35
8/14/2019 TAYL16-001-035.I http://slidepdf.com/reader/full/tayl16-001-035i 1/35 SUBATOMIC PHYSICS Chapter 16 The Structure of Atomic Nuclei Chapter 17 Radioactivity and Nuclear Reactions Chapter 18 Elementary Particles In Parts I and II we developed the two theories, relativity and quantum me- chanics, that revolutionized twentieth-century physics. In Part III we described some of the many applications of these theories to systems containing several atoms — molecules, solids, and some of the systems treated by statistical me- chanics. Now, in Part IV we discuss subatomic systems, namely, atomic nuclei and subnuclear particles. As far as practical, we have made the chapters of Parts III and IV independent of one another so that you can select just those topics that interest you. In particular, you do not have to have read Part III at all before reading Part IV. Part IV contains three chapters. Chapters 16 and 17 describe the proper- ties of atomic nuclei, while Chapter 18 treats the subnuclear,“elementary” particles. The division between the two chapters on nuclear physics is that Chapter 16 is concerned with the properties of an isolated,static nucleus,while Chapter 17 treats nuclear reactions, in which nuclei change their energy and often their identity as well. These two chapters are not completely indepen- dent of one another, but if your main interest is in nuclear reactions (Chapter 17),you need to read only Sections 16.1 to 16.3 before skipping to Chapter 17.If your main interest is in particle physics (Chapter 18),you could read just Sections 16.1 to 16.3 and 17.1 to 17.4 before skipping to Chapter 18. PART IV F F P P O O 1 TAYL16-001-035.I 1/22/03 12:46 PM Page 1

TAYL16-001-035.I

  • Upload
    orion20

  • View
    218

  • Download
    0

Embed Size (px)

Citation preview

Page 1: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 1/35

SUBATOMICPHYSICS

Chapter 16 The Structure of Atomic NucleiChapter 17 Radioactivity and Nuclear ReactionsChapter 18 Elementary Particles

In Parts I and II we developed the two theories, relativity and quantum me-chanics, that revolutionized twentieth-century physics. In Part III we describedsome of the many applications of these theories to systems containing several

atoms — molecules, solids, and some of the systems treated by statistical me-chanics. Now, in Part IV we discuss subatomic systems, namely, atomic nucleiand subnuclear particles. As far as practical, we have made the chapters of Parts III and IV independent of one another so that you can select just thosetopics that interest you. In particular, you do not have to have read Part III atall before reading Part IV.

Part IV contains three chapters. Chapters 16 and 17 describe the proper-ties of atomic nuclei, while Chapter 18 treats the subnuclear, “elementary”particles. The division between the two chapters on nuclear physics is thatChapter 16 is concerned with the properties of an isolated, static nucleus,while

Chapter 17 treats nuclear reactions, in which nuclei change their energy andoften their identity as well. These two chapters are not completely indepen-dent of one another, but if your main interest is in nuclear reactions(Chapter 17), you need to read only Sections 16.1 to 16.3 before skipping toChapter 17. If your main interest is in particle physics (Chapter 18), you couldread just Sections 16.1 to 16.3 and 17.1 to 17.4 before skipping to Chapter 18.

PARTIV

F

FP

PO

O

1

TAYL16-001-035.I 1/22/03 12:46 PM Page 1

Page 2: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 2/35

2

C h a p t e r 16The Structure of Atomic Nuclei

16.1 Introduction16.2 Nuclear Properties16.3 The Nuclear Force16.4 Electrons Versus Neutrons as Nuclear Constituents16.5 The IPA Potential Energy for Nucleons16.6 The Pauli Principle and the Symmetry Effect16.7 The Semiempirical Binding-Energy Formula16.8 The Shell Model★

16.9 Mass Spectrometers★

Problems for Chapter 16

★These sections can be omitted without loss of continuity.

16.1 Introduction

In discussing the properties of atoms, we needed to know surprisingly little

about the atomic nucleus. In fact, for most purposes we needed only threefacts: The nucleus carries a positive charge the nucleus is so much heavierthan the electrons that we can usually treat it as fixed; and the nucleus is sosmall that — for the purpose of atomic physics — we can usually treat it as apoint particle.

In this and the next chapter we turn our attention to the internal struc-ture of atomic nuclei. We will find that in some ways the motion of the protonsand neutrons inside a nucleus is analogous to the motion of the electrons in anatom. Many of the ideas developed for atomic physics (the Schrödinger equa-tion, the Pauli principle, spin, etc.) are equally useful in nuclear physics.

Nevertheless, there are significant differences between atomic and nu-clear physics, of which perhaps the most important is the difference betweenthe forces involved. The force that holds electrons in their atomic orbits is thefamiliar electrostatic force, which was well understood long before the adventof atomic physics. In nuclear physics, the situation was quite different, at leasthistorically. As we will argue in Section 16.3, the force that holds protons andneutrons together in a nucleus is not a force that was known in classicalphysics, and even today we do not have a complete theoretical understandingof nuclear forces. Therefore, the history of nuclear physics during the last 80years is the story of a simultaneous effort to elucidate the nuclear force and to

explain the properties of nuclei. In this sense, the development of nuclear

Ze;

TAYL16-001-035.I 1/22/03 12:46 PM Page 2

Page 3: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 3/35

Section 16.2 • Nuclear Properties 3

*This arrangement has changed from time to time, and many older books use differentconventions.On those relatively rare occasions when we need to use the right subscriptto indicate the neutron number, be careful not to confuse it with the right subscriptused to indicate the number of atoms in a molecule.

physics has been more difficult than that of atomic physics. Nevertheless, we

will see that nuclear physicists have achieved a remarkable understanding of nuclear properties. We will find that this understanding is both interesting forits own sake and of great practical importance.

16.2 Nuclear Properties

We begin our study of nuclear physics with a survey of several nuclear proper-ties, some of which we met briefly in Chapter 3. In later sections we will see

that many of these properties can be understood with the help of the samequantum principles developed in Chapters 6 through 10.

Constituents

We have already stated that nuclei are made up of protons and neutrons. Aswe will describe in Chapter 18, these two kinds of particles are themselvesmade up of quarks, but for the present we can treat protons and neutrons as“elementary” particles.We denote the number of protons by Z , and the num-ber of neutrons by N . Since each proton has charge while neutrons areuncharged, the total charge of the nucleus is

The mass of a nucleus is

(16.1)

where B is the total binding energy, the energy required to pull the nucleusapart into its separate constituents.As discussed in Chapter 3, the binding-energy term, is usually less than 1% of Furthermore, the differ-ence between the masses and is only 1 part in 1000. Hence,we can oftenneglect the last term in (16.1) and set to give the approximation

where, as before, we have defined the mass number, or nucleon number,

which is the total number of nucleons (protons and neutrons) in the nucleus. Itis sometimes convenient to write some (and occasionally all) of the numbers

 A, Z ,and N explicitly with the chemical symbol. The agreed placement around

the chemical symbol X is now*

Some examples are

(16.2)11H0 , 1

2H1 , 24He2 , 6

12C6 , 2963Cu34 , 92

238U146

ZAXN

A = Z + N

mnuc L 1Z + N2mp = Amp

mn L mp

mnmp

mnuc .-B>c2,Z + N

mnuc = Zmp + Nmn -B

c2

+Ze.+e,

TAYL16-001-035.I 1/22/03 12:46 PM Page 3

16 001 035 1/22/03 12 46 4

Page 4: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 4/35

4 Chapter 16 • The Structure of Atomic Nuclei

Fortunately, it is usually not necessary to give all of the three numbers on thesesymbols, and we will generally include only as many as seem necessary for thediscussion at hand. (This may be a good time to remind you that the first threenuclei in (16.2) have their own special names: is, of course, the proton, de-noted p; the nucleus of heavy hydrogen, is called the deuteron,and is some-times denoted D; and the nucleus is called the alpha particle, denoted )

Two different nuclei, such as and that have the same numberof protons, are called isotopes. As discussed in Chapter 3, the correspondingatoms are chemically indistinguishable, or very nearly so. Two nuclei, such as

and with the same mass number are called isobars; and two nuclei,

such as and with the same number of neutrons are called isotones.The naturally occurring elements have atomic numbers in the range

and neutron numbers with The values of Z and N for all naturally occurring nuclei are plotted in Fig. 16.1, from which one seesthat for light nuclei N and Z are nearly equal; while for heavier nu-clei, N is a little larger than Z , with in the heaviest nuclei. The ten-dency to have is called the   symmetry effect . We will show inSection 16.6 that the symmetry effect results from the Pauli principle and theso-called charge independence of nuclear forces and that the tendency for N 

to be somewhat bigger than Z is due to the electrostatic repulsion of the pro-

tons in nuclei with large Z .Another important tendency among nuclei is the observed preference

for Z and N to be even, as illustrated in Table 16.1.The stable nuclei in whichboth Z and N are even outnumber comfortably those in which either Z or N isodd, and there are only four stable nuclei in whichboth Z and N are odd.Thistendency is the result of the so-called “pairing effect,” which is described inSection 16.8.

Sizes and Shapes of Nuclei

We described in Chapter 3 how Rutherford probed the atomic nucleus using

energetic alpha particles. (The main reason that he used alpha particles was sim-ply that they were readily available, since they are ejected with various different

N L Z

N L 1.5Z

1A f 402 0 … N … 146.1 … Z … 92

  714N7 ,613C7

  714N7 ,6

14C8

  817O9 ,8

16O8

a.4He

2H,

1H

20

20

40

60

80

40 60

   Z

80 100 120 140

  Z      N

FIGURE 16.1

The values of N and Z for allnaturally occurring nuclei. Stablenuclei are shown by black dots;unstable, by colored dots.

TAYL16-001-035.I 1/22/03 12:46 PM Page 4

TAYL16 001 035 I 1/22/03 12:46 PM Page 5

Page 5: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 5/35

Section 16.2 • Nuclear Properties 5

TABLE 16.1Numbers of stable nuclei.

Number of Z N  stable nuclei

Even Even 148Even Odd 51Odd Even 49Odd Odd 4

*Of course, because of their wave functions, the electrons’ probability distribution isspread out through the atom. This does not change the fact that the electron itself isvery small compared with the atom. In fact, current theory holds that the electron is apoint particle — unlike nucleons,which have a measurable size.

energies by many naturally occurring radioactive substances.) Rutherford’s ex-periments established that nuclei are much smaller than atoms,with radii of justa few femtometers, or fermis More recent experiments haveused artificially accelerated beams of electrons, protons, and alpha particles toprobe the size and shape of atomic nuclei. The results of these experiments aresummarized in the next few paragraphs.

The majority of nuclei are spherical.A number are slightly nonspherical,some being prolate (elongated, like an American football), and some oblate(flattened, like most pumpkins).In nonspherical nuclei the longest diameter istypically a few percent greater than the shortest diameter, although in a fewcases this difference can be as large as 30%. It is usually a reasonable approxi-mation to treat nonspherical nuclei as spherical, and, for the most part, we willconsider nuclei to be spherical.

The radii of nuclei increase steadily from about 2 fm for light nuclei (heli-um, for example) to about 7 fm for heavy nuclei (uranium, for example). It isfound that the radii of all nuclei are given approximately by the simple formula

(16.3)

where A is the number of nucleons and is a constant,

(16.4)

The significance of (16.3) is clearer if we consider the volume of the nucleus,

(16.5)

where is the constant Evidently, (16.3) implies that the volume of a

nucleus is proportional to A, the number of constituent nucleons. This meansthat the volume per nucleon is approximately the same in all nuclei. In otherwords, the density of nucleons inside the nucleus is the same in all nuclei.

The volume of a single isolated nucleon turns out to be about Thuswe can say that nucleons in a nucleus take up half of the total volume. Thismeans that the nucleons are quite closely packed inside a nucleus, quite likethe molecules in a liquid — and entirely unlike the electrons in an atom.*

V0>2.

