Nuclear Physics A483 (1988) 92-108 North-Holland, Amsterdam
ELECTRON
SCA’ITERING
FROM **Ne AND BOSON
‘*Mg IN A MICROSCOPIC
MODEL
R. KUCHTA’ Laboratory of Theoretical Physics, JINR, Dubna, Head Post O@ce, POB 79, Moscow, USSR Received (Revised
2 November 1987 18 January 1988)
It is shown that a mean-field approximation applied to the microscopically derived boson hamiltonian yields a reasonable description of the form factors for both elastic and inelastic electron scattering from some sd-shell nuclei (aoNe, 24Mg). The results agree well with experimental data for the 0’ + 0’ and O++ 2+ transitions but much less so for the Oc + 4+ transitions. Possible sources of the observed discrepancies are suggested. Comparison with other approaches is also given.
Abstract:
1. Introduction
Form factors of electron scattering are known to provide a deeper insight into the structure of nuclear states than the energy spectra and electromagnetic transition probabilities alone ‘). This is so because the momentum-transfer dependence of the associated nuclear matrix elements contains information about the spatial structure of nuclear states, thereby yielding a more stringent test on the reliability of the model wave functions. Various methods have been proposed to study the electron scattering form factors, both in heavy *“) and in light 4-6) nuclei. In this paper we present a new approach which consists in applying the mean-field (MF) approximation ‘) to a boson hamiltonian derived microscopically from the underlying shellmodel hamiltonian by means of the Belyaev-Zelevinsky-Marshalek (BZM) mapping procedure “). Due to the obvious simplicity, the search for a bosonic description of nuclear spectra has been a long one 9, and it has even more intensified lo) after the success of the phenomenologi~al interacting boson model (IBM) of Arima and Iachello I’). Recently, there have also appeared some attempts toward a boson treatment of the electron scattering form factors 3,5*6).These attempts are based either on the IBM 3,5) or on a specific realization of the symplectic sp(3, R) algebra in terms of the harmonic oscillator boson operators 6). Our approach is conceptually closer to that using the IBM, in the sense that it treats the bosons as counterparts of nucleon pairs “), However, we go beyond the conventional IBM ‘l,‘*) because the MF approximation enables us to take into account the bosons with all possible angular momenta as well as those corresponding to proton-neutron pairs. ’ Permanent
address:
Institute
of Mechanical
Engineering,
0375-9474/88/$03.50 @ Elseviet Science Publishers (North-Holland Physics Publishing Division)
B.V.
360 00 Plzeii, Czechoslovakia.
93
R. Kuchta / Electron scattering The
organization
relevant
of the present
formalism.
(Coulomb)
Sect.
form factors
24Mg. A summary
paper
3 contains
is as follows.
its application
for elastic and inelastic
In sect. 2 we describe to investigating
electron
scattering
the
the charge from “Ne
and
is given in sect. 4.
2. The model 2.1. DERIVATION
We consider nuclear
OF THE BOSON
HAMILTONIAN
a system of n, protons
hamiltonian
and nV neutrons
is taken to have the general &=
C T,pc& @
+t
(n, = n, = ZV, N even). The
form (1)
C Va,+:c;csc,, 4-Y~
where the indices (Y,/I, -y, 6 run through a suitable complete set of single-particle states and cz(c,) are the corresponding creation (annihilation) operators of nucleons. The quantities Tap stand for the matrix elements of a one-body operator, matrix such as the kinetic energy, while the Vapvs represent the antisymmetrized elements of an effective two-body interaction (1) is assumed to have the usual hermitian relation
between the nucleons. The hamiltonian and time-reversal properties. Using the
c~c;c~c, =spsc~cy - c~csc;c,, we can rewrite
(2)
(1) as
where
If we choose the single-particle states to form the Hartree-Fock one-body term (3b) is clearly diagonal because
(HF)
basis,
the
by definition of the HF basis. However, when describing realistic light nuclei, in which the single-particle states can be restricted to those taken from neighbouring shells only, &, may be diagonal in the harmonic oscillator basis as well 13). We make this choice here, so that fi, acquires the form I& = c E,C& a
-a
1 vo~vsc~c&;c~ nPr8
(5)
94
R. Ku&a
/ Elecrron scattering
and the single-particle states (Y,/3, . . . are characterized by the oscillator quantum numbers in the isospin formalism 14), a?= (n,, I,, ja,& m,, T,), where T, =+4(-f) for neutrons (protons). We will also use the letter a to denote the same set except m, and 7,. Now, we introduce the boson creation and annihilation operators b& and hop, which correspond to fermion pair operators czci and cBc,, respectively. These boson operators are assumed to satisfy the following antisymmetry and commutation relations b& = -b& IIbe,,
,
(64
bawl = [ b:p , bbpl = 0 ,
[hap, b&y] = &&I,,,-
6,,3,,,.
