0gg)=2αs2|Rns(0)|23mQ2(1+4.4αsπ)\displaystyle\Gamma(n^{1}S_{0}\rightarrow gg)=\frac{2\alpha_{s}^{2}|R_{ns}(0)|^{2}}{3m_{Q}^{2}}\left(1+\frac{4.4\alpha_{s}}{\pi}\right) (24)
Γ(n3P0gg)\displaystyle\Gamma(n^{3}P_{0}\rightarrow gg) =6αs2|RnP(0)|2mQ4(1+10.0αsπ)\displaystyle=\frac{6\alpha_{s}^{2}|R^{{}^{\prime}}_{nP}(0)|^{2}}{m_{Q}^{4}}\left(1+\frac{10.0\alpha_{s}}{\pi}\right) (25)
Γ(n3P2gg)\displaystyle\Gamma(n^{3}P_{2}\rightarrow gg) =8αs2|RnP(0)|25mQ4(10.1αsπ)\displaystyle=\frac{8\alpha_{s}^{2}|R^{{}^{\prime}}_{nP}(0)|^{2}}{5m_{Q}^{4}}\left(1-\frac{0.1\alpha_{s}}{\pi}\right) (26)
Γ(n1D2gg)\displaystyle\Gamma(n^{1}D_{2}\rightarrow gg) =2αs2|RnD′′(0)|23πmQ6\displaystyle=\frac{2\alpha_{s}^{2}|R^{{}^{\prime\prime}}_{nD}(0)|^{2}}{3\pi m_{Q}^{6}} (27)

Terms in brackets are the next to leading-order QCD radiative corrections. α=1/137\alpha=1/137 is fine structure constant. RnPR^{{}^{\prime}}_{nP} is the first order derivative of the normalized reduced wave function at the origin in PP wave formulas. Similar RnD"R^{"}_{nD} is the second order derivative of the normalized reduced wave function at the origin in DD wave formulas. The calculated values of digluons are mentioned in Table 12.

4 Results and Discussions

4.1 Mass Spectroscopy

The spectroscopic masses of S,P,D,FS,P,D,F waves of bottomonium mesons are calculated by employing a non-relativistic potential given in Eq. (1) and solving the Schrodinger wave equation numerically. Our calculated mass for S,P,D,FS,P,D,F waves of bottomonium mesons is listed in Table 1, Table 2, Table 3, Table 4 respectively and compared with experimental data as well as different theoretical predictions.