V0

43

 pR03 .V0

  = V0  A

 V =43

 pR3=

43

 pR03

 A

R0 = 1.07 fm

R0

R = R0  A1>3

11 fm = 10-15 m2.

TAYL16-001-035.I 1/22/03 12:46 PM Page 5

TAYL16-001-035.I 1/22/03 12:46 PM Page 6

Page 6: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 6/35

6 Chapter 16 • The Structure of Atomic Nuclei

3 1017

2 1017

1 1017

0

5

27Al 97Mo 238U

10

(r ) 

r (fm)

FIGURE 16.2

The density, in as a functionof r in the nuclei of aluminum,molybdenum, and uranium. Thesegraphs can be accurately fitted byan analytic expression called theFermi function (see Problems 16.15,16.63, and 16.64).

kg>m3,

The radius of a nucleus is not an exactly defined quantity.This is evident

in Fig. 16.2, which shows the distribution of mass in three representative nu-clei. In no case is there a unique radius at which the density falls exactly tozero.The usual definition of R and the definition assumed in (16.3) and (16.4)is the radius at which the density falls to half its maximum value No-tice that, as expected, the central density is the same in all three pictures.Notice also that all three graphs fall to zero at about the same rate; that is, thethickness of the surface region is about the same in all three nuclei.

The actual density inside a nucleus is seen from Fig. 16.2 to be aboutsome 14 orders of magnitude greater than the density of ordi-

nary liquids or solids. (Remember that water has density ) The

enormous density inside nuclei is easy to understand if we recall that almost allthe mass of an atom is in the nucleus,whose volume is some times smallerthan the atom (Problem 16.9). Therefore, the nuclear density is times big-ger than the atomic density, which is itself about the same as that of ordinaryliquids or solids.

Nuclear Energy Levels

Like atoms, nuclei are found to have quantized energy levels. However, thespacing of the levels in nuclei is many times larger than that in atoms.In heavy

nuclei the level spacing is usually several tens of keV, while in light nuclei itcan range up to 10 MeV and more. Because their excitation energies are somuch larger than the energies involved in chemical reactions, nuclei remainlocked in their ground states during normal chemical processes. This is why itis often possible in atomic physics to treat the nucleus as a structureless bodywith no internal motion.

If a nucleus is raised to one of its excited states (when struck by an ener-getic proton, for instance), it can return to the ground state by emitting a pho-ton.This is analogous to the emission of photons in atomic transitions, the onlyimportant difference being that the photons emitted by nuclei are usually far

more energetic. Photons emitted in nuclear transitions (with energies rangingfrom keVs to MeVs) are called Measurement of their energy tells usthe separation of the two nuclear energy levels involved.

 Angular Momenta of Nuclei

Like the electron, both the neutron and proton have spin half.Thus the totalangular momentum J of a nucleus is the sum of the orbital angular momentaof all of its nucleons plus all of their spins. As with all angular momenta, themagnitude of  J is quantized; the possible values are labeled by a quantumnumber j , with

(16.6)J = ƒ J ƒ = 4  j1 j + 12 U

g-rays.

10141014

103 kg

>m3.

3 * 1017 kg

>m3,

r 102.r 1r2

TAYL16 001 035.I 1/22/03 12:46 PM Page 6

TAYL16-001-035.I 1/22/03 12:46 PM Page 7

Page 7: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 7/35

Section 16.3 • The Nuclear Force 7 

16.3 The Nuclear ForceThe force that holds electrons in their atomic orbits is the famil iar electrostaticattraction between the positive nucleus and the negative electrons. On theother hand, it should be clear that electrostatic attraction isnot what holds thenucleons in the nucleus. First, the neutron is uncharged and feels no electro-static force at all.Second, the protons are all positive and the effect of the elec-trostatic force is therefore to push the protons apart.

Since the gravitational force acts on all matter, charged or not, and isalways attractive,one might wonder if gravity could be the force that holds the

nucleus together. However, gravity is many orders of magnitude too weak, asthe following example illustrates.

Example 16.1

Compare the gravitational attraction between two protons with their electro-static repulsion.

The gravitational attraction is

while the repulsive electrostatic force is

The ratio of these forces is independent of r and is equal to

Clearly, the gravitational force is far too weak to overcome the electrostaticrepulsion of the protons and hold the nucleus together.

Since neither the electrostatic force nor gravity holds the nucleus togeth-er, there must be some other force between nucleons, and we call it the nuclearforce. The nuclear force must be stronger than the electrostatic force, in thesense that it must more than balance the electrostatic repulsion of the protons.

For this reason, the nuclear force is often called the strong force.A useful way to view the strength of the nuclear force is in terms of the

energy of a nucleon in the nucleus.As a crude model, we can imagine any onenucleon moving inside a cubical box whose side a is of order 6 fm (the ap-proximate diameter of a fairly light nucleus such as aluminum, ).We sawin Chapter 8, Equation (8.32), that the minimum kinetic energy for such aparticle is

(16.7)

  L

3p2

1200 MeV # fm

22

2 * 1940 MeV2 * 16 fm22 L 20 MeV

 Kmin L3 U

2 p2

2ma2=

3p21 Uc22

2mc2 a2

27Al

Fgrav

Felec

=Gmp

2

ke2=

16.7 * 10-11 N # m2

>kg2

2*

11.7 * 10-27 kg

22

19.0 * 109 N # m2>C22 * 11.6 * 10-19 C22L 10-36

Felec =ke2

r2

Fgrav =Gmp

2

r2

TAYL16-001-035.I 1/22/03 12:46 PM Page 8

Page 8: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 8/35

~1 fm ~2 fm

~100 MeV

   U r 

FIGURE 16.3

The nuclear potential energy of twonucleons as a function of theirseparation r .

8 Chapter 16 • The Structure of Atomic Nuclei

*In fact, the constancy of the density in nuclei involves several other effects in additionto the short-range repulsion. Nevertheless, the conclusion stated is correct.† Strictly speaking, the potential energy sketched in Fig. 16.3 is that for two nucleonswith zero angular momentum.The force between nucleons depends somewhat on their

angular momentum, but for our present purposes, the potential energy shown is themost important.

That is, the minimum kinetic energy of any one nucleon in a nucleus is of order

20 MeV. In order that the nucleus hold together, its total energy must be nega-tive.Therefore, the nuclear force on any nucleon due to the other nucle-ons must produce a potential-energy well of depth at least several tens of MeV.

While the nuclear potential well must be very deep, it must also be of very short range; that is, the force between nucleons must fall rapidly to zero astheir separation increases. Perhaps the simplest evidence for this comes fromexperiments in which a beam of protons is scattered by nuclei. Those protonsthat pass by a nucleus at large distance are influenced only by the Coulomb re-pulsion between the proton’s charge and the nuclear charge. Even protonspassing within 2 or 3 fm of the nuclear surface scatter in the pattern predicted

on the basis of a simple Coulomb repulsion. Only when the protons approachwithin 1 or 2 fm of the nuclear surface does the pattern of scattering change,indicating the presence of a strong attractive nuclear force. We conclude thatthe nuclear force between nucleons is negligibly small when they are morethan 2 or 3 fm apart.

At separations of 1 or 2 fm, the force between nucleons is strongly at-tractive, as we have just argued. However, at separations less than about 1 fmthe nuclear force becomes strongly repulsive. The simplest argument for thislast claim is that, as we saw in Section 16.2, the density inside a nucleus is ap-proximately constant and is the same for all nuclei. If the nuclear force were

purely attractive, we would expect nucleons at the center of a large nucleus tobe forced inward more tightly than in a small nucleus. The easiest way to ex-plain the observed constancy of nuclear densities is to assume that at veryshort distances the nucleons are held apart by a repulsive force, and thisproves to be the case.*

Today, much of the best evidence for the details of the nuclear forcecomes from nucleon–nucleon scattering experiments, in which a beam of neu-trons or protons is scattered by the protons in a hydrogen target, much asRutherford’s alpha particles were scattered by the nuclei in a metal foil. Themain conclusions of these experiments are summarized in Fig. 16.3, which

shows the potential energy of the nuclear force between two nucleons,plottedas a function of their separation r . You can see that the main features of thisgraph agree with our qualitative conclusions:The nuclear potential energy ap-proaches zero rapidly beyond about 2 fm. Inside 2 fm the force is attractiveand the potential energy becomes negative, dropping to a minimum of about

at about 1 fm. Inside this minimum, the force is repulsive and thepotential energy rapidly becomes positive and large.†

Charge Independence of Nuclear Forces

Two final properties of the nuclear force should be mentioned.First, it is foundthat for a given quantum state, the nuclear force between two nucleons is thesame whether they are both protons, or both neutrons, or one proton and oneneutron. This property is called the charge independence of nuclear forces —the nuclear force on a nucleon is independent of whether the nucleon ischarged (a proton) or uncharged (a neutron). It is charge independence that is

-100 MeV

A - 1

TAYL16-001-035.I 1/22/03 12:46 PM Page 9

Page 9: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 9/35

Section 16.4 • Electrons Versus Neutrons as Nuclear Constituents 9

responsible for the observed tendency for nuclei to have equal numbers of neutrons and protons as we describe in Section 16.6.

Much detailed evidence for the charge independence of nuclear forces

comes from studies of the interactions between two nucleons (p–p, p–n, n–n)in collision experiments. Further evidence comes from the study of energylevels in nuclear isobars, that is, two or more nuclei that have the same totalnumber of nucleons and differ only in their numbers of protons and neutrons.Consider, for example, the pair of nuclei and The only difference be-tween these is that one neutron in has been replaced by a proton inFigure 16.4 shows the lowest five energy levels of each of these two nuclei. Be-cause has one extra charge, all of its levels are raised by 1.7 MeV due tothe additional Coulomb repulsion. Apart from this obvious difference, the lev-els of the two nuclei are almost identical. The close correspondence of the

energy levels (both in excitation energy and angular momentum) is strongevidence that the nuclear force acts equally on neutrons and protons.

Electrons and the Nuclear Force

Although the nuclear force is so crucially important for neutrons and protons,it does not act on electrons at all — in just the same way that the electrostaticforce does not act on particles that are electrically neutral. The principal evi-dence for this claim comes from experiments in which high-energy electronscollide with nuclei; even when the electrons penetrate deep into the nucleus,

there is no evidence that they are affected by the strong nuclear force. Thismakes the electron a very useful probe of nuclei, since its interaction with thenucleons is just the well-understood electromagnetic interaction.

16.4 Electrons Versus Neutrons as Nuclear Constituents

With our general knowledge of nuclear properties and forces, we are nowready to examine several questions in more detail. The first question that weaddress concerns the constituents of nuclei.

Although we now know that nuclei consist of protons and neutrons, thiswas not always obvious. That nuclei contain protons was established in 1919

7Be

47Be3 .3

7Li4

47Be3 .3

7Li4

7.46 ( j   )25

7.21 ( j   )2

5

6.73 ( j   )25

4.57 ( j   )27

6.68 ( j   )25

4.63 ( j   )27

0.48 ( j   )21

0.53 ( j   )21

0 ( j   )23

0 ( j   )23

1.7 MeV

7Li 7Be

FIGURE 16.4

Energy levels of the isobars

and The numbers labelingeach level are its excitation energyin MeV and its angular-momentumquantum number j . Correspondinglevels are connected by a dashed

line; they have angular momentathat are exactly equal andexcitation energies that are verynearly so. This close agreementis evidence for the chargeindependence of nuclear forces.

47Be3 .

37Li4

TAYL16-001-035.I 1/22/03 12:46 PM Page 10

Page 10: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 10/35

10 Chapter 16 • The Structure of Atomic Nuclei

*The minimum, in (16.9) reflects that the maximum possible wavelengthis l = 2a.

lpx = h>2a,

when Rutherford observed protons ejected from collisions between alpha par-

ticles and nitrogen nuclei. The neutron was not discovered for another 12years, and at first it was natural to suppose that nuclei were made of protonsand electrons. Recall that the two basic facts about nuclei are that they havecharge and mass (approximately). Today we explain these by sayingthat the nucleus has Z protons, to give it charge , and neutrons (with

), to give it total mass However, the same two facts could beequally well explained by supposing that the nucleus has  A protons and

electrons. Because is so small, the mass would still be aboutand because the electron has charge the electrons would reducethe total charge to , as observed.