(6b) (6~)
We also define the boson vacuum IO), by the condition b,, 1% = 0.
(7)
In the BZM approach ‘), the boson images of the bifermion operators czc$, cBc, and czc, are expressed as power series in the boson operators b$, b,, and the unknown coefficients of these expansions are determined by the requirement that the original fermion commutation relations remain valid in the boson space. In the case of the particle-hole operators c& this leads to a very simple finite form (c;tc,),
= C K&&y, Y
(8)
where the suffix B on the left-hand side denotes the boson image of the corresponding fermion operator. On the other hand, the operators (czci), and (c$c,)~ are in general represented by infinite expansions, which naturally give rise to the problem of convergence. Nevertheless, these complicated operators are not needed to find the boson image of the fermion hamiltonian if the latter is conveniently expressed in the particle-hole form (5). In this case only the simple boson operators (8) are relevant. Replacing czc, in (5) by (czca), according to (8) and arranging the boson operators in the interaction term into the normal order with respect to the boson vacuum IO),, we obtain an exact boson image fiB of the fermion hamiltonian fir in the form ci,=
C e,&;tPb,&
C C &&&b&&J+, ww
MY6
where the quantities eaprs are given by
[
(9)
P”
P 1 9
Ea@yfi= &7, ecJ,y -a c v,,,,
+4 V&S
(10)
The bosons associated with the dynamics of the underlying fermion system can be expressed as functions (in general rather complicated) of the operators b$ , bnp.
R. Kuchta / Electron scattering
In this paper we assume that the corresponding
95
transformation
is linear and unitary,
i.e.
JTah
+(iMI)
(iM7) JJ-~~l/f~f~.~‘bf= SJJ,S,,{S,,,~,,,+(-~)‘,+‘~+~+~S,,,S,,.)
4 c
.
(llc)
The quantities (. . . .I. .) in (11 a ) are the usual Clebsch-Gordan coefficients for the angular momentum and isospin coupling. We characterize the dynamical bosons BTMrby the isospin projection T( r = -1, 0, + 1 for the proton-proton (TUT), protonneutron (TV) and neutron-neutron ( VV) bosons, respectively), the angular momentum projection M and the label i which distinguishes between independent configurations with the same T and M. Using (11) and a standard angular momentum algebra, one gets the boson hamiltonian fi, in the form fin = C J%B:Bz+ where
l=(i,M,~,),
(12)
C W,D&B:&&,
12
1234
2~(i*M~7~) ,...
and (13a)
W 1234
=
C J,J,J1J4
x
The functions
C T, TzTxT,
C ahcdef
W~k$$z~,T3J,T,(M171,
M2r2,
M373,
(i M T )+cI::‘m;c;‘)+~p~)+~~(%;p). (CIJ,‘T,L’
M474)
(13b)
E yThcd’and W$$$zqzJ3T3J4T4(Ml
T, , kt2~2,
M3~3, M4~4) are the angular
momentumand isospin-coupled analogues of the quantities E+,~, VaPrs appearing in (9) and their explicit expressions can be found in ref. 15).
2.2. APPROXIMATION
Up to this point,
METHOD
no approximations
have been made.
