~p~ ) the lowest surface wave dominates the sum of all of them according to eq. (55). For pT sufficiently small as compared to p~), we may approximate the exponent in the spheroidal angular functions eqs. (52) and (53): 0
I d0'x/(~ + ½)2+ (k AR sin 0') 2-~ (c~ + 5)0 ~-
+1
(82)
0
Hence we obtain do- 1 4 2 1 ~ = ~ T r R x ] f t . . . . ,,i~g(k, 0)1 o c ~ e -Ap',
(83)
B. Schrempp, F. Schrempp / Tunnelling phenomenon
423
with A = 2 Im al(k)/k. A is a constant as we recall from eq. (38) or from the approximation (39) for a,n and approximately given in terms of pt.(O)/k 2, where pL(O) = R ~ / R T oc k 2. Thus an energy-independent Orear-type exponential behaviour [38] exp (-APT) is predicted, the slope A of which is sensitive to the longitudinal extension RL of the interaction region as well as to the transverse one. Figs. 7 and 8a show elastic pp ~ pp data for a comparison with the predicted Orear behaviour. Indeed, for 1.5 ~
o
Im(O ) ~- f dotal(olin q-l)2 q'- ( k A e sin 0') 2 ,
(84)
o
which is an incomplete elliptical integral of the second kind. Using well-known identities of elliptical integrals, we obtain
=
2kAR sin 2 ½0
+ ~:,,[ 1 +log 4khR+log (tan ~ ½0)] + o(1)
Krn
(85)
B. Schrempp, F. Schrempp / Tunnelling phenomenon
424
10 2
I
I
I
l
o
,,~o
\o exp. (-6.75 pj. )
",,o ,..o
100
o
~o &o •
o •
o
~ 10-2
o
•
o ,~
o
,~,, oc p p ~ p p,19.2 GeVlc
'.9
10-4
t~
E
..\
"U
10-6
O=gO
°
i"
,lO-e
-/ PP~PP
-
Io -'° 0
{:
290 GeVIc 1480 GeVlc
t
I
i
I
l
I
0.5
1.0
1.5
20
25
30
3.5
p~ [ GeV ] Fig. 7. Test for an Orear-type behaviour exp (--APT) in p p ~ p p and ~--po r/n at intermediate PT as predicted from the hadronic tunnelling amplitude. Data from refs. [29, 30, 43].
B. $chrempp, F. Schrempp / Tunnelling phenomenon
425
10-t, ~I.
P P ~ pp p
10-s
d~rmb ]
%
::TL\ 2J
10-6
10-7 10 -8
10-9 10-1o 10-11
PLab=400 GeVlc' FNAL 1 ] ~ V-~ = 23.4-62.1 GeV, ISR i
I
I
=
2
3
I
I
2 PT [GeV/c]
PT [GeVlc]
I
3 b
Fig. 8. p p - ~ p p data at six energies: PLab = 400 G e V / c , FNAL [30] and x/s = 23.4, 30.5, 44.6, 52.8, 62.1 GeV, ISR [29] demonstrating energy-independence of d~r/dt for pT ~> 1.5 GeV/c. (a) The plot of log (PT do'/dt) versus PT exhibits the Orear-type behaviour exp (-APT) for intermediate PT; (b) the plot of log (do-/dt) versus log (PT) indicates a transition to a powerlaw behaviour (p2)-N, N = 8-10, for large pT>~2 G e V / c with a considerable overlap region between the two behaviours. The straight lines are to guide the eye.
with
K,~ = 4 - ~ 1 - - - - ~ )
.
(86)
Thus we find the following power-law behaviour .m) . . . Uing(k,
0)[2oc (k sin1 0) 2 e-2 Im (I(0))
oc(p~.)-~g,.(8)
(87)
where the powers Arm are given by Im
N,. -~ 2 Im m. + 1 -
12
(a,. +~) +1 2kAR
(88)
426
B. Schrempp, F. Schrempp / Tunnellingphenomenon
and the angular distributions g,,,(O) by
gin(O) = D,,(1
+cos
0) 2Nm 2,
(89)
with constants Din. The powers Nm do, in fact, not depend on the energy k: from eq. (37) we see that Im (a,, +½)2/(kAR) is asymptotically, for large k, given by (kR~/RL) Im (1 + y2); Y,,, depends only on the product (m-])RL/(kR~) according to eq. (38) and RL/(kR2)~-c is constant. Thus for small values of m the power N,,, depends on c and m; however, for sufficiently large values of c(m-¼)~r we find ~2m +½, Nm c(m 1/4)'rr large
(90)
depending exclusively on the index m, which counts the surface wave orbits. To the extent that the numbers c and m enter the power Nm as well as the fixed PT slope A,,, = 2 Im a,,(k)/k, these two quantities are related (in contrast to parton models!). The functions gin(O) look reasonable; they are roughly constant for 0 ~<45 °. Their precise form cannot, however, be compared immediately to data at very large angles. One first has to clarify the question of s - u crossing properties, which differ for different reactions. Incidentally, it is, however, interesting that the s - u crossed version of the tunnelling amplitude (87) becomes almost angle independent for fixed Pv
~,m) "k 2 2Nm(l +COS 0) 3 , ] f t . . . . llingt , 0)[ Is.~u~PT
(91)
where (1 +cos 0) 3 varies only by a factor 8 from 0 = 0 to 0 = 90 °. Finally let us recall that also in this kinematical region of large PT the lowest surface wave contribution dominates the cross section. We emphasize that the requirements Rx ~ const and RLCC k are at the same time responsible for energy independence at fixed pT in region (iii) and the powerlaw behaviour for fixed angles in region (iv). Can we understand intuitively how a power law can result from our approach? We notice that for diffraction into any fixed angle 0 and for k ~ oo, the incoming ray of near grazing incidence will have travelled along the surface (as a surface creep wave) all the way up to the tip of the spheroid; the coordinates of the point P(xo, Zo; 0) where it is radiated off tangentially (see fig. 6), are easily calculated: 1 xo(O, k)---cot 0~-~ k ~ 0,
Zo(0, k)~-RL 1 - 0
k~oo
and the longitudinal radius of curvature in P(xo, Zo; 0) is [see eq. (70)] 1 1 0. 0L(0, k ) - s i n 3 0 ck k ~
(93)
B. Schrempp, F. Schrempp / Tunnelling phenomenon
427
This means, that for fixed 0 the outgoing rays will be radiated off the tip of the spheroid, where the transverse coordinate Xo, as well as the longitudinal radius of curvature pL, have natural dimension and tend to zero like 1/k for k ~ co. Thus, in the fixed-angle limit the complex surface ray (wave) probes the pointlike tip of the spheroid before being radiated off. Let us remember from eq. (65) that the integral I,,(0), eq. (85), is in fact approximately given in terms of the integral o
k
1/3
dO'(pL(O')) 1/3 "
(94)
0
We then realize that the integration between 0o # 0 and 0 leads to a scale-invariant result [~ gin(0)]; while the contribution from the lower integration limit retains some memory of the region 0'~ 0, where the scale Rx is important, resulting in breaking the scale invariance by the powers (RZp2) -u". To the extent that there exists a possible relation between the spheroidal interaction region and the long intermediate colour electric flux tube in bag-bag scattering (as speculated on in subsect. 4.3), we can vaguely reconcile our geometric ideas with parton ideas: probing the pointlike tip of the spheroid means probing the tip of the intermediate flux tube, where indeed the quarks are situated. In this sense, the scale-invariant portion of the integral I,,(0) obtained from integrating over 0'~> Oo> 0 simulates a scale free pointlike interaction with quarks. If this is a correct interpretation, we may infer that the mechanism which confines the quarks at large energies to the pointlike tips of the intermediate fat string is of the same crucial important at large px as it is at small pw. In summary, it is the geometry of the stringlike confining walls with pointlike tips and not the interaction of quasi-free pointlike constituents, which gives rise to the power-law behaviour in our approach.