Table 1: The predicted values of nSnS bottomonium masses (GeVGeV) compared with some other model predictions.
JPC(n2S+1Lj)J^{PC}(n^{2S+1}L_{j}) Ours Ref.[32] Ref.[26] Ref.[66] Ref.[27] Ref.[67] Ref. [68] Ref.[69] PDG [3]
0+(11S0)0^{-+}(1^{1}S_{0}) 9.43297 9.423 9.398 9.402 9.41222 9.392 9.398 9.393 9.399
1(13S1)1^{--}(1^{3}S_{1}) 9.44816 9.463 9.463 9.465 9.46075 9.460 9.460 9.460 9.460
0+(21S0)0^{-+}(2^{1}S_{0}) 10.00660 9.983 9.989 9.976 9.99548 9.991 9.990 9.987 9.999
1(23S1)1^{--}(2^{3}S_{1}) 10.02020 10.001 10.017 10.003 10.02622 10.024 10.023 10.023 10.023
0+(31S0)0^{-+}(3^{1}S_{0}) 10.50310 10.342 10.336 10.336 10.33900 10.323 10.329 10.345 -
1(33S1)1^{--}(3^{3}S_{1}) 10.51650 10.354 10.356 10.354 10.36465 10.346 10.355 10.364 10.355
0+(41S0)0^{-+}(4^{1}S_{0}) 10.96910 10.638 10.597 10.623 10.57249 10.558 10.573 10.364 -
1(43S1)1^{--}(4^{3}S_{1}) 10.98250 10.650 10.612 10.635 10.59447 10.575 10.586 10.643 10.579
0+(51S0)0^{-+}(5^{1}S_{0}) 11.41690 10.901 10.810 10.869 10.74676 10.741 10.851 - -
1(53S1)1^{--}(5^{3}S_{1}) 11.43030 10.912 10.822 10.878 10.76614 10.755 10.869 - 10.885
Table 2: The predicted values of bottomonium nPnP meson masses (GeVGeV) compared with some other model predictions.
JPC(n2S+1Lj)J^{PC}(n^{2S+1}L_{j}) Ours Ref.[26] Ref.[70] Ref.[27] Ref. [31] Ref. [66] Ref.[68] Ref. [71] PDG[3]
0++(13P0)0^{++}(1^{3}P_{0}) 9.75011 9.858 9.845 9.84961 9.806 9.847 9.859 9.865 9.859
1+(11P1)1^{+-}(1^{1}P_{1}) 9.75938 9.894 9.881 9.87456 9.819 9.876 9.892 9.897 9.899
1++(13P1)1^{++}(1^{3}P_{1}) 9.75671 9.889 9.875 9.87147 9.821 9.882 9.900 9.903 9.893
2++(13P2)2^{++}(1^{3}P_{2}) 9.76608 9.910 9.896 9.88140 9.825 9.897 9.912 9.918 9.912
0++(23P0)0^{++}(2^{3}P_{0}) 10.25270 10.235 10.225 10.25254 10.205 10.226 10.233 10.226 10.233
1+(21P1)1^{+-}(2^{1}P_{1}) 10.26000 10.259 10.250 10.27000 10.217 10.246 10.255 10.251 10.260
1++(23P1)1^{++}(2^{3}P_{1}) 10.25840 10.255 10.246 10.26786 10.220 10.250 10.260 10.256 10.255
2++(23P2)2^{++}(2^{3}P_{2}) 10.26670 10.269 10.261 10.27477 10.224 10.261 10.268 10.269 10.269
0++(33P0)0^{++}(3^{3}P_{0}) 10.72320 10.513 10.521 10.51288 10.540 10.552 10.521 10.502 -
1+(31P1)1^{+-}(3^{1}P_{1}) 10.72930 10.530 10.540 10.52650 10.553 10.538 10.541 10.524 -
1++(33P1)1^{++}(3^{3}P_{1}) 10.72840 10.527 10.537 10.52484 10.556 10.541 10.544 10.529 10.514
2+(33P2)2^{+-}(3^{3}P_{2}) 10.73630 10.539 10.549 10.53021 10.560 10.550 10.550 10.540 10.524
0++(43P0)0^{++}(4^{3}P_{0}) 11.17450 10.736 10.773 10.70356 10.840 10.775 10.781 10.732 -
1+(41P1)1^{+-}(4^{1}P_{1}) 11.17990 10.751 10.790 10.71480 10.853 10.788 10.802 10.753 -
1++(43P1)1^{++}(4^{3}P_{1}) 11.17950 10.749 10.787 10.71344 10.855 10.790 10.804 10.757 -
2++(43P2)2^{++}(4^{3}P_{2}) 11.18700 10.758 10.797 10.71786 10.860 10.798 10.812 10.767 -
0++(53P0)0^{++}(5^{3}P_{0}) 11.61260 10.926 10.998 10.85338 11.115 11.014 - 10.933 -
1+(51P1)1^{+-}(5^{1}P_{1}) 11.61730 10.938 11.013 10.86300 11.127 11.014 - 10.951 -
1++(53P1)1^{++}(5^{3}P_{1}) 11.61740 10.936 11.010 10.86183 11.130 11.016 - 10.955 -
2++(53P2)2^{++}(5^{3}P_{2}) 11.62460 10.944 11.020 10.86562 11.135 11.022 - 10.965 -
Table 3: The predicted values of bottomonium nDnD meson masses (GeVGeV) compared with some other model predictions.
JPC(n2S+1Lj)J^{PC}(n^{2S+1}L_{j}) Ours Ref.[26] Ref.[70] Ref. [27] Ref.[67] Ref. [71] Ref. [68] Ref.[31] PDG [3]
1(13D1)1^{--}(1^{3}D_{1}) 9.99397 10.153 10.137 10.14499 10.147 10.145 10.154 10.074 -
2+(11D2)2^{-+}(1^{1}D_{2}) 9.99238 10.163 10.148 10.15380 10.162 10.151 10.161 10.075 -
2(13D2)2^{--}(1^{3}D_{2}) 9.99423 10.162 10.147 10.15277 10.166 10.152 10.163 10.074 10.164
3(13D3)3^{--}(1^{3}D_{3}) 9.99386 10.170 10.155 10.15831 10.177 10.156 10.166 10.073 -
1(23D1)1^{--}(2^{3}D_{1}) 10.46920 10.442 10.441 10.45023 10.428 10.432 10.435 10.423 -
2+(21D2)2^{-+}(2^{1}D_{2}) 10.46740 10.450 10.450 10.45660 10.437 10.438 10.443 10.424 -
2(23D2)2^{--}(2^{3}D_{2}) 10.46990 10.450 10.449 10.45586 10.440 10.439 10.445 10.424 -
3(23D3)3^{--}(2^{3}D_{3}) 10.47040 10.456 10.455 10.45985 10.447 10.442 10.449 10.423 -
1(33D1)1^{--}(3^{3}D_{1}) 10.92470 10.675 10.698 10.65968 10.637 10.670 10.704 10.731 -
2+(31D2)2^{-+}(3^{1}D_{2}) 10.92210 10.681 10.706 10.66470 10.645 10.676 10.711 10.733 -
2(33D2)2^{--}(3^{3}D_{2}) 10.92550 10.681 10.705 10.66412 10.646 10.677 10.713 10.733 -
3(33D3)3^{--}(3^{3}D_{3}) 10.92640 10.686 10.711 10.66725 10.652 10.680 10.717 10.733 -
1(43D1)1^{--}(4^{3}D_{1}) 11.36630 10.871 10.927 10.81883 10.805 10.877 10.949 11.013 -
2+(41D2)2^{-+}(4^{1}D_{2}) 11.36290 10.876 10.934 10.82300 10.811 10.882 10.957 11.016 -
2(43D2)2^{--}(4^{3}D_{2}) 11.36710 10.876 10.934 10.82252 10.813 10.883 10.959 11.015 -
3(43D3)3^{--}(4^{3}D_{3}) 11.36810 10.880 10.939 10.82512 10.817 10.886 10.963 11.015 -
1(53D1)1^{--}(5^{3}D_{1}) 11.79730 11.041 11.137 10.94901 10.945 11.060 - - -
2+(51D2)2^{-+}(5^{1}D_{2}) 11.79300 11.046 11.143 10.95260 10.952 11.066 - - -
2(53D2)2^{--}(5^{3}D_{2}) 11.79820 11.045 11.143 10.95159 10.950 11.065 - - -
3(53D3)3^{--}(5^{3}D_{3}) 11.79920 11.049 11.148 10.95442 10.955 11.069 - - -
Table 4: The predicted values of bottomonium nFnF meson masses (GeVGeV) compared with some other model predictions.
JPC(n2S+1Lj)J^{PC}(n^{2S+1}L_{j}) Ours Ref.[26] Ref.[70] Ref. [66] Ref. [31] Ref. [68] Ref. [69]
2++(13F2)2^{++}(1^{3}F_{2}) 10.20690 10.362 10.350 10.358 10.283 10.343 10.353
3+(11F3)3^{+-}(1^{1}F_{3}) 10.20130 10.366 10.354 10.355 10.288 10.347 10.356
3++(13F3)3^{++}(1^{3}F_{3}) 10.20430 10.366 10.354 10.355 10.287 10.346 10.356
4++(13F4)4^{++}(1^{3}F_{4}) 10.20050 10.369 10.358 10.358 10.291 10.349 10.357
2++(23F2)2^{++}(2^{3}F_{2}) 10.66600 10.605 10.615 10.615 10.604 10.610 10.610
3+(21F3)3^{+-}(2^{1}F_{3}) 10.66080 10.609 10.619 10.619 10.607 10.647 10.613
3++(23F3)3^{++}(2^{3}F_{3}) 10.66500 10.609 10.619 10.619 10.607 10.614 10.613
4++(23F4)4^{++}(2^{3}F_{4}) 10.66340 10.612 10.622 10.622 10.609 10.617 10.615
2++(33F2)2^{++}(3^{3}F_{2}) 11.11170 10.809 10.849 10.850 10.894 - -
3++(31F3)3^{++}(3^{1}F_{3}) 11.10530 10.812 10.853 10.853 10.897 - -
3++(33F3)3^{++}(3^{3}F_{3}) 11.11120 10.812 10.853 10.853 10.896 - -
4++(33F4)4^{++}(3^{3}F_{4}) 11.11040 10.815 10.856 10.856 10.898 - -
2++(43F2)2^{++}(4^{3}F_{2}) 11.54670 10.985 11.063 - - - -
3+(41F3)3^{+-}(4^{1}F_{3}) 11.53850 10.988 11.066 - - - -
3++(43F3)3^{++}(4^{3}F_{3}) 11.54640 10.988 11.066 - - - -
4++(43F4)4^{++}(4^{3}F_{4}) 11.54590 10.990 11.066 - - - -

