Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Nuclear, Particle and Many Body Physics
Nuclear, Particle and Many Body Physics
Nuclear, Particle and Many Body Physics
Ebook1,310 pages12 hours

Nuclear, Particle and Many Body Physics

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Nuclear, Particle, and Many Body Physics, Volume I is a collection of scientific papers dedicated to the memory of Amos de-Shalit, a distinguished physicist. The book contains chapters that discuss various studies in the field of nuclear and particle physics. The text covers such topics as chemical binding in classical Coulomb lattices; algebraic treatment of subsidiary conditions in dual resonance models; the complete Schwarzschild solution; the investigation of nucleon-nucleon correlations by means of high-energy scattering; and some aspects on the magnetic-dipole moments of states in near-spherical nuclei. Theoretical, experimental, and nuclear physicists, researchers, and students in the field of physics will find the book invaluable.
LanguageEnglish
Release dateDec 2, 2012
ISBN9780323143035
Nuclear, Particle and Many Body Physics

Related to Nuclear, Particle and Many Body Physics

Related ebooks

Physics For You

View More

Related articles

Reviews for Nuclear, Particle and Many Body Physics

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Nuclear, Particle and Many Body Physics - Philip Morse

    1972.

    Chemical Binding in Classical Coulomb Lattices

    FREEMAN J. DYSON,     Institute for Advanced Study, Princeton, New Jersey 08540

    Received February 9, 1970

    Publisher Summary

    This chapter describes the calculation of Coulomb energies of classical Coulomb attics that are composed of two species of charge. Some of the compound lattices are chemically stable as compared with pure lattices at the same density. In particular, the sodium chloride structure — with a ratio around 0.07 between the two species of charge — is unusually stable. This chapter explains whether those results are relevant to the behavior of real matter at the densities existing in neutron-star crusts. The chapter additionally describes that the screening of the nuclear Coulomb potentials by electrons does not change the chemical binding appreciably. The Fermi sea of relativistic electrons is approximated by a uniform fluid of constant density. After the electron screening, the next quantum-theoretical contribution to the energy of a crystal comes from the zero-point motions of the nuclei.

    The Coulomb energies of classical Coulomb lattices composed of two species of charge are calculated. It is shown that some of these compound lattices are chemically stable as compared with pure lattices at the same density. In particular, the NaCl structure with a ratio around 0.07 between the two species of charge is unusually stable. The question is raised, whether these results are relevant to the behavior of real matter at the densities existing in neutron–star crusts. It is shown that the screening of the nuclear Coulomb potentials by electrons does not change the chemical binding appreciably. Other real–matter effects such as zero–point energies are discussed but not calculated. Thus the existence of chemical compounds such as FeHe in neutron–star crusts is made plausible but not reliably established.

    I COMPLETELY CLASSICAL SYSTEMS

    In connection with the external crustal layers of neutron stars (see Ruderman [1]), it is of interest to consider the behavior of matter at densities in the range 10⁸−10¹² g cm−3 and temperatures in the range 10⁶−10⁸ K. Under these conditions it may be reasonable to approximate the Fermi sea of relativistic electrons by a uniform fluid of constant density, and the nuclei by a crystal lattice of classical point Coulomb charges. We will examine in Sections II and III the lowest-order corrections to these approximations. In this section we take the classical Coulomb lattice as a starting point and calculate the chemical binding effects which occur in it.

    We confine our attention to lattices with cubic symmetry and to binary chemical compounds. Consideration of other types of lattice would require a larger computing effort but might well lead to the discovery of interesting new compounds. Suppose then that there are two types of positive point charges of magnitude Ze and Ye. We consider three possible compounds: (1) ZY, NaCl structure; (2) ZY, CsCl structure; (3) ZY3, body-centered structure. Compound (1) is a simple cubic (sc) lattice with the Z and Y sites forming face-centered cubic (fcc) sublattices. Compound (2) is a body-centered cubic (bcc) lattice with the Z and Ysites forming (sc) sublattices. Compound (3) is a sc lattice with the Z sites forming a bcc sublattice. We take a to be the edge of the sc lattice in cases (1) and (3), while 2a is the edge of the sc sublattice in case (2). The electron density is then

    (1.1)

    (1.2)

    and

    (1.3)

    in the three cases.