5.2. Blurring the boundary Up until now we have only discussed the idealized case of a caustic generated by a discontinuity of the potential at its boundary ~ = s%, i.e., at p = PB = AR sinh s% = Ra-. In this section let us investigate the more general case of a complex caustic due to a complex pole surface of the potential, which allows the potential to have a smooth drop along the real ~, viz., p axis (blurred boundary). We consider a potential U(k, x) exhibiting the same spheroidal symmetry as our previous sharp-boundary potential with focal distance
2AR = 2ckR~
, ce,
k~oo
with Rx ~ 1 fm,
c << 1.
(95)
The wave equation (1) is separable with the following form of the potential
u(k, AR sinh () u(k, AR sinh ( ) |--//l+l\ -U(x, k) = (AR)2(sinh 2 ~C+sin2 0) = 2AR cosh ( \ r l r 2 /
'
(96)
B. Schrempp, F. Schrempp / Tunnelling phenomenon
428
where rl and r~ are the distances of the point x from the two foci, respectively [thus U(k, x) exhibits a Coulomb-type behaviour with respect to each focus at short distances], u(k, AR sinh sc) is some short-range, spheroidally symmetric, reduced potential. The resulting angular equation remains equal to the free equation (19), while the radial equation leads to a radial action
Se = I d(#(khR sinh ~)2
-- (/~ +1)2 __ u(k,
AR sinh sc)
or,
S=_So=_~f
1
[ 2
2 u(k,p)
(97)
in terms of the transverse variable p = AR • sinh sc. It is obvious that, in general, the potential u(k, p) will introduce (if at all) turning points in p, which do not only depend on the geometrical extensions but also on the coupling strength of the potential. In order to obtain a turning point close to the periphery of the potential independent of its coupling, we need a strong variation of the potential nearby, as typically provided by a (complex) singularity (or discontinuity). The simplest example for a singularity, admitting a blurred, fuzzy boundary is a complex pole in the variable s¢ or equivalently p. Let the pole position be at = sCo,
with
AR sinh £o = po,
(98)
situated close to the periphery of the potential Re po = RT = 1 fm Im Po ~ RT,
(99)
where RT is now the mean transverse range of the potential and Im po a measure for the fuzziness of the boundary. We then may write near the pole 2
u(k, p ) ~ -
UoPo
2
2. P -Po
(100)
For the sake of simplicity let us assume for the time being that Uo = uo(k) is independent of p, even though this is not really a short-range potential. A potential with a complex pole has been discussed extensively in ref. [41] for the spherical case. We proceed analogously and present here only a sketch of the arguments. The potential introduces two turning points pl, p2 with 2 # +po+ pl,2 = 2
f12 - -
-po~-~ •
(101)
429
B. $chrempp, F. Schrempp / Tunnelling phenomenon
We are interested in the case ,
=
ip21_p2o[ =
p2_p2o
< ~
~
.
(102)
Then the turning point at p2 is close to the pole at po and pl close to/3, i.e., the positions of the two turning points become independent of the strength of the potential. One obtains again a semiclassical series involving complex multiple reflections between the two turning points P2-Po and pl =/3, in complete analogy to the series obtained from multiple reflections between RT and fl in the sharp boundary case (cf. subsect. 3.2). Hence tunnelling (generalized surface creep waves) will occur along a complex caustic p(x)=po, which is again situated close to the periphery of the potential. This tunnelling phenomenon corresponds again to the contributions from complex poles in the A-plane, or equivalently the/3 = (a + ½)/k plane. The radial action between pl and p2 takes the form P2
do p2 /32.~ 2
Sp(P2,Pl)~
2 P -Po
forlPl,2] 2<
Pl P2
k lap
AR
P -p
P -Po
Pl Do
= k [ dPx/-~--~-P~+O(E21°ge2)+S°(P2'P°) ~R
O1
,-,free l
=~
tPo,/3) + O(e 2 log e 2) + O(e2),
(103)
where Po
8
The Bohr-Sommerfeld type quantization condition analogous to eq. (36), which determines the complex pole positions a,, in the a-plane, is given by [41] t
k
ofre¢
Sotpz, p l ) = a o
(po,B)
~-(m-1)rr+½ilog[iF(1So(po, p2))] ,
(105)
for m = 1, 2 ..... Compared to the quantization condition (36) for the sharp boundary case, we find (i) the real position of the discontinuity, p = RT, replaced by the complex pole
B. Sehrempp,F. Schrempp / Tunnellingphenomenon
430
position p =Po; (ii) m _1 replaced by m _1; and (iii) the additional correction term log iF, which is due to the presence of a pole in addition to the two turning points. F(rt) is the following function [41] F(r/) =
{
27fir/exp (-27/log (e-i~rl)/e) F(1 - rt) 2
=-i, forlr/l>>l and ~ 2~'ir/, for r/-->0.
0
Furthermore, using eq. (102), we obtain Uo
Po
1
sp(po, p2)-~~AR 4/32-p~4~''
(107)
For very small values of this action, the log term in eq. (105) becomes important; (this leads to the Born approximation, where the amplitude becomes proportional to the coupling strength Uo of the potential). In order to obtain independence of the coupling Uo, we shall require
~llog iFI << (m -
½)~r.
(108)
This is the non-perturbativity requirement analogous to (62) and (69) in case of the sharp boundary. With this additional constraint the pole positions am(k) in the complex A-plane are situated at
a,,,(k)+½=k/3,,,
-2 ~- kpox/l+ym,
k la r ge
(109)
where ym is defined like y~ in eq. (38) with (m -¼)RL/(kR 2) substituted by ( m ½)RL/(kp2o) with RE ='JAR 2 +po2. With this equation the two inequalities (102) and (108) for Uo become tightest for m = 1. Assuming that the bound (108) is well fulfilled by 0.25 <<(m -½)zr, we obtain (for a favourable choice of phases and for realistic values of k]polE/AR and 3;x) 0.015 ~<
<<1.
(110)
Within this range of coupling strength uo the tunnelling amplitudes, given by the contributions of the Regge-like poles at am, are non-perturbative; they depend only on the parameters characterizing the complex pole position Po of the potential close to its periphery and not on Uo. Roughly speaking, the complex pole at p = po acts like a complex sharp boundary at p = po. It is obvious from this analysis that a genuine short-range potential with a complex pole at p = po will lead to similar results (pl ~/3 for Re/3 > RT will even be obtained without any conditions on the potential strength). After this exercise the following generalization is very plausible: given a shortrange potential which
B. Schrempp, F. Schrempp / Tunnelling phenomenon
431
(i) exhibits spheroidal symmetry as given in eq. (96), and (ii) generates a turning point at p = R with Re R = RT-~ 1 fm, which only depends on geometrical properties of the potential (like extension and fuzziness of the boundary given in terms of Re R and Im R, respectively) and not on its strength, then we obtain a tunnelling phenomenon along a caustic p(x)----R, which is due to complex multiple reflection between the complex turning points R and [3 = (~ + ½)/k. This tunnelling phenomenon is non-perturbative if a lower limit for the potential strength is observed. This limit depends on the nature of the singularity of the potential. The intuitive interpretation of tunnelling by means of surface creep waves can be maintained. All our results obtained in the sharp-boundary case - eqs. (45), (51), and (52) as well as their qualitative features extracted in subsect. 5.1 (Orear-type behaviour, power-law behaviour, etc.) - remain v ~ R remains unchanged, RT is substituted by R and RL is substituted by ~/AR2+ R 2. This in fact justifies our extensive treatment of the sharp-boundary case. Furthermore, we have seen that the dependence of the tunnelling amplitude on the quantization index m may vary slightly according to the generating mechanism of the caustic, so we should not take it too literally in a comparison with data. Of course this whole discussion may be extended to potentials with more than one complex pole in p; however, the one situated closest to the real axis will dominate the behaviour for pT ¢" 0.
