FZJ-IKP(TH)-1999-09
arXiv:nucl-th/9903046v2 25 Mar 1999
ISOSPIN VIOLATION IN THE TWO–NUCLEON SYSTEM
E. Epelbaum, Ulf-G. Meißner
FZ Jülich, IKP(TH), D-52425 Jülich, Germany
E-mail:
[email protected]
I discuss isospin violation in an effective field theory appraoch of the nucleon–
nucleon interaction. The observed leading order charge independence breaking
is explained in terms of one–pion exchange together with a four–nucleon contact
term. A general classification scheme of corrections to this leading order results
is presented and an explanation is offered why charge–dependent pion–nucleon
coupling constants can not play a role in the observed isospin violation in the NN
scattering lengths. Charge symmetry breaking, the size of πγ–exchange graphs
and the pionless theory are also discussed.
1
Introduction
Heisenberg introduced the notion of isospin in 1930 as a label to distinguish
between neutrons and protons when he was studying atomic nuclei. He noticed
that the clustering of stable nuclei around N ≃ Z implied a new (internal)
symmetry. This was later extended to the pion triplet, which is believed to
mediate the long range part of the nuclear force. Heisenberg’s ideas were
used by Breit and collaborators (and others) to establish the so–called charge
independence, observing that after removal of the Coulomb effects between
protons, the pp, np and nn force were equal to a good precision. A weaker
condition is the so–called charge symmetry, which states that the pp and nn
force should be equal with no assertion to the np force. Since that time,
QCD has emerged as the theory of the strong interactions. In the absence of
electroweak interactions, isospin symmetry is observed in the sector of the two
lightest quark flavors for equal masses, mu = md , i.e. the QCD Hamiltonian is
invariant under a global (flavor) transformation of the type
u
u
q=
→ q′ = V q = V
, V ∈ SU (2) .
(1)
d
d
Of course, in nature the light quark masses are not equal. In fact, they differ
considerably, mu /md = 0.55. Still, isospin is a good approximate symmetry
because what counts is not the ratio of the quark masses but rather their difference compared to the typical scale of strong interactions, (md − mu )/ΛQCD ≪
1. QCD thus provides a reason why isospin is such a good approximate symmetry. This can hardly be considered “accidental”. In the presence of electro1
magnetism, further isospin violation is induced since the charges of the quarks
are unequal. These effects are in general small due to the explicit apperance of
the fine structure constant α ≃ 1/137. In fact, in many cases the strong and
electromagnetic isospin–violating effects are of the same size. A good example
is the neutron–proton mass difference. On the quark level, charge symmetry
is a rotation about the 2–axis in isospin space
u
−d
u
2
(iπ/2)τ2
=
, Pcs
=1 .
(2)
=e
Pcs
d
u
d
Our aim is get some insight about isospin symmetry (charge independence)
and charge symmetry in the two–nucleon system based on an effective field
theory approach. On the level of the two–nucleon force, charge independence
implies a potential of the form
Vij = A + Bτ(i) · τ(j) ,
(3)
where A and B are functions of the nucleon spin and momentum (or space)
coordinates and ‘(i), (j)’ labels the interacting nucleons. Charge symmetry
allows for a more general two–body force,
3
3
.
· τ(j)
Vij,cs = A + Bτ(i) · τ(j) + Cτ(i)
(4)
If one works out the Coulomb potential in terms of the nucleon charge matrix
Q = e(1 + τ 3 )/2, it is obvious the electromagnetic effects lead to breaking of
charge independence and charge symmetry,
Vij,Coulomb =
e2
3
3
(1 + τ(i)
)(1 + τ(j)
).
4r2
(5)
I now wish to show how the question of isospin violation can be addressed
in an effective field theory approach to the NN interactions. There are two
equivalent ways to proceed. One is to follow the suggestion of Weinberg 1
and perform the chiral expansion on the level of the potential (i.e. for the
irreducible diagrams) and iterate this potential in the Lippmann–Schwinger
equation to generate the bound and scattering states. The other is to directly
expand the amplitude after performing a nonperturbative resummation for the
S–waves.2 The first path has been followed by van Kolck 3 and I will consider
the second option here.4 Pions are included and treated perturbatively in this
approach. For the topic addressed here, the discussion whether or not pions
should be treated in such a way is inessential. In fact, as shown in section 6,
one can even work with a pionless theory. In that case, however, the physical
interpretation of CIB and CSB becomes much less transparent.