    We are interested in the Coulomb energy Ec of the lattice of N charges Ze together with N or 3N charges Ye and the uniform negative background. This energy is

    (1.4)

    (1.5)

    and

    (1.6)

    where x = (Y/Z). The numerical coefficients P1, P2, P3 are lattice sums, and α1, α2, γ are ratios of lattice sums. All the required sums have been evaluated by Coldwell–Horsfall and Maradudin [2], namely,

    (1.7)

    (1.8)

    (1.9)

    (1.10)

    (1.11)

    (1.12)

    In assessing the chemical stability of these compound lattices, we must compare them with separated pure lattices of elements Z and Y at the same pressure. In the approximation in which the electron-density is uniform, the pressure is a function of electron-density only. Therefore we must compare Ec with the energy Em of a mixture of pure lattices containing the same number of charges at the same electron density. The most stable pure lattice has the bcc structure. The mixed pure lattice energies are then

    (1.13)

    (1.14)

    (1.15)

    The degree of chemical stability is measured by the ratio

    (1.16)

    which is a function of the charge ratio x only and is independent of density. The compound will be stable at zero temperature whenever

    (1.17)

    The ratios are explicitly

    (1.18)

    (1.19)

    with

    (1.20)

    (1.21)

    The coefficient η1 is the ratio between the Coulomb energies of pure fcc and bcc lattices at the same density, while η2 is the ratio between pure sc and bcc lattices.

    In the first three columns of Table I, numerical values of R for the three lattices are displayed. The main conclusion is that each of the three lattices gives stability for a certain range of values of x, but that the magnitude of the chemical binding energy is in general extraordinarily small. For each lattice the range of stability is x1 < x < x2, with the values of x1 and x2 given in the first three columns of Table II. Also in Table II is shown the maximum value of R and the value of x at which the maximum is attained. The only case in which a substantial chemical binding exists is the NaCl lattice with a ratio x in the neighbourhood of 0.07. Two elements which may be expected to be abundant in neutron-star crusts are iron and helium, which give

    TABLE I

    R1, R2, R3 are ratios between the Coulomb energies of binary compounds and those of pure elements at the same electron density, R1 for NaCl structure, R2 for CsCl structure, R3 for ZY3 structure; x is the ratio between the two species of charge. Rs is the same as R1 including effects of electron screening, letting x = (Y/Z) vary while Y + Z = 28 is held constant.

    TABLE II

    For each structure, x1 < x < x2 is the range of charge ratio x for which a binary compound is chemically stable. Rmax is the maximum value of the stability ratio, xmax the value of x at the maximum, and RFeHe is the value of R for x = (2/26). In the last column Y + Z = 28 is held constant.

    (1.22)

    close to the maximum of the binding-energy curve. The last row of Table II gives the values of R for iron–helium lattices of all three types. We shall give primary attention in what follows to the stable compound FeHe with the NaCl structure.

    From Eq. (1.22), (1.13) and (1.1), the chemical binding energy per He atom in the compound FeHe is

    (1.23)

    Converted from electron density n to mass densit ρ, this becomes

    (1.24)

    Thus EB = 4.43 keV at ρ = 10⁸, and EB = 95.5 keV at ρ = 10¹². These energies are small compared with nuclear reaction energies or electron Fermi energies, but may nevertheless be significant in certain regions of a neutron–star crust.

    Because of the delicate behavior of the small difference (R1 − 1), we can have no confidence that the result (1.24) of the purely classical approximation has any validity at all for real matter. Quantum corrections which are in a perturbation-theoretic sense small might nevertheless change the sign of (R1 − 1) and destroy the whole chemical binding effect. Some of these corrections will be considered in Sections II and III.