5.3. Hadronic tunnelling: the shadow of copious particle production We are now ready to make our hypothesis of maximal importance of particle production (see sect. 2) more concise. So far we have almost exclusively discussed the tunnelling component of the quantum mechanical scattering amplitude f(k, O) and established a very promising agreement with the characteristics of high-energy hadronic diffraction data out to large pT. Two points are crucial for the following discussion. (i) Remember that the tunnelling component is due to incoming rays with (quantized) impact parameters Re [3m ~ R T
(111)
corresponding to complex poles with positions [3,. = (am +~)/k in the [3 plane. RT is the mean transverse range of the potential, i.e., the mean transverse extension of the interaction region. (ii) The remaining classical component of the amplitude f(k, O) arises from contributions of saddle points [3s(0) with Re[3s(0) (<- ) R-r
(112)
where Re [3s(0) specifies the impact parameter of the classical ray scattered into the angle 0. More precisely, in the case of a spheroidal cut-off potential (and refraction
B. Schrempp, F. Schrempp / Tunnellingphenomenon
432
index N-~ ioe), we have a real saddle point at /3s(0)-
RT cos ½0
< RT,
(113)
x/1 + (AR/RT)= sin 210 (
being near RT only for 0 = 0. For a complex pole in the potential at p = po the saddle point becomes complex; it is obtained from/3s [eq. (113)], by substituting RT by Po (as has been shown in ref. [41] for the spherical case, AR = 0). In a strict quantum mechanical framework with complex potentials, even for so-called total absorption (Re N ~ 1, Im N ~ 0 such that k Im N ~ oc for k ~ oo; N = refraction index), this "classical" term dominates the amplitude for sufficiently large angles 0 (i.e., outside the Fraunhofer diffraction domain). Our hypothesis of maximal importance of particle production, which leads beyond the quantum mechanical framework, states: whenever the two incoming hadrons interpenetrate each other classically, i.e., whenever an incoming ray impinges on the scatterer with the shape of the hadronic interaction region, the intensity goes into particle production. This means that the (real and complex) semiclassical contributions due to incoming rays with impact parameter Re/3 < RT are totally absorbed. This leaves us with the following amplitude for elastic hadron scattering: (i) it consists of the full (spheroidal) tunnelling component at all angles, as rigorously calculated in the quantum mechanical framework; (ii) it obtains a contribution from the saddle-point component only near 0-~ 0 (where Re/~s = RT) to make up for a Fraunhofer-type behaviour by conspiring with the tunnelling component (cf. subsect. 3.2). There is an immediate consistency check. We have seen in subsect. 3.2 that as long as the saddle and the tunnelling components are of comparable order of magnitude they interfere and produce Fraunhofer diffraction minima at angles, where
JI(kRTO)=O,
1 i.e. pT~--kO~-~(3.84+nrc),
for n = 1 , 2 , 3 .....
(114)
Diffraction minima are then, according to our interpretation, signals of a residual effect of the saddle point contribution, which is dying off rapidly for increasing angles. High-energy pp ~ pp data do indeed show a single dip [15, 29] at p r 1.15 GeV/c, which is compatible with being the first diffraction minimum for RT 0.7 fm. This is consistent with an amplitude, where the saddle-point term starts from its full Contribution at 0 = 0, then decreases at least as fast as the tunnelling component, which fully takes over beyond PT = 1.5 GeV/c; hence the absence of further minima in the data [15, 29, 30]. Let us finally emphasize very strongly that our absorption prescription to account for multiparticle channels is completely different from the traditional one in
B. Schrempp, F. Schrempp / Tunnelling phenomenon
433
hadron physics, where impact parameter amplitudes oo
b)oz j d(kO)Jo(kOb)f(k, O)
(115)
O
are absorbed for real impact parameters b < RT. (Both, the classically allowed saddle-point component and the classically forbidden Regge-like pole component, contribute to [(k, b) for all real values of b in an untransparent manner; their separation is only achieved by going to complex values of b by means of a Sommerfeld-Watson transformation).
6. Semiquantitative comparison with data
6.1. Scattering off nuclei at i'ntermediate energies." tunnelling along a spherical caustic In proton scattering off heavy nuclei the effective radii are much larger than in hadron-hadron scattering. Typically RA-- 1.13 fm' A 1/3 ,
(116)
where A is the atomic number of the heavy nucleus. Correspondingly we might expect three energy regions with different physics: (i) k ~ 100 MeV, where k R g ~ 1, such that the semiclassical approximation is not yet applicable (k = c.m.s, momentum); (ii) k - a few GeV, where kRA >>1; the interaction region will still be about spherical (with radius RA) due to the low energy; (iii) k ~ 50-100 GeV; the long hadronic interaction region will have developed [36] as in hadron-hadron scattering. Region (i) certainly has nothing in common with high-energy hadron-hadron scattering; region (iii) should show similar features as hadron-hadron physics, which is so far nicely borne out by the data [37]. Region (ii), where particle production starts to become important, should show a significant tunnelling component coming up at larger angles. However, the singlar surface responsible for such a "nuclear tunnelling" should be spherical with radius RA. Assuming to first approximation a surface caustic, leaves us with a parameter-free description of the tunnelling amplitude as obtained from eqs. (45), (51) and (52) in the spherical limit, AR-~0, with RT=--RL=--RA given by eq. (116). There exist high-precision data [42] for pA ~, pA scattering at one energy E l a b = 1.98 GeV for A = 1=C, 4°Ca, 48Ca, 58Ni, 2°sPb, where kRA varies roughly from 20 to 60. Fig. 9 shows the elastic data plotted versus kRAO. The forward peak is followed by an exponential decrease in this variable, modulated by regular oscillations, which follow the diffraction pattern (114): minima at kRAO = 3.84+ nTr for n = 0, 1, 2 ..... They signalize an interference between the tunnelling and the
B. Schrempp, F. Schrempp / Tunnelling phenomenon
434
10 ~
10~
10: 10: m
E 10: •o
I(31 10c
10-I
10-2 0
5
10
15
20
kRecm
Fig. 9. Data [42] of elastic scattering of protons off heavy nuclei (with atomic n u m b e r A) at ELab = 1.98 G e V versus kRAO are shown (RA = 1.13 fm A 1/3 radius of the nucleus). The oscillations follow a diffraction pattern with minima at the positions of the zeros of JI(kRAO)/(kRAO). T h e forward peak is, on the average, followed by an exponential decrease which exhibits strong antishrinkage with increasing kRA. •
=
saddle-point contributions; the decrease in amplitude of these oscillations for increasing kRAO is, according to our interpretation, due to the tunnelling component becoming dominant. It is quite consistent that we should find a saddlepoint component, which is non-negligible, since at these energies absorption into particle production should be weaker than at very high energies. A crucial test for a strong presence of the tunnelling component is for the peculiar (kRA) dependence of the exponential slope predicted from eqs. (39) and (54) to be e -c°nstant
(kRA)I/30
(117)
in the domain of dominance of the lowest (or the lowest few) surface wave*, i.e., • This behaviour is expected for all sufficiently large angles; for a spherical caustic there is, of course, no power-law behaviour.
B. Schrempp, F. Schrempp / Tunnelling phenomenon I0 S
' •
10z'
I
I
I
I
I
,.b
-~"~,
10~
""
'
I
'
p+A-- p+A
" ~ E
.
I
435
=1.98GeV fermi
R= 1.13A 113
•
%,
103
,~,
-
10
•
..'x.
10 2
'.,
m
SeNi
•
E 10~
•
%
10~ "ID
•.
1010 °1
1~2
-..."~.