Our findings for the masses concur with experimentally observed masses for respective states. Our calculated masses for 13S11^{3}S_{1} state 9.448169.44816 GeVGeV, which is found to be in good agreement with experimental value 9.460±0.000269.460\pm 0.00026 GeVGeV and for its spin partner 11S01^{1}S_{0} state, our estimated mass is roughly 11 MeVMeV higher than experimental mass mentioned in PDG. On observing Table 1, we found masses of nS wave up to n = 5 agree with the experimental data and other theoretical predictions with a deviation of ±5%\pm 5\%. For nSnS states, our predicted masses are close to theoretical estimations of Ref. [32, 66, 67]. In Ref. [32], Kher et al. employed a variational method with a single Gaussian trial wave functions, Ref. [66], Godfrey et al. employed relativized quark model and Ref. [67], Shah et al. employed the non-relativistic potential model to investigate bottomonia mass spectroscopy. Specifically, on comparing with Ref. [26], we observed our masses are very close to their predictions up to n=4n=4 SS wave states, but our results slightly deviate for 5S5S states from them. This may happen due to screening effects employed by Ref. [26] as screening effects contribute significantly to higher excited states and affects mass values, wave functions, and decay behaviors. Similarly, for PP waves, our results are in good agreement with experimental values and theoretical estimations. For n=1n=1 PP wave masses, our mass estimations are 1.45%1.45\% below experimental values. Our computation for state 21P12^{1}P_{1} is 10.260010.2600, which is in excellent agreement with the experimental value mentioned in PDG (10.260±1.2×10410.260\pm 1.2\times 10^{-4}). Also, its spin partners, i.e 23P0,23P1,23P22^{3}P_{0},2^{3}P_{1},2^{3}P_{2}, our results differ from experimental values by 1919 MeVMeV, 33 MeVMeV, 22 MeVMeV respectively which reflects reliability of our model. One more important parameter to check the authenticity of any model is hyperfine splittings. For 1P1P state, ΔMhfs(1P)\Delta M_{hfs}(1P) = 4.9 MeVMeV and 2P2P state ΔMhfs(2P)\Delta M_{hfs}(2P) = 2.4 MeVMeV, which consistent to experimental values [3]. This splitting decreases more and becomes negligible for higher excited states. This shows that spin-spin contribution is negligible for higher excited states and thus reflects the vanishing of long-range chromomagnetic interactions in quarkonium. The calculated masses of n=4,5n=4,5 for PP are found to be overestimated in comparison to other theoretical estimations. However, masses for n=5n=5 P wave are closer to outcomes of Ref. [66, 70, 31] than other estimations. In Ref. [70, 66], Godfrey et al. implemented a relativistic approach, while Ref. [31] used non-relativistic approach. This shows relativistic factors have very small contributions that do not much affect the mass spectroscopy of bottomonia. In Table 3, we listed predicted masses for nDnD wave (n=1,2,3,4,5n=1,2,3,4,5) and compared them with other model predictions and experimental masses, which are also in good accord with them. Only one state 11D21^{1}D_{2} of DD waves is experimentally available and the measured value of mass is 10.16410.164 GeVGeV. Our predicted value for this state deviates by 1.59%1.59\% from the experimental value. Our result for 2D2D masses is found to be consistent with other theoretical predictions, while masses of 1D,3D,4D1D,3D,4D states are higher than predictions of other models. However, masses for 5D5D states are closer to predictions of Ref. [70], which employed relativized quark model with chromodynamics and included the concept of one-gluon-exchange-plus-linear confinement potential. This is also notable in Ref. [70], authors have not taken spin-spin, spin-orbit interaction into account, which reflects in mass degeneracy in spin states. Also, computed masses for nFnF wave (n=1,2,3,4n=1,2,3,4) listed in Table 4 are very close to other models’ predictions. For the 1F1F state, our results are more consistent with Ref. [31], which used Cornell’s potential to study bottomonium spectra. This is a similar kind of method that we used in this paper. Such comparison enhances the validity of our framework. Other results of FF states are also in reasonable agreement with theoretical estimates.

4.2 Decay constants

The study of leptonic decay constants is important for investigating non - perturbative QCD dynamics. Using calculated masses for S-wave collected in Table 1 and applying Van Royen-Weisskopf formula, extracted in eqns 8, 9, we computed values of pseudoscalar decay constants fPf_{P} and vector decay constants fVf_{V} listed in Table 5 and Table 6 respectively.

Table 5: Pseudoscalar decay constants of bottomonium (MeVMeV) without and with QCD corrections
States fPf_{P} fPCf_{P}^{C} Ref.[32] Ref.[27] Ref.[31] Ref.[72] Ref. [73] Ref. [74]
1S1S 632.626 549.879 529 578.21 646.025 756 744 498
2S2S 551.298 479.188 317 499.48 518.803 285 577 366
3S3S 528.840 459.668 280 450.35 474.954 333 511 304
4S4S 516.337 448.800 264 413.93 449.654 40 471 259
5S5S 507.351 440.990 255 385.68 432.072 - 443 228
Table 6: Vector decay constants of bottomonium (MeVMeV) without and with QCD corrections
States fVf_{V} fVCf_{V}^{C} Ref.[32] Ref.[27] Ref.[31] Ref. [74] Ref. [75] Ref. [73] PDG [3]
1S1S 666.036 535.517 530 551.53 647.250 498 705.4 706 715
2S2S 564.438 466.707 317 477.05 519.436 366 554.9 547 498
3S3S 542.290 447.715 280 430.42 475.440 304 436.8 484 430
4S4S 529.483 437.141 265 395.80 450.066 259 332.4 446 336
5S5S 520.281 429.544 255 368.91 432.437 228 286.5 419 -

Our results for pseudoscalar decay constants are in tune with Ref. [32, 27, 31], while lower from Ref. [72, 73]. It is noticed that Ref. [32, 27, 31] have employed a similar kind of framework with different potential forms, which enhance the authenticity of our work. The estimated values of vector decay constant for states 1S,2S,3S,4S1S,2S,3S,4S differ by 179 MeVMeV, 31 MeVMeV, 17 MeVMeV, and 101 MeVMeV from experimental values, respectively. In comparison with theoretical approaches, values for vector decay constants without QCD corrections are consistent with Ref. [31], while with QCD corrections, our values concur with Ref. [27]. Ref. [27] employed instanton-induced potential by including confining terms to study bottomonium spectroscopy.

4.3 Radiative Transitions

The numerical result of radiative transitions in electric diploe (E1) and magnetic dipole (M1) expansion are listed in Table 7and Table 8, respectively.

