Fundamental properties of short-lived subatomic particles S. Ceci,1, ∗ M. Hadˇzimehmedovi´c,2 H. Osmanovi´c,2 A. Percan,3 and B. Zauner1 2

1 Rudjer Boˇskovi´c Institute, Bijeniˇcka 54, HR-10000 Zagreb, Croatia University of Tuzla, Univerzitetska 4, 75000 Tuzla, Bosnia and Herzegovina 3 University of Zagreb, Bijeniˇcka 34, HR-10000 Zagreb, Croatia (Dated: August 24, 2016)

arXiv:1608.06485v1 [hep-ph] 23 Aug 2016

Two distinct sets of properties are used to describe short-lived particles: the pole and the BreitWigner parameters. There is an ongoing decades-old debate on which of them is fundamental. All resonances, from excited hydrogen nuclei hit by ultra-high energy gamma rays in deep space, to new particles produced in Large Hadron Collider, should be described by the same fundamental physical quantities. In this study of nucleon resonances we discover an intricate interplay of the parameters from the both sets, and realize that neither set is fundamental on its own. PACS numbers: 11.55.Bq, 14.20.Gk, 13.75.Gx, 12.40.Yx Keywords: Resonance mass, Scattering amplitude poles, Breit-Wigner parameters

All particle scattering processes are described by the scattering amplitude, a complex function of energy. It is also an analytic function. That means it is expandable to the experimentally unreachable complex energies. An infinite value of the amplitude at some complex energy indicates the existence of a short-living particle, i.e. the resonance [1]. This infinity, the first order pole, may produce experimentally observable signal in a measured probability of the reaction, the cross section. In such cases, the cross section increases rapidly as the energy approaches the resonance mass, and then suddenly drops producing a characteristic bell-shaped peak. Main resonant features come from the mathematical properties of the pole. Its mass M is given by the real part of the pole position in the complex energy plane, and its total decay width Γ is directly determined from the imaginary part. Generally, the peak position and its width do not correspond to M nor Γ. Two other properties, the magnitude |r|, and especially the phase θ of the complex residue seem to be purely mathematical objects. There is no physical interpretation for them. Experimental physicists, therefore, describe resonances using a different set of parameters. Instead of the pole parameters, they use Breit-Wigner mass MBW , width ΓBW , and branching fraction xBW . These Breit-Wigner parameters are also used in some theoretical studies: the quark models [2], the effective-field theories [3], and the lattice quantum chromodynamics [4]. It is, however, important to distinguish the Breit-Wigner parameters from the quantities in the Breit-Wigner formula [5] that can be found in most textbooks [6, 7]. Parameters MBW , ΓBW , and xBW collected by the Particle Data Group (PDG) [8] are not extracted using the Breit-Wigner formula, but rather elaborate functions that are fundamentally different for Z boson [8], ∆ resonance [9], or ρ meson [10]. The debate whether having two sets of resonant properties is redundant lasts for decades now, and the BreitWigner parameters seem to be losing [11]. That is the case particularly since their mere physicality came into

question. Namely, the Breit-Wigner masses of Z boson [12] and ∆ resonance [13], calculated using the standard definition, change when otherwise unobservable field transformations are imposed on a quantum-field level. Here we show that for the prominent isolated resonances their pole residue phase can be predicted with known reaction threshold, the resonant pole position, and the corresponding Breit-Wigner mass. This is, as far as we know, the first time anyone provided the physical meaning for the residue phase. For other resonances, pole residue ceases to be a fundamental property of a single resonance and, due to the unitarity, becomes a collective property strongly influenced by all particles having the same quantum numbers. We begin by reminding the reader that the resonant cross section [6–8] is given by σ=

4π 2J + 1 2 |A| , 2 q (2s1 + 1)(2s2 + 1)

(1)

where q is the center-of-mass momentum of incident particles, s1 and s2 are their spins, J is the spin of the resonance, and A is the key object in this relation, the resonant amplitude. The simplest resonant amplitude is the Breit-Wigner formula [5–8] ABW =

x Γ/2 , M − W − i Γ/2

(2)

where W is the center-of-mass energy, M − i Γ/2 is the pole position, and x Γ/2 is the residue magnitude |r|. Here, the residue phase θ is taken to be zero. (Mathematically it is −180◦ , but in the resonance physics this odd convention is used.) It is useful to rewrite the amplitude in this form with explicitly written complex phase ABW = x eiρ sin ρ,