"~'~~Ca
I
I
G2
I
[
O~
I
[
O~
(kR) v3
I
l
lIB
em
I
I
tO
i
I
12
I
I
I~
Fig. 10. Same data as in fig. 9 plotted versus (kRA)1/30. They exhibit, on the average, an exponential behaviour exp (-B(kRA)I/30) with universal slope B as predicted by the nuclear tunnelling amplitude (corresponding to a spherical caustic with radius RA). The straight parallel lines are to guide the eye.
more or less right after the diffraction peak; the constant has to be universal for all data. In a plot log d e / d O versus the variable (kRA)I/30, fig. 10, we see that each data set may be averaged by a straight line with a universal slope! In fig. 9, versus kRAO, the same data showed a considerable antishrinkage; so this is a highly nontrivial result. A last test is to compare the average behaviour of the data with the parameterfree prediction for the tunnelling amplitude with respect to shape and normalization. Fig. 11 shows the data in a plot of RA 2 dtr/df~ versus (kRA)I/30, where,
B. Schrempp, F. Schrempp / Tunnelling phenomenon
436
10 2
p*A
~
p*A
E~ab : 198 GeV R : 113 A 113 f e r m i 101
"i I0 o
b ill
~A
10-2
°°°° o °o
10-3
o
p b z°a
• o
NI ~ C a ~e
• •
C a t,O C 12
e I o o o Oo °
10 -(
02
0.4
0.6
0.8
1.0
1.2
1.4
(kR) ~30cm
Fig. 11. The same data as in fig. 9 are shown. The parameter-free prediction of the nuclear tunnelling amplitude (solid line) is compared with the average behaviour of all data with respect to shape and normalization. according to our prediction, the different data sets should (on the average) fall on top of each other [this prediction is true to the extent that (am +I)/(kRA) ~independent of kRA]. Fig. 11 shows the prediction for 12C of surface waves due to a surface caustic with radius rn~2
which perfectly accommodates the average behaviour of all data! (The lowest surface wave was excluded, since it gave a somewhat too flatfish behaviour at large angles; we remember, however, from subsect. 5.2, that the starting value for the surface wave index is not sacred, since it may vary slightly with the nature of the mechanism generating the caustic.) 6.2. High-energy pp ~ pp scattering
A qualitative comparison with key features in pp-* pp scattering has already been performed in subsect. 5.1. Let us now compare the pp ~ pp data [29, 30] at representative energies Plab>~ 400 G e V / c with the predictions for the tunnelling amplitude on a semiquantitative level. We ignore the contribution from the classical
B. Schrempp, F. Schrempp / Tunnelling phenomenon
437
saddle-point component completely, since its parametrization would have to depend on all details of the interaction (and thus pollute this comparison); correspondingly, we cannot expect to describe the px = 0 behaviour, nor the dip region (see subsect. 5.3). We do expect, however, a quantitative description of the data beyond the second maximum, i.e., for pv~> 1.5 GeV/c. For a spheroidal sharp boundary there are only two free parameters of purely geometrical origin, RT and c, where the constant c is a measure for the eccentricity of the spheroid (remember that the focal distance is 22~R with ~R = ckR2). For a complex spheroidal pole surface at p(x)=R as discussed in subsect. 5.2, we have three free parameters: Re R = RT, Im R and c as defined by AR = ckR 2. We typically expect R T = I fm,
c<<1,
_<1
arg R -,:~rr.
(118)
We performed eye-ball fits to the data. The more realistic prediction from a complex spheroidal pole surface (for the sum over all surface wave contributions) gives an excellent solution for pv ~> 1.5 GeV/c and reproduces the strong forward peak semiquantitatively, as displayed
102
PP-'- PP lo o
~
-
-
Tunnelling cu'nplitude
~
10-2 r'&---i
PLab = 4 0 0 GeVlc, F N A L
:¢
(.9
Vg= 52.8 GeV } V'~ = 62 GeV,pT< 19 GeV/c ISR
10-4 r
"o
i0-6 r
10-8
10-lo
z
0
16
_t [G.v2] Fig. 12. Semiquantitative comparison of the prediction (solid line) from the hadronic tunnelling amplitude (corresponding to a fat string-like caustic with RT = 1 fm and RLOC~/s) with ISR data [29] at ~/s = 52.8 and 62.1 G e V and the F N A L data [30] at PLab = 400 G e V / c for pp ~ pp. The amplitude exhibits an Orear-type behaviour for intermediate PT and turns automatically into a power law at large PT.
B. Schrempp, F. Schrempp / Tunnelling phenomenon
438
in fig. 12. Thus shape and normalization of the data may be understood just in terms of three geometrical parameters, which come out to be of the expected order of magnitude RT = 1.1 fm,
argR =0.45,
c =0.086.
(119)
The asymptotic power 1 Im (aa +1) 2 N1
2
kAR
+1
(120)
in the power law do-/dtoc (p~)-Ulg(O) for PT >>2.5 GeV/c comes out to be N1 =
10.2
(121)
(surprisingly close to its value N = 10, expected from the constituent counting rule [40]). An example for an eye-ball fit to the predictions for the real spheroidal boundary has already been presented in ref. [10]; it is of almost comparable quality, if the first two surface waves are omitted, and RT and c are of similar order of magnitude as above (RT = 0.81 fm and c = 0.158). A satisfactory fit may also be obtained with only the lowest surface wave omitted. Let us emphasize that RT and c are of the same order of magnitude as the corresponding parameters calculated for a colour flux tube with colour triplets at its ends [9] [cf. subsect. 4.3, eqs. (78) and (79)]. In summary, one can state that the tunnelling amplitude has just the right kind of angular decrease and normalization as shown by the data; it provides a uniform semiquantitative description from small PT out to the largest PT available.
7. Further interesting applications 7.1. Extension to quantum-number exchange reactions An extension to 2-~ 2 quantum-number exchange reactions is most direct and intuitive. Each such reaction amplitude may be approximately written as difference of two or more elastic amplitudes [using isospin invariance, invariance under SU(3) or some higher symmetry], which become identical for k - ~. Those elastic amplitudes are given as sums over all surface waves corresponding to spheroidal interaction regions with essentially the same dimensions. Slight differences in the elastic amplitudes arise, since they are shadows of slightly different inelastic channels. Inelastic channels obtain contributions from rays with impact parameters/3 <~RT, surface waves from those with/3 ,> RT. Thus the differences of the elastic amplitudes will naturally be in the amount of contribution from the lowest surface wave (or the lowest few ones) coming from/3 ~ RT. Hence the exchange reaction will
B. Schrempp, F. Schrempp / Tunnelling phenomenon
439
receive (energetically suppressed) contributions from the first (or the first few) surface waves. Correspondingly, one expects that (i) it does not show the enormous forward peak of elastic reactions, which is due to the conspiration of infinitely many surface waves; (ii) instead the Orear-type behaviour of the cross section, exp (-2 Im (al/k)pT)/PT continues from intermediate PT down to small values of PT; and (iii) for sufficiently large values of PT, where also the elastic amplitudes are dominated by the lowest surface wave, elastic and exchange cross sections should show the same PT dependence (apart from normalization). If one compares the data [43] for 7r-p~ on at 40 GeV/c with the elastic ISR data for p p ~ p p in fig. 7, one finds all these predictions confirmed. Thus we have made plausible that the 2 ~ 2 quantum-number exchange reactions are dominated by the lowest surface wave, i.e., by one single direct channel Reggelike pole al(k) (in the A-plane and not in the J plane!), which is peripheral: Re (a l(k) + ½)/ k = Re fl 1---RT ~ 1 fm. This le ads to a peripheral impact parameter profile for the imaginary part of the amplitude. A similar reaction mechanism has already been proposed by us [44] and other authors [45] and has been successfully compared with the data of a large number of reactions.