Table 7: E1 transition widths of bottomonia (in keVkeV)
Decays Ours Ref.[31] PDG[3] Ref.[25] Ref. [76] Ref. [68] Ref. [71]
23S113P02^{3}S_{1}\rightarrow 1^{3}P_{0} 1.38000 2.370 1.220 1.09 1.15 1.65 1.67
23S113P12^{3}S_{1}\rightarrow 1^{3}P_{1} 3.87511 5.689 2.210 2.17 1.87 2.57 2.54
23S113P22^{3}S_{1}\rightarrow 1^{3}P_{2} 5.84652 8.486 2.290 2.62 1.88 2.53 2.62
23S011P12^{3}S_{0}\rightarrow 1^{1}P_{1} 10.3487 10.181 - 3.41 4.17 3.25 6.10
33S123P03^{3}S_{1}\rightarrow 2^{3}P_{0} 1.44755 3.330 1.200 1.21 1.67 1.65 1.83
33S123P13^{3}S_{1}\rightarrow 2^{3}P_{1} 4.10768 7.936 2.560 2.61 2.74 2.65 2.96
33S123P23^{3}S_{1}\rightarrow 2^{3}P_{2} 6.29319 11.447 2.660 3.16 2.80 2.89 3.23
33S113P03^{3}S_{1}\rightarrow 1^{3}P_{0} 6.89619 0.594 0.055 0.097 0.03 0.124 0.07
33S113P13^{3}S_{1}\rightarrow 1^{3}P_{1} 2.54430 1.518 0.018 0.0005 0.09 0.307 0.17
33S113P23^{3}S_{1}\rightarrow 1^{3}P_{2} 3.91653 2.354 0.200 0.14 0.13 0.445 0.15
31S011P13^{1}S_{0}\rightarrow 1^{1}P_{1} 6.96943 3.385 - 0.67 0.03 0.770 1.24
33S021P13^{3}S_{0}\rightarrow 2^{1}P_{1} 11.39010 13.981 - 4.25 - 3.07 11.0
13P213S11^{3}P_{2}\rightarrow 1^{3}S_{1} 37.69970 57.530 34.380 31.8 31.2 29.5 38.2
13P113S11^{3}P_{1}\rightarrow 1^{3}S_{1} 34.72050 54.927 32.544 31.9 27.3 37.1 33.6
13P013S11^{3}P_{0}\rightarrow 1^{3}S_{1} 32.70690 49.530 - 27.5 22.1 42.7 26.6
11P111S01^{1}P_{1}\rightarrow 1^{1}S_{0} 38.71140 72.094 35.770 35.8 37.9 54.4 55.8
23P223S12^{3}P_{2}\rightarrow 2^{3}S_{1} 11.05290 28.848 15.100 15.5 16.8 18.8 18.8
23P123S12^{3}P_{1}\rightarrow 2^{3}S_{1} 10.07620 26.672 19.400 15.3 13.7 15.9 15.9
23P023S12^{3}P_{0}\rightarrow 2^{3}S_{1} 9.43571 23.162 - 14.4 9.90 11.7 11.7
21P121S02^{1}P_{1}\rightarrow 2^{1}S_{0} 11.01620 35.578 - 16.2 - 23.6 24.7
23P213S12^{3}P_{2}\rightarrow 1^{3}S_{1} 9.41195 29.635 9.800 12.5 7.74 8.41 13.0
23P113S12^{3}P_{1}\rightarrow 1^{3}S_{1} 9.20092 28.552 8.900 10.8 7.31 8.01 12.4
23P013S12^{3}P_{0}\rightarrow 1^{3}S_{1} 9.05706 28.552 - 10.8 6.69 7.36 11.4
21P111S02^{1}P_{1}\rightarrow 1^{1}S_{0} 9.44253 26.769 - 5.4 - 9.9 15.9
13D113P01^{3}D_{1}\rightarrow 1^{3}P_{0} 16.24530 34.815 - 16.1 - 24.2 23.6
13D113P11^{3}D_{1}\rightarrow 1^{3}P_{1} 11.27790 9.670 - 19.8 - 12.9 12.3
13D113P21^{3}D_{1}\rightarrow 1^{3}P_{2} 0.67091 0.394 - 13.3 - 0.67 0.65
13D213P11^{3}D_{2}\rightarrow 1^{3}P_{1} 20.36500 11.489 - 1.02 19.3 24.8 23.8
13D213P21^{3}D_{2}\rightarrow 1^{3}P_{2} 6.05832 3.583 - 7.23 5.07 6.45 6.29
13D313P21^{3}D_{3}\rightarrow 1^{3}P_{2} 24.12070 14.013 - 32.1 21.7 26.7 26.4
11D211P11^{1}D_{2}\rightarrow 1^{1}P_{1} 25.03720 14.821 - 30.3 - 30.2 42.3
Table 8: M1 transition widths of bottomonia (in eV)
Decays Ours Ref.[32] Ref.[25] PDG[3] Ref. [76] Ref. [77] Ref. [78]
13S111S01^{3}S_{1}\rightarrow 1^{1}S_{0} 0.01436 37.668 10.00 - 4.00 5.8 15.36
23S121S02^{3}S_{1}\rightarrow 2^{1}S_{0} 0.01601 5.619 0.59 - 0.05 1.4 1.82
23S111S02^{3}S_{1}\rightarrow 1^{1}S_{0} 1.25374 77.173 66.00 12.5±4.912.5\pm 4.9 - 6.4 -
33S131S03^{3}S_{1}\rightarrow 3^{1}S_{0} 0.01629 2.849 3.90 - - 0.8 -
33S121S03^{3}S_{1}\rightarrow 2^{1}S_{0} 3.12243 36.177 11.00 <14<14 - 1.5 -
33S111S03^{3}S_{1}\rightarrow 1^{1}S_{0} 1.72572 76.990 71.00 10±210\pm 2 - 10.5 -