(3)

where phase ρ is defined by tan ρ =

Γ/2 . M −W

(4)

2 This amplitude is not very realistic. The cross section diverges at the threshold, since q = 0 there, and the residue phase is hardly ever zero [8]. A more general resonant amplitude would be Am.g. =

V (W ) , m0 − W + Σ(W )

(5)

where m0 is the real-valued bare mass, while the vertex function V and the self-energy term Σ are nontrivial complex functions of energy. (Due to the relativity, everything should be a function of energy squared, but poles and residues are still defined with W by convention. Nevertheless, all our formulas can be easily generalized to relativistic forms.) Assuming there are no other resonances or thresholds nearby, V and Σ can be expanded in polynomial series. We need just a few terms in the vicinity of the resonant pole. By keeping only constant terms, one gets the Breit-Wigner formula. In Ref. [14] two terms are kept instead. Here we use the result of Ref. [14] rewritten in the same way as the Breit-Wigner formula in equation (3) A = x ei(ρ+β) sin(ρ + δ).

(6)

The meanings of x and ρ have already been explained, while β and δ are parameters that build the residue phase θ = β + δ.

(7)

To get the Breit-Wigner mass, we rewrite equation (6) as   cos δ Γ x cos β 2 + (M − W ) tan δ   . (8) A= M − Γ2 tan β − W − i Γ2 + (M − W ) tan β

the complex zero of the denominator. The Breit-Wigner mass is the renormalized mass of a resonance [6] defined as the real energy at which the real part of the denominator vanishes [9, 12, 13, 17] Re [m0 − MBW + Σ (MBW )] = 0.

(11)

However, there is a serious problem with this definition. In order to get equation (8) from equation (6), at some point we divided both numerator and denominator by eiβ . Any such transformation that does not change the pole position nor the residue (or any observable) will generally change the Breit-Wigner mass defined by equation (11). Something similar could have happened in Refs. [12, 13], where a quantum-field transformation that did not change any observable or pole position, changed MBW defined as the real part of a denominator. A mathematically more consistent definition of the Breit-Wigner mass could be Re A (MBW ) = 0.

(12)

It would be interesting to see the calculations in Refs. [12, 13] repeated using equation (12) instead of equation (11). Graphical representation of our model is shown in Fig. 1. The phase of the resonant amplitude plotted in the complex energy plane is produced by equation (6) with parameters of ∆(1232) resonance from PDG [8]. There, we also show geometrical meaning of the phases

���°

This formula belongs to a large family of equations [8, 15, 16] whose general form is A=

Γpar (W )/2 , MBW − W − i Γtot (W )/2

�°

(9) -��°

where MBW is the Breit-Wigner mass, Γpar is partial decay width function, and Γtot is total decay width function. The latter is usually considered to be a real function. However, that is not the case for subthreshold resonances [15, 17], which may produce dubious conclusions regarding the value of MBW . From comparison of Eqs. (8) and (9) we see that the Breit-Wigner mass is MBW = M − Γ/2 tan β.

��°

(10)

The Breit-Wigner width ΓBW , defined as Γtot (MBW ), is Γ/ cos2 β. Both results are consistent with Ref. [9]. In addition, dividing Γpar (MBW ) with Γtot (MBW ) gives us the Breit-Wigner branching fraction xBW as x cos(δ −β). If function Σ in equation (5) is known, there is another somewhat more elaborate way to determine fundamental resonant parameters. The pole position is simply

-���°

FIG. 1. The resonant amplitude phase of our model (OM) plotted in the complex energy plane. The amplitude is given by equation (6) with parameters of ∆(1232) from PDG [8]. The black disk shows the position of the zero at the threshold, the white disk is at the pole, and the white with a black eye at the Breit-Wigner mass. Solid white curve goes where the amplitude is purely imaginary, and dashed white where it is real. Residue phase θ is the angle between real axis and the tangent to the dashed white line at the pole.

β and δ. Both are negative here: δ is measured from the real axis, and β form the vertical line crossing the pole position.

3 The amplitude in Fig. 1 has a zero at threshold W0 = M + Γ/2 cot δ,

(13)

which enables us to estimate δ from known M , Γ, and W0 . We study nucleon resonances observed in the elastic πN scattering with W0 = 1077 MeV, the sum of pion and nucleon masses. To calculate β we use equation (10) with PDG estimates [8] for M , Γ, and MBW , and then predict θ using equation (7) for the four-star nucleon resonances with mass below 2 GeV. The results are compared to experimental values in Table I.