The picture of the nucleus as made of protons and electrons had two veryappealing features.First, it was economical: If correct, it would have meant thatall matter was made of just two kinds of particles, protons and electrons. Sec-ond, since electrons and protons have opposite charges, it would have beenconceivable that the nucleus was held together by electrostatic attraction.

The question that we address here is this: How do we know that the pic-ture of nuclei as made of protons and neutrons is correct and the proton–electron picture incorrect? We present two arguments.

The Energy of Nuclear Constituents

We have seen that the kinetic energy of a proton or neutron inside a nucleus isat least several MeV. This requires that the nuclear force be strong enough toconfine nucleons with these energies, and we have seen that the observed nu-clear force is sufficiently strong. Suppose, however, that the nucleus containedelectrons. It is a simple matter to estimate the minimum kinetic energy of anelectron in a nucleus, as in the following example.

Example 16.2

Assuming that an electron could be confined inside a nucleus of radius about3 fm (for example, ),estimate the electron’s minimum kinetic energy.

To estimate the kinetic energy of nucleons in a nucleus, we treated thenucleus as a rigid cubical box.We can do the same thing for an electron, butwe need to be careful because the electron mass is only so wemust anticipate that its motion may be relativistic. Therefore, we will startfrom the general relation

(16.8)

and find the minimum possible value of We saw in Section 8.3 that the requirement that the wave function vanish

at the walls of the box led to a minimum for each component of momentum:

(16.9)

where a is the length of the cube’s edge,and the same minimum applies toand as well.* Therefore, the minimum possible value of is

p2= px

2+ px

2+ px

2=

3h2

4a2

p2pz

py

px = Ukx =h

2a

p2.

E = 4 1pc22+ 1mc222

0.5 MeV>c2,

27Al

Ze

A - Z-e,Amp ;meA - Z

Amp .mn L mp

A - ZZe

AmpZe

TAYL16-001-035.I 1/22/03 12:46 PM Page 11

Page 11: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 11/35

Section 16.4 • Electrons Versus Neutrons as Nuclear Constituents 11

*Since the neutron is chargeless, one might expect it to have no magnetic moment.However, we now know that although its total charge is zero, there are charges inside

the neutron, and it is the motion of these charges that gives the neutron its magneticmoment.

Setting (the diameter of the nucleus), we find that

Substituting into (16.8), we obtain for the minimum energy

Note that this answer is much larger than and hence thatthe electron would indeed be moving relativistically. Therefore, the kinetic

energy, is close to the total energy and

(16.10)

This example shows that an electron confined inside a nucleus wouldhave kinetic energy of order at least 100 MeV. However, there is no knownforce that could confine an electron with this much kinetic energy. In the 1920sthe only known possibility was the electrostatic Coulomb force, and since wenow know that the nuclear force does not act on electrons, the Coulomb forceis still the only candidate. It is easy to estimate the electrostatic potential ener-

gy of an electron inside any given nucleus. For for example, one finds(Problem 16.25)

This energy is negative and would tend to bind the electron, but comparedwith (16.10) it is hopelessly too small to overcome the minimum kinetic ener-gy of the electron, and we conclude that there can be no electrons boundinside a nucleus.

 Magnetic Moments of Nuclei

Our second argument against nuclei made of protons and electrons concernstheir magnetic moments.We saw in Chapter 9 that the electron has a magneticmoment whose magnitude is given by the Bohr magneton,

(16.11)

where is the electron’s mass. (The exact magnetic moment depends on theorbital quantum number l but is always equal to times a number of order1.) In the same way we would expect the nucleons’ magnetic moments to be of 

order

(16.12)

where denotes the mass of the nucleon.This proves to be correct: Both theproton and neutron* have magnetic moments whose magnitudes are close tothe value (16.12), which is therefore called the nuclear magneton.

mN

mN =e U

2mN

mB

me

mB =e U

2me

U L -9 MeV

27Al,

Kmin L 180 MeV

K = E - mc2

,

mc2= 0.5 MeV,

E = 4 1pc22+ 1mc222

L 4 13 * 1042 + 10.252 MeV L 180 MeV

1pc22=

31hc22

4a2L

3 * 11240 MeV # fm22

4 * 16 fm22L 3.2 * 104 MeV2

a L 6 fm

TAYL16-001-035.I 1/22/03 12:46 PM Page 12

Page 12: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 12/35

12 Chapter 16 • The Structure of Atomic Nuclei

*The IPA is often called the independent-particle model , abbreviated IPM, or themean-field approximation.

The important difference between in (16.11) and in (16.12) is that,

since is about 2000 times smaller than the Bohr magneton is 2000times greater than Thus if there were any electrons in the nucleus, theirlarge magnetic moment would dominate, and the total magnetic momentwould be of order In fact, however, the measured magnetic moments of allnuclei are far smaller than this and are all of order — in keeping with theproton–neutron theory.

For these and other reasons (Problems 16.23 and 16.24) it is clear thatthere are no electrons inside the atomic nucleus, which consists, instead, of protons and neutrons.

16.5 The IPA Potential Energy for Nucleons

In studying atomic structure we relied heavily on the independent-particleapproximation, or IPA.* In this approximation each of the Z electrons wasassumed to move independently in the field produced by the average distribu-tion of the other electrons. Working within this approximation, wefound the possible states of each separate electron and, hence, those of theatom as a whole.In this way, we were led to the shell structure of atoms and toa remarkably successful account of the periodic properties of the elements.

The question naturally arises whether the corresponding approximationwould be useful in nuclear physics. In the early days of nuclear physics it wasgenerally believed that the answer would be no. The main reasons were thatthe nuclear force is so strong and of such short range. Figure 16.5(a) is aschematic representation of the potential energy of a nucleon moving alongthe x axis and interacting with just one other nucleon.(This is just the graph of Fig. 16.3 folded over, because x — unlike r — can be positive or negative, cor-responding to the nucleon being to the right or left of its fixed companion.)Figure 16.5(b) shows the potential energy of one nucleon in a one-dimensionalnucleus containing four other nucleons. The dashed curve shows a smoothed,

average potential energy, and at first sight it would seem an extremely badapproximation to replace the rapidly fluctuating solid curve by the smoothdashed curve.

In fact, however, the effect of the rapidly fluctuating nuclear potentialenergy is approximated surprisingly well by its smoothed IPA average. Thereare several reasons for this, some of which depend on subtle cancellations be-tween the attractive and repulsive parts of the nuclear force. However, onesimple reason is just that the system in question is a quantum system.Figure 16.5(b) represents the potential energy of a nucleon in the field of fourclassical nucleons, each at a definite position in the nucleus. But quantum me-

chanics tells us that the four nucleons are themselves spread out in accordance

Z - 1

mN

mB .

mN .mBmN ,me

mNmB

   P  o   t  e  n   t   i  a   l  e  n  e  r  g  y

 x x

(a) (b)

FIGURE 16.5

(a) The nuclear potential energy of one nucleon confined to the x axisand interacting with a secondnucleon on the x axis. (b) Thepotential energy of one nucleon inthe proximity of four othersfluctuates rapidly. Nevertheless, itseffect is well approximated by thesmoothed average, shown as adashed curve.

TAYL16-001-035.I 1/22/03 12:46 PM Page 13

Page 13: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 13/35

Section 16.5 • The IPA Potential Energy for Nucleons 13

R (force range)

50 MeV

FIGURE 16.6The nuclear IPA potential well,seen by one nucleon under

the influence of the othernucleons in a nucleus. The forcerange R can be defined as theradius at which the potential energyis half its maximum depth and isfound to be somewhat larger thanthe radius of the nucleus.

A - 1U1r2,

with their wave functions. This quantum spreading of the other particles is an

important factor in the success of the IPA in describing the motion of nucleonsin nuclei, as it is also for the case of electrons in an atom.

Figure 16.5 is one-dimensional, while the real nucleus is, of course, three-dimensional. We have seen that the nucleons are distributed nearly uniformlythrough a sphere of radius

(16.13)

Therefore, the average potential energy seen by each nucleon is found by av-eraging the nucleon–nucleon potential energy over this sphere. The resultingIPA potential energy, is spherically symmetric and is sketched in

Fig. 16.6. The radius of the potential well is given approximately by

(16.14)

Comparing this with (16.13), we see that, as one would expect, the force rangeis a little larger than the nuclear radius since the force can “reach out” a littlebeyond the distribution of nucleons. The depth of the IPA well is about50 MeV and is approximately the same for most nuclei.

The potential-energy function in Fig. 16.6 is the averaged nuclear poten-tial energy seen by each nucleon. Because of the charge independence of thenuclear force,this IPA potential energy is the same for neutrons as for protons.In the case of protons, however, we must add the potential energy of the elec-trostatic Coulomb force.To a good approximation this is the potential energyof a proton in the electric field of a uniform sphere of charge Out-side the nucleus this is

(16.15)

Inside the nucleus it has the shape of the inverted parabola shown in Fig. 16.7and is maximum at the center, where it is 1.5 times its value at the nuclear

surface (Problems 16.28 and 16.30):

(16.16)

The Coulomb potential energy (16.16) is easily calculated for any givennucleus (Problems 16.29 and 16.30).The answer varies from 1 or 2 MeV for the

UCoul102 =3

2 1Z - 12 

ke2

R

UCoul1r2 = 1Z - 12 

ke2

r

1r 7 R2 1Z - 12e.

R1force range2 = 11.17 fm2A1>3U

1r

2,

R1nuclear surface2 = 11.07 fm2A1>3

R

U Coul

(Z   1)32

ke2

R

FIGURE 16.7

The electrostatic potential energyof a proton in the electric field of 

the other protons in anucleus.

Z - 1

TAYL16-001-035.I 1/22/03 12:46 PM Page 14

Page 14: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 14/35

Proton Neutron

FIGURE 16.9

The IPA potential-energy functionsfor a proton and a neutron in amedium-mass nucleus.

14 Chapter 16 • The Structure of Atomic Nuclei

lightest nuclei to about 30 MeV for the heaviest. For protons in light nuclei, it is

therefore a fair approximation to ignore compared to the nuclear poten-tial energy of Fig. 16.6. For protons in heavy nuclei, this is certainly not a rea-sonable approximation,and we will have to use the total potential energy,

This function is plotted in Fig. 16.8, which shows the IPA potential well for aproton in the middle-mass nucleus (tin).

It is often convenient to display the IPA potential-energy functions forprotons and neutrons side by side. One convenient way to do this is to drawthem “back to back” as in Fig. 16.9. It is important to remember that althoughwe have drawn these two functions separately, they describe the potentialenergy of a proton and a neutron in one and the same nucleus.

Using the IPA potential-energy functions, one can calculate the energylevels of each nucleon in a nucleus. In this way we can construct a shell modelof the nucleus, quite analogous to the atomic shell model described inChapter 10.As we describe in Section 16.8, the nuclear shell model was devel-oped about 1950 and is very successful in explaining many (though certainlynot all) nuclear properties. Before taking up the shell model, we discuss twoother ideas that depend on the existence, though not the details, of the energylevels of the wells shown in Fig. 16.9.

16.6 The Pauli Principle and the Symmetry Effect

We saw in Fig. 16.1 that there is a tendency for stable nuclei to have nearlyequal numbers of protons and neutrons, especially when A is small. This ten-dency is often called the symmetry effect. Using the ideas of Section 16.5, wecan now show that the symmetry effect is a simple consequence of the Pauliprinciple and the charge independence of nuclear forces.