Provided
the sums in (12)
extend over all possible values of the labels 1~ ( il MI T1), 2 3 ( i2M2r2), . . . , the boson hamiltonian (12) is a precise image of the original fermion hamiltonian (5). However, the diagonalization of (12) in the entire boson space is practicable only for very special types of interaction and for a substantially limited number of bosons. In addition, such an exact diagonalization will also produce unphysical (spurious) states which have nothing to do with the underlying fermion system because of the overcompleteness of the boson basis. These spurious states must be identified and removed, which has long been a difficult problem 16). Fortunately, it has recently been shown “) that
96
R. Kuchta / Electron seatterhg
(i) this identification and removal of spurious states can be accomplished in a rigorous and systematic way; (ii) the remaining boson states, identified as not spurious, emerge unharmed by the admixtures of unphysical components, and consequently, they have the correct energies associated with the original fermion system. It is rather reassuring to see this, but actually one has not achieved much, because the complications of calculating the fermion matrix elements have only been transferred to other place, namely to the procedure for identifying the spurious boson states. It therefore appears to make more sense not to insist on the exact diagonalization of (12) but rather to look for an approximate solution which would be free from spurious states. This amounts to truncating the hamiltonian to a few bosons which must be chosen so that the truncated part of fiB is only weakly coupled to the neglected rest. No general consensus seems to exist as to how one should find the optimal bosons ‘*). In our opinion, the mean field approach is well suited for this purpose because ‘,14) (i) it is based on a variational criterion which ensures that the states built by the desired bosons are maximally separated from all other states, (ii) it is able to treat important physical properties in terms of few relevant parameters, (iii} it may include many different degrees of freedom in a single intrinsic state, (iv) it has a largely established power to deal successfully with many-body systems. The mean field method starts from the assumption that the transformation coefficients +$$’ in (11) are determined so as to minimize the ground-state expectation value of fis. In the present case of rr,_,protons and n, neutrons (n, = ny = N = even) we can take the ground-state wave function /GSfB to be a condensate consisting of only one kind of bosons (i = g; M = 0; r = 0) (14) where IO), is the boson vacuum defined by eq. (7). Note that IGS)n is properly normalized because [B,,, , B&J = 1 as a consequence of (llb). In writing (14) we have assumed that the ground state is axially symmetric (M = 0) and all the 2N = n, + n, nucleons are paired so as to give rise to N gv-bosons (T = 0). This concept of building the GS wave function is to be compared with the usual approach i9) in which rather the identical nucleons are paired to form the +zr=-and vu-bosons. The latter approach is certainly well justified in heavy nuclei where protons and neutrons occupy very different single-particle orbits. However, for applications to lighter nuclei with protons and neutrons filling the same shell, the explicit inclusion of rrv-bosons is expected to be important. With the assumed form (14) of IGS&, the variational requirement &(GS/I;l,/GS)a
=0
(1%
R. Kuchta / Electron scattering
under
97
the restriction
(16) leads to the following
system
~(@x
-6
JTt,b,J’T’a’b’-
of non-linear
JJ’ 6 ‘“-
eigenvalue
equations
E$‘$'b"+2(N-1)86,b 1
for $$jjy:
c J+@$j
J, T, J2 T2 ted
(17) We use standard iterative techniques to solve (17) and together with (ci%i’, corresponding to the lowest eigenvalue A,, we also obtain the solutions 1,5:‘:zi orthogonal to $9;;’a ,
(18) JTab
Once the structure of the ground state has been fixed, we turn our attention to the excited states, which can conveniently be generated in the framework of the equations-of-motion operator
formalism
*“). In this method
one attempts
to find an excitation
Qz satisfying
Qd=o,
(19)
b#dCQo,,& Q:lh) = ~&o~[Qc+,Q%#d
(20)
,
where Q, = (Qii)‘, I& hamiltonian excitation
re P resents a suitable reference state, fi stands for the of the system, the subscript (+ denotes a particular solution of (20) with
energy
w, and the symmetrized [A, B,
It is also useful
double
commutator
Cl = ;{[A, [4 Cll+ [[A, Bl, Cl) .
(but not necessary)
to demand
state the boson
as
(21)
that
[Q,,, Q:] = 6,,, . Here we take as reference to verify that the operator
is defined
ground
(22) state (14). It is then not difficult
Q:, = (I+ ~~o&oo)-“‘%m~,oo satisfies (19), (22) and its action on ]GS), produces in which one of the Bc,,,, bosons in the condensate boson BkO, i.e.
(23)
a normalized boson state (UC), (14) is replaced by an excited
(24)
R. Kuchra
98
/ Electron scattering
By successive application of Q& on /GS), one can generate higher (two-boson, three-boson, . . .) excitations but in this paper we will restrict our discussion to the lowest excited states of the form (24). Substituting (23), (14), (12), (13) into (20) and making use of (22), (llc), we obtain
ytw1 JTQh,J’T’a’b’ = 6,J&&$yh”
--~d*~l?*l&)+2(~--1)
x wz;?;::*,*T(~o,
0% 00, ~Owg~!l*+
JITs*T2; 1CI!I$% cfi!!$~L1~
(25)
Solutions of (2.5) give the structure coefficients $JTah tiKO) 0 f the excited boson I?&,, as well as the excitation energies wrK of the states (24). Since the matrix I%C$$&,~~~~~, is fully specified in terms of the quantities arising from the ground-state calculation (4ti$+tT, h ) eqs. (25) are linear with respect to the amplitudes I,!&$, Note that the lowest (fL’1) solutions of (25) with K = 0 and I( = 2 correspond to the so called p.. and y-bands, respectively ‘). We can also convince ourselves that the K = 0 solutions coincide with the J/$7$ obtained from (17), and consequently, they satisfy the relation (18). In the following we will restrict our discussion only to the K =I0 and K = 2 states of the type (24), and therefore, we will not encounter the problem of identifying and removing the one-boson excitations associated with the rotation of the nucleus as a whole j9>. To proceed further, we must restore the spherical and isospin symmetries which are obviously broken in the boson wave functions of the form (14) and (24). This is accomplished by angular momentum and isospin projection according to /GS; J’I’; MAc&-)~= ~;“+‘,,i;;t;,O]GS)B, (26)
/I; JT; ~~~)~
= P&B”,,&&.,lZK>, ,
where PZ~‘,*Eb are the normalization constants and 3”,,+,,,,, @;;rb, stand for the projection operators onto states with definite angular momentum and isospin, respectively 14). The energies of different nuclear states IJT) are then obtained by simply taking the expectation value of I?, in the states (26).