7.2. Relation to dual models In dual models 2 ~ 2 quantum-number exchange reactions are built up by a sum over a parent Regge pole a(s) in the direct channel angular momentum (J) plane with approximately linear trajectory and infinitely many parallel daughter trajectories. In subsect. 7.1 we have seen that in our framework such an exchange reaction is described by one single Regge-like pole in the A-plane, al(k) (remember that k-~ ½x/s). It is very instructive to investigate (asymptotically for k -* oo) the spectrum in J generated by one such Regge-like pole, a(k) say, in the A-plane. To this end we need the expansion [25] of the spheroidal angular functions C(1) ,,on in terms of Legendre polynomials co
(1) Son (kAR, cos O) =
~t
d°n(kAR)Pj(cosO)
J=0,1
1
where the sum }~' extends over [ odd / J for
,
(122)
{even
odd / n. Using eqs. (22) and (122)
we obtain the relation between the spherical (f(J, k)) and the spheroidal (fA(n) (kRT, kRL)) partial-wave amplitudes +1
f(~, k)=l(kRT) 2 J dzPj(z)f(k, O)
[z =cos 8]
-1 =
E'
oo ,,=o,~
(A(n)+½)d J+½ ~nn) fx(,,)(kRT ' kRe)d°"(kAR),
(123)
B. Schrempp, F. Schrempp / Tunnelling phenomenon
440
/even/ /even/ where the sum Y~' runs over t odd J n for ! odd J J' A single pole of fx~,~ at A = c~(k) may be written as -
- ~(k)'
(124)
where r(k) is the residue as defined in eq. (46). Expanding ,~ near the pole ,~ = c~+ (n - v(~))d-~-~ (In
,
(125)
I n = v(c,)
we obtain oo
(J+½)f(J, k ) =
.
2' d°"(k AR) =o,1
n -
F.
~,(a (k))
"
(126)
Using eqs. (47) and (85) (for 0 = lzr) we obtain
.(~ (k))---- ~,(s) k =~./%-~o .o + .'s + 2 ( R e K) log (v's)+i N - 1 log (v's),
(127)
where x (= Xl) is given in eq. (86) and the large PT power N (=N1) according to eq. (88) by N - 1 = 2 Im K.
(128)
The intercept vo is 1 2 27re vo = - ~ + ~ K l o g - - ' K
(129)
the slope v' v'=
AR = R 2 c T O.5GeV_2 2¢rk 2rr
(130)
[using the best-fit values for c and RT as obtained in subsect. 6.2, eq. (119)]. Thus the n-trajectory v(s) is asymptotically a linear trajectory with reasonable slope and with an imaginary part increasing like log s. From ref. [25] one knows that the series of coefficients d o" decreases like (lkaR)J
d°"(kAR)-const F(j+3------~,
for J > n .
(131)
Hence we obtain an approximately linear parent trajectory at J ~ v(s) accompanied by infinitely many parallel even daughter trajectories. Although there are ancestors, they decouple stronger than exponentially!
B. Schrempp, F. Schrempp / Tunnelling phenomenon
441
Thus one single Regge-like pole in the a-plane at the same time (i) simulates [according to eqs. (28) and (51)] for sufficiently small angles a single, effective, peripheral Regge-pole in the J plane with a (s)cc x/s; (ii) leads to a power-law behaviour of the amplitude at large angles; (iii) gives approximately the kind of spectrum as expected from dual models.
7.3. Glory effect in hadron physics? In the scattering of visible light with wave number k off water droplets with radius R (in a cloud or in fog) one observes, for kR = 30-1000, the beautiful natural phenomenon [26] of the so-called glory effect. The glory effect is a very strong enhancement of the backscattered intensity (by several orders of magnitude) resulting in a steep narrow backward peak in the observed angular distribution. This phenomenon has been understood as late as 1969 by Nussenzveig [21] in terms of a subtle enhancement due to surface waves creeping around the surface of the water droplet. Unfortunately, the normalization of the backward peak depends on the refraction index [21], i.e., the strength of the potential; however, its angular dependence (in the spherical case) is given by [21] intensity cc 0~
IJo(ke(Tr-O))l 2,
(132)
only depending on kR. Backward peaks in scattering off atoms and nuclei have already been attributed to glory and have been successfully compared with the predicted angular dependence (132) (see, for example, ref. [46]). In hadron reactions we have backward peaks as well [and many of them show in fact a dip at - u = k2(Tr- 0) 2 -~ 0.2 (GeV/c) 2 which, for R -~ 1 fm, coincides nicely with the position of the first zero of the Jo Bessel function], More recently data [47] at rather high energies are available with kR ~ 30. They show very steep backward peaks which cannot at all be understood by conventional u-channel Regge exchanges. It is quite likely [48] that one is observing hadronic glory there. We are very grateful to T.T. Wu for many inspiring discussions and his constant encouragement at various stages of this work. Furthermore, we thank H.B. Nielsen, G. Preparata and K.A. Ter-Martirosyan for fruitful and critical discussions. We are also indebted to T. Ericson and R. Viollier for helpful advice on the nuclear physics aspects of our approach. Finally, we thank K. Winter and K. Schubert for having made available to us the latest pp + pp data of the CHOV collaboration prior to publication. Appendix A In this appendix we shall summarize the essential steps in the derivation of the exact partial-wave solution of the problem treated in subsect. 3.2: scattering of a
B. Schrempp, F. Schrempp / Tunnellingphenomenon
442
plane wave off an infinitely deep (prolate) spheroidal potential well (with half axes RT and RL). The appearing spheroidal wave functions [24, 25] will be defined and the Sommerfeld-Watson transformation used in subsect. 3.2 will also be given. The stationary wave (Schr6dinger) equation [A+k 2 - U(x, k)]O(x, k)=0
(A.1)
reduces for this case to the free equation (U(x, k)=-0) outside the spheroidal boundary and the boundary condition O(x, k)=-0 on the boundary. For the sake of generality, in the following derivation of the spheroidal partial wave expansion, we shall allow for a plane wave of arbitrary incidence. The special case of axial incidence - relevant to subsect. 3.2 - will be considered at the end. In terms of the (prolate) spheroidal coordinates (17) the (free) wave equation separates into the following radial and angular equations:
(kAR sinh~)2-(,~ +½)2 s i - - ~ J x ~/s~nh~R(kAR, cosh ~:)= 0,
(A.2)
[d~2+(kA R sin 0)2+(a + ½)2- si---~m2-¼]j
× 4s-~n0 S(k±R, cos 0) = 0.
(A.3)
The equation for the azimuthal angle (&) dependence is identical to the spherical case with solutions cos m~b,sin m4~,
m = 0, 1, 2 .....
(A.4)
The quantity A + ½is a separation parameter analogous (but not identical) to the angular momentum in the spherical case. For an infinite, discrete set of real noninteger eigenvalues A,,,(kAR), n = 0, 1. . . . . Iml ~
(A.5)
regular at 0 = 0 and 0 = ~r. We choose to normalize them to the Legendre functions at 0 = 0 lim S ~ ( k A R , cos O)/P'~(cos 0) = 1.
0~O
(A.6)
Then, multiplying eq. (A.3) by S(1~, and integrating, we find the orthogonality relation +1
f -1
(1) (1) (n+m)! 1 dz S,,n(kAR, z)Smn,(kAR, z) = 8n,,(n - rn)! (,~m~+½) dhmn/dn"
(A.7)
B. Schrempp, F Schrempp
/
Tunnelling phenomenon
443
Three types of solutions of the radial equation (A.2) are of interest here: the solution R ~ ( k A R , cosh ,f)regular at ,f = 0 and the solutions R ~ ( k A R , cosh ~) / outgoing/ behaving as purely LincomingJ spherical waves for kAR • cosh ~ co, with corresponding normalization to the spherical Hankel functions h ~)(kAR • cosh ~) at infinity [24, 25]: lira
k A R cosh ~5~c~
R .(3) ~ / P. .A. D. . .
cosh ~)/h~ )(khR
.
cosh ~:) = 1 ,
(A.8)
with ,1, ,,,,,) + R , ,(4) ,,), R,,,, (= ~1{o,(3)
(A.9)
and wronskian determinant [24, 25] ~ ~ ran, 1~ rnn / c o s h ,~
kAR sinh 2 ~"
(A.10)
We are now ready to expand the sum of the incident plane wave ~ i = e ikx ,
withwavevectork=k(sinTlcos¢',sin~sin¢',cos~)
(A.11)
and the scattered wave @~ in terms of product solutions of eq. (A.1) q)
~./cos m6
R,,,,,(kAR, cosh ¢)S~)~(kAR, cos v)] sin me.