Like electrons, both protons and neutrons obey the Pauli exclusion prin-

ciple.That is, no two protons can occupy the same quantum state, and similar-ly, no two neutrons can occupy the same quantum state. It is important tounderstand that the Pauli principle restricts the states of two identical particlesonly; it places no restrictions on the states occupied by two particles of differ-ent type. Thus it is perfectly possible to have a proton and neutron both occu-pying exactly the same quantum state.

The ground state of a nucleus is found by assigning its Z protons and N 

neutrons to their lowest possible states consistent with the Pauli principle. Tosee how this works, let us first ignore the Coulomb repulsion of the protons, sothat the “wells” in which the protons and neutrons move are identical. (This

will be a good approximation for light nuclei, but not as good for heavier nu-clei.) Figure 16.10(a) shows the two wells and the lowest few levels in each,

120Sn

U1r2 = Unuc1r2 + UCoul1r2UCoul

   E  n  e  r  g  y    (   M  e   V    )

U Coul

U nuc U Coul

U nuc

10 r (fm)

40

20

0

20

Protons Neutrons

(2)

(4)

(2)

(a)

(2)

(4)

(2)

12

6C6

(b)

FIGURE 16.10

(a) Schematic diagram of the IPAwells and levels for protons andneutrons in a light nucleus. The twowells are essentially identicalbecause of charge independence.The numbers in parentheses arethe observed degeneracies. (b) The

ground state of is found byputting its six protons in the lowestavailable proton levels and its six

neutrons in the lowest availableneutron levels.

  612C6

FIGURE 16.8

The IPA potential energy of any oneproton in the nucleus (tin) isthe sum of two parts, and

A neutron feels only Unuc.UCoul .Unuc

120Sn

TAYL16-001-035.I 1/22/03 12:46 PM Page 15

Page 15: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 15/35

Section 16.6 • The Pauli Principle and the Symmetry Effect 15

along with their degeneracies. Since the wells are identical, the same is true of the levels.The Pauli principle requires that we place theZ protons in the low-

est proton levels, with the number of protons in a level never exceeding thatlevel’s degeneracy; and similarly with the neutrons. Figure 16.10(b) shows theresulting ground state of the nucleus (The precise degeneracies of thelevels do not matter for the present discussion, but we have shown their actualobserved values — two for the lowest level and four for the next.)

The example shown in Fig. 16.10(b), is a nucleus with Letus now consider a nucleus with the same number of nucleons, but thatis, one of the isobars of For example, consider the nucleus with fiveprotons and seven neutrons. The ground state of is shown on the left of Fig. 16.11, which also shows the ground states of and for comparison.

Comparing the ground state of with that of we see that, because of thePauli principle, the seventh neutron in has to occupy a higher level thanany of the neutrons or protons in Therefore, the nucleus has highertotal energy than

If we look on the other side of at we find a similar situation.The seventh proton of must occupy a higher level than any of the nucleonsin and has higher total energy than If we move still farther awayfrom (to for example), the total energy has to be even higher.

Similar arguments can be applied to any set of isobars. Based on thesearguments (in which we have so far ignored the Coulomb energy of the pro-

tons), we can state the following general conclusion:Among any set of isobars(nuclei with the same total number of nucleons) the nucleus (or nuclei) withZ 

closest to N will have the lowest total energy.To understand what this conclusion implies, we need to anticipate a prop-

erty of radioactive decay, which we discuss in Chapter 17. If a nucleus has ap-preciably more energy than a neighboring isobar, the nucleus with greaterenergy is unstable and will eventually convert itself into the isobar with lowerenergy. The process by which this occurs is called decay and is discussed inSection 17.4, but for now the important point is this:Any nucleus withZ muchdifferent from N  is unstable and will convert itself into a neighboring isobar

with Z closer to N and hence lower total energy. Therefore, the stable nuclei,and hence the nuclei normally found to occur naturally, should all have

In discussing the symmetry effect, we have so far ignored the electrosta-tic energy of the protons. This is easily taken into account: As discussed inSection 16.5, the Coulomb energy simply raises the proton well relative to theneutron well. This is shown in Fig. 16.12(a), which shows the two wells and aschematic set of energy levels. If the neutron well is appreciably lower than theproton well, there may be several neutron levels below any of the proton lev-els. In this case the most stable isobar will have more neutrons than protons, asindicated schematically in the figure.

For light nuclei the difference between the proton and neutron wells issmall, and our original conclusion that stable nuclei should have isZ L N

Z L N.

b

8O4 ,Z = N6C6 .7N56C6 ,

7N5

  712N5 ,6

12C6

6C6 .5B76C6 .

5B7

6C6 ,5B7

7N56C6

5B7

  512B76

12C6 .Z Z N,Z = N.6

12C6 ,

  612C6 .

Z   N Z   N 

E

125B7

126C6

Z   N 

127N5

FIGURE 16.11

The ground states of the threeisobars and Becauseof the Pauli principle, the two nucleiwith have higher energy bythe amount shown as ¢E.

Z Z N

12N.12C,12B,

TAYL16-001-035.I 1/22/03 12:46 PM Page 16

Page 16: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 16/35

16 Chapter 16 • The Structure of Atomic Nuclei

Z   N 

(a) (b)

FIGURE 16.12

(a) In nuclei with many protonsthe Coulomb repulsion pushes theproton well appreciably higherthan the neutron well. For a givennumber of nucleons, the lowestenergy is obtained when the highestoccupied proton and neutron levelshave the same energy (as indicatedby the arrow). This requires that Z be somewhat less than N . (b) Thesituation is closely analogous to apair of buckets of water, whose

total energy is lowest when thetop surfaces are level and thehigher bucket has less water.A pipe between the buckets letsthe water levels adjust to this stateof lowest energy, whatever theinitial configuration. In nuclei thecorresponding adjustment occursthrough radioactive decay.

*There is, of course, a small correction due to the binding energy of the electrons.Since

average electron binding energies are at most several keV, while nuclear binding ener-gies are several MeV, this correction is almost always negligible.

unaltered. However, as we move to heavier nuclei, the Coulomb energysteadily rises and the ratio should increase slowly, as is confirmed by thedata in Fig. 16.1.

It is perhaps worth reiterating the important role of the Pauli principle inthe symmetry effect. In Fig. 16.12(a) it is clear that without the Pauli principle,the state of lowest energy would consist in all (or almost all) the nucleonsbeing neutrons, all occupying the lowest neutron level. Under these conditionsthe majority of elements that make up our world would not exist.

16.7 The Semiempirical Binding-Energy Formula

The binding energy B of a nucleus is the total energy needed to separate com-pletely its  A individual nucleons. Binding energies of different nuclei are auseful measure of their relative stabilities as we will see later. In this sectionwe discuss how the binding energy depends on the nuclear parameters Z , N ,and We start with a brief discussion of the measurement of bind-ing energies.

As mentioned in Section 3.5, nuclear binding energies are large enoughto cause an observable difference (of order 1%) between the mass of a nucleusand the sum of the separate masses of its Z protons and N neutrons:

(16.17)

where denotes the mass of the nucleus concerned. Given accurate mea-surements of the masses involved, we can use this relation to findB as

(16.18)

Most mass tabulations, including that in Appendix D, give the mass of theatom, rather than that of the bare nucleus, since it is that is usually mea-sured. The mass is equal to plus the mass of the Z atomic electrons*

matom = mnuc + Zme

mnucmatom

matom

matom

B = 1Zmp + Nmn - mnuc2c2

mnuc

mnuc = Zmp + Nmn -B

c2

A = Z + N.

N>Z

TAYL16-001-035.I 1/22/03 12:46 PM Page 17

Page 17: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 17/35

Section 16.7 • The Semiempirical Binding-Energy Formula 17 

Now, we can add to in (16.18) if we also add to the term

This replaces by

where is the mass of the hydrogen atom This means that we canrewrite (16.18) as

(16.19)

and work entirely with atomic masses.

Example 16.3

Use the atomic masses in Appendix D to find the binding energy of thenucleus

According to (16.19), the required binding energy is

or, with the data from Appendix D (rounded to five significant figures),

(16.20)

(16.21)

Notice in (16.20) that the mass of the separate constituents is greater than thatof confirming that we must add energy to pull the nucleus apart. Noticealso that (16.20) gives B as the difference of two masses that are very nearly

equal. This difference is always of order 1% or less.Thus, we must know themasses to at least two significant figures more than are required in our an-swer. Here we needed five significant figures in the masses to get three inB.

As this example shows, one can find the binding energy of a nucleus if one knows the corresponding atomic mass accurately. Using a mass spectrom-eter, it is possible to measure atomic masses very accurately (see Section 16.9),and this is therefore one of the best ways to find nuclear binding energies.

The direct measurement of binding energies, by complete separation of anucleus, is seldom possible. However, it is often possible to find the binding

energy of one nucleus in terms of the known binding energy of some othernucleus. For example, one can measure the energy needed to separate oneneutron from a nucleus.This is called the neutron separation energy,

It is easy to see that the binding energy of the nucleus is

(16.22)

that is, the energy to dismember completely equals the energy to pulloff one neutron plus the energy to dismember the remaining nucleus 1Z, N - 12.1Z, N2

B

1Z, N

2= Sn

1Z, N

2+ B

1Z, N - 1

2

1Z, N2Sn1Z, N2 = energy to separate one neutron from nucleus 1Z, N2

35Cl

  = 298 MeV

  = 10.329 u2 *931.5 MeV>c2

1 u

  = 335.289 u - 34.969 u4c2 B = 31711.0078 u2 + 1811.0087 u2 - 134.969 u24c

2

B = 317mH + 18mn - matom135Cl24c2

1735Cl18 .

B = 1ZmH + Nmn - matom2c2

(1H).mH

Z1mp + me2 = ZmH

Zmp

Zmp .ZmemnucZme

TAYL16-001-035.I 1/22/03 12:46 PM Page 18

Page 18: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 18/35

2 fm

2 fm

FIGURE 16.13

A nucleon in a nucleus is bound

only to those neighbors within asphere of radius about 2 fm (uppersmall sphere); thus all interiornucleons are bound to aboutthe same number of neighbors.A nucleon near the surface isbound to fewer neighbors (lowersmall sphere), and this slightlyreduces the total binding energy.

18 Chapter 16 • The Structure of Atomic Nuclei

Thus, if the binding energy of is known, we can use (16.22) to find

the binding energy of or vice versa.

Example 16.4

The binding energy of is known to be 127.6 MeV. It is found that4.2 MeV is needed to separate a neutron from What is the bindingenergy of 

According to (16.22), the required binding energy is

In the remainder of this section we derive a formula that gives the bind-ing energy of nuclei as a function of the numbers Z , N , and All of the terms in this formula have at least partial theoretical justifications, but thevalues of the coefficients that appear are generally found by fitting the formu-la to the measured binding energies. For this reason the formula is called thesemiempirical binding-energy formula, or the Weizsacker semiempiricalformula, after its inventor. If we insert the formula for B into the equation

the resulting equation is called the semiempiri-cal mass formula.To an excellent approximation, the binding energy of a nucleus is found

to be proportional to the number, A, of its nucleons. This result is easily under-stood by the following classical argument: We have seen that the nuclear forcehas a range of about 2 fm, which is smaller than most nuclear sizes. For mostnuclei, then, each nucleon is bound to only a fraction of the other nucleons,namely, those neighbors within a sphere of radius about 2 fm, as shown inFig. 16.13. We have also seen that the density of nucleons is about the same in-side all nuclei. It follows that the number of neighbors to which each nucleon

is bound is approximately the same for all nuclei.That is, the average bindingenergy of each nucleon is the same in all nuclei, so that the total binding ener-gy is proportional to A. Since A is proportional to the volume of the nucleus,this approximation is called the volume term and is written

(16.23)

where is a positive constant.There is an immediate correction to the approximation (16.23). Those

nucleons near the nuclear surface have fewer neighbors to attract them

(Fig.16.13) and are less tightly bound than those in the interior.The number of these surface nucleons is proportional to the surface area of the nucleus,Accordingly, (16.23) must be reduced by an amount pro-

portional to and our next approximation is

where the surface correction is negative and has the form

(16.24)

with a positive constant.asurf 

Bsurf  = -asurf  A2

>3

Bsurf 

B L Bvol + Bsurf 

A2>3,4pR2= 4pR0

2 A2>3.

avol

B L Bvol = avol  A

mnuc = Zmp + Nmn - B>c2

,

A = Z + N.