2.3. PHYSICAL
BOSON
STATES
For the procedure described in the preceding section to be reliable, we must be sure that the basic boson states (14) and (24) are indeed physical, i.e. that they are in one-to-one correspondence with actual states of the underiying fermion system. We therefore construct the following fermion analogues (not images!) of the states (14), (24):
/GSb= &~G’;oo)~ 1%t
(27)
Ii%=
(28)
J&~~O(~&)N-‘lO)F,
R. Kuchta / Electron scattering
where NGs, .N[, are the normalization (c,IO&= 0) and I%-,, = ,g,
,C,
99
constants,
IO), denotes
~~~~‘(i,m~~mpIJM)(~7,~7pI
T+&.~_c&,
the fermion
,
vacuum
i = g, 1
(29)
T@TP that with the same $y$$’ as ’ in (11). At the present stage, it should be emphasized the fermion states (27) and (28) cannor, in general, be considered as counterparts of the boson states (14) and (24), respectively. This is because the former may be linearly dependent (as a consequence of the Pauli principle), while the latter are always linearly independent. It is thus necessary to check explicitly whether or not the linear dependence among fermion states occurs. This is done by diagonalizing the norm matrix F(GS (GS),; ~=
F(GS 1IK),;
F( I’K 1GS)F . . . . . * : . . . . . . . . . . . . . . :
(30)
(K)
HI’,/ ,...............
with H.$’ = ,(1’K 1IK), and looking at the eigenvalues. For the lowest excitations with K = 0,2 and for the parameters described in sect. 3 we have found that none of the eigenvalues is zero, which implies linear independence of the fermion states IGS),, IIK),. Consequently, the boson states (14) and (24) can be put into one-to-one correspondence with the fermion states (27) and (28), thereby showing that they indeed represent certain physical states of the fermion system considered. 2.4. THE ELECTRON
SCATTERING
FORM FACTORS
Our aim is to calculate the charge (Coulomb) electron scattering form factor for a transition from an initial nuclear state characterized by angular momentum Ji and isospin Ti to a final state with Jr and Tf. In the plane-wave Born approximation (PWBA)
this form factor
squared
at momentum
transfer
q is given by ‘)
(31) where 2 is the atomic number of the target nucleus, has the second quantized form gL( q) = e
I
mdr 5 (aIjL(qr)YLM(t-
the transition
t3)b)chp
operator
g=(q)
(32)
( . . . 111. .I/.. . j” means that the matrix element is reduced both in the ordinary and isospin spaces. In (32), e represents the charge of a nucleon, jL(qr) and YLM stand for the spherical Bessel functions and the spherical harmonics, respectively, and t, is the operator of the third isospin component, t,)n) = -b/m), &J)= +$Iv). and the symbol
100
R. Kuchta / Electron scattering
In the present boson formalism, the states JJi~i), IJfTr) are taken to be those given in (26) and the boson image of the transition operator (32) is obtained by replacing czcB with (c~c,), according to (8). 2.5. THE SHELL-MODEL
HAMILTONIAN
In order to carry out explicit calculations we have to specify the characteristics of the nucleon system considered, namely the single-particle states and the two-body interaction. As already remarked in sect. 2.1.) instead of performing a self-consistent HF procedure, we choose as the single-particle basis the three-dimensional harmonic oscillator basis restricted to neighbouring shells only (the Op and Odls shells in the present case) and observe that the HF hamiltonian (3b) is diagonal in this limited single-particle basis. As for the nucleon-nucleon interaction, we combine a phenomenological form 2*) Y,(i)*
V,,,(U) = v0 ‘(riiir,)x
Y,~){l-~+~u(i)~o(j))(l-~i+~“l(i)~7(j)}
L
(33) with a two-body spin-orbit
interaction 22)
to get the correct spin-orbit splitting of the single-particle levels. As is well known r4), any calculation in a space larger than one major oscillator shell introduces the problem of identifyng and removing the centre-of-mass (c.m.) excitations, associated with the oscillations of the nucleus as a whole in the shellmodel potential. In the case of an approximate treatment, these (spurious) c.m. excitations can mix with the actual intrinsic excitations and their mutual separation is often very difficult or even impossible. In this paper we try to diminish the coupling between the c.m. and intrinsic excitations by performing a unitary transformation of the shell-model hamiltonian
A = i$l g+f f {V,,,(i,j)+ V,.,.(i,j)l
(35)
i#j
to
13)
2 = UHU-’
=
(36)
where A stands for the nucleon number, E? are the harmonic-oscillator energies and V,,,,( i, j) =
1
1 -A
Gpi
I
*
pj
+$t?WJ2F~ * Fj
>
.