2
2
Facts, explanations and mysteries
It is well established that the nucleon–nucleon interactions are charge dependent (for a review, see e.g.5 and a recent summary is given in 6 ). For example,
in the 1 S0 channel one has for the scattering lengths a and the effective ranges
r
1
(ann + app ) − anp = 5.7 ± 0.3 fm ,
2
1
= (rnn + rpp ) − rnp = 0.05 ± 0.08 fm .
2
∆aCIB =
(6)
∆rCIB
(7)
These numbers for charge independence breaking (CIB) are based on the Nijmegen potential 7 and the Coulomb effects for pp (nn) scattering are subtracted
based on standard methods (for a treatment of this in an EFT framework see
refs.8,9 ). The charge independence breaking in the scattering lengths is large,
of the order of 25%, since anp = (−23.714 ± 0.013) fm. Of course, it is magnified at threshold due to kinematic factors (as witnessed by the disapperance
of the effect in the effective range). In addition, there are charge symmetry
breaking (CSB) effects leading to different values for the pp and nn threshold
parameters,
∆aCSB = app −ann = 1.5±0.5 fm ,
∆rCSB = rpp −rnn = 0.10±0.12 fm . (8)
Combining these numbers gives as central values ann = −18.8 fm and app =
−17.3 fm. Both the CIB and CSB effects have been studied intensively within
potential models of the nucleon–nucleon (NN) interactions. In such approaches,
the dominant CIB comes from the charged to neutral pion mass difference in
the one–pion exchange (OPE), ∆aOPE
CIB ∼ 3.6 ± 0.2 fm. Additional contributions come from γπ and 2π (TPE) exchanges. According to some calculations,
their size is approximately 1/3rd of the leading OPE contribution. For the
TPE, significantly smaller results can also be found in the literature, see e.g.
table 3.3 in ref.5 In case of the πγ contribution, a recent calculation10 gives a
value which is more than a factor of ten smaller than the leading OPE contribution. We will come back to these issues later (see section 5). Note also that
the charge dependence in the pion–nucleon coupling constants (if existing) in
OPE and TPE almost entirely cancel. Naively, one would expect a possible
charge dependence of the pion–nucleon coupling constants to be of a similar
importance as the pion mass difference in OPE. This strong suppression of
charge–dependent couplings has so far been eluded a deeper understanding.
We will come back to this later on. In the meson–exchange picture, CSB originates mostly from ρ − ω mixing, ∆aρ−ω
CSB ∼ 1.2 ± 0.4 fm. Other contributions
3
due to π − η, π − η ′ mixing or the proton–neutron mass difference are known to
be much smaller. It will be shown in what follows that EFT gives novel insight
concerning the size of the TPE and πγ contributions as well as the suppression
of possible charge–dependent πN coupling constants.
3
Effective Lagrangian and power counting
The effective Lagrangian underlying the analysis of isospin violation in the
two–nucleon system consists of a string of terms,
Leff = Lππ + LπN + LN N .
(9)
Each of these terms splits into a strong and an electromagnetic one. The
strong contributions include (besides many others) the isospin–violation due
to the quark mass difference mu − md whereas the em terms stem from the
dynamics of the virtual photons (photon loops and em counterterms). To
include virtual photons in the pion and the pion–nucleon system is by now a
standard procedure.11−15 We note that in the pion and pion–nucleon sector,
one can effectively count the electric charge as a small momentum
√
√ or meson
mass. This is based on the observation that Mπ /Λχ ∼ e/ 4π = α ∼ 1/10
since Λχ ≃ 4πfπ = 1.2 GeV. It is thus possible to turn the dual expansion in
small momenta/meson masses on one side and in the electric coupling e on
the other side into an expansion with one generic small parameter. We also
remark that from here we use the fine structure constant α = e2 /4π as the em
expansion parameter. Because of the additional 1/4π factor (from integrating
out hard photons) one expects the corresponding em constants to be of order
one (of natural size). A dimensional analysis of such terms in the pion–nucleon
sector is given in ref.16 and similarly for the pion sector in refs.11,14 .