    To conclude this classical discussion, we mention a naive geometrical argument which may help to explain why the NaCl lattice with x ˜ 0.07 should be the most strongly bound. Suppose that we are trying to pack space as densely as possible with spheres of two different sizes. It seems to be true (though it has never been mathematically proved) that the highest packing density is achieved when the spheres form an NaCl lattice with lattice constant a, the radii of the two types of sphere being 2−1/2a and (1 − 2−1/2) a. The larger spheres lie in a close–packed fcc configuration, and the smaller spheres are arranged so that each touches six of the larger spheres. The double packing has a density of 0.793, and the ratio between the volumes of the two types of sphere is

    (1.25)

    Now it is plausible that a double lattice of point charges will give a maximum Coulomb binding when they are arranged at the centers of a double sphere–packing of maximum density, and when the charges are proportional to the volumes of the spheres in which they lie. It is perhaps not entirely accidental that the volume ratio (1.25) turns out to be very close to the charge ratio xmax = 0.0720 of the most stable Coulomb lattice.

    II ELECTRON SCREENING

    We calculate the effect of electron screening on the Coulomb lattice in the Thomas–Fermi approximation, following Salpeter [3]. Let n the average value of n, and

    (2.1)

    which is given by Eqs. (1.1), (1.2), or (1.3) as the case may be. The local energy density of the Fermi distribution of electrons is F(n), where

    (2.2)

    is the local Fermi energy, and the local Fermi momentum p is related to n by

    (2.3)

    The total energy density expanded to second order in y is

    (2.4)

    where E is the electric field, ϕ the electrostatic potential, σ the charge density of the nuclei plus the uniform background, and

    (2.5)

    by (2.2) and (2.3). In the Thomas–Fermi approximation, y everywhere adjusts itself to minimize (2.4), and therefore

    (2.6)

    (2.7)

    and the potential ϕ satisfies the Yukawa equation

    (2.8)

    The only effect of electron screening is to replace all Coulomb potentials by Yukawa potentials with the screening length

    (2.9)

    given by (1.1), then (2.9) becomes

    (2.10)

    with the electron Fermi velocity given by

    (2.11)

    10⁶). Then, as Ruderman [4] observed, λ is proportional to a, and the screened potential of the lattice has a shape independent of density.

    Equation (2.8) is a linear approximation to the Thomas–Fermi equation, and is in fact the first term in a perturbation expansion in powers of k²a². It is, therefore, consistent to calculate the screening effects only to first order in k²a². According to (2.10), this means that we shall be calculating to first order in the parameter

    (2.12)

    By expressing the Yukawa potential as a Fourier series summed over the vectors of the reciprocal lattice, it is easy to derive the expression

    (2.13)

    for the screening correction to be added to the Coulomb energy (1.4) of the NaCl lattice. Here

    (2.14)

    and

    (2.15)

    where A1, A2 are the lattice–sums

    (2.16)

    summed over integers (n1, n2, n3) which are all even (for A1) or all odd (for A2). Numerical values of these sums have been tabulated by Jones and Ingham [5],

    (2.17)

    Thus

    (2.18)

    As in Sections I, we have to compare the screening energy ΔEc1 with the corresponding quantity ΔEm1 calculated for a mixture of pure bcc lattices of elements Y and Z at the same electron density. We find the correction to (1.13) to be

    (2.19)

    where

    (2.21)

    and A3 is the lattice sum (2.16) extended over integers (n1, n2, n3) with (n1 + n2 + n3) even. According to Jones and Ingham [5],

    (2.22)

    The ratio between (2.13) and (2.19) is

    (2.23)

    with

    (2.24)

    and β given by (2.17).

    The ratio between the total Coulomb energies of the NaCl lattice and the separated bcc lattices, including the screening corrections, is

    (2.25)

    with

    (2.26)

    Numerical values of RS as a function of x with Y + Z = 28 are shown in the last column of Table I. The last column of Table II shows, also for Y + Z = 28, the range of x for which the compound lattice is stable, the maximum value of Rs, and the value of Rs for the FeHe compound.

    The main result of this calculation of screening effects is that the screening does not change qualitatively the results of Sections I. Although the chemical binding of the classical lattice is produced by a delicate cancellation of terms, this delicate cancellation persists in the screening effects. The screening correction to the binding energy is, in spite of the cancellations, only 1% of the binding energy. The fractional magnitude of the screening correction is of the order (Z²/³/137) with a numerical coefficient smaller than unity, in accordance with the most naive perturbation–theoretical expectations.