���° ��° �° -��° -���°

TABLE I. Test of the model on the four-star nucleon resonances. Here, δ is calculated using equation (13) assuming W0 is 1077 MeV, β using equation (10), and θ using equation (7). Experimental parameters (exp) are from PDG [8]. Their errors are ignored since they are not statistical, but estimates determined from several analyses using different models Resonance Name J π + ∆(1232) 32 − N (1520) 23 − N (1675) 25 5+ N (1680) 2 + ∆(1950) 72 + N (1440) 12 − N (1535) 21 − N (1650) 21 1− ∆(1620) 2 − ∆(1700) 32 + N (1720) 32 + ∆(1905) 52 1+ ∆(1910) 2

xexp BW (%) 100 60 40 68 40 65 45 60 25 15 11 12 23

M exp MeV 1210 1510 1660 1675 1880 1365 1510 1655 1600 1650 1675 1820 1855

Γexp MeV 100 110 135 120 240 190 170 135 130 230 250 280 350

δ (◦ ) −21 −7 −7 −6 −8 −18 −11 −7 −7 −11 −12 −11 −13

exp MBW MeV 1232 1515 1675 1685 1930 1430 1535 1655 1630 1700 1720 1880 1890

β (◦ ) −24 −5 −13 −9 −23 −34 −16 0 −25 −23 −20 −23 −11

θ (◦ ) −44 −12 −19 −15 −31 −53 −27 −7 −32 −34 −32 −34 −24

θexp (◦ ) −46+2 −2 −10+5 −5 −25+6 −6 −10+10 −10 −31+8 −8 −85+10 −15 −15+15 −15 −70+20 −10 −101+9 −9 −20+20 −20 −130+30 −30 −40+10 −10 −162+83 −83

Table I is divided in three groups. First, there are five resonances that are in excellent agreement with our model. Their residue phases are correctly predicted using the known resonant parameters, and their Breit-Wigner masses fully consistent with equation (12). The latter is evident from the plots of realistic L+P [18] amplitudes’ phases close to the resonances shown in Fig. 2. In the second group there are three remaining resonances with large xBW for which we fail to predict θ and MBW . Since the used model is strictly single resonance, − it will not work for strongly overlapping 12 resonances N(1535) and N(1650). To tackle this problem we combine the amplitudes of the overlapping resonances, while trying to preserve unitarity of the scattering matrix, defined as S = 1 + 2iA. For that, we use a unitary recipe S = S1 S2 S3 · · · , where the indices indicate resonances. We ignore other contributions coming from the fact that S is really a matrix. As a result, the residue phase of N(1535) becomes somewhat smaller, and of N(1650) much larger. Both of them are now substantially closer

FIG. 2. Phase of the amplitude near the first five resonances in Table I calculated using L+P results [18]. Breit-Wigner masses are consistent with PDG estimates [8] (white bars).

to the experimental values in Table I. Incidentally, the phases are practically the same as those in L+P [18], as can be seen in Fig. 3.

FIG. 3. Phase of the amplitude in which N(1535), N(1650), and N(1895) emerge. a) The L+P result [18] with residue phases (θ) for the first two resonances. b) Our model (OM) using only PDG estimates [8] and the unitary recipe.

4 We built the contribution of each resonance using equation (6), but did not calculate x by its definition (by dividing |r| with Γ/2) because when resonant terms are combined in a unitary amplitude, it is not only residue phase θ that is changed, but also its magnitude |r|. Instead, we calculated each x by dividing xBW with cos(δ − β), where xBW , β, and δ are taken from Table I. − Here we also included the third 21 resonance, N(1895), for which we estimate parameters from PDG [8]. It has no significant effect; we get roughly the same result when we completely omit it. Still, it is interesting that even though the unitarity constraint mixes all the resonances in the amplitude, distant resonances with small branching fractions, as is N(1895), will have a nearby zero of the amplitude, and this pole-zero pair can be completely detached from other resonances (i.e. not connected by solid nor dashed white lines). This is important because that is exactly the case with the isolated resonances with small xBW in the third group of Table I: ∆(1620), N (1720), and ∆(1910). We show them in Fig. 4.

crepancy of N(1440) residue phase because the closest 1+ 2 resonance, N(1710), is too far and with too small xBW + [8] to affect it at all. However, the nucleon itself is 12 particle and unitarity will mix the two. Nucleon has a pole in the subthreshold region, at 938 MeV. We estimate its contribution assuming the unitary recipe and equation (6) with β = 0◦ , δ = −180◦ , Γ → 0 MeV, and x → ∞. We use |r| = 59 MeV because with that value the amplitude at the real axis (its real and imaginary part) roughly resembles the realistic one in Ref. [18] close to N(1440). In Fig. 5 we see that the residue phase of N(1440) is now consistent with its experimental value.