We now turn to the two–nucleon sector. One can either expand the potential (irreducible diagrams) or directly the amplitudes. A complication arises
due to the unnaturally large S–wave scattering lengths, 1/a ≪ Mπ . A systematic power counting scheme to deal with this directly on the level of the
scattering amplitude is the recently proposed power divergence subtraction
method (PDS) of Kaplan, Savage and Wise (KSW) 2 .a Essentially, one resums the lowest order local four–nucleon contact terms ∼ C0 (N † N )2 (in the
S–waves) to generate the large scattering lengths and treats the remaining effects perturbatively, in particular also pion exchange. This means that most
a There
exist by now modifications of this approach and it as been argued that it is equivalent
to cut–off schemes. We do not want to enter this discussion here but rather stick its original
version.
4
low–energy observables are dominated by contact interactions. The chiral expansion for NN scattering entails a new scale ΛN N of the order of 300 MeV,
so that one can systematically treat external momenta up to the size of the
pion mass. There have been suggestions that the radius of convergence can
be somewhat enlarged 17 , but in any case ΛN N is considerably smaller than
the typical scale of about 1 GeV appearing in the pion–nucleon sector. It is
straigthforward to extend the underlying effective Lagrangian to include the
strong isospin–violating and electromagnetic four–fermion contact interactions.
Consider first the strong terms. Up to one quark mass insertion, the effective
Lagrangian takes the form (note that terms involving the Pauli isospin matrices
~τ can be obtained by Fierz reordering)
†
2
†
Lstr
σ N )2 + l3 (N † hχ+ iN )(N † N ) + l4 (N † χ̂+ N )(N † N )
N N = l1 (N N ) + l2 (N ~
+ l5 (N †~σ hχ+ iN )(N † ~σ N ) + l6 (N † ~σ χ̂+ N )(N † ~σ N ) + . . . ,
(10)
where h. . .i denotes the trace in flavor space and χ+ contains the light quark
masses. χ̂+ = χ+ − hχ+ i/2 is only non–vanishing for mu 6= md and the ellipsis
denotes terms with two (or more) derivatives acting on the nucleon fields.
Similarly, one can construct the em terms. The ones without derivatives on
the nucleon fields read (possible terms involving the Pauli isospin matrices are
of no relevance for the following)
o
n
†
r1 hQ2+ − Q2− i + r2 Q̂+ hQ+ i N (N † N )
Lem
NN = N
o
n
+ N †~σ r3 hQ2+ − Q2− i + r4 Q̂+ hQ+ i N (N †~σ N )
+ N † {r5 Q+ + r6 hQ+ i} N (N † Q+ N )
+ N †~σ {r7 Q+ + r8 hQ+ i} N (N † ~σ Q+ N )
+ r9 (N † Q+ N )2 + r10 (N †~σ Q+ N )2 ,
(11)
with Q± = uQ† u ± u† Qu† and Q the nucleon charge matrix. All these terms
are due to the integration of the “hard” virtual photons. The “soft” ones are
contained in the pertinent covariant derivatives acting on the pion or nucleon
fields, see e.g. refs.13,15 . We remark that since ΛN N is significantly smaller than
Λχ , it does not pay to treat the expansion in the generic KSW momentumb
Q̃/ΛN N ∼ µ/ΛN N ∼ Mπ /ΛN N
(12)
b So that no confusion with the charge matrix can arise, we denote this parameter with an
extra “tilde” contrary to common use. A pedagogical discussion of this expansion parameter
can be found in ref.18 .
5
(with µ the scale of dimensional regularization) simultaneously with the one
in the fine structure constant (as it is done e.g. in the pion–nucleon sector).