(A.12)
The boundary condition ~ = $i + ~ - - 0 on the spheroidal boundary cosh (B = RL/AR fixes the coefficients. Taking the limit Ix] ~ oo, we then arrive at the following spheroidal partial-wave expansion of the scattering amplitude f(k, O, ¢) [using eqs. (2) and (20)] 2
~
+
f~m.(kRT '
f(k, O, ~b)- ( k ~ ) 2 ~=o,,~m (2-6o,,)(,L,,, +1) × dn
(a)
(1) x Sm~(kAR, cos 0) cos m ( ¢ - ¢ ' ) ,
(A.13)
with partial-wave amplitudes
R ~ ( k A R , RL/AR) fx,,,~(kRT, kRL) = 2i (3) . _ . Rm,(kAg, RL/AR)
(A.14)
Let us now specialize to the case of a plane wave incident in the z-direction (subsect. 3.2), i.e., we put rt --0 in eq. (A.13). Then, only terms with m = 0 survive in eq. (A.13) and the form (22) of subsect. 3.2 results.
B. Schrempp, F. Schrempp / Tunnelling phenomenon
444
For this case the Sommerfeld-Watson transformation of eq. (22) into the complex A-plane is obtained in the standard way, using eq. (A.5):
f(k' O)=
f (h +zrn½) £ (kg ' Cx
X S(12(A)(khR, -cos 0),
(A.15)
where CA encircles the real eigenvalues h , ( n ( h , ) = 0, 1 .... ) in clockwise direction. It may be shown from the wave equation (A.2) - in complete analogy to the spherical case [49] - that fA(kRT, kRID is a meromorphic function of h and also that S(1~ On(h) (kAR, -cos 0) and sin 7rn(h) are entire functions of h (cf. ref. [24]). The Regge-like moving poles of Ix (kRx, kRL) in the complex A-plane at h = a,, (kRx, kRL), m = 1, 2 ..... are the solutions of
=0 ROv(a~)( (khR, 3 RL/,SR) )
(A.16)
according to eq. (A.14).
Appendix B In this appendix we use the method of uniform approximations [17, 18] to derive an approximate expression for the spheroidal partial waves fA (kRw, kRL), eq. (23), - valid near its h-poles - for the case of a sharp spheroidal boundary with RL oCk ~ oe. We take special account of the peculiarity of a strongly k-dependent boundary, not treated in the literature so far. Moreover, the uniform method provides more accurate approximations in general than the semiclassical one, allows for the estimation of errors, and may finally serve as an independent verification of the results (31) and (33) obtained from the multiple reflection method using the rules of ref. [12]. The first problem one encounters in approximating the radial wave functions R~,(a)(kAR, RL/AR) in eq. (23) on the boundary with RL = AR cc k ~ oo, is that the boundary point sinh ~B = RT/AR oc1/k tends to zero and hence approaches a singularity of the radial equation (18). Therefore, as a first step, we perform a Langer-type substitution [17], which for spherical problems is well-known to give improved approximations near the origin. The substitution
1{ RT'~ 2 cosh ~:- 1--~k~-~} e '~,
-co~o-
(B.1)
Ron(khR, cosh sc) = ----7-~ Y(khR, o-),
(B.2)
1
cosn ~¢
B. Schrempp,F.Schrempp/ Tunnellingphenomenon
445
transforms eq. (18) into
(B.3)
[d-~2+p2(o')] Y = O , with
(B.4)
1 + (RT/2AR) 2 impact parameter fl -- (,~ +½)/k and e(o')= 4--~
e
1+ ~ - ~
e"
(1)
,
for all realo'.
(B.5)
We observe (i) the singularity at sc = 0 is mapped into a turning point at crl = -oo; (ii) the image of the boundary ~ , at ~rB = log (2AR/R2T)(RL-AR) ,0 is k~oo now well-separated from o"1 = -oo; (iii) the second relevant turning point of P2(~r) is located at 6~2=1og 2 ~
(41+(~/AR) -1)
=21og--RT
(assuming IZ/ARI<< 1 and larg/31< ½~r); (iv) in the limit e"<< 4AR2/R 2 oc k 2, with eccentricity parameter kR2 /AR = 1/c held fixed, eq. (B.3) approaches the confluent hypergeometric equation
I ~ + [ k R 2 x ' ~ 2 2,, [kR2x\ k 2 ~2--~) e - ~-~--R--] ( 4 - - ~ / 3 ) e'~] Yas = 0,
(B.6)
with solutions in terms of the well-studied Whittaker functions [50]
Y~s =
t
N1 ' M-(i/4)(k/aR)#2,o (e
2
N2 " W-u/a)(k/ag)t3 ,o (e
4kAR sinh 2 ½~)/sinh ½¢,
--i'n'/2 --i'n'/2
4k&R
being regular at ¢ = 0,
sinh 2 ½~)/sinh ½sc,
being purely outgoing for large kAR sinh 2 ½(. (B.7) This limiting equation (B.6), having the same qualitative structure as the original one (relevant turning points, etc.), serves as an excellent comparison equation [17, 18] for deriving a uniform approximation of the radial functions R~ol,) and R(o32 over the whole range of ~, viz., ~ for AR oc k ~ oo. For space reasons we cannot, however, present the derivation in full mathematical rigour here. An excellent discussion of the uniform approximation method may be found in refs. [17, 18].
B. Schrempp, F. Schrempp / Tunnellingphenomenon
446
We map the original radial equation (B.3) onto a confluent hypergeometric equation of the type (B.6)
,
(B.8)
r(o-) = ~/--Z(r(o-)),dr/do-
with
2, , _ [ k R ~ tr)- ~ ~ ]
2e 2, kR2T e" ---~K
(B.9)
and x left as a parameter to be determined later. The exact mapping r = r(o-) of eq. (B.3) onto eq. (B.6) is the solution of the differential equation [17, 18] (with 7:=dr/do') ~/~ d e 1 e(o-) / ÷202(r ) =p2(o-) 14 p2(o-) do2 4 i p2(o-)j,
(B.10)
which is hard to solve in general. However, since our comparison function 02(r) was chosen to be qualitatively very similar to the original p2(o-), r(o-) will be slowly varying and hence from eq. (B.10) the mapping is approximately determined by
I
d r ' 4 ~ 7 ) - - - I do-'4P--~'),
g2
O"2
(B.11)
and
y(o-) = ~=Z(r(o-))= [ Q2(r(o-))]j 1/4z ( r ( o - ) ) , •
(B.12)
with range of validity specified by the error control function [17] o-
E(o-)-½fdo-'
I
d2
p
1 e(o-') I
(B.13)
,) do-,2 @ p2(o-,) <<1.
With the convenient change of variables in eq. (B.11)
v \AR]e,
p = R T e ~/2 1 + \ 2 - - ~ / e"
=ARsinhG
(B.14)
the transformed mapping v(() reads* v(/D
AR sinh
dv v2
dp ~p-p-p-p-p-p-p-p-p~~=---gp(p, fl) ,
-4K=kv
(B.15)
/3
* We fix all appearing square roots in eq. (B.15) to be real and positive if their arguments are real and positive. Moreover, we require [arg v/4K[ < rr.