  = 14.2 MeV2 + 1127.6 MeV2 = 131.8 MeV

 B

117O

2= Sn

117O

2+ B

116O

2

  817O9 ?

  817O9 .

  816O8

1Z, N2 1Z, N - 1

2

TAYL16-001-035.I 1/22/03 12:46 PM Page 19

Page 19: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 19/35

Section 16.7 • The Semiempirical Binding-Energy Formula 19

A further correction is needed because of the electrostatic repulsion of 

the protons.This reduces the binding energy by an amount equal to the poten-tial energy of the total nuclear charge. The potential energy of a uniformsphere of charge is (Problem 16.39)

(16.25)

Since this implies a correction of the form

(16.26)

where

Since the nucleus is not precisely a uniform sphere of charge, this is not the

exact value of The next correction arises from the symmetry effect discussed in

Section 16.6, where we saw that for any set of isobars, the nucleus withhas highest binding energy. Since B gets less as increases in either di-rection, this suggests a negative term proportional to The observedbinding energies of all nuclei are best fitted by a correction of the form

(16.27)

The factor of  A in the denominator reflects that for larger nuclei, the levelspacing is smaller and the difference of energy between neighboring isobars(shown as in Fig. 16.11) is therefore smaller.

The final term in our formula for the binding energy reflects a phenome-non that we will discuss in more detail in the next section. We mentioned inSection 16.2 that the majority of stable nuclei (148, in fact) have bothZ and N 

even; there are 100 nuclei with Z even but N odd or vice versa; and there areonly 4 in which both Z and N are odd.This preference for Z and N to be evenis explained by a property of the nuclear force called the pairing effect. For

now, we simply state that this effect gives a correction to the binding energywith the form

(16.28)

where is some constant and

(16.29)e=

c1  if  Z and N are both even

if  Z or N is even and the other is odd-1  if  Z and N are both odd

apair

Bpair = e 

apair

A1>2

¢E

Bsym = -asym 

1Z - N22

A

1Z - N22.Z - N

Z = N

aCoul .

aCoul L3

5 ke2

R0

BCoul = -aCoul Z2

A1>3R = R0 A1>3,

UCoul =3

5 k1Ze22

R

Ze

TAYL16-001-035.I 1/22/03 12:46 PM Page 20

Page 20: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 20/35

20 Chapter 16 • The Structure of Atomic Nuclei

0

2

50 100 150 200

4

6

8

10

1H

2H

3He

6Li

4He

56Fe

208Pb

238U

 A

   B   /   A    (   M  e   V    )

FIGURE 16.14

The binding energy per nucleon,as a function of  A. The points

are measured values. (Black dotsindicate stable nuclei; colored,unstable.) The curve is theprediction of the binding-energyformula (16.30). Values of areshown for the most stable isobar of each given A. For onlyeven–even nuclei are shown, andfor clarity, only half the data areincluded. The curve was calculated

from (16.30), assuming that both Z and N are even.

A 7 7,

B>AB>A,

Putting all five of these contributions together, we obtain the

semiempirical binding-energy formula

(16.30)

where is defined in (16.29). The five coefficients are generally adjusted to fitall the measured binding energies as well as possible. One set of values thatgives a good fit is as follows:

(16.31)

Because the dominant behavior of B is it is usual to discussthe binding energy per nucleon. Figure 16.14 shows the measured value of 

for a large number of nuclei. The smooth curve is the value of pre-dicted by (16.30) and is seen to give an excellent fit for all but the lightest nu-clei (those with  A below about 20). It is not surprising that the smooth

function (16.30) does not work well for the light nuclei, in which the quantiza-tion of energy levels is proportionately more important. Nevertheless, evenwhen A is small, the general trend of is correctly given by (16.30).

The most obvious feature of Fig. 16.14 is that is close to 8 MeV foralmost all nuclei, with the maximum value of 8.8 MeV for iron (Fe). Beyondiron, slopes gently down to about 7.6 MeV for this decrease is duemainly to the increasing importance of the Coulomb repulsion of the protons.On the other side of when A decreases below about 20, falls rapidlyto zero for (which has no binding energy); this decrease occurs because al-most all nucleons in a small nucleus are close to the surface and the negative

surface correction is proportionately large.The lower values of when A is large imply that energy is released

when a heavy nucleus splits, or fissions, into two lighter nuclei.The lower valuesB>A

1HB>A56Fe,

238U;B>A B>AB>AB>AB>A B>A,B r A,

apair = 11.2 MeV

 aCoul = 0.711 MeV asym = 23.7 MeV

 avol = 15.75 MeV asurf  = 17.8 MeV

e

  = avol  A - asurf  A2>3

- aCoul Z2

A1>3 - asym 

1Z - N22

A+ e 

apair

A1>2 B = Bvol + Bsurf  + BCoul + Bsym + Bpair

TAYL16-001-035.I 1/22/03 12:46 PM Page 21

Page 21: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 21/35

Section 16.8 • The Shell Model 21

of when A is very small imply that energy is also released when two light

nuclei fuse to form one heavier nucleus. Nuclear fission is the source of energyin nuclear power stations, while fusion provides the energy of the stars, includ-ing our own sun. Both fission and fusion are used in nuclear weapons. Clearly,the behavior of as shown in Fig.16.14, is of great practical significance.

Example 16.5

Use the binding-energy formula (16.30) to predict the binding energy of and compare the result with the measured value found in (16.21).

Inserting the values and the coefficients

(16.31) into the formula (16.30), we find that

which agrees very well with the measured value of 298 MeV. Notice that thepairing term is zero in this example becauseZ is odd and N is even.

16.8 The Shell Model★

★The shell model is an important part of our understanding of nuclei and is a beautifulapplication of many of the ideas of the atomic shell model.Nevertheless,we will not beusing the ideas of this section again, and you could omit it without loss of continuity.

The semiempirical binding-energy formula gives a remarkably accurate pic-ture of the general trends of nuclear binding energies. Nevertheless, it is asmoothed average picture and shows none of the fluctuations that one would

expect if nucleons, like atomic electrons, occupy energy levels that fall into dis-tinct shells. Recall that atomic binding energies fluctuate periodically withZ ,being greatest when the electrons exactly fill a closed shell and droppingsteeply just beyond each closed shell. In this section we discuss the corre-sponding shell properties of nuclei.

The Magic Numbers

The total numbers of electrons in the closed-shell atoms are known to be

(16.32)

By the late 1940s, substantial evidence had accumulated that nuclei have anal-ogous closed-shell numbers which lead to unusual stability, and that thesemagic numbers, as they came to be called, were

(16.33)

A nucleus is especially stable if Z or N is equal to one of these magic numbers,

and even more so if both are. The magic numbers are the same for protons asfor neutrons, as one would expect from the charge independence of nuclear

2,  8,  20,  28,  50,  82,  126

nuclear magic numbers:

2,  10,  18,  36,  54,  86

atomic closed-shell numbers:

  = 297 MeV

  = 1551.2 - 190.5 - 62.8 - 0.7 + 02 MeV

 B = avolA - asurf  A2>3

- aCoul Z2

A1>3 - asym 1Z - N22

A+ e 

apair

A1>2

N = 18Z = 17,A = 35,

35Cl,

B>A,

B

>A

TAYL16-001-035.I 1/22/03 12:46 PM Page 22

Page 22: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 22/35

22 Chapter 16 • The Structure of Atomic Nuclei

1110987

20 40 60

(a)

(b)

80 100 120

50

82

28

20

Sp(Z ) (MeV)

101112

13

9876

20 40 60 80 120100

50

82

126

2820

Sn(N ) (MeV)

FIGURE 16.15

(a) The proton separation energy

averaged over all stableisotopes for each Z . (b) Theneutron separation energyaveraged over all stable isotones foreach N . The dashed curves showthe predictions of the binding-energy formula (16.30). Only evenvalues of Z were included in (a) andonly even N in (b).

 S n1N2S p1Z2

*Note, however, that nuclei with have not been observed. Therefore, there isno experimental evidence for a proton magic number in this region.Z = 126

forces.* On the other hand, the nuclear magic numbers are quite different

from the atomic closed-shell numbers (16.32), reflecting the great differencebetween the dominant forces in nuclei and atoms.

The evidence for the magic numbers (16.33) consists of many observa-tions, few of which are conclusive by themselves. Nuclei in which Z is magictend to have larger numbers of stable isotopes than average. For example, tin,with Z equal to the magic number 50, has 10 stable isotopes, more than anyother element. Similarly, nuclei in which N is magic tend to have more stableisotones than average; holds the record, with 7 stable isotones. Thebinding energy per nucleon, in each of the “doubly magic” nuclei

and is markedly higher than the prediction of the

binding-energy formula. (This is easily discernible in the case of on thegraph of in Fig. 16.14.)

Perhaps the best evidence for the magic numbers (16.33) concerns the sep-aration energies and These are the energies required to remove one protonor one neutron from a nucleus and are analogous to the ionization energy of anatom. Recall that a plot of atomic ionization energy against number of electrons(Fig. 10.11) shows a maximum at each of the closed shells, with an abrupt dropimmediately thereafter. The nuclear separation energies depend on two vari-ables, Z and N , and are harder to plot. One simple procedure is to eliminate oneof these variables by an appropriate averaging. If we consider, for example, the

proton separation energy then for each value of Z we can computethe average for all stable isotopes with this Z . A plot of against Z 

is shown in Fig. 16.15(a) and, as expected, has pronounced drops immediatelyafter each of the magic numbers indicated. (We have not included data for

since our averaging procedure obscures shell effects for very light nu-clei.) The dashed curve in Fig. 16.15(a) shows the proton separation energyZ 6 20,

 S p1Z2 S p1Z2 Sp1Z, N2,

Sn .Sp

B>A 2He2

82Pb12620Ca28 ,20Ca20 ,8O8 ,2He2 ,B>A,

N = 82

TAYL16-001-035.I 1/22/03 12:46 PM Page 23

Page 23: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 23/35

1 g

2 p

1 f 

1 p

1 s

2 s1d

18

6

14

210

6

2

40

58

20

8

2

Z or N 

(for closed shells)

FIGURE 16.16

The (incorrect) filling order of thelevels for a proton or neutron,calculated for the nuclear potentialwell of Fig. 16.6. The number to theright of each level is its degeneracy.The levels fall into distinct energyshells, but the closed-shell numbersdo not agree with the observedmagic numbers (16.33). It followsthat these levels are not correct.

Section 16.8 • The Shell Model 23

*Recall that in atomic physics the lowest  p state is called , the lowest d state and

so on. This scheme is very convenient in the hydrogen atom (where it means that allstates of given n have the same energy) but has no particular merit in nuclear physics.

3d,2p

predicted by the binding-energy formula.This reproduces the general trend very

well but entirely misses the fluctuations associated with the magic numbers.Figure 16.15(b) shows the average neutron separation energies

each averaged over all stable isotones of a given N . This showsabrupt drops beyond the same magic numbers, 20, 28, 50, and 82, as inFig. 16.15(a), and also at

Origin of the Magic Numbers

Having established the existence of the nuclear magic numbers (16.33), physi-cists naturally hoped to explain them in terms of nucleon energy levels and thegrouping of those levels into energy shells. The IPA potential energy felt byany one proton or neutron was described in Section 16.5 and sketched inFig. 16.6. The energy levels for this well can be calculated numerically, and theordering of the levels is shown in Fig. 16.16.