���° ��° �° -��° -���°

���° ��° �° -��° -���°

FIG. 5. Phase of the amplitude in which N(1440) emerges. a) Our model (OM) with nucleon pole and N(1440). b) Our model with only N(1440). c) The realistic L+P result [18].

FIG. 4. Phase of the L+P amplitude [18] close to the isolated resonances with small xBW looks somewhat like the ∆(1232), just entirely below the real axis. Interestingly enough, if we draw a triangle similar to the one in Fig. 1 and define angles β 0 and δ 0 , residue phase θ is almost exactly a sum of the two.

For the remaining two resonances in the third group, ∆(1700) and ∆(1905), we get reasonable θ estimates but not the Breit-Wigner masses because the resonances − strongly overlap with unusually broad ∆(1940) 32 and + nearby ∆(2000) 52 , respectively. The final and hardest challenge for this model is the Roper resonance N(1440). We cannot use the overlapped unitary version of the model to explain the strong dis-

In conclusion, we have shown that for prominent nonoverlapping resonances the residue phase crucially depends on the Breit-Wigner mass. For the other resonances, the pole residue becomes a collective property influenced by all resonances with the same quantum numbers. In a way, that concludes the old debate. If we want to describe the scattering amplitude close to the resonance, we need the amplitude parameters: poles, residues, and zeros. If we, however, want to use or calculate the physical properties of the resonance, we must have its intrinsic parameters: the pole position, but also the Breit-Wigner parameters. During extensive discussions between S.C., B.Z. and A.P. the main idea of this work was conceived. S.C. did the calculations, plotted the figures, and prepared the manuscript which B.Z. and A.P. substantially edited. H.O. and M.H. provided the crucial L+P results. S.C. thanks Lothar Tiator for valuable comments and suggestions. The authors have no conflict of interest to declare.

5



[email protected] [1] R. H. Dalitz and R. G. Moorhouse, Proceedings of the Royal Society of London, Series A 318, 279-298, (1970). [2] S. Capstick and W. Roberts, Progress in Particle and Nuclear Physics 45, Supplement 2, S241-S331 (2000). [3] V. Pascalutsa, M. Vanderhaeghen, and S. N. Yang, Phys. Rept. 437, 205 (2007). [4] S. D¨ urr et al., Science 322, 1224 (2008). [5] G. Breit and E. Wigner, Phys. Rev. 49, 519 (1936). [6] M. Maggiore, A Modern Introduction to Quantum Field Theory (Oxford University Press, 2005). [7] J. M. Blatt and V. F. Weisskopf, Theoretical Nuclear Physics (Springer-Verlag, 1979). [8] K. A. Olive et al. (Particle Data Group), Chin. Phys. C 38, 090001 (2014) and update for 2015.

[9] D. M. Manley, Phys. Rev. D 51, 4837 (1995). [10] G. S. Gounaris and J. J. Sakurai, Phys. Rev. Lett.21, 244 (1968). [11] G. H¨ ohler, “II. Against Breit-Wigner parameters — A pole-emic”, in Caso et al. (Particle Data Group: Review of Particle Physics) Eur. Phys. J. C3, 624 (1998). [12] A. Sirlin, Phys. Rev. Lett. 67, 2127 (1991). [13] D. Djukanovic, J. Gegelia, and S. Scherer, Phys. Rev. D 76, 037501 (2007). [14] S. Ceci, M. Korolija, and B. Zauner, Phys. Rev. Lett. 111, 112004 (2013). [15] B. C. Liu and B. S. Zou, Phys. Rev. Lett. 96, 042002 (2006). [16] S. M. Flatt´e, Phys. Lett. B63, 224 (1976). ˇ [17] S. Ceci, A. Svarc, and B. Zauner, Phys. Rev. Lett. 102, 209101 (2009). ˇ [18] A. Svarc, M. Hadˇzimehmedovi´c, R. Omerovi´c, H. Osmanovi´c, and J. Stahov, Phys. Rev. C 89, 045205 (2014).