eigen-
(37)
R. Kuchta / Electron scattering
This transformation among
various
states becomes
13,23) to distribute
the strength
states in such a way that the amount enhanced
task we are then (c.m. excitations) evaluating
is expected
while in others
it turns
101
of c.m. excitations
of c.m. components
in certain
out to be suppressed.
The only
left with is to identify which states belong to the first category and which to the second (intrinsic excitations). This is done by
the quantities %.nl.(JT) = (-WL.IJ~>
3
(38)
where fl km.
is the operator the respective
=
c k=-1
(39)
a:4
that counts the number of c.m. motion quanta 23) and IJT) represents nuclear state. The creation operator u: in (39) is defined as (40)
where
Fk and &
are the c.m. momentum
and coordinate,
respectively.
3. Results The procedure described in sect. 2 was applied to the study of the electron scattering form factors in “Ne and 24Mg. The calculations were performed using the oscillator model single-particle space consisting of the 0~312, Op,,,, Od,,,, l~i,~ and Od3,* shells with energies 24) ~(0~312)
=
-21.8
MeV, ~(lsi,~)
e(Op,,,) = -3.28
= -15.65
MeV,
MeV,
e(Od&
= -4.15
MeV,
e(Od3,2) = 0.93 MeV.
The oscillator length parameter b = mw/ h, as well as the nucleon-nucleon interaction parameters V,, r], p, 5,. have been determined in such a way that the diagonal elements of the HF matrix hiI = T,, + $ & VaPaP [cf. eq. (3b)] provide the best fit to the above single-particle energies and the ground-state energy calculated in the present model reproduces the experimental binding energy of a given nucleus. The resulting values are listed in table 1. In fig. 1 we compare the experimental excitation spectra with those obtained in the present approach. An overall agreement is observed, both for 20Ne and 24Mg. TABLE
Calculated
parameters
Nucleus
V, (MeV . fm3)
“Ne
13.6 74.1
=Mg
1
characterizing the nucleon-nucleon oscillator wave functions 17 0.254 0.257
interaction (33), (34) and the single-particle (b = mo/ h) b (fm)
P 0.509 0.490
-1.26 -1.26
0.16 0.16
1.65 1.68
102
R. Kuchtn
/ Electron
scattering
MNe (T=O)
(T=O) 2’
__-__;: -____
-2'
O'-___ O'--
Z4MCJ
--0' --__
-0'
2*-__
-_2*
o+_-_-t'-__
-m
0'
3+-__
__-2’
4’ ---
EXP
Fig. 1. Positive-parity
3+ 2+
2'_____&a--
2'
2'-___-
--
1'
--__L'
2' ______2+
CALC
EXP
T = 0 levels of “‘Ne and 24Mg. Experimental
CALC
values are taken from ref. ‘*).
ELASTIC
10-l
. . . . . . SP(3,R)
0.5
1.0
1.5
2.;
4 IfmT Fig. 2. Calculated form factor for the elastic electron scattering by ” Ne , compared with experiment. The solid curve shows the result obtained in the present boson mean-field (MF) approach. The dashed curve represents the result of the shell-model calculation in the restricted sd-shell subspace. For comparison, we also show the earlier result obtained by Vassanji and Rowe [ref.6)] in their microscopic boson sp (3, R) model (dotted curve). Experimental data are taken from ref. 25).
R. Kuchta / Electron scattering
103
*'Ne
-
MF BOSON
. ..**
SP(3.R)
-.-._
IBM
0.5
15
1.0
2.0
q[fm-'1 Fig. 3. Calculated form factor for inelastic electron scattering to the first 2+ state in “Ne. The solid, dashed and dotted curves have the same meaning as explained in the caption to fig. 2. The dash-dotted curve is taken from ref. ‘) and it has been added to compare our results with those obtained in the IBM. Experimental data are taken from ref. 25).