The levels shown in Fig. 16.16 are labeled according to the schemenormally used in nuclear physics. This scheme is nearly, although not quite,the same as that used in atomic physics. Just as in atomic physics, each levelis identified by two quantum numbers, n and l . The number l identifies the

magnitude of the nucleon’s orbital angular momentum,The possible values 1,2, 3, are indicated by the code letters s, p,

d, For each value of  l , there are possible values of (with ) and two possible values of The de-generacy of each level is therefore Evidently, all of the parame-ters associated with angular momentum are the same as in atomic physics.On the other hand, nuclear physicists define the quantum numbern so thatthe levels of a given l are simply numbered 2, Thus the levelswith are labeled in order of increasing energy; the lev-els with are and so on.*

The levels (or subshells) shown in Fig. 16.16 are grouped into well-defined energy shells,and the total numbers of protons or neutrons needed to

produce closed shells are indicated on the right. The first three of these num-bers agree exactly with the observed magic numbers, but beyond this, all of thepredicted closed-shell numbers turn out to be wrong.

At first it would seem natural to assume that a different shape for theIPA potential well might produce the correct magic numbers. In fact, howev-er, no reasonable well gives the observed numbers. The correct explanationwas discovered independently in 1949 by Maria Goeppert-Mayer and byJohannes Jensen and coworkers. It turns out that the nuclear force dependsstrongly on the relative orientation of each nucleon’s spin and orbital angularmomentum. Specifically, the IPA potential well of Fig. 16.6 is the correct

potential-energy function for a nucleon with but if the nucleonhas an additional potential energy that depends on the orbital angular mo-mentum L and its orientation relative to the spin S. This additional energyis found to have the form

¢U r SL cos u

¢U

l Z 0,l = 0,

3d, Á ,2d,1d,l = 23p, Á2p,1p,l = 1

3, Á .n = 1,

212l + 12.Sz = ;  

12

  U.- ll - 1, Á ,m = l,Lz = m U2l + 1g,

Á. f,

4, Ál = 0,L = 4 l1l + 12 U.

N = 126.

 S n1N2,

TAYL16-001-035.I 1/22/03 12:46 PM Page 24

Page 24: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 24/35

24 Chapter 16 • The Structure of Atomic Nuclei

where is the angle between S and L; that is,

(16.34)

For this reason, the energy is called the spin-orbit energy.Since the spin-orbit energy of each nucleon is proportional to each

energy level of Fig. 16.16 (with ) is split into two levels according to thetwo possible orientations of S relative to L. (Recall that S has just two possibleorientations relative to any direction.) A spin-orbit splitting of this kind is fa-miliar in atomic physics,where it produces the so-called fine structure of atom-ic spectra. As described in Section 9.7, an orbiting electron sees a magnetic

field since the electron has a magnetic moment this leads to amagnetic energy However, the nuclear spin-orbit splitting isnot a magnetic effect. Rather, it is a new property of the nuclear force. The nu-clear splitting is much larger than the atomic effect and is in the opposite di-rection: For an electron in an atom, the state withS and L parallel has slightlyhigher energy; for a nucleon in a nucleus, the corresponding state has muchlower energy. The first clear evidence for this nuclear spin-orbit energy camefrom studies of nuclear shell structure, but scattering experiments with sepa-rate spin-up and spin-down beams of nucleons have since confirmed its exis-tence and properties.

The effect of the spin-orbit energy on the levels of Fig. 16.16 is shownin Fig. 16.17, which shows the resulting levels (or subshells) with their de-generacies and their grouping into energy shells.As one would expect, theeffect of the spin-orbit energy increases with increasing l : The level, likeall  s levels, is not split at all; the level, like all p levels (and also d lev-els), is split, but the resulting two levels are still in the same shell; for the

level and all levels with the splitting is so large that the resultingtwo levels belong to different shells. It can be seen that as a result of thesesplittings, the closed-shell numbers agree perfectly with the observedmagic numbers.

To understand the labeling and degeneracies of the levels in Fig. 16.17,we must discuss briefly the total angular momentum J of a nucleon,

The magnitude of L is with 1, and that of  S is

with s always equal to It is found that the magnitude of Jis given by a similar formula,

(16.35)

If its orbital angular momentum is zero the nucleon’s total angular

momentum is just the spin, and hence If it is found that there aretwo possible values of  j , depending on the orientation of S relative to L: If S isparallel to L, then

(16.36)

but if S is antiparallel to L, then

(16.37) j = l -12 (S and L antiparallel)

 j = l +12 (S and L parallel)

l Z 0, j =12 .1l = 02,

J = 4  j

1 j + 1

2 U

12 .S = 4 s1s + 12 U,

2, Á ,l = 0,L = 4 l1l + 12 UJ = L + S

l Ú 3,1 f

1p

1s

-M # B r S # L.M

rS,B

rL;

l Z 0S # L,

¢U

¢U r S # L

uMaria Goeppert-

Mayer (1906–1972,German–US)

After getting her PhD under MaxBorn at Gottingen in 1933, Goep-pert-Mayer moved to the UnitedStates. During most of her career,nepotism rules at universitieswhere her husband worked pre-vented Mayer from holding officialpositions.In 1949,while at the Uni-versity of Chicago, she proposedthe nuclear shell model, for which

she shared the 1963 Nobel Prize inphysics with Jensen. AlthoughGoeppert-Mayer was an activemember of the Chicago faculty — teaching courses, advising graduatestudents,and doing groundbreakingresearch — she had no official ap-pointment and no salary. In 1960she and her husband moved to theUniversity of California, San Diego,where she was appointed profes-sor of physics, her first full-timepaid position. The picture showsher talking with Enrico Fermi at asummer school at the University of Michigan in the 1930s.

 Johannes Jensen(1907–1973,German)

  Jensen was a professor first atHamburg and then at Heidelberg.In 1949 he and Goeppert-Mayerindependently proposed the nu-clear shell model. Together they

wrote a book on the model in1955, and they shared the NobelPrize in physics in 1963.

TAYL16-001-035.I 1/22/03 12:46 PM Page 25

Page 25: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 25/35

Section 16.8 • The Shell Model 25

3 s1/22d3/2

1h11/2

1 g7/2

1 g9/2

1 f 5/2

1 f 7/2

1d3/2

1d5/2

1 p1/2

1 s1/2

2d5/2

2 p1/2

2 p3/2

2 s1/2

1 p3/2

2 (82)4 (80)

12 (76)6 (64)

8 (58)

10 (50)2 (40)

6 (38)4 (32)

8 (28)

4 (20)2 (16)

6 (14)

2 (8)

4 (6)

2 (2)

82

50

28

20

8

2

1h

3 s

2d

1 g

2 p

1 f 

2 s1d

1 p

1 s

Z or N 

(closed shells)

FIGURE 16.17

Filling order of the levels through Z or for a single nucleon in anucleus, including the spin-orbit

energy proposed by Goeppert-Mayer and Jensen. The levels onthe left are the corresponding levelsin the absence of the spin-orbitenergy. The numbers to the rightof each level are the level’sdegeneracy and (in parentheses)the running total of protons orneutrons needed to fill throughthat level. On the far right are theclosed-shell numbers, which agreeperfectly with the observed magic

numbers (16.33). The ordering of certain nearby levels is ambiguous(just as it is in atoms) and can bedifferent for protons and neutrons.Beyond 82, where the proton wellis strongly distorted by Coulombrepulsion, the level orderings forprotons and neutrons aresignificantly different.

N = 82

In Fig. 16.17 the value of  j in each level is shown by a subscript following thecode letter for l . Thus the original level, shown on the left, is really twolevels with

These are shown on the right of Fig. 16.17 as and withsomewhat lower in energy.

For each magnitude of the total angular momentum, there are several

possible orientations of J, all with the same energy. These correspond to thedifferent possible values of which are given by

where can have any of the values

Thus, each of the levels of Fig. 16.17 has a degeneracy of asindicated to the right of each level. 12 j+

12,

m j = j,  j - 1, Á , - j

2 j + 1m j

Jz = m j   U

Jz ,

1p3>21p1>2 ,1p3>2  =

32  or  1

2

  j = l ;12 = 1 ;

12

1p

TAYL16-001-035.I 1/22/03 12:46 PM Page 26

Page 26: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 26/35

26 Chapter 16 • The Structure of Atomic Nuclei

Example 16.6

Find the degeneracies of the and levels. Show that the sum of these is the same as the total degeneracy of the original level on the left of Fig. 16.17.

The required degeneracies are

(16.38)

and

(16.39)

In the absence of the energy, there would have been a single levelwhose degeneracy would have been (that is, two orientationsof the spin times three of L). This is equal to the sum of (16.38) and (16.39).

It can be shown quite generally (Problem 16.55) that for any givenvalues of n and l , the sum of the degeneracies of the two corresponding levels(with ) is equal to the total degeneracy of the single level, nl , thatwould have existed in the absence of the energy. Thus, the effect of the

term is simply to redistribute the degeneracies so as to produce theobserved magic numbers.

 Angular Momenta of Nuclei

In Section 10.7 we saw that one can predict the angular momentum of manyatoms when one knows the order in which the electron energy levels are fi lled.We can now apply similar reasoning to nuclei. We first note that any filled levelof given j contributes zero total angular momentum. (This is because for everynucleon with there is a second with ) This implies thatany nucleus in which all of the occupied levels are closed (no partially filledlevels) should have zero total angular momentum. That is,

in this case. For example, in the protons and neutrons both completely filltheir and levels. Therefore, should have, and does have, zerototal angular momentum, Referring to Fig. 16.17,we can predict in thesame way that the closed-subshell nuclei and should all have

and in every case this is found to be correct.Let us consider next a nucleus with a single nucleon outside otherwise

filled subshells. For example, in the eight protons and eight of the nine

neutrons fill the levels and but the ninth neutron must occu-py the level all by itself Since the closed levels all have zero angular mo-mentum, the angular momentum of the whole nucleus is just that of the final,odd neutron. Thus we predict that should have and this provesto be correct.

Throughout this section we have been discussing the ground states of nuclei, but we can also predict the properties of some excited states. For ex-ample, we just saw that the ground state of has its last neutron in thelevel. In Fig. 16.17 we see that just above the level is the Hence wewould expect that the first excited state of would be obtained by lifting

the last neutron to the level, and should have This is observedto be the case.  jtot=

1>2.2s1>28O9

2s1>2 .1d5>2 1d5>28O9

 jtot = 5>2,8O9

1d5>2 1p1>2 ,1p3>2 ,1s1>2 ,

8O9

 jtot = 0,20Ca2814Si16 ,8O8 ,

 jtot = 0.6C61p3>21s1>2 6C6

Jtot = 4  jtot1 jtot + 12 U = 0

Jz = -m j   U.Jz = m j   U,

S # LS # L

 j = l ;12

212l + 12 = 61pS # L

1p1

>2: degeneracy = 2 j + 1 = A2 *

12 B + 1 = 2

1p3>2: degeneracy = 2 j + 1 = A2 *32 B + 1 = 4

1p

1p1>21p3>2

TAYL16-001-035.I 1/22/03 12:46 PM Page 27

Page 27: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 27/35

Section 16.8 • The Shell Model 27 

*The difficulty is to predict which relative orientation of the angular momenta of thelast proton and last neutron gives the lowest energy.

Pairing Energy 

Let us next consider a nucleus with two identical nucleons (two protons or twoneutrons) outside otherwise filled subshells. We could, for example, consider

with two neutrons in the level. The filled levels all have zero angularmomentum, but since each of the last two neutrons has there are sev-eral possible values of depending on the relative orientation of the twoseparate angular momenta. However, it turns out that the state in which thetwo angular momenta are antiparallel has significantly lower energy.Therefore, the ground state of has

It is found quite generally that any pair of identical nucleons in the samelevel has lowest energy when their total angular momentum is zero.The extra

binding energy in the state is called the pairing energy and is typicallya few MeV. This energy is the origin of the term (16.28), in the semiem-pirical binding-energy formula.