Instead, one has to assign to each term a double expansion parameter,
(Q̃/ΛN N )n αm ,
Q̃/ΛN N ≃ 1/3 ,
α = 1/137 ,
(13)
with n and m integers (m ≥ 1 and n can be negative). Clearly, even if we
work at next–to–next–to leading order (NNLO) which entails a supression by
two powers in Q̃, we need only be concerned with the leading em terms ∼ α
since α ≪ 1. Lowest order charge independence breaking is due to a term
∼ (N † τ 3 N )2 whereas charge symmetry breaking at that order is given by a
structure ∼ (N † τ 3 N )(N † N ). In the KSW approach, it is customary to project
the Lagrangian terms on the pertinent NN partial waves. Denoting by β the
1
S0 partial wave for a given cms energy Ecms , the Born amplitudes for the
lowest order CIB and CSB operators between the various two–nucleon states
takes the form
Mp
(1)
(2)
hβ, pp|Lem
|β,
ppi
=
−
α
E
(µ)
+
E
(µ)
,
NN
0
0
2π
Mp
(1)
(2)
α
E
(µ)
−
E
(µ)
,
(14)
hβ, nn|Lem
|β,
nni
=
−
NN
0
0
2π
(1,2)
where the coupling constant E0 (µ) will be determined later on. From the
renormalization group equations (RGE) it will also be possible to derive the
(1,2)
scaling properties of the E2n (µ) for four–nucleon contact operators with 2n
derivatives. The terms with the superscript ’(1)’ refer to CIB whereas the
second ones relevant for CSB are denoted by the superscript ’(2)’. Higher
order operators are denoted accordingly. There is, of course, also a CIB contribution to the np matrix element. To be consistent with the charge symmetric calculation of ref.2 , one has to absorb its effect in the constant D2 ,
i.e. it amounts √
to a finite renormalization of D2 and is thus not observable.
In eq.(14), p = M Ecms is the nucleon cms momentum and M the nucleon
mass.
4
Isospin violation in nucleon–nucleon scattering
Let me now work out the pertinent 1 S0 phase shifts. The np case has already
been calculated in ref.2 , the leading amplitude being of order Q̃−1 . We want to
calculate the leading em corrections (here, “Cs” means Coulomb–subtracted
and we neglect the small magnetic corrections for the nn system),
∆A = Ann − Anp = ACs
pp − Anp .
6
(15)
2
Figure 1: Relevant graphs contributing to charge independence breaking at leading order
αQ̃−2 . The blob stands for the resummation of the lowest order (N † N )2 contact term ∼ C0 .
The open (filled) circle denotes a pion mass insertion ∼ δm2 (an insertion of the leading
(1)
four–nucleon operators ∼ αE0 ). For charge symmetry breaking, the leading contribution
(2)
is given by the last diagram with the filled circle denoting an insertion ∼ αE0 .
The pertinent diagrams are shown in fig. 1. Besides the contact terms written
down in eq.(14), we have to consider the pion mass difference in one–pion
exchange. OPE diagrams with different pion masses have the isospin structure
3
3
O12 = τ(1)a ∆ab
τ(1) ~τ(2) − b τ(1)
τ(2)
and obviously lead to CIB.
π τ(1)b = (a + b) ~
From the first graph in fig. 1 one easily concludes that these effects are of order
αQ̃−2 since the diagram contains two derivative πN couplings of order Q̃ and
two pion propagators of order Q̃−2 . The notation follows closely the one of
KSW in that the corresponding three amplitudes are called AII,III,IV
. Here,
2,−2
the first (second) subscript refers to the power in α (Q̃) and the superscripts
to the first three diagrams of the figure. This gives
1
4p2
1
−
,
ln
1
+
4p2
m2
m2 + 4p2
"
#
1 + 2ip
1
2p
i
4p2
III
∆A1,−2 = Γ F
arctan
+
ln 1 + + 2 − 2 m 2
pm
m
2pm
m
m + 4p
#
"
2
2ip
m + 4p2
2p
1
1 1+ m
1
i
IV
2
arctan
−
ln
∆A1,−2 = Γ F
+ 2−
m2
m
2m2
µ2
m
2 m2 + 4p2
2
mM A−1
2 gA
, δm2 = m2π± − m2π0 ,
(16)
, F=
Γ = −δm
2fπ2
4π
∆AII
1,−2 = Γ
7
with A−1 ≡ A0,−1 the leading term in the expansion of the np 1 S0 amplitude,
A1 =
1+
C0
C0 M
4π (µ
+ ip)
.
(17)
Furthermore, fπ = 92.4 MeV is the pion decay constant, gA = 1.26 the axial–
vector coupling constant and m2 = 2m2π± − m2π0 the “average” pion mass.