With the corresponding wave functions, we have calculated the electron scattering form factors for various transitions in the above nuclei. The results for 20Ne are displayed
(as solid lines) in figs. 2-5. As is seen, the agreement with experimental (fig. 5). For comparison, we data is quite good, except for the 0: +4: transitions also show the shell-model results obtained in the restricted sd-shell single-particle subspace (dashed lines). While for the elastic scattering (fig. 2) as well as for the excitation of 2: and 2: states (figs. 3 and 4a, respectively) the shell-model results do not differ substantially from the mean field (MF) boson ones and both compare well with experimental data, the same is not true for the excitation of the 2: state (fig. 4b) and the 4: state (fig. 5). Nevertheless, it is worthwhile to notice that in both cases the MF boson approach provides a better agreement with experimental data than the shell-model calculation, though for the 4: state the discrepancies are still large. This improvement is most likely due to the fact that the MF boson calculation takes into account the excitations of the I60 core (though in an approximate way), whereas the present shell-model calculation does not. In fact, there is some evidence 26) that the O:, 2: and 2: states in *‘Ne are built mainly by four nucleons outside the 160 core while the 2:, 4: states contain non-negligible excitations from the Op-shell.
scattering
R. Kuchta f Electron
104
r
1o-z-
al
20N~ 0; -2;
10-3_ Nd + 1o-4_ Lb A A //
lci5.,!I 0.5
-
MFBO.SON
----
SHELL
\
1.0
MODEL
\
1.5
21;
\
9~fd'I
20-l
I
b1
/
'*Ne
\ \
I lo+.
\\
/’
1G7_
I
0.5
2
,/--
-.
5.
\-
I’
SHELL MODEL --=i 1 \
It I
1.0
1.5
2.0
stfm-'1 Fig. 4. Electron scattering form factors for the excitation of the 2: (a) and 2: (b) states in *‘Ne. The solid and dashed curves represent the results of the MF boson and shell-model calculations, respectively. For experimental data see the caption to fig. 2.
It is also interesting to compare our results with those obtained in other approaches. In figs. 2 and 3 we therefore reproduce (as dotted curves) the recent results of Vassanji and Rowe 6), and in fig. 3 we also show the IBM results of Park and Elliott “) (dash-dotted curve). Needless to say that this comparison is only qualitative because of the conceptual and computational differences between various approaches. Much work is still required before quantitative conclusions can be drawn. In fig. 6 we present our results for the elastic and inelastic electron scattering form factors of 24Mg. We again observe that the MF boson form factors for the O+ + O+ and 0’ + 2+ transitions (figs. 6a, b, c) reproduce the experimental data quite well (in general better than the shell-model calculations) but those for the 0:+4:
105
R. Kuchta / Electron scaitering
I
"Ne
0.5
10
1.5
2.0
2.5
q [fm-’ 1 Fig. 5. Calculated
form
factor
for the excitation of the first 4+ state in *‘Ne. For further caption to fig. 4.
details
see
transitions (fig. 6d) disagree with the data both in shape and magnitude. Possible sources of the observed discrepancies are the following. First of all, we have performed the variation before projection in this work. It is well known 14) that such an approach does not allow for changes of the self-consistent internal field within a rotational band. Since there exists no a priori decoupling between rotational and intrinsic motion, the method of variation before projection can be expected to work well only in cases when the coupling terms are relatively small, which occurs most likely for not-too-large angular momenta. How large they may be depends in an essential way on the choice of the intrinsic system. It is indeed possible that in the present case the values J = 0, J = 2 (for which we have obtained satisfactory results) still permit a sufficiently good decoupling between intrinsic and rotational motion, whereas
the value J = 4 is already
too large to allow the variation
before projection
to be correct. The second reason for the failure of our approach in reproducing 0: + 4: form factors can be traced back to the occurrence of c.m. components
the in
the 4: state. In table 2 we give the quantities (38), which measure the amount of c.m. excitations in a given state, for the relevant cases (i.e. the O:, 2:, 2:, 4: states). It is seen that the values of these quantities for the 4: states are by two orders of magnitude larger than those for the 0: and 2: (i = 1,2) states, irrespective of the fact that the former are still very small (-10m4). It remains to be investigated to what extent such an admixture of c.m. excitations destroys the structure of a given state. When looking at the spectra displayed in fig. 1, we cannot say that the energies of the 4+ states are reproduced qualitatively worse than those of the 2+ states. However, the electron scattering form factors are known to be more sensitive to the actual form of the wave function than the energy spectrum, and therefore, even a
R. Kuchta
106
f
b a)
10”
2.94
\o
/ Electron
b)
9
,(j2
%
“= ,(j2
\
cr
,/ko‘*,,
I
:$i3
%
b-
24Mg
%1/ a\\
ELASTIC
t
scattering
'$0-3
,,b-
-
16”
Iti ,
d
0.5
1.0
1.5 2.0 slfm-ll
0.5
2.5
IO
d)
o;-2;
‘P
:+ 4 in
-\
‘\
I/
If
‘\\
1.5 2.0 qlfm-‘1
25
24Mg 0;-4;
Ii”.