Our estimations for radiative transitions widths Γ(23S113P0γ)\Gamma(2^{3}S_{1}\rightarrow 1^{3}P_{0}\gamma), Γ(23S113P1γ)\Gamma(2^{3}S_{1}\rightarrow 1^{3}P_{1}\gamma), Γ(23S113P2γ)\Gamma(2^{3}S_{1}\rightarrow 1^{3}P_{2}\gamma) differ by 0.16 keVkeV, 1.66 keVkeV, 3.5 keVkeV from experimental values, respectively, and are also compatible with theoretical models estimations. These transitions (Γ(23S113P0γ)\Gamma(2^{3}S_{1}\rightarrow 1^{3}P_{0}\gamma), Γ(23S113P1γ)\Gamma(2^{3}S_{1}\rightarrow 1^{3}P_{1}\gamma), Γ(23S113P2γ)\Gamma(2^{3}S_{1}\rightarrow 1^{3}P_{2}\gamma)) are closer to results of the potential model with v2/c2v^{2}/c^{2} corrections [76], quasi potential approach [68], the potential model with screened potential [71] shown in Table 7. The E1 transitions Γ(33S123P0γ)\Gamma(3^{3}S_{1}\rightarrow 2^{3}P_{0}\gamma), Γ(33S123P1γ)\Gamma(3^{3}S_{1}\rightarrow 2^{3}P_{1}\gamma), Γ(33S123P2γ)\Gamma(3^{3}S_{1}\rightarrow 2^{3}P_{2}\gamma) are consistent with experimental data while Γ(33S113P0γ)\Gamma(3^{3}S_{1}\rightarrow 1^{3}P_{0}\gamma), Γ(33S113P1γ)\Gamma(3^{3}S_{1}\rightarrow 1^{3}P_{1}\gamma), Γ(33S113P2γ)\Gamma(3^{3}S_{1}\rightarrow 1^{3}P_{2}\gamma) are lower by 7 keVkeV, 3 keVkeV, 4 keVkeV from experimental results respectively. It is also inferred from Table 7, radiative transition widths for 3P1S3P\rightarrow 1S are suppressed than other E1E1 transitions. This shows a general feature that E1E1 transition between two states differs by two radial numbers that are highly suppressed and have very less contribution in E1E1 transitions. The calculated electric dipole transitions for Γ(13P213S1γ)\Gamma(1^{3}P_{2}\rightarrow 1^{3}S_{1}\gamma), Γ(13P113S1γ)\Gamma(1^{3}P_{1}\rightarrow 1^{3}S_{1}\gamma), Γ(11P111S0γ)\Gamma(1^{1}P_{1}\rightarrow 1^{1}S_{0}\gamma) are very close to experimental data. Also, the E1E1 transitions for 1P1S1P\rightarrow 1S give higher contributions relative to 2P1S2P\rightarrow 1S transition, which is expected. For E1E1 transition 1D1P1D\rightarrow 1P, there is no experimental data, but compared with theoretical predictions, it is found that our results have a high degree of similarity with Ref. [68, 71], while lower values than Ref. [31]. We also calculated M1M1 transitions and listed them in Table 8. Generally, M1M1 transitions are weaker than E1E1 transitions but play an important role in finding spin-singlet states, which is back-breaking in other ways. It is inferred from Table 8 that predicted values are very small relative to E1E1 transitions. Our results for M1M1 transitions are lower than other model’s predictions.

4.4 Annihilation Decays

The study of annihilation decays, i.e., decays of quarkonium states into photons, leptons, and gluons, sheds light on the perturbative aspect of QCD. Investigation of these decays is also helpful in the production, identification of conventional or non-conventional mesons, multiquark structures, etc. Using predicted masses and applying formulas obtained from Van Royen-Weisskopf formula, we estimate leptonic, digamma, tri gamma, digluon decay widths. These calculated decay widths are listed in Table 9, Table 10, Table 11, Table 12.