The somewhat unusually looking expressions for m2 and δm2 are derived
in appendix A. Note that while the diagrams II and III are convergent, IV
diverges logarithmically. Therefore, the Lagrangian must contain a coun(1)
terterm of the structure E0 (µ)α(N † τ 3 N )(N † τ 3 N ) to make the amplitude
scale–independent. The leading CIB contribution thus scales as Q̃−2 and for
higher order operators with 2n derivatives we can establish the scaling prop(1)
erty E2n ∼ Q̃−2+n . This does not contradict the KSW power counting for the
isospin symmetric theory since α ≪ 1. The insertion from this contact term is
shown in the last diagram of fig. 1 and leads to an additional contribution to
∆A. In complete analogy, we can treat the leading order CSB effect which is
(2)
due to an operator of the form αE0 (N † τ3 N )(N † N ). This term is, however,
finite. Putting pieces together, this gives for the nn and pp case
A 2
−1
(1)
(2)
I
,
(18)
∆AV1,−2,nn
= −α E0 − E0
C0
A 2
−1
(1)
(2)
VI
,
(19)
∆A1,−2,pp = −α E0 + E0
C0
(1,2)
where the coupling constants E0
tions,
(µ) obey the renormalization group equa-
(1)
µα
2
dE0 (µ)
M (1)
M gA
C 2 (µ) ,
= α E0 (µ)C0 (µ)µ − δm2
2
dµ
2π
48π fπ2 0
µα
M (2)
dE0 (µ)
= α E0 (µ)C0 (µ)µ .
dµ
2π
(2)
(20)
The relation of the pp and nn scattering lengths to the np one reads (of course,
in the pp system Coulomb subtraction is assumed),
(1)
(2)
(1)
(2)
1
1
4πα(E0 + E0 )
=
−
+∆ ,
app
anp
M C02
1
ann
=
1
4πα(E0 − E0 )
−
+∆ ,
anp
M C02
8
(21)
∆ = δm2
2
gA
(−C0 M (m − 2µ) + 8π + C0 M m ln(m2 /µ2 ))
.
12πmC0 fπ2
For the effective ranges, we have only CIB
rnn = rpp = rnp + δm2
2
gA
(C0 M µ + 4π)(C0 M (3µ − 2m) + 12π)
.
6πM m4 C02 fπ2
(22)
Note that the relations in eqs.(21,22) are perturbatively scale–independent and
that the last one does not contain any unknown parameter. We remark that for
the CIB scattering lengths difference the pion contribution alone is not scale–
independent and can thus never be uniquely disentangled from the contact
(1)
term contribution ∼ E0 . While the leading OPE contribution resembles the
result obtained in meson exchange models, the mandatory appearance of this
contact term is a distinctively new feature of the effective field theory approach.
Matters are somewhat different for CSB. Here, to leading order there is simply
(2)
a four–nucleon contact term proportional to the constant E0 . Its value can
be determined from a fit to the empirical value given in eq.(8). Corrections
to the leading order CIB and CSB effect are classified below. We add the
following remarks. Eqs.(21) and (22) are not exact but hold perturbatively.
Similarly, the phase shifts are also defined perturbatively and therefore one
can not have an exact effective range expansion of δ = δ0 + δ1 (referring to the
isospin symmetric case for the moment). If one works e.g. at next–to–leading
order, the norm of the S–matrix is not exactly one, S † S = 1+O(Q̃2 ) 6= 1. From
this it follows that if one adjusts the parameters so as to reproduce the phase
shifts in a certain momentum range, one can only approximately reproduce
the scattering length and the effective range, a ≃ aexp and r ≃ rexp , although
the phase shifts may be very well described. This should be kept in mind.
We are now in the position to calculate the phase shifts. In total, we have
five coupling constants, three of them related to the isospin–symmetric case,
called C0 , C2 and D2 . We fit the paramters C0,2 , D2 to the np phase shift
under the condition that the scattering length and effective range are exactly
(1,2)
reproduced. The two new parameters E0
are determined from the nn and
pp scattering lengths. The resulting parameters at µ = m are of natural size,
C0 = −3.46 fm2 ,
(1)
E0
= −6.47 fm2 ,
C2 = 2.75 fm4 ,
(2)
E0
= 1.10 fm2 .
D2 = 0.07 fm4 ,
(23)
The np and nn 1 S0 phase shifts are shown in fig. 2. To arrive at these curve, the
physical masses for the proton and the neutron have been used. We stress again
(1,2)
that the effect of the em terms ∼ E0
is small because of the explicit factor of
9
α not shown in eq.(23). Having fixed these parameters, we can now predict the
nn 1 S0 phase shift as depicted by the solid line in fig.2. It agress with nn phase
shift extracted from the Argonne V18 potential (with the scattering length and
effective range exactly reproduced) up to momemta of about 100 MeV. The
analogous curve for the pp system (not shown in the figure) is close to the solid
line since the CSB effects are very small.