0.5
1.0
1.5
2.0
25
0.5
1.0
4[fm-‘I
1.5
2.0
2!
qlfm-‘I
Fig. 6. Electron scattering form factors of a4Mg calculated with the MF boson (solid curves) and with the sd-shell model (dashed curves) wave functions. In (a) is the elastic (0: + 0:) form factor, in (b), (c) and (d) are the inelastic form factors for the excitation of the 2:, 2: and 4: states, respectively. Experimental data are taken from ref. 29).
TABLE 2 Expectation
values
of the c.m. quanta counting operator states of *‘Ne and *4Mg
(39) in some I&‘, T = 0)
3f
0:
2+ 1
2’*
4:
‘“Ne *4Mg
0.000 001 0.000 001
0.000 001 0.000 000
0.000 002 0.000 001
0.000 135 0.000 172
R. Kuchta
small
admixture
of the
c.m.
/ Eleciron
excitation
107
scaftering
may
have
large
influence
on the
final
result. We have between
thus
suggested
the experimental
two possible and calculated
sources
of the discrepancies
form factors
for the excitation
observed of the 4:
states in *‘Ne and 24Mg. While the implementation of the first point (i.e. projection before variation) is rather straightforward (though laborious), an answer to the question of how to remove the c.m. excitations from the physical states is still far from clear. Of course, we could resort to an exact treatment in a restricted singleparticle space and use the special techniques for isolating the c.m. excitations, which are available in this case 27). However, the main advantage of our approximate method, namely the possibility of including a larger single-particle space without serious difficulties would then be lost. We are therefore currently engaged in an effort to develop a sound and reliable method for dealing with this problem. 4. Summary In this paper we have studied the electron scattering form factors in some sd-shell nuclei using a second quantized boson representation of the relevant operators. Starting from the BZM mapping procedure, an exact boson image of the microscopic shell-model framework
hamiltonian has been obtained and then solved approximately in the of the MF approach and the equations-of-motion method. The resulting
boson states were shown to be free from spurious components due to the violation of the Pauli principle in the boson space. The angular momentum and isospin eigenstates were projected out of them and used to calculating the electron scattering form factors, both for the elastic scattering (0: + 0:) and for the excitation of the 2t(i = 1,2) and 4: states. The results obtained in this way suggest the potential power and utility of the present boson approach for describing such complicated and model-sensitive physical quantities as the electron scattering form factors. At the same time, however, they indicate the urgent need for a careful and detailed treatment of the effects associated with the symmetry restoration and the interplay between the intrinsic and c.m. excitations. In this article, we have restricted our discussion to nuclei with an equal number of protons and neutrons. An extension of the formalism to nuclei with nrr Z nV, together with an application to “0, is currently under way and some preliminary results have been reported in ref. 30). The author
thanks
Dr. M. Gmitro
for his stimulating
interest
in this work.