Table 9: Leptonic decay widths of bottomonia (in keVkeV)
n2S+1Ljn^{2S+1}L_{j} Γl+l\Gamma_{l^{+}l^{-}} Γl+lC\Gamma_{l^{+}l^{-}}^{C} Ref.[27] PDG[3] Ref.[32] Ref. [75] Ref. [26](eV) Ref. [79]
13S11^{3}S_{1} 0.91227 0.59379 0.7700 1.340±0.0181.340\pm 0.018 0.582 1.300 1.65 0.71
23S12^{3}S_{1} 0.65325 0.42526 0.5442 0.612±0.0110.612\pm 0.011 0.197 0.760 0.821 0.37
33S13^{3}S_{1} 0.57278 0.37288 0.4288 0.443±0.0080.443\pm 0.008 0.149 0.450 0.569 0.27
43S14^{3}S_{1} 0.48505 0.34039 0.3549 0.272±0.0290.272\pm 0.029 0.129 0.260 0.431 0.21
53S15^{3}S_{1} 0.48505 0.31578 0.3035 0.310±0.070.310\pm 0.07 0.117 0.180 0.348 0.18
13D11^{3}D_{1} 10.96020 (eV) 7.13748 (eV) 0.0050 - 1.65 0.106 (eV) 1.880 (eV) 1.40 (eV)
23D12^{3}D_{1} 12.69930 (eV) 8.26971 (eV) 0.0058 - 2.42 0.078 (eV) 2.810(eV) 2.50(eV)
33D13^{3}D_{1} 18.73760 (eV) 12.2004 (eV) 0.0059 - 3.19 0.051 (eV) 3.000(eV) -
43D14^{3}D_{1} 24.99260 (eV) 16.2709 (eV) 0.0058 - 3.97 0.042(eV) 3.000(eV) -
53D15^{3}D_{1} 32.20100 (eV) 20.9611 (eV) 0.0057 - - 0.028 (eV) 0.003(eV) -
Table 10: Digamma decay widths of SS and PP waves bottomonia (in keVkeV) without and with QCD corrections
n2S+1Ljn^{2S+1}L_{j} Γγγ\Gamma_{\gamma\gamma} ΓγγC\Gamma_{\gamma\gamma}^{C} Ref.[31] Ref.[27] Ref.[66] Ref. [32] Ref. [26] Ref. [71]
11S01^{1}S_{0} 0.30499 0.23719 0.3870 0.3035 0.940 0.2361 1.050 0.527
21S02^{1}S_{0} 0.24568 0.19106 0.2630 0.2122 0.410 0.0896 0.489 0.263
31S03^{1}S_{0} 0.23729 0.18454 0.2290 0.1668 0.290 0.0726 0.323 0.172
41S04^{1}S_{0} 0.23624 0.18372 0.2120 0.1378 0.200 0.0666 0.237 0.105
51S05^{1}S_{0} 0.23739 0.18462 0.2010 0.1176 0.170 0.0636 0.192 0.121
13P01^{3}P_{0} 0.08412 0.08522 0.0196 0.1150 0.150 0.0168 0.199 0.037
13P21^{3}P_{2} 0.07952 0.01467 0.0052 0.0147 0.093 0.0024 0.011 0.007
23P02^{3}P_{0} 0.07959 0.08056 0.0195 0.1014 0.150 0.0172 0.205 0.037
23P22^{3}P_{2} 0.08078 0.01386 0.0052 0.0131 0.012 0.0024 0.013 0.006
33P03^{3}P_{0} 0.08216 0.08063 0.0194 0.0875 0.130 0.0192 0.180 0.037
33P23^{3}P_{2} 0.02252 0.01387 0.0051 0.0114 0.013 0.0027 0.004 0.006
43P04^{3}P_{0} 0.02128 0.08184 0.0192 0.0768 0.130 - 0.157 -
43P24^{3}P_{2} 0.02129 0.01408 0.0051 0.0100 0.015 - 0.014 -
53P05^{3}P_{0} 0.02162 0.08323 0.0191 0.0686 - - 0.146 -
53P25^{3}P_{2} 0.02198 0.01431 0.0050 0.0090 - - 0.014 -
Table 11: Trigamma decay widths of bottomonia (in unit of 10310^{-3} eVeV) without QCD corrections
n2S+1Ljn^{2S+1}L_{j} Γγγγ\Gamma_{\gamma\gamma\gamma} Ref.[31] Ref.[26] Ref.[66] Ref. [32] Ref. [79] Ref.[28]
13S11^{3}S_{1} 30.36900 33.560 19.40 17.0 33.560 19.40 16
23S12^{3}S_{1} 24.46270 12.670 10.90 9.8 12.670 10.90 3
33S13^{3}S_{1} 23.62710 10.261 8.04 7.6 10.261 8.04 1
43S14^{3}S_{1} 23.52240 9.400 6.36 6.0 9.400 6.36 -
53S15^{3}S_{1} 23.63810 8.979 5.43 - 8.979 5.43 -
Table 12: Digluon decay widths of bottomonia (in MeVMeV) without and with QCD corrections
n2S+1Ljn^{2S+1}L_{j} Γgg\Gamma_{gg} ΓggC\Gamma_{gg}^{C} Ref.[31] Ref.[27] Ref.[66] Ref. [32] Ref. [26] Ref. [79]
11S01^{1}S_{0} 4.33520 5.58201 5.448 6.8520 16.600 8.219 17.9 20.180
21S02^{1}S_{0} 3.49207 4.49639 3.710 5.2374 7.200 3.121 8.33 10.640
31S03^{1}S_{0} 3.37280 4.34282 3.229 4.3182 4.900 2.529 5.51 7.940
41S04^{1}S_{0} 3.35785 4.32357 2.985 3.6829 3.400 2.317 4.03 -
51S05^{1}S_{0} 3.37436 4.34483 2.832 3.2196 2.900 2.214 3.26 -
13P01^{3}P_{0} 1.20151 0.10870 0.276 1.4297 2.600 0.721 3.37 2.000
13P21^{3}P_{2} 0.32040 0.31831 0.073 0.2370 0.147 0.192 0.165 0.084
23P02^{3}P_{0} 1.13537 0.08409 0.275 1.2358 2.600 0.741 3.52 2.370
23P22^{3}P_{2} 0.30276 0.30078 0.073 0.2064 0.207 0.198 0.220 0.104
33P03^{3}P_{0} 1.13611 0.07036 0.273 1.0539 2.200 0.828 3.10 2.460
33P23^{3}P_{2} 0.30296 0.30098 0.072 0.1767 0.227 0.221 0.243 0.112
43P04^{3}P_{0} 1.15308 0.06057 0.271 0.9175 2.100 - - -
43P24^{3}P_{2} 0.30749 0.30547 0.072 0.1543 0.248 - - -
53P05^{3}P_{0} 1.17255 0.05283 0.269 0.8127 - - - -
53P25^{3}P_{2} 0.31268 0.31063 0.071 0.1370 - - - -
11D21^{1}D_{2} 5.93167 (keVkeV) - - - 1.800 (keVkeV) 0.489 (keVkeV) 0.657 (keVkeV) 0.370 (keVkeV)
21D22^{1}D_{2} 7.53656 (keVkeV) - - - 1.530 (keVkeV) 0.764 (keVkeV) 1.22 (keVkeV) 0.670 (keVkeV)
31D23^{1}D_{2} 12.10030 (keVkeV) - - - 1.839 (keVkeV) 1.0006 (keVkeV) 1.59 (keVkeV) -
41D24^{1}D_{2} 17.45940 (keVkeV) - - - - 1.380 (keVkeV) 1.86 (keVkeV) -
51D25^{1}D_{2} 24.21910 (keVkeV) - - - - - 2.13 (keVkeV) -

On observing Table 10, we found that calculated partial decay widths for nSnS states agree with other models’ predictions. But our results for states nPnP are slightly off from other models’ predictions. They differ from other predictions by ±65\pm 65 eVeV. We also calculated tri gamma decay widths mentioned in Table 11, and the results are in good accord with estimations deduced by other theoretical approaches.

The calculated digluon decay widths are listed in Table 12 and compared with predictions of other models. On comparing, it is found that our estimations are consistent with the results of other models for nSnS states. However, our results for nPnP states are higher than the predictions of other models. A similar trend can be seen in the case of nDnD states, i.e., our results are slightly higher than other model estimations.

The dileptonic decay widths of quarkonium play an important role in the estimation of strong coupling constants, decay constants and in checking the validity of theoretical models as its decay amplitude carries quarkonium wave function (can be seen in eqn.18). We calculated leptonic decay width and mentioned it in Table 9. On comparing, it is found that our predictions for nSnS states agree with the outcomes of other models. For nDnD states, our results agree with Ref. [27] and are higher in comparison to Ref. [75, 26, 79]. For higher nn, our decay widths for DD states show a monotonically increasing trend, which agrees with some of the compared models. It is also noticed on observing tables of annihilation decays that for DD wave states, values of leptonic decays and digluon decays are highly suppressed as compared to SS waves, which is expected.

As we know, the total decay width of a particular state is the sum of strong, radiative, and weak decay widths. But strong decay gives more contribution to the total decay width of that state relative to radiative and weak decay. This is inferred from Table 7, 8, 9, 10, 11, 12, which reflects the value of digluon decays (strong decays) are much more relative to radiative transitions and leptonic decays (weak decays).

5 Conculsion

In this paper, we have presented a comprehensive analysis of the mass spectra of bottomonium states and their decay properties by employing a non-relativistic potential model. Using adopted quark-antiquark confining potential, we solved Schrodinger wave equations numerically with Runge -Kutta method using the Mathematica notebook [80]. Our calculated mass spectrum (nS,nP,nD,nF,n=1,2,3,4,5nS,nP,nD,nF,n=1,2,3,4,5) of bottomonia is fairly close to available experimental data and other theoretical model predictions. The hyperfine splittings of 1P1P and 2P2P states are in fair agreement with the experiment observation. We have computed annihilation decays - di-leptonic, di-gamma, tri-gamma, di-gluons decay widths. Also, the electromagnetic transition widths by Van Royen - Weisskopf formula are calculated. All these decays are also estimated with first-order QCD corrections. Our findings agree with the available experimental observations and theoretical estimations. We have also explored leptonic decay constants, which are consistent with the experimental values mentioned in PDG. The present results may be helpful in upcoming experimental information in the near future.