Whenever a level is partially filled, the angular momenta of its nucleonscan combine to give various different total angular momenta. However, thepairing effect implies that in the state of lowest energy the nucleons will bearranged, as far as possible, in pairs with zero angular momentum.If the num-ber of nucleons in the level is even, this means that the state of lowest energyhas if the number of nucleons is odd, the state of lowest energy has allbut one nucleon arranged in pairs and the total angular momentum is

 just that of the odd,unpaired nucleon.These properties lead us to the following useful rules:

The ground state of any nucleus in which both Z and N are even(an “even–even”nucleus) has zero total angular momentum. (16.40)

and

If a nucleus has A odd (that is, Z is even and N odd or vice versa),the total angular momentum of its ground state is just that of thelast unpaired nucleon. (16.41)

Nuclei in which both Z and N are odd are harder to analyze,* but there areonly four such stable nuclei in any case.

The rule (16.40) is true for all known even–even nuclei. The rule (16.41),together with the level ordering shown in Fig. 16.17, correctly predicts the an-gular momentum of most odd- A nuclei; the only exceptions involve nonspher-ical nuclei, for which the ordering of Fig. 16.17 is not always applicable. (SeeProblem 16.53.) The use of both rules is illustrated in the following example.

Example 16.7

What are the quantum numbers for the total angular momenta of thenuclei and

Since has and it is an even–even nucleus and hencehas The nucleus has and Since Z is even, theprotons have zero total angular momentum. The total angular momentum of the neutrons is just that of the twenty-third neutron, which is in the level(Fig.16.17). Hence the nucleus should,and does,have jtot = 7>2.43Ca

1 f7>2N = 23.Z = 2043Ca jtot = 0.

N = 12,Z = 1022Ne

43Ca?22Ne jtot ,

 j = 0 jtot = 0;

Bpair , jtot = 0

 jtot = 0.8O10

1 jtot = 02 jtot , j = 5>2

1d5>28O10

TAYL16-001-035.I 1/22/03 12:46 PM Page 28

Page 28: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 28/35

28 Chapter 16 • The Structure of Atomic Nuclei

Ion source

V 0

B out of page

Photographicfilm

FIGURE 16.18

A mass spectrometer. Positive ionsare produced in the ion source by

an electric discharge and areaccelerated by a potential differenceThey then enter a magnetic

field and follow a curved path. Ionswith different masses follow pathsof different curvature and strike thephotographic film at differentplaces. The entire apparatusoperates in a vacuum.

V0 .

*We are treating the motion as nonrelativistic, which is usually an excellent approxi-mation (see Problem 16.60).

16.9 Mass Spectrometers★

★As we describe in this section, the mass spectrometer can measure atomic and nu-clear masses with extraordinary precision and has become a powerful tool of chemicalanalysis.Nevertheless, we will not be using the ideas of this section again,and you couldomit it if pressed for time.

In this section we describe mass spectrometers, the devices used to measurenuclear masses. As we saw earlier, measurements of nuclear masses con-tributed to our understanding of nuclear structure in two important ways.First, early measurements, which could distinguish masses differing by a fewpercent, revealed the existence of isotopes. Second, more precise measure-

ments were able to determine the small relativistic shifts (about 1%) in nu-clear masses due to the binding energy. The variation of binding energy withmass number, A, revealed several properties of the nuclear force, as we dis-cussed in Section 16.7. Today, mass spectrometers can measure masses with anaccuracy of order 1 part in which allows binding energies of typical nucleito be calculated within 1 keV or so. Most recently, the mass spectrometer hasdeveloped into a powerful tool of chemical analysis, as we describe shortly.

In practice, it is generally the mass of an ionized atom (or molecule) thatis measured in a mass spectrometer. The substance to be analyzed is intro-duced into an electric discharge, which causes some atoms to lose one electron

and become positive ions of charge e. In the arrangement shown in Fig. 16.18,the ions are next accelerated by a potential difference Finally, they arebent into a circular path, of radius R, by a magnetic field B. Measurement of 

R, and B allows one to calculate the ions’ mass, m, as we will see directly.Since m is the mass of the ion, the mass of the corresponding neutral atom iscalculated by adding one electron mass, that of the bare nucleus is foundby subtracting

To calculate the mass, we note that when the ions are accelerated by thepotential difference they gain a kinetic energy,

Their momentum p is then*

(16.42)p = 2 2mK = 2 2mV0 e

K = V0 e

V0 ,

1Z - 12me .me ;

V0 ,

V0 .

108,

TAYL16-001-035.I 1/22/03 12:46 PM Page 29

Page 29: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 29/35

200 202 204 206 208 210

Mass (u)

   N  u  m   b  e  r  o   f   i  o

  n  s

Lead, Pb

FIGURE 16.19

Results of analyzing natural lead in amass spectrometer. The vertical

axis shows the number of ionsdetected, and the horizontal axis iscalibrated in atomic mass units. Thedifferent heights of the four peaksreflect the different natural abun-dances of the four isotopes of lead.

Section 16.9 • Mass Spectrometers 29

Hydrogen chloride, HCI36

38

0 10 20 30 40

Mass (u)

   N  u  m   b  e  r  o   f

   i  o  n  s

FIGURE 16.20

Mass spectrum obtained from HCl.The peaks labeled 36 and 38 wereproduced by the molecules

and containing the twonaturally occurring isotopesand The smaller peaks at 35and 37 were produced by atoms of 

and that were detached inthe ion source.

37Cl35Cl

37Cl.

35Cl1H37Cl,

1H35Cl

where m is the mass of the ion. When the ions enter the magnetic fieldB, they

follow a circular path with radius given by (2.48) as

(16.43)

Substituting (16.42) and solving for m, we find that

(16.44)

If B and are fixed, different masses will have different radii, and measure-ment of R determines m from (16.44). In most early mass spectrometers,R wasmeasured by locating the spots where the ions struck a photographic film, as inFig. 16.18. In modern spectrometers a computer monitors the acceleratingvoltage the field B, and the location of an electronic ion detector; after ini-tial calibration with ions of known mass, the computer can display the resultsdirectly in atomic mass units.

When an element is analyzed in a mass spectrometer, its various iso-topes are separated because of their different masses. Figure 16.19 shows thespectrum of masses obtained when natural lead is analyzed in a modern spec-trometer, which displays its results directly as a graph of number of particlesagainst mass.

Today, mass spectrometers are widely used in chemical analysis to helpdetermine the composition of molecules.This is illustrated in Fig. 16.20, whichshows the spectrum obtained when the simple diatomic compound HCl is in-troduced into the ion source of a mass spectrometer. Since chlorine has twostable isotopes, and there are two large peaks with masses close to36 u and 38 u In addition, smaller peaks can be seen at the

positions one would expect for single atoms of and These occur be-cause the electric discharge in the ion source dissociates some of the moleculesinto H and Cl atoms. If the HCl analyzed in the spectrometer were an un-known substance, its composition could be deduced from the following obser-vations: (1) the total molecular mass has two values, 36 and 38; (2) themolecule has fragments with masses of 35 and 37; and (3) the ratio of the peakheights at 36 and 38 (and those at 35 and 37) is the same as that of naturallyoccurring chlorine.

The mass spectrum obtained for a more complex molecule is shown inFig. 16.21, together with a diagram of the molecule’s structure. The moleculeshown, methionine, has mass 149 u and produces the peak labeled 149 in thespectrum. As was the case with HCl, the molecule can be broken into smallerfragments in the ion source, and it is these fragments that produce the severalpeaks at lower masses. For example, the peaks labeled 61 and 88 arise becausethe molecule breaks easily at the point indicated by the arrow in Fig. 16.21(b),producing fragments of masses 61 u and 88 u. The masses of the various seg-ments of this complex organic molecule provide valuable clues to its structure.

Close inspection of the major peaks in Fig. 16.21 shows that each is ac-companied by several smaller peaks.These smaller peaks differ from the dom-inant peaks by one or two mass units and come about in two ways: (1) amolecule or molecular fragment can gain or lose one or two hydrogen atomsin the ion source; and (2) about 1% of the carbon atoms in any natural sampleis the isotope (as opposed to the dominant ). The details of these sub-sidiary peaks are additional clues to the composition of the molecule.

12C13C

37Cl.35Cl

(1H37Cl).(1H35Cl)

37Cl,35Cl

V0 ,

V0

m =eB2

 R2

2V0

R =p

eB

TAYL16-001-035.I 1/22/03 12:46 PM Page 30

Page 30: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 30/35

30 Chapter 16 • The Structure of Atomic Nuclei

50 100 150Mass (u)

(a)

   N

  u  m   b  e  r  o   f   i  o  n  s

45

74

75

61

88

104 132

149

Methionine (mass 149)

CH3 CH2 CH2 CH

NH2

COOHS

(61) (88)

(b)

FIGURE 16.21

(a) Mass spectrum obtained from methionine, with mass 149 u. The lower

mass peaks are produced by fragments of the molecule that were detached in the ionsource. (b) The molecule’s structure. The arrow shows a point where the molecule caneasily be broken in the ion source; the resulting fragments have masses of 61 u and 88 uand produce the peaks labeled 61 and 88 in the spectrum.

C5H11O2NS,

Current research in molecular structure has extended mass spectroscopyto molecules with masses of thousands of atomic mass units. Mass spectrome-ters with this capability are commercially available and can quickly and auto-matically produce spectra like those in Figs. 16.20 and 16.21 for almost anysubstance that needs to be analyzed.

CHECKLIST FOR CHAPTER 16

CONCEPT DETAILS

Isotopes Different nuclei (or atoms) with the same atomic number Z 

Isobars Different nuclei (or atoms) with the same mass number A

Isotones Different nuclei (or atoms) with the same neutron number N 

Nuclear radius (16.3)

Nuclear (or strong) force Range and strength; charge independence (Sec. 16.3)

IPA for nucleons Average potential energy seen by an individual nucleon in a nucleus(Sec.16.5)

Symmetry effect Pauli principle tendency for nuclei to have(Sec.16.6)

Semiempirical binding-energy formula Eq. ( 16.30)

Binding energy per nucleon, Measure of nuclear stability, Fig. 16.14

Shell model★ Magic numbers, spin-orbit splitting, total angular momentum(Sec.16.8)

Mass spectrometers★ Measure atomic masses by observing curvature in magneticfield (Sec. 16.9).

B>AZ L NQ

R = R0  A1>3

TAYL16-001-035.I 1/22/03 12:46 PM Page 31

Page 31: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 31/35

Problems for Chapter 16 31

PROBLEMS FOR CHAPTER 16

SECTION 16.2 (Nuclear Properties)

16.1 • Use Appendix D to find the numbers of protonsand neutrons in the following nuclei:

16.2 • Use Appendix D to find the numbers of protonsand neutrons in the following nuclei:

16.3 • Use Appendix D to find the chemical symbolsand the natural percent abundances of the isotopeswith (a) (b)(c) (d)

16.4 • Use Appendix D to find the chemical symbolsand the half-lives of the four radioactive nucleiwith (a) (b)(c) (d)

16.5 • (a) Use Appendix D to make a list of all the stableisotopes, isobars, and isotones of  (b) Do thesame for and (c)

16.6 • Make lists of all the stable isotopes, isobars, and

isotones of each of the following:

16.7 • Find the mass of the atom in Appendix D. Whatfraction of its mass is contained in its atomic electrons?

16.8 • The mass of a lead atom’s nucleus isand its half-density radius R is 6.5 fm.Estimate its densi-ty by assuming its volume to be Compare

your result to the density of solid lead

16.9 • Atomic radii range from a little less than 0.1 nm to0.3 nm, while nuclear radii range from about 2 fm toabout 7 fm. Taking the representative values

and find the ratio of the volume of an atom to that of a nucleus. What isthe ratio of the corresponding average densities?

16.10 • Use the formula (16.3), withto find the approximate radii of the most abundantisotopes of carbon,iron, and lead.(Mass numbers andabundances are in Appendix D.)