447
B. Schrempp, F. Schrempp / Tunnelling phenomenon
where the r.h.s, just equals the radial part of the classical action, as defined in eq. (32) of subsect. 3.2. Next, we fix the lower integration limit v2 and the free parameter x in eq. (B.9) by the requirement that the factor
1
[Q2(T(O'))] 1/4 = [
~/-~=[ ~
J
_/)(~)
11/4[
L4khRtanh2½scJ
V(~)-eg
]1/4
LkARsinh2~-(k/hR)~ZJ (B.16)
in eq. (B.12) remains continuous across (i) the turning point p2 = kAR sinh s¢2= fl, and (ii) the singular point ~ = 0: (i) requires v2 = v(~2)= 4K, (ii) v(0)= 0, or using eq. (B.15)
4~ 1 --
dv
,
v
=x
o B
oO~/~
it31~aR 4AR
,
(B.17)
0 thus fixing K to k 2 x---~-~-/3 ,
(I/31<
(B.18)
v(O I ~ 4k AR sinh 2 ~
(B.19)
both as expected from the limiting eq. (B.6). Using eqs. (B.2), (B.12) and (B.15) it is now a simple matter to write down the radial functions ,,~'(1)onand R(o3,) in terms of the uniform Whittaker approximants
M_i,,,o(e-"/2v(s¢)) ,
W_,,,.o(e-i'~/Zv(()),
(B.20)
respectively, with limit ]El << 1 given in eq. (B.7). Moreover, a complete set of simple uniform approximation formulae for the Whittaker functions has been derived by Taylor [50, 51] in terms of the variable [cf eq. (B.15)]
v(O r
~ ,
,
i13 - - 4 K
4~
W_iK.o(e-i'~/% ) - G(e-i~/2K' e-"~/2v) F(½- &)
p = AR sinh s¢,
(B.21)
448
B. Schrempp, F. Schrempp / Tunnelling phenomenon
2 e -i~/4 cosh (i~v+ ¼icr) , / ei£r ,
x]2" e i~/4 cosh (i~T--lirc) ,
-2Ir < arg ~x< 0,
(B.22a)
- ~ ' < arg ~:r< 2~r,
(B.22b)
7r < arg ~T< 3~r,
(B.22c)
2~'< arg ~x< 5~r,
(B.22d)
/
L e itJT+i'rt/2 ,
valid for sufficiently large v and/or x, [arg v/4x[ < rr, with e-iK log
G(-iK,-iv)=
(-iK/e)[ 4K]-1/4
F(~--~K)
[
= 1-
[l-vj ,
for [arg(-iK)[<~-.
(B.23)
Using these results, we easily obtain the correct normalization of R(o3,) in terms of W_i,,,o(e-i'~/2v) for s¢, viz., v ~ oo without a tedious matching procedure. The normalization of R(ol,) then follows from the wronskians (A.10) and [50] 1 { W~,,o(Z), M~,,o(Z)}z = l-'(½-/z)"
(B.24)
Finally, the following uniform approximation for the spheroidal partial waves results •
fA (kRT, k R L .
(1)
.
--~rK
=-e
.
.
M_,,.o(e-i~/2Va) W-i,,.o( e-i'~/2vB)/F(½ - iK)'
(B.25)
with k 2 (a+½)2 K =-$-~/3 = 4kAR ,
kR2T 1 vB = v(~B)--- AR = c '
(B.26)
for I/3[<
(B.27)
449
B. Schrempp, F. Schrempp / Tunnelling phenomenon
The equation for the Regge-like poles in the complex ,~-plane now takes on the form (I/31<
-ilr/2
VB) = 0.
(B.28)
Finally, let us demonstrate, how from this pole condition the simpler, semiclassical one, eq. (36), is obtained using eqs. (B.22) for vB = kR2T/AR = 1/c >>1. The phase of SeT(V)=Sp(RT, [3) in the a, viz., [3 domain of interest 0
~-~
= I[3[~>RT,
(B.29)
crucially determines which of the approximations (B.22) has to be used in eq. (B.28). Thus, let us continue Sp(RT, [3) from real [3
/ R~ , \312 I / R T \ 2 s°(gT'[3)°Ct-fl'2-1) / t ~ ) "
(B.30)
We choose to go around this branch point in the upper -~-/ - 1 ~ e"r(1
(R~/[32) plane
implying
RT 2
to end up in region (B.29), i.e., -zr < arg region (B.29)
(R~/[32) < 0.
Hence, from eq. (B.30) in
3
(B.32)
and eq. (B.22c) applies, verifying eq. (36) of subsect. 3.2: cosh
(iSp(RT, [3,,)-~i~r) = 0,
or
So(RT, ___.~)am +½ = (m -~)~-, m = 1, 2, 3 .....
(B.33)
with arg 5'p = 21r or Sp > 0.
Appendix C In this appendix we shortly outline the derivation of a uniform approximation of the spheroidal angular functions S(1) on(~)[~l .aA nl x , cos 8) (cf. appendix A) for complex n, viz., A, 0 ~<0 ~<~r, using exactly the same mapping techniques as in appendix B. In the 8-range of interest 0 <~O ~<~r, the only relevant exceptional points are the singularities at O = 0 and 8 = It. For 0 < arg (A + ½)< ½zr the turning points are
450
B. Schrempp, F. Schrempp / Tunnelling phenomenon
unimportant here. Hence, we approximately map the spheroidal angular eq. (19) [a__~2+(kz~R sm . 0) 2 +(A +~) ~2 + ~ 1J
1
X x / s i ~ s(ld(h)(kAR, cos 0) = 0,
(C.1)
on the Legendre equation (f~ = l~(0))
d2 12 1 3,. ~-~-~+ (n +$1 + ~ J 4 s m
1) P,,(cos 121= 0,
(C.21
which has the same singularities at f~ = 0 and [I = ¢r as eq. (C.1). The analogue of the approximate mapping eq. (B.11) now reads
8
(n + 1)~'~(0) = ~ d0"J(/~ -{-1)2 -t- (k AR sin O') 2 -= So(O, 0),
(C.3)
0 where the r.h.s, is just the angular part of the classical action and 0 = 0 has already been mapped on 12 = 0. The free parameter n + 1 is fixed by requiring that the second singular point at 0 = rr corresponds to fl = ~r, i.e.,
,r/2 n+1=2
dO'#(A+l)2+(kARsin8')2"
I
(C.4)
o Then we obtain as in appendix B the uniform approximation for O~O~
InloCkAR
>>1,
S(ol),(,)(kAR, cos O)
11/=
_~[sinO(o)
(n+½)
t sin 0
[(a +½)2+(kAR sin 0)211/2j
P,(x)(cosIl(0))
(C.5)
in terms of the Legendre functions, with f~(0) given by eq. (C.3), and proper normalization
S(1)(kAR On
~,
~
1)=P,(1)= 1
(C.6)
Since clearly fl(~r - O) = ~r - f~(O)
(C.7)
e(1) is continuous at 0 = ¢r and satisfies eq. (A.5), are the eigenvalues an, for which oo, determined by integer values of n in eq. (C.4) [remembering P,(-z) = (-1)"P,(z), n =0, 1 .... ]. Thus, indeed, n(A) in eq. (C.5) equals the index n(a) of the angular functions o(1) ,,on(a).