    The binding energy of the FeHe compound including screening is

    (2.27)

    (2.28)

    The difference between these numbers and Eq. (1.23) and (1.24) is from a practical point of view completely insignificant. Since the screening effects are roughly proportional to (Z + Y)²/³, they remain insignificant even when elements heavier than iron are considered.

    It is of some interest to examine the behavior of Rs as a function of Z when x = 0. This is the ratio between the Coulomb energies of fcc and bcc lattices composed of pure element Z at equal density. According to Eqs. (2.25), (2.24), and (1.20),

    (2.29)

    (2.30)

    Although the screening correction tends to reduce the margin of stability of the bcc lattice, the bcc remains stable (Rs < 1) until

    (2.31)

    It is again remarkable to see that the very delicate balance between the fcc and bcc energies is not disturbed by screening for any realistic value of Z.

    Another by-product of this analysis is the following remark. The screening corrections which we calculated all arose from the term proportional to k² in the Yukawa potential

    (2.32)

    Thus the screening corrections are simply the Coulomb energy in a lattice of point charges interacting with forces independent of distance. If we imagine a lattice of positive charges repelling each other with forces independent of distance in a uniform negative background, the Coulomb energy is positive and so R1′ < 1 for a stable chemical compound. We find in fact from Eq. (2.23) that R1′ > 1 for all x. Therefore, in a world with distance–independent Coulomb forces, no stable binary lattice with the NaCl structure can exist.

    III OTHER QUANTUM–MECHANICAL EFFECTS

    After the electron screening, the next quantum–theoretical contribution to the energy of a crystal comes from the zero-point motions of the nuclei. The zero-point energy of a lattice of identical nuclei was calculated by Coldwell-Horsfall and Maradudin [2] and also by Carr [6]. It would not be difficult to extend these calculations to binary compound lattices, but unfortunately the precision of the results would not be sufficient for our purposes. Both Coldwell-Horsfall and Carr-used methods of approximate integration with errors which are not carefully controlled. The two results are stated to be in good agreement when they differ by about 1%. Since we are concerned with differences between compound and pure lattices which are of the order of a few parts per thousand, we need to calculate the energies reliably to one part in 10⁴. In order to do this it would be necessary to carry through an accurate numerical integration over the three dimensional phonon spectrum.

    The order of magnitude of the zero-point energy is easy to estimate. The zero-point energy per nucleus, in a binary compound of elements Y and Z, is roughly

    (3.1)

    where ω is the plasma frequency, a is the lattice spacing, and M is the reduced mass

    (3.2)

    This is to be compared with the Coulomb energy

    (3.3)

    The ratio between the two energies is

    (3.4)

    where a0 is the Bohr radius and m is the electron mass. For the FeHe compound, Eq. (3.4) becomes

    (3.5)

    For densities in the range 10⁸−10¹², the zero-point energy is less than the Coulomb energy but not by a large factor.

    It appears from this discussion that the zero-point energy might easily upset the small margins of chemical stability which we calculated in Sections I and II. However, it is possible that the zero-point energies of compound and pure lattices will again cancel almost exactly, just as the electron screening contributions did, in which case the FeHe compound will remain stable. The precise calculation of these zero-point energies remains a task for the future.

    Another important factor which we have not calculated in detail is the effect of a compound lattice in stabilizing a light nucleus against pycnonuclear burning. Kirzhnits [7] long ago suggested that hydrogen might be preserved by this mechanism in crystalline white-dwarf cores. In neutron-star crusts where the density is above 10⁸, hydrogen certainly cannot exist, and the survival of helium is problematical.

    If helium nuclei are imbedded in a FeHe compound lattice with NaCl structure, the iron nuclei create a considerable Coulomb barrier which the helium must penetrate in order to interact. A dimensional analysis of the barrier indicates that the helium reaction rate will be slowed down by the Gamow factor

    (3.6)

    as compared with the burning rate of pure helium at the same density, with r a numerical coefficient of the order of unity. The exponent in (3.6) is just the reciprocal of the ratio (3.4). Thus the lattice structure can preserve helium for times much longer than its normal lifetime, provided that the density is low enough for the zero-point energy of the lattice to be a small fraction of the Coulomb energy. An exact calculation of the reaction rates, showing under what conditions the FeHe lattice can actually survive for given periods of time, would be a difficult undertaking. It would be necessary to describe in detail the multidimensional space of distortions of the lattice by which two helium nuclei may be brought together. A beginning in this direction has been made by Salpeter and Van Horn [8].