6 Acknowledgement

The authors gratefully acknowledge the financial support by the Department of Science and Technology (SERB/F/9119/2020), New Delhi and for Junior Research Fellowship (09/0677(11306)/2021-EMR-I) by Council of Scientific and Industrial Research, New Delhi.

References

  • [1] S.W. Herb, D.C. Hom, L.M. Lederman, J.C. Sens, H.D. Snyder, J.K. Yoh, J.A. Appel, B.C. Brown, C.N. Brown, W.R. Innes et al., Phys. Rev. Lett. 39, 252 (1977)
  • [2] W.R. Innes, J.A. Appel, B.C. Brown, C.N. Brown, K. Ueno, T. Yamanouchi, S.W. Herb, D.C. Hom, L.M. Lederman, J.C. Sens et al., Phys. Rev. Lett. 39, 1240 (1977)
  • [3] R.L. Workman, Others (Particle Data Group), PTEP 2022, 083C01 (2022)
  • [4] K. Han, T. Böhringer, P. Franzini, G. Mageras, D. Peterson, E. Rice, J.K. Yoh, J.E. Horstkotte, C. Klopfenstein, J. Lee-Franzini et al., Phys. Rev. Lett. 49, 1612 (1982)
  • [5] G. Eigen, G. Blanar, H. Dietl, E. Lorenz, F. Pauss, H. Vogel, T. Böhringer, P. Franzini, K. Han, G. Mageras et al., Phys. Rev. Lett. 49, 1616 (1982)
  • [6] C. Klopfenstein, J.E. Horstkotte, J. Lee-Franzini, R.D. Schamberger, M. Sivertz, L.J. Spencer, P.M. Tuts, P. Franzini, K. Han, E. Rice et al., Phys. Rev. Lett. 51, 160 (1983)
  • [7] F. Pauss, H. Dietl, G. Eigen, E. Lorenz, G. Mageras, H. Vogel, P. Franzini, K. Han, D. Peterson, E. Rice et al., Physics Letters B 130, 439 (1983)
  • [8] D. Besson, J. Green, R. Namjoshi, F. Sannes, P. Skubic, A. Snyder, R. Stone, A. Chen, M. Goldberg, N. Horwitz et al., Phys. Rev. Lett. 54, 381 (1985)
  • [9] D.M.J. Lovelock, J.E. Horstkotte, C. Klopfenstein, J. Lee-Franzini, L. Romero, R.D. Schamberger, S. Youssef, P. Franzini, D. Son, P.M. Tuts et al., Phys. Rev. Lett. 54, 377 (1985)
  • [10] B. Aubert, M. Bona, Y. Karyotakis, J.P. Lees, V. Poireau, E. Prencipe, X. Prudent, V. Tisserand, J. Garra Tico, E. Grauges et al. (BABAR Collaboration), Phys. Rev. Lett. 101, 071801 (2008)
  • [11] P. del Amo Sanchez, J.P. Lees, V. Poireau, E. Prencipe, V. Tisserand, J. Garra Tico, E. Grauges, M. Martinelli, A. Palano, M. Pappagallo et al. (BABAR Collaboration), Phys. Rev. D 82, 111102 (2010)
  • [12] J.P. Lees, V. Poireau, E. Prencipe, V. Tisserand, J. Garra Tico, E. Grauges, M. Martinelli, D.A. Milanes, A. Palano, M. Pappagallo et al. (The BABAR Collaboration), Phys. Rev. D 84, 091101 (2011)
  • [13] I. Adachi, H. Aihara, K. Arinstein, D.M. Asner, T. Aushev, T. Aziz, A.M. Bakich, E. Barberio, K. Belous, V. Bhardwaj et al. (Belle Collaboration), Phys. Rev. Lett. 108, 032001 (2012)
  • [14] G. Aad, B. Abbott, J. Abdallah, A.A. Abdelalim, A. Abdesselam, O. Abdinov, B. Abi, M. Abolins, O.S. AbouZeid, H. Abramowicz et al. (ATLAS Collaboration), Phys. Rev. Lett. 108, 152001 (2012)
  • [15] V.M. Abazov, B. Abbott, B.S. Acharya, M. Adams, T. Adams, G.D. Alexeev, G. Alkhazov, A. Alton, G. Alverson, M. Aoki et al. (D0 Collaboration), Phys. Rev. D 86, 031103 (2012)
  • [16] R. Aaij, , B. Adeva, M. Adinolfi, A. Affolder, Z. Ajaltouni, S. Akar, J. Albrecht, F. Alessio, M. Alexander et al., Journal of High Energy Physics 2014, 088 (2014)
  • [17] A.M. Sirunyan, A. Tumasyan, W. Adam, F. Ambrogi, E. Asilar, T. Bergauer, J. Brandstetter, M. Dragicevic, J. Erö, A. Escalante Del Valle et al. (CMS Collaboration), Phys. Rev. Lett. 120, 231801 (2018)
  • [18] R. Mizuk, D.M. Asner, A. Bondar, T.K. Pedlar, I. Adachi, H. Aihara, K. Arinstein, V. Aulchenko, T. Aushev, T. Aziz et al. (Belle Collaboration), Phys. Rev. Lett. 109, 232002 (2012)
  • [19] Q. Wu, D.Y. Chen, T. Matsuki, Phys. Rev. D 102, 114037 (2020)
  • [20] Z.G. Wang, T. Huang, Nucl. Phys. A 930, 63 (2014)
  • [21] D.Y. Chen, X. Liu, T. Matsuki, Chin. Phys. C 38, 053102 (2014)
  • [22] R. Mizuk, A. Bondar, I. Adachi, H. Aihara, D. Asner, V. Aulchenko, T. Aushev, R. Ayad, I. Badhrees, S. Bahinipati et al., Journal of High Energy Physics 2019 (2019), cited by: 33; All Open Access, Gold Open Access, Green Open Access
  • [23] B. Chen, A. Zhang, J. He, Phys. Rev. D 101, 014020 (2020)
  • [24] Z.G. Wang, Chin. Phys. C 43, 123102 (2019)
  • [25] W.J. Deng, H. Liu, L.C. Gui, X.H. Zhong, Phys. Rev. D 95, 074002 (2017)
  • [26] J.Z. Wang, Z.F. Sun, X. Liu, T. Matsuki, Eur. Phys. J. C 78, 915 (2018), 1802.04938
  • [27] B. Pandya, M. Shah, P.C. Vinodkumar, Eur. Phys. J. C 81, 116 (2021)