References 1) H. tiberall, Electron scattering from complex nuclei (Academic, New York, T.W. Donnelly and J.D. Walecka, Ann. Rev. Nucl. Sci. 25 (1975) 329 2) G. Mukherjee and S.K. Sharma, Phys. Rev. C29 (1984) 2101;
1971);
108
R. Kuchta / Electron
scatter&g
P. Czerski, H. Miither, P.K. Rath, A. Faessler and W.H. Dickhoff, Nucl. Phys. A465 (1987) 189; A.E.L. Dieperink and E. Moya de Guerra, Phys. Lett. B189 (1987) 267 3) A. Van Egmond, K. Allaart and G. Bonsignori, Nucl. Phys. A436 (1985) 458 4) A. 1. Steshenko, Nucl. Phys. A445 (1985) 462; M. Bouten and M.C. Bouten, Nucl. Phys. A459 (1986) 253 5) P. Park and J.P. Elliott, Nucl. Phys. A448 (1986) 381 6) M.G. Vassanji and D.J. Rowe, Nucl. Phys. A466 (1987) 227 7) J. Dukelsky, G.G. Dussel, R.P.J. Perazzo, S.L. Reich and H.M. Sofia, Nucl. Phys. A425 (1984) 93 8) S.T. Belyaev and V.G. Zelevinsky, Nucl. Phys. 39 (1962) 582; E.R. Marshalek, Nucl. Phys. A224 (1974) 221; A224 (1974) 245 9) T. Holstein and H. Primakoff, Phys. Rev. 58 (1940) 1098; F.J. Dyson, Phys. Rev. 102 (1956) 1217; T. Marumori, M. Yamamura and A. Tokunaga, Prog. Theor. Phys. 31 (1964) 1009; D. Janssen, F. D&au, S. Frauendorf and R.V. Jolos, Nucl. Phys. AI72 (1971) 145 10) Y.K. Gambhir, P. Ring and P. Schuck, Nucl. Phys. A384 (1982) 37; M.R. Zirnbauer, Nucl. Phys. A419 (1984) 241; H.B. Geyer, Phys. Rev. C34 (1986) 2373; R.L. Wang, J.Z. Pan and C.G. Bao, Phys. Rev. C3S (1987) 324; H. Tsukuma, H. Thorn and K. Takada, Nucl. Phys. A466 (1987) 70 11) A. Arima and F. Iachello, Ann. of Phys. 99 (1976) 253; 111 (1978) 201; 123 (1979) 468 12) T. Otsuka, A, Arima and F. lachelio, Nucl. Phys. A309 (1978) I; 5 Pitted, D. Duval and B.R. Barrett, Phys. Rev. C25 (1982) 2834; A. Van Egmond and K. Allaart, Nucl. Phys. A425 (1984) 274 13) R.P. Hirsekon, Ph.D. thesis, University of Braunschweig (1975); R.P. Hirsekorn and L.J. Weigert, Nucl. Phys. A288 (1977) 439 14) P. Ring and P. Schuck, The nuclear many-body problem (Springer, New York, 1980) 15) R. Kuchta, Ph.D. thesis, Charles University, Prague (1987) 16) P. Ring and P. Schuck, Phys. Rev. Cl6 (1977) 801; H.B. Geyer and S.Y. Lee, Phys. Rev. C26 (1982) 642; G.K. Kim and CM. Vincent, Phys. Rev. C32 (1985) 1776 17) H.B. Geyer, CA. Engelbrecht and F.J.W. Hahne, Phys. Rev. C33 (1986) 1041; P. Park, Phys. Rev. C35 (1987) 807 18) T. Marumori, F. Sakata, T. Une and Y. Hashimoto, Prog. Theor. Phys. Suppl. 74,75 (1983); J. Dukelsky, S, Pitted, H.M. Sofia and C. Lima, Nucl. Phys. A456 (1986) 7s; W. Pannett and P. Ring, Nucl. Phys. A465 (1987) 379; C.H. Druce, S. Pittel, B.R. Barrett and P.D. Duval, Ann. of Phys. 176 (1987) 114 19) S. Pittel, J. Dukelsky, R.P.J. Perazzo and H.M. Sofia, Phys. Lett. BI44 (1984) 145; B.R. Barrett and P. Halse, Phys. Lett. B155 (1985) 133; A. Van Egmond and K. Altaart, Phys. Lett. B164 (1985) 1 20) D.J. Rowe, Rev. Mod. Phys. 40 (1968) 153; D.J. Rowe, Nuclear collective motion (Methuen, London, 1978) 21) V. Gillet and N. Vinh Mau, Nucl. Phys. 54 (1964) 321 22) J.M. Eisenberg and W. Greiner, Nuclear theory, vol. 3 (North-Holland, Amsterdam, 1972) 23) S. Gatienhaus and C. Schwartz, Phys. Rev. 108 (1957) 482 24) H.P. Jolly, Phys. Lett. 5 (1963) 289 25) Y. Horikawa, Phys. Lett. B36 (1971) 9; S. Mitsunohu and Y. Torizuka, Phys. Rev. Lett, 28 (1972) 920 26) J.M. Irvine et ai, Phys. Lett. B44 (1973) 16; R.P. Singhal et at., Phys. Lett. B76 (1978) 170 27) E. Baranger and C.W. Lee, Nuct. Phys. 22 (1961) 157; M. Gmitro, E. Tinkova, A. Rimini and T. Weber, Czech. J. Phys. B29 (1979) 155 28) P. M. Endt and C. Van der Leun, Nucl. Phys. A214 (1973) 1; A310 (1978) 1 29) A. Nakada and Y. Torizuka, J. Phys. Sot. Jpn 32 (1972) 1 30) R. Kuchta, JINR Dubna report E4-87-753 (1987)