Figure 2: 1 S0 phase shifts for the np (dashed line) and nn (solid line) systems versus the
nucleon cms momentum. The emipirical values for the np case (open squares) are taken
from the Nijmegen analysis. The open octagons are the nn “data” based on the Argonne
V18 potential.
5
Classification scheme
In the framework presented here, it is straightforward to work out the various leading (LO), next–to–leading (NLO) and next–to–next–to–leading order
(NNLO) contributions to CIB and CSB, with respect to the expansion in Q̃,
to leading order in α and the light quark mass difference. Here, I will simply
enumerate the pertinent contributions which appear at a given order. This naturally also gives an estimate about their relative numerical importance based
on simple dimensional analysis. Based on such an analysis, one can shed some
light on the puzzles discussed in section 2, in particular the size of the πγ
graphs and the role of charge–dependent πN coupling constants. Let me first
consider CIB. In the classification scheme given below, TPE/3PE stands for
two/three–pion–exchange.
10
LO
αQ̃−2
NLO
αQ̃−1 , εQ̃−1
NNLO
αQ̃0
Pion mass difference in OPE,
four–nucleon contact interaction with no derivatives.
Pion mass difference in TPE,
four–nucleon contact interaction with two derivatives,
insertions proportional to the np mass difference,
Pion mass difference in 3PE,
four–nucleon contact interaction with four derivatives,
CIB in the pion–nucleon coupling constants,
πγ–exchange.
Note that there are no CIB effects due to the light quark mass difference linear
in ε = mu − md from OPE and from the four–nucleon contact terms. These
type of terms can only appear to second order in ε. Effects due to the charge
dependence of the pion–nucleon coupling constants, i.e. isospin breaking terms
0
from Lem
πN , only start to contribute at order αQ̃ . Such effects are therefore
suppressed by two orders of Q̃ compared to the leading terms, i.e. by a numerical factor of about (1/3)2 . This finding is in agreement with the various
numerical analyses performed in potential models (if one assumes that there
is indeed some charge dependence in the pion–nucleon coupling constants, see
e.g. the discussion in ref.5 .) We remark that the TPE contribution is nominally suppressed by one order in the small parameter, i.e. one would expect
its contribution to be one third of the leading OPE with different pion masses.
This agrees with some (but not all) meson–exchange model calculations.
a
b
c
Figure 3: πγ graphs. For CIB, the leading contribution comes from the three diagrams a),
b) and c). Due to the Kroll–Ruderman vertex in a) and c) the exchanged pion is charged.
In case of CSB, the dominant contribution comes from b) with neutral pion exchange. c)
vanishes due to parity and a) is suppressed by two additional powers in Q̃. Crossed graphs
are not shown. Also omitted are the initial and final–state interactions between the nucleon.
Some remarks on the γπ graphs (some of them shown in fig. 3) are in order.
First, we note that their contribution for CIB is suppressed by two powers
in Q̃ with respect to the leading terms (the power counting for these graphs
is detailed in appendix B). The larger contribution found in some meson–
exchange model calculations (approximately 1/3rd of the leading OPE con11
tribution) is therefore at first sight at variance with dimensional analysis
since one would expect a suppression of about (1/3)2 .5 However, the much
smaller value obtained in the more recent calculation of ref.10 (performed
in the static limit for the nucleons and Coulomb gauge for the photons),
∆aCIB (πγ)/∆aCIB (OPE) ≃ −1/14.6, is in much better agreement with the
suppression by two powers in Q̃. This deserves further study. Second, the γπ
graphs evaluated in heavy baryon chiral perturbation theory have been used
in a refined potential model calculation.19 The 1 S0 np low–energy parameters
were found to be very little affected. It is also instructive to see how the corrections due to the np mass difference come about. For that, let us write the
nucleon mass as M + δM , where δM subsumes the em and strong contribution
to the np mass splitting, i.e. these terms are O(α) and O(ε). Differentiating
the leading isospin–symmertic amplitude eq.(17) with respect to the nucleon
mass shift gives
A2
∂A−1
= − −1 (µ + ip) ,
(24)
∂M
4π
from which it follows that
δA−1 = −δM
A2−1
(µ + ip) ∼ εQ̃−1 , αQ̃−1 .