    In conclusion, we may say that the results of Sections I and II make it plausible that stable chemical compounds exist among the elements occurring in neutron-star crusts. But the additional effects mentioned in Sections III must be evaluated in detail before the existence and properties of these compounds can be firmly established.

    ACKNOWLEDGMENT

    I am grateful to Dr. E. E. Salpeter, Dr. M. Ruderman, and Dr. P. A. G. Scheuer for information about previous work on this subject.

    REFERENCES

    1. RUDERMAN, M. Nature. 1968; 218:1128.

    2. COLDWELL-HORSFALL, R.A., MARADUDIN, A.A. J. Math. Phys. 1960; 1:395.

    3. SALPETER, E.E. Astrophys. J. 1961; 134:669.

    4. M. RUDERMAN Nyu Technical Report No. 6/69 (1969).

    5. JONES, J.E., INGHAM, A.E. Proc. Roy. Soc. A. 1925; 107:636.

    6. CARR, W.J. Phys. Rev. 1961; 122:1437.

    7. KIRZHNITS, D.A. Sov. Phys. JETP. 1960; 11:362.

    8. SALPETER, E.E., VAN HORN, H.M. Astrophys. J. 1969; 155:183.

    Algebraic Treatment of Subsidiary Conditions in Dual Resonance Models*

    S. FUBINI† and G. VENEZIANO††,     Center for Theoretical Physics and Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139

    Received March 13, 1970

    Publisher Summary

    The dual formulation of the resonance model in order to study the algebraic structure of the generalized Ward identities, within the context of the projective operator language, is discussed in this chapter. It explains that by introducing amplitudes with different types of trajectories, it is possible to construct new amplitudes, in which the unphysical requirement of imaginary external masses can be avoided. This chapter presents a procedure to understand the presence of Ward identities that can lead to a complete cancellation of all ghosts. It explains the importance of the Ward-like identities, which provides a mechanism for the partial cancellation of the unphysical states. This chapter also describes the existence of ghosts, which is just a consequence of the algebraic structure of the operators and of the properties of the lowest state.

    WE BOTH REMEMBER AMOS DE–SHALIT AS A WONDERFUL FRIEND AND TEACHER

    The Ward-like identities previously proposed in order to eliminate unphysical states (ghosts) from dual amplitudes and generalized by Virasoro, are investigated within the context of the projective operator language. We are able to generalize further Virasoro’s result to the case of amplitudes with different internal trajectories and to release his unphysical condition on the external masses. A purely algebraic treatment of the problem of ghost cancellation is also presented and shown to provide cancellation of ghosts up to the third excited level.

    I INTRODUCTION

    One of the difficulties of dual resonance models in their present form [1] is the appearance of unphysical states of negative norm within the complete set of resonances which factorize the general amplitude [2, 3]. The existence of those ghosts is directly related to the indefinite metric of the Lorentz group.

    Indeed the resonances which appear in the models can be obtained by means of an infinite set of operators [4] which are four–dimensional vectors. The unphysical states are related to the presence of time like operators with the bad sign in their commutation relations.

    It has been shown [2] that at least some of the unphysical states are not linearly independent so that it is possible to obtain for them a cancellation with states of positive norm. This is somewhat analogous to the cancellation, taking place in quantum electrodynamics, between longitudinal and time–like photons.

    The linear relations or Ward identities, of the dual resonance model, which can be expressed in a nice compact way in the framework of the operator formalism[5], were however too few to provide the possibility of a complete elimination of unphysical states.

    Recently, Virasoro [6] has shown the possibility of extending, in a special case, the Ward identities into a larger class of linear relations. This enlarged class of relations contains enough identities to give some hope for a complete ghost cancellation. It is unfortunate that in the case of external neutral particles the Virasoro extension of the Ward identities exists only in the unphysical case of space–like external momenta.¹

    The problem is therefore to find the general class of models in which the Virasoro relations are valid and to see whether the unphysical requirement of imaginary masses can in some way be dropped.