B. Schrempp, F. $chrempp / Tunnelling phenomenon
451
References [1] A. Chodos, R.L. Jaffe, K. Johnson, C.B. Thorn and V.F. Weisskopf, Phys. Rev. D 9 (1974) 3471. [2] W.A. Bardeen, M.S. Chanowitz, S.D. Drell, M. Weinstein and T.M. Yan, Phys. Rev. D 11 (1975) 1094; R.C. Giles, Phys. Rev. D 13 (1976) 1670. [3] P, Gn/idig, P. Hasenfratz, J. Kuti and A.S. Szalay, Proc. Neutrino 75 IUPAP Conf., (1975) vol. 2, p. 251; P. Gn/idig, P. Hasenfratz, J. Kuti and A.S. Szalay, Phys. Lett. 64B (1976) 62. [4] G. Preparata and N. Craigie, Nucl. Phys. B102 (1976) 478; N. Craigie and G. Preparata, Nucl. Phys. B102 (1976) 497; G. Preparata, in Proc. 1975 Erice Summer School, ed. A Zichichi. N. Craigie, in Proc. l l t h Rencontre de Moriond (1976), ed. Tran Thanh Van. [5] P. Hasenfratz and J. Kuti, Phys. Reports 40 (1978) 75, and references therein. [6] J. Kogut and L. Susskind, Phys. Rev. D 9 (1974) 3501. G. 't Hooft, preprint CERN TH-1902 (1974). [7] F.E. Low, Phys. Rev. D 12 (1975) 163. [8] S. Nussinov, Phys. Rev. Lett. 34 (1975) 1286. [9] K. Johnson and C.B. Thorn, Phys. Rev. D 13 (1976) 1934. [10] B. Schrempp and F. Schrempp, Phys. Lett. 70B (1977) 88; Proc. 12th Rencontre de Moriond, Flaine 1977, ed. J. Tran Thanh Van, vol. II, p. 67; Nuovo Cim. Lett. 20 (1977) 95. [11] R. Balian, G. Parisi and A. Voros, Saclay preprint DPh-T-78-20 (1978). [12] J. Knoll and R. Schaeffer, Ann. of Phys. 97 (1976) 307. [13] R. Balian and C. Bloch, Ann. of Phys. 85 (1974) 514; 63 (1971) 592. [14] B. Schrempp and F. Schrempp, CERN TH. 2604 (1979). [15] K. Winter, Proc. European Conf. on Particle physics, Budapest, Hungary (1977), vol. I, p. 55. [16] R.C. Fuller, Nucl. Phys. A216 (1973) 199; J. Knoll and R. Schaeffer, Phys. Reports 31 (1977) 159, and references therein; K.W. McVoy, Lectures at Escuela Latino-Americana de Ffsica, University of Mexico, University of Wisconsin preprint (August, 1977). [17] M.V. Berry and K.E. Mount, Rep. Prog. Phys. 35 (1972) 315. [18] F.W.J. Olver, Asymptotic solutions of differential equations and their applications, ed. C.H. Wilcox (Wiley, London, 1964). [19] B.R. Levy and J.B. Keller, Can. J. Phys. 38 (1960) 128. [20] H.M. Nussenzveig, Ann. of Phys. 34 (1965) 23. [21] H.M. Nussenzveig, J. Math. Phys. 10 (1969) 82; 129. [22] W. Franz, Z. Naturforschung 9a (1954) 705; T.T. Wu, Phys. Rev. 104 (1956) 1201; P. Beckmann and W. Franz, Z. Naturforschung 12a (1957) 257; S.I. Rubinow, Ann. of Phys. 14 (1961) 305; J.J. Bowman, T.B.A. Senior and P.L.E. Uslenghi, Electromagnetic and acoustic scattering by simple shapes (North-Holland, Amsterdam, 1969). [23] J.B. Keller, A geometrical theory of diffraction, Calculus of variations and its applications, Proc. Symp. on Appl. Math. 8 (McGraw-Hill, New York, 1958), p. 27; J.B. Keller, J. Opt. Soe. Am. 52 (1962) 116; B.R. Levy and J.B. Keller, Comm. Pure Appl. Math. 12 (1959) 159; R.M. Lewis, N. Bleistein and D. Ludwig, Comm. Pure Appl. Math. 20 (1966) 295. [24] J. Meixner and F.W. Schiifke, Mathieusche Funktionen und Sph~iroidfunktionen (Springer-Verlag, Berlin, 1954).
452
B. Schrempp, F. Schrempp / Tunnelling phenomenon
[25] C. Flammer, Spheroidal wave functions (Stanford Univ. Press, Stanford, Calif., 1957). [26] H.C. Bryant and N. Jarmie, Scientific American, (July, 1974) p. 60. [27] W.G. Neubauer, in Physical acoustics, ed. W.P. Mason and R.N. Thurston, (Academic Press, New York, 1974), vol. X, p. 61; H. Uberall, ibm p. 1. [28] A. Sommerfeld, Partial differential equations in physics (Academic Press, New York, 1949) p. 279. [29] CERN-Hamburg-Heidelberg-Annecy-Vienna Collaboration: H. De Kerret et al., Phys. Letters 62B (1976) 363; 68B (1977) 374; E. Nagy et al., Paper contributed to the 19th Int. Conf. on High-energy physics, Tokyo, 1978; CERN preprint (August, 1978) Nucl. Phys. B, submitted. [30] Cornell-McGill-Northeastern-Lebedev Collaboration: Paper contributed to the European Conf. on Particle physics, Budapest, 1977; Phys. Rev. Lett. 39 (1977) 975; paper contributed to 19th Int. Conf. on High-ene~'gy physics, Tokyo, 1978. [31] R.C. Giles, MIT-preprint, MIT-CTP 706 (1978). [32] O.V. Kancheli, ZhETF Pis'ma 18 (1973) 469. [33] L. Stodolsky, 6th Int. Coll. on Multiparticle reactions, Oxford, 1975, p. 577. A. Biatas and L. Stodolsky, Acta Phys. Pol. B7 (1976) 845. [34] J. Koplik and A.H. Mueller, Phys. Rev. D 12 (1975) 3638; [35] F.E. Low and K. Gottfried, Phys. Rev. D 17 (1978) 2487. [36] K. Gottfried, Phys. Rev. Lett. 32 (1974) 957. [37] W. Busza, Proc. 12th Rencontre de Moriond, Flaine, 1977, ed. J. Tran Thanh Van, p. 129. [38] J. Orear, Phys. Rev. Lett. 12 (1964) 112; C. Hojvat and J. Orear, preprint CLNS-346 (1976). [39] J.V. Allaby et al., Phys. Lett. 25B (1967) 156; E.J. Sacharidis, Rutherford Lab. Report R H E L / R 186 (1969). [40] V.A. Matveev, R.M. Muradyan and A.N. Tavkelidze, Nuovo Cim. Lett. 7 (1973) 719; S.J. Brodsky and G.R. Farrar, Phys. Rev. Lett. 31 (1973) 1153. [41] A.Z. Patashinskii, V.L. Pokrovskii and I.M. Khalatnikov, JETP (Soy. Phys.) 18 (1964) 522; 683; V.L. Pokrovskii and I.M. Khalatnikov, JETP (Soy. Phys.) 13 (1961) 1207. [42] R. Bertini et al., Phys. Lett. 45B (1973) 119; J. "Fhiron, private communication. [43] W.D. Apel et al., Serpukhov preprint IHEP 77-38; Yad. Fiz. 27 (1978) 154. [44] B. Schrempp and F. Schrempp, Nucl. Phys. B54 (1973) 525; B60 (1973) 110; B77 (1974) 453; Proc. 11th Rencontre de Moriond, M~ribel, 1974. [45] M. Imachi, T. Oroguchi, S. Otsuki and F. Toyoda, Prog. Theor. Phys. 45 (1971) 1849; 47 (1972) 1958; 48 (1972) 210; S.Y. Chu and A.W. Hendry, Phys. Rev. Lett. 25 (1970) 313; Phys. Rev. D 4 (1971) 2743; 3282; Y. Hara, T. Kuroda and Y. Takaiwa, Prog. Theor. Phys. 51 (1974) 840. [46] H.C. Bryant and N. Jarmie, Ann. of Phys. 47 (1968) 127. [47] A. Babaev et al., Phys. Lett. 67B (1977) 351. [48] E. Di Salvo and G.A. Viano, INFN-AE 77-11 (1977). [49] R.G. Newton, The complex J-plane (Benjamin, New York, 1964). [50] H. Buchholz, The confluent hypergeometric functions (Springer-Verlag, Berlin, 1969). [51] W.C. Taylor, J. Math. Phys. Massachusetts 18 (1939) 34.