16.11 • The nuclear radius R is well approximated by theformula (16.3), whereFind the approximate radii of the most abundant iso-

topes of each of the following elements: Al, Cu, andU. (Mass numbers and abundances are given inAppendix D.)

16.12 • The first excitation energy of the nucleus is0.48 MeV. What is the wavelength of a photon emit-ted when a nucleus drops from its first excitedstate to the ground state? Find the wavelength for thecorresponding transition in the Li atom (excitationenergy, 1.8 eV).

16.13 • The first excitation energy of the helium nucleus,is about 20 MeV. What is the wavelength of a

photon that can just excite the helium nucleus fromits ground state? Compare this with the wavelength

4He,

7Li

7Li

R0 = 1.07 fm.R = R0 A1>3,

R0 = 1.07 fm,R = R0 A1>3,Rnuc L 4.5 fm,Ratom L 0.2 nm

111 g>cm32.

14>32pR3.

3.5 * 10-25 kg,

12C

208Pb.36S,56Fe,

88Sr.27Al

70Ge.

A = 244.Z = 94,A = 40;Z = 19,A = 14;Z = 6,A = 3;Z = 1,

A = 206.Z = 82,A = 56;Z = 26,A = 13;Z = 6,A = 3;Z = 2,

232Th.209Bi,64Zn,29P,

9Be,4He,2H,

206Pb.63Cu,40Ar,20Ne,

7Li,4He,1H,

of a photon that can just excite a helium atom (excita-tion energy 19.8 eV).

16.14 •• (a) Treating the nucleus as a uniform sphere of ra-dius calculate the density (in ) of 

the nucleus and compare it to that of solid iron(b) Calculate the electric charge density

of this nucleus in Comparethis charge density to that which can be uniformlydistributed in an insulating sphere of 1 m radius inair without sparking. The maximum possible electric

field strength at the surface of the sphere withoutsparking is about

16.15 •• The density of mass inside a nucleus is shown inFig. 16.2. There is no simple theory that predicts theexact shape shown, but it is found that the shape canbe approximated by the following mathematicalform, known as the Fermi function:

(16.45)

where R, and t are positive constants, with

(a) Prove that at (b) Show that themaximum value of occurs at and is very closeto given that (c) Sketch as a function of r .(d) Prove that as r  increases, the density falls from90% to 10% of in a distance (Thus t characterizes the thickness of the surface region.)

SECTION 16.3 (The Nuclear Force)

16.16 • If there were no nuclear attraction, what would bethe acceleration of either of two protons released at aseparation of 4 fm? Compare your answer with g, theacceleration of gravity. (As large as your answer is, itis still small compared to nuclear accelerations. SeeProblem 16.20.)

16.17 • Use Eq. (16.7) to estimate the minimum kineticenergy of a nucleon in a nucleus of diameter 8 fm.

16.18 •• Use Eq. (16.7), taking the size a to be twice the ra-dius given by (16.3), to find the minimum kinetic ener-gy of a nucleon inside each of the nuclei and

16.19 •• (a) The minimum kinetic energy of a nucleonin a rigid cubical box of side a is given by (16.7). By

setting the nuclear diameter, and

(with ), derive a formula for as afunction of the mass number A. (b) Find for

and

16.20 •• A representative value for the kinetic energy of anucleon in a nucleus is 20 MeV.To illustrate the greatstrength of the nuclear force, do the following classi-cal calculation: Suppose that a nucleon in a nucleusoscillates in simple harmonic motion with amplitudeabout 4 fm and peak kinetic energy 20 MeV. (a) Findits maximum acceleration. (b) By how many ordersof magnitude does this exceed the acceleration of gravity, g? (c) Find the force required to produce thisacceleration.

238U.27Al,

9Be,KminKminR0 = 1.07 fm

R = R0 A1

>3a L 2R,

Kmin

244Pu.12C

¢r = 4.4t.r 0

r t V R.r 0 ,r = 0r 

r = R.r  = r 0>2

t V R.r 0 ,

r 1r2 =r 0

1 + e1r- R2>t

3 * 106 V>m.

coulombs>m3.1Z = 262(7800 kg>m3).

56Fe

kg>m3(1.07 fm)A1>3,

TAYL16-001-035.I 1/22/03 12:46 PM Page 32

Page 32: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 32/35

32 Chapter 16 • The Structure of Atomic Nuclei

16.21 •• The charge independence of nuclear forces impliesthat in the absence of electrostatic forces, the energy

levels of and would be the same.The main ef-fect of the electrostatic forces is simply to raise all the

levels of compared to those of Approximat-ing both nuclei as uniform spheres of chargeand the same radius R, estimate the difference in thebinding energies for any level of and the corre-sponding level of (The electrostatic energy of auniform charged sphere is — see Problem16.45. The observed radius R is about 2.5 fm.) Com-pare your rough estimate with the observed differ-ence of about 1.7 MeV. (Note that the true chargedistribution is not a uniform sphere with a well-defined radius R, but is spread out somewhat beyondR. Therefore, the observed electrostatic energieswould be expected to be somewhat smaller than yourestimates.)

16.22 •• Do the same calculations as in Problem 16.21, butfor the nuclei and The observed differenceis 4.8 MeV; use the formula to get thenuclear radius.

SECTION 16.4 (Electrons Versus Neutrons asNuclear Constituents)

16.23 • One argument against the proton–electron modelof the nucleus concerns the total spins of nuclei.The

proton, electron, and neutron all have spin and the

total spin of any number of spin-half particles takes

the familiar form If there is just one

particle, then of course, With two particles, the

spins can be parallel or antiparallel, giving or 0.It can be shown that the general rule is this: For anodd number of particles, the total spin has some half-

odd-integer value for For an even

number of particles, the total spin has s equal to someinteger In light of this rule, consider

the nucleus (a) Assuming that is made of 7protons and 7 neutrons, predict the character of itstotal spin (integer or half-odd integer). (b) Repeat forthe proton–electron model of  (c) The observedtotal spin of is integer; which model does thissupport?

16.24 • Repeat Problem 16.23 for the deuteron whoseobserved total spin is given by

16.25 •• Supposing that nuclei could contain electrons,esti-mate the potential energy of one such electron at the

center of a nucleus of Compare your answerwith the minimum kinetic energy (16.10) of an elec-tron in a nucleus. [Hint: The required potential energyis where is given by Eq. (16.49) below.The radius R is given by (16.3),and the total charge act-ing on the electron is ]

16.26 •• (a) Following the suggestions in Problem 16.25,find the potential energy that an electron would haveat the center of a nucleus. (b) Repeat for

(c) Compare your answers with the minimum kineticenergy, which is of order 100 MeV.

  82208Pb.6

12C

Q = 1Z + 12e.

V1r2-eV102,

27Al.

s = 1.

2H,

14N

14N.

14N14N.1s = 0, 1, 2, Á 2.

5

2, Á

B.3

2,s

As =

1

2,

s = 1

s =12 .

4 s1s + 12 U.

12 ,

R = R0 A1>323Mg.23Na

3kQ2>5R37Li.

47Be

Q = Ze

7Li.7Be

7Be7Li

16.27 •• Because of the proton’s magnetic moment, everyenergy level of the hydrogen atom, as calculated pre-

viously, actually consists of two levels, very close to-gether. The proton’s magnetic moment creates amagnetic field, which means that the energy of the Hatom is slightly different, depending on the relativeorientations of the electron and proton moments —an effect known as hyperfine splitting. You can esti-mate the magnitude of the hyperfine splitting as fol-lows: (a) The magnetic field at a distance r  from amagnetic moment is

(16.46)

where is the permeability of space. [For simplicity, we have given the field at apoint on the axis of At points off the axis the fieldis somewhat different, but is close enough to (16.46)for the purposes of this estimate.] Given that themagnetic moment of the proton is roughly equal tothe nuclear magneton defined in (16.12), estimatethe magnetic field at a distance from a proton.(b) Taking your answer in part (a) as an estimate forthe B field “seen” by an electron in the state of ahydrogen atom, show that the atom’s energy differs

by roughly for the cases that the proton andelectron spins are parallel or antiparallel. [The energyof a magnetic moment in a B field is given by (9.10).](c) This means that the state of hydrogen is reallytwo very closely spaced levels. Compare your roughestimate with the actual separation of these levelsgiven that the photon emitted when a hydrogen atommakes a transition between them has(This 21-cm radiation is used by astronomers to iden-tify hydrogen atoms in interstellar space.)

16.28 ••• It is often a useful approximation to treat a nucle-

us as a uniformly charged sphere of radius R andcharge Q. In this problem you will find the electrosta-tic potential inside such a sphere, centered atthe origin. (a) Write down the electric field atany point a distance r  from the origin with(Remember that this is the same as the field of apoint charge Q at the origin.) (b) Use the definition

(16.47)

to find the potential difference between points at

and (both greater than R). It is usual to defineso that approaches 0 as By choosingand show that

(16.48)

(c) Now find for (Remember that tells us

Gauss’s law this is where is the totalcharge inside the radius r .) Check that your answersfor parts (c) and (a) agree when (d) UseEq. (16.47) to find for any two points

inside the sphere.Now choose and anduse the value of from part (b) to prove thatV1R2 r1=

R,r2=

r

V

1r2

2- V

1r1

2r = R.

Q¿kQ¿>r2,

r … R.E1r2V1r2 =kQ

r  1r Ú R2

r2 = r,r1 = q

r: q .V1r2 V1r2r2

r1

¢V = V1r22 - V1r12 = -Lr2

r1

 E1r2 dr 

r 7 R.E1r2V1r2

l = 21 cm.

1s

10-

6 eV

1s

aB

mN

M.

m0 = 4p * 10-7 N>A2

B =m0M

2pr3

M

TAYL16-001-035.I 1/22/03 12:46 PM Page 33

Page 33: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 33/35

Problems for Chapter 16 33

(16.49)

In Problems 16.25, 16.26, 16.29, and 16.30 this resultis needed to find the potential energy of a chargeinside a nucleus.

SECTION 16.5 (The IPA Potential Energyfor Nucleons)

16.29 • (a) Use the result (16.49) of Problem 16.28 to estimatethe electrostatic potential energy of a proton at the cen-ter of a nucleus. (b) Repeat for a nucleus of 

16.30 •• The electrostatic potential energy of a protonin a nucleus can be approximated by assuming that itmoves in a uniform sphere of charge andradius R. (a) Use the results (16.48) and (16.49) of Problem 16.28 to write down and sketch(b) Prove that (c) Setting

estimate for a proton in

and for a proton in

SECTION 16.6 (The Pauli Principle and theSymmetry Effect)

16.31 • Enlarge Fig. 16.11 to include the nuclei andBy how much (in terms of the quantity

shown in Fig. 16.11) does the energy of each isobar

with exceed the energy of 

SECTION 16.7 (The Semiempirical Binding-Energy Formula)

16.32• Use the data of Appendix D to find the bindingenergy of 

16.33 • (a) Find the mass (in u) of the atom inAppendix D. (b) Find the mass of the nucleus toseven figures (but ignore corrections due to the atom-ic electrons’ binding energy). (c) Do any of theseseven figures change if you take into account theelectrons’ binding energy (about 80 eV total)?

16.34 • Use the data of Appendix D to find the bindingenergies of and

16.35• Use the data of Appendix D to deduce the energyrequired to remove one neutron from to pro-duce and a neutron at rest, well separated from

the nucleus.16.36 •• (a) Use conservation of energy to write an equa-

tion relating the mass of a nucleus and its neu-tron separation energy to the masses of thenucleus and the neutron. (b) Use the data

in Appendix D to find for and

16.37 •• (a) The proton separation energy (energy to

remove one proton) for is 7.1 MeV. Given that

the total binding energy of is 1559.4 MeV, findthe total binding energy of  (b) Compare youranswer with the answer obtained directly from themass of given in Appendix D.198Hg

198Hg.

197Au

198Hg

Sp

200Hg.120Sn,13C,Sn

1Z, N - 12 Sn1Z, N