  • [28] R. Chaturvedi, A.K. Rai, N.R. Soni, J.N. Pandya, Journal of Physics G: Nuclear and Particle Physics 47, 115003 (2020)
  • [29] I. Adachi, L. Aggarwal, H. Ahmed, H. Aihara, N. Akopov, A. Aloisio, N. Anh Ky, D.M. Asner, T. Aushev, V. Aushev et al. (Belle II Collaboration), Phys. Rev. Lett. 130, 091902 (2023)
  • [30] E. Eichten, K. Gottfried, T. Kinoshita, K.D. Lane, T.M. Yan, Phys. Rev. D 21, 203 (1980)
  • [31] N.R. Soni, B.R. Joshi, R.P. Shah, H.R. Chauhan, J.N. Pandya, Eur. Phys. J. C 78, 592 (2018)
  • [32] V. Kher, R. Chaturvedi, N. Devlani, A.K. Rai, Eur. Phys. J. Plus 137, 357 (2022)
  • [33] A. Martin, Physics Letters B 93, 338 (1980)
  • [34] S. Jena, T. Tripati, Physics Letters B 122, 181 (1983)
  • [35] C. Quigg, J.L. Rosner, Physics Letters B 71, 153 (1977)
  • [36] J.L. Richardson, Physics Letters B 82, 272 (1979)
  • [37] X.t. Song, H.f. Lin, Z. Phys. C 34, 223 (1987)
  • [38] S. Patel, P.C. Vinodkumar, S. Bhatnagar, Chin. Phys. C 40, 053102 (2016)
  • [39] A. Mittal, A.N. Mitra, Phys. Rev. Lett. 57, 290 (1986)
  • [40] K.K. Gupta, A.N. Mitra, N.N. Singh, https://link.aps.org/doi/10.1103/PhysRevD.42.1604Phys. Rev. D 42, 1604 (1990)
  • [41] A. Sharma, S.R. Choudhury, A.N. Mitra, Phys. Rev. D 50, 454 (1994)
  • [42] N.N. Singh, A.N. Mitra, Phys. Rev. D 38, 1454 (1988)
  • [43] S. Chakrabarty, K. Gupta, N. Singh, A. Mitra, Progress in Particle and Nuclear Physics 22, 43 (1989)
  • [44] A. Mittal, A.N. Mitra, Phys. Rev. Lett. 57, 290 (1986)
  • [45] R. Alkofer, L. von Smekal, Physics Reports 353, 281 (2001)
  • [46] R. Alkofer, P. Watson, H. Weigel, Phys. Rev. D 65, 094026 (2002)
  • [47] G. Cvetič, C. Kim, G.L. Wang, W. Namgung, Physics Letters B 596, 84 (2004)
  • [48] S. Bhatnagar, S.Y. Li, Journal of Physics G: Nuclear and Particle Physics 32, 949 (2006)
  • [49] S. Bhatnagar, J. Mahecha, Y. Mengesha, Phys. Rev. D 90, 014034 (2014)
  • [50] M. Voloshin, Progress in Particle and Nuclear Physics 61, 455 (2008)
  • [51] V. Berestetskii, E. Lifshits, L. Pitaevskii, Quantum Electrodynamics (Pergamon, 1982)
  • [52] S.S. Gershtein, V.V. Kiselev, A.K. Likhoded, A.V. Tkabladze, Phys. Rev. D 51, 3613 (1995)
  • [53] R. Van Royen, V.F. Weisskopf, Nuovo Cim. A 50, 617 (1967), [Erratum: Nuovo Cim.A 51, 583 (1967)]
  • [54] E. Braaten, S. Fleming, Phys. Rev. D 52, 181 (1995)
  • [55] A.V. Berezhnoy, V.V. Kiselev, A.K. Likhoded, Z. Phys. A 356, 89 (1996)
  • [56] E. Eichten, K. Gottfried, T. Kinoshita, J. Kogut, K.D. Lane, T.M. Yan, Phys. Rev. Lett. 34, 369 (1975)
  • [57] E. Eichten, K. Gottfried, T. Kinoshita, K.D. Lane, T.M. Yan, Phys. Rev. D 17, 3090 (1978)
  • [58] N. Brambilla, Y. Jia, A. Vairo, Phys. Rev. D 73, 054005 (2006)
  • [59] D.M. Li, P.F. Ji, B. Ma, Eur. Phys. J. C 71, 1582 (2011)
  • [60] W. Kwong, P.B. Mackenzie, R. Rosenfeld, J.L. Rosner, Phys. Rev. D 37, 3210 (1988)
  • [61] L.D. Landau, Dokl. Akad. Nauk SSSR 60, 207 (1948)
  • [62] C.N. Yang, Phys. Rev. 77, 242 (1950)
  • [63] J.P. Lansberg, T.N. Pham, Phys. Rev. D 79, 094016 (2009)
  • [64] R. Barbieri, M. Caffo, R. Gatto, E. Remiddi, Nucl. Phys. B 192, 61 (1981)
  • [65] M.L. Mangano, A. Petrelli, Phys. Lett. B 352, 445 (1995)
  • [66] S. Godfrey, K. Moats, Phys. Rev. D 92, 054034 (2015)
  • [67] M. Shah, A. Parmar, P.C. Vinodkumar, Phys. Rev. D 86, 034015 (2012)
  • [68] D. Ebert, R.N. Faustov, V.O. Galkin, Eur. Phys. J. C 71, 1825 (2011)
  • [69] S.F. Radford, W.W. Repko, Nucl. Phys. A 865, 69 (2011)
  • [70] S. Godfrey, N. Isgur, Phys. Rev. D 32, 189 (1985)
  • [71] B.Q. Li, K.T. Chao, Commun. Theor. Phys. 52, 653 (2009)
  • [72] A. Krassnigg, M. Gomez-Rocha, T. Hilger, J. Phys. Conf. Ser. 742, 012032 (2016)
  • [73] B. Patel, P.C. Vinodkumar, J. Phys. G 36, 035003 (2009)
  • [74] G.L. Wang, Phys. Lett. B 633, 492 (2006)
  • [75] T. Bhavsar, M. Shah, P.C. Vinodkumar, Eur. Phys. J. C 78, 227 (2018)
  • [76] S.F. Radford, W.W. Repko, Phys. Rev. D 75, 074031 (2007)
  • [77] D. Ebert, R.N. Faustov, V.O. Galkin, Phys. Rev. D 67, 014027 (2003)
  • [78] Bhaghyesh, K.B.V. Kumar, A.P. Monteiro, Journal of Physics G: Nuclear and Particle Physics 38, 085001 (2011)
  • [79] J. Segovia, P.G. Ortega, D.R. Entem, F. Fernández, Phys. Rev. D 93, 074027 (2016)
  • [80] W. Lucha, F.F. Schoberl, Int. J. Mod. Phys. C 10, 607 (1999)