4π
(25)
We now turn to CSB. The pattern of the various contributions looks different:
LO
αQ̃−2 , εQ−2
NLO
αQ̃−1 , εQ−1
Electromagnetic four–nucleon contact interactions
with no derivatives.
Em four–nucleon contact terms with two derivatives,
strong isospin–breaking contact terms w/ two derivatives,
insertions proportional to the np mass difference.
We remark that the leading order CSB effects do not modify the effective range
but the corresponding scattering length. Also, the leading πγ graphs (with the
exchanged pion now being neutral) are of order αQ̃0 , i.e. suppressed by two
orders in the KSW expansion parameter. This can be used as an argument to
simply ignore such contributions in the study of CSB. Within meson–exchange
models, their contribution comes out small but there seems to be no consensus
about the actual size.5 A similar hierarchy can also be derived based on the
Weinberg power counting (which applies to the potential) as discussed in ref.3 .
6
The pionless theory
It is also of interest to consider a simplified theory in which even the pions are
integrated out and one only deals with four–nucleon contact terms. Such an
12
approximation is clearly only justifiable in a KSW–like approach as followed
here. We do not want to enter the discussion about how to treat the pions
(perturbative versus nonperturbative) but rather wish to see how accurate the
pionless theory works for the question of isospin violation. In such an approach
the connection to the chiral aspects of QCD is, of course, lost. Integrating out
the pions modifies the coupling constants of the various contact terms like
EFT with pions:
EFT without pions:
C0 , C2 , D2
C̃0 , C̃2
and
and
(1)
(2)
E0 , E0 ,
(1)
(2)
Ẽ0 , Ẽ0 .
In the pionless theory, there is no more D2 term. One has thus four parameters
which can be fixed from the three scattering lengths and the np effective range.
The values of the parameters are
C̃0 = −3.42 fm2 ,
(1)
Ẽ0
= 8.09 fm2 ,
C̃2 = 6.05 fm4 ,
(2)
Ẽ0
= 1.07 fm2 .
(26)
Figure 4: 1 S0 phase shifts for the np (dashed line) and nn (solid line) systems versus the
nucleon cms momentum in the pionless theory. The emipirical values for the np case (open
squares) are taken from the Nijmegen analysis.
Compared to the values for the theory with pions given in eq.(23), we note
that the value of C̃0 almost the same as the one of C0 . This is expected since
(1)
we are close to a fix point. Noticeably, there is a dramatic difference in Ẽ0
(1)
compared to E0 , which is also expected since the former pion contribution
is now completely absorbed in the contact term. The resulting np and nn
13
1
S0 phase shifts shown in fig. 4. The description of the phase shifts has clearly
worsened, they start to deviate from the data for momenta larger than 50 MeV
and are considerably off the data for higher momenta, say about 200 MeV.
These results are not surprising, but it is comforting to see that the pionless
theory indeed describes the small momentum region correctly. For example,
to this order the nn effective range equals the one of the np system which is in
agreement with the data (partial wave analysis). However, the interpretation of
the CIB is now very different since it is entirely given by the isospin–violating
contact interaction. This makes a comparison with the results obtained in
potential models very difficult, if not impossible.
7
Summary
We have shown that isospin violation can be systematically included in the
effective field theory approach to the two–nucleon system in the KSW formulation. For that, one has to construct the most general effective Lagrangian
containing virtual photons and extend the power counting accordingly. We
have shown that this framework allows one to systematically classify the various contributions to charge indepedence and charge symmetry breaking (CIB
and CSB). In particular, the power counting combined with dimensional analysis allows one to understand the suppression of contributions from a possible charge–dependence in the pion–nucleon coupling constants. Including
the pions, the leading CIB breaking effects are the pion mass difference in
OPE together with a four–nucleon contact term. These effects scale as αQ̃−2 .
Power counting lets one expect that the much debated contributions from
two–pion exchange and πγ graphs are suppressed by factors of 1/3 and (1/3)2 ,
respectively. This is in agreement with some, but not all, previous more model–
dependent calculations. The leading charge symmetry breaking is simply given
by a four–nucleon contact term. One can also use the pionless theory to describe the observed CIB and CSB, but the physical interpretation is much
less clear. It would certainly be interesting to investigate such effects in the
presence of external sources.
Acknowledgments
I thank the organizers for providing a stimulating atmosphere and Jerry Miller
for some useful comments.