    In this paper we shall use the dual formulation² of the model in order to study the algebraic structure of the generalized Ward identities (Sections II and III). It will be shown (Section IV) that, by introducing amplitudes with different types of trajectories it is possible to construct new amplitudes (containing an extra set of scalar operators) in which the unphysical requirement of imaginary external masses can be avoided. The model is not yet physically acceptable since it has an unphysical pole for negative s coming from an internal trajectory with positive intercept. It does however suggest that a more thorough investigation might lead to physically satisfactory models in which the whole set of Ward identities is present. We finally sketch (Section V) a procedure by which one can try to answer the still open question of whether the presence of Ward identities can indeed lead to a complete cancellation of all ghosts.

    Although a general treatment is not yet available it is possible to show that all ghosts appearing in the first three excited levels are completely compensated.

    II PROJECTIVE INVARIANT FORM OF FACTORIZED AMPLITUDES

    We now follow the procedure of I in order to recast the general n-body amplitude in an operator form which exhibits directly the duality properties. In the present treatment we shall gain further generality since we have to drop the symplifying assumption of zero external masses and we allow one of the external particles to be in an excited state. We shall use the notation of I apart from the following changes:

    We start from the operator form of Ref. [4] which gives for the amplitude depicted in Fig. 1.

    FIG. 1 Kinematics of the process considered in this paper.

    (2.1)

    where all external particles have the same mass ki² = −μi² = α(0) = a. Introducing the zero frequency operators p0 and q0 of I we can rewrite (2.1) as

    (2.2)

    We have also defined

    (2.3)

    where

    (2.4)

    Finally

    (2.5)

    We observe that, because of the fact that ki² are different from zero we are forced to use the ordered form (2.3). The commutator

    is indeed singular for x y. In order to express U as a single exponential we have to use a limiting procedure like, e.g.,

    (2.6)

    where, as in I

    (2.7)

    Thus we have to investigate again the commutation relations between the vertex Uj(z) and the fundamental generators. This is done in the Appendix. The result is

    (2.6)

    (2.6′)

    and

    (2.7)

    We note that whereas Eqs. (2.6), (2.6′) are those expected from analogy with the treatment in I, the extra term in k² appearing in the right hand side of Eq. (2.7) originates from the ordered form of Eq. (2.3).

    By repeated use of (2.6′) we can shift all the xL0 operators to the left and obtain:

    (2.8)

    Equation (2.8) in the particularly simple case k² = a = 1 will be the starting point of the derivation (given in Sections III) of the generalized Ward identities. We want to conclude this introductory section by exhibiting explicitly the projective in variance in the case where | λ〉 is also in the unexcited level | 0〉.

    As in I we associate each external line with a projective variable ρ including the lines of momentum k0 and kn+1. The amplitude now becomes

    (2.9)

    where

    (2.10)

    and dV(ρ) is the usual invariant volume introduced in I.

    It is now easy to prove the invariance of ϕ(kiρi) under the simultaneous projective transformation of the variables ρi. It is enough to show that ϕ(ki, ρi) does not vary under infinitesimal transformation. We thus need to verify the equations

    and

    (2.12)

    This can be achieved by considering the identities

    (2.13)

    and

    (2.14)

    which follow from the fact that³Li| 0〉 = 0 (i = −1, 0, +1). Using Eq. (2.13) and the standard commutator (2.6) between L0 and Uj one obtains immediately Eq. (2.11). In the same way, Eq. (2.14) together with the commutator (see Eq. 2.7),

    (2.15)

    leads directly to Eq. (2.12). The projective invariance of the operator formulation of the dual amplitude can be seen more directly by using the operator Q(ρ, σ) defined in Eq. (2.7).

    As already pointed out in I one obtains

    (2.16)

    This form will turn out to be very useful when we shall extend our considerations to the case of a dual amplitude with different types of trajectories.

    III GENERALIZED WARD–IDENTITIES

    In this section we shall give a new derivation of the whole set of Ward–like identities. We shall use both the Li operators of I and their generalization including the new operators considered by Virasoro [6].

    All these operators belong to the class of operators Lf defined⁴ for each function f(z) through the commutation

    Enjoying the preview?
    Page 1 of 1