14
Appendix A: The average pion mass
Here, I briefly show how the average pion mass used in the np system in
section 4 comes about. Consider the graphs shown in fig. 5. The strong
interaction potential for OPE in the nn and pp system is given by the standard
Yukawa form
Vnn = Vpp =
2 (~
~ σ(2) · ∇)
~ e−m0 r
σ(1) · ∇)(~
gπN
,
2
4π
4M
r
(27)
where m0 is a shorthand notation for the neutral pion mass.
n
n
π
n
n
p
π
0
n
p
π
0
n
n
p
+
n
p
Figure 5: OPE graphs for the nn (left) and the np system (middle and right).
For the np case, we also have charged pion exchange,
Vnp = Vnn −1 + 2e−δm r ≃ Vnn [1 − 2δm r] ,
(28)
with δm = m± − m0 and m± denotes mass of the charged pion. So effectively
one has for the np OPE a mass
m = m0 + 2δm or m2 = m20 + 2δm2 ,
(29)
since m2 ≃ m20 + 4δm m0 = m20 + 4(m± − m0 )m0 ≃ m20 + 2δm2 (note that
δm2 6= (δm)2 ). These are exactly the formulae used in section 4.
Appendix B: Power counting for the πγ graphs
In this appendix, we will work out the explicit power counting of the various
πγ–exchange diagrams (cf. fig. 3). First we need the counting rules for the
propagators, vertices etc.: Pion propagator ∼ Q̃−2 , nucleon propagator ∼
Q̃−1 , pion-nucleon vertex ∼ Q̃, photon-nucleon vertex ∼ Q̃0 , charged pionphoton-nucleon vertex ∼ Q̃0 , neutral pion-photon-nucleon vertex ∼ Q̃ and
loop integration ∼ Q̃4 . This gives (displayed is the respective power in Q̃ for
all vertices, propagators, etc for the specified diagram):
15
Graph
3a (π ± )
3a (π 0 )
3b (π ± )
3b (π 0 )
3c (π ± )
3c (π 0 )
Vertices
0
2
2
2
1
2
Propagators
-4
-4
-6
-6
-5
-5
Loops
4
4
4
4
4
4
Total
0
2
0
0
0
1
For CIB, we have charged and neutral pion exchange and all πγ diagrams 3a,b,c
start contributing Q̃0 . In the case of CSB, matters are somewhat different.
Graph 3b is leading at Q̃0 , while 3c is formally of order Q̃ but vanishes due to
parity. The contribution of graph 3a to CSB is suppressed by two additional
powers of the small momentum.
References
1. S. Weinberg, Nucl. Phys. B363 (1991) 3.
2. D.B. Kaplan, M.J. Savage and M.B. Wise, Phys. Lett. B424 (1999) 329;
Nucl. Phys. B534 (1998) 329.
3. U. van Kolck, Ph.D. Thesis, University of Texas, Austin (1993);
U. van Kolck, Few–Body Syst. Suppl. 9 (1995) 444.
4. E. Epelbaum and Ulf-G. Meißner, nucl-th/9902042.
5. G.A. Miller, B.M.K. Nefkens and I. Slaus, Phys. Rep. 194 (1990) 1.
6. S.A. Coon, nucl-th/9903033.
7. V.G.J. Stoks et al., Phys. Rev. C49 (1994) 2950.
8. X. Kong and F. Ravndal, nucl-th/9811076.
9. B.R. Holstein, nucl-th/9901041.
10. J.L. Friar and S.A. Coon, Phys. Rev. C53 (1996) 588.
11. R. Urech, Nucl. Phys. B433 (1995) 234.
12. H. Neufeld and H. Rupertsberger, Z. Phys. C68 (195) 91.
13. Ulf-G. Meißner, G. Müller and S. Steininger, Phys. Lett. B406 (1997);
(E) ibid B407 (1997) 434.
14. M. Knecht and R. Urech, Nucl. Phys. B519 (1998) 329.
15. Ulf-G. Meißner and S. Steininger, Phys. Lett. B419 (1998) 403.
16. G. Müller and Ulf-G. Meißner, hep-ph/9903375.
17. T. Mehen and I.W. Stewart, Phys. Lett. B445 (1999) 378;
nucl-th/9809095.
18. D.B. Kaplan, nucl-th/9901003.
19. U. van Kolck et al., Phys. Rev. Lett. 80 (1998) 4386.
16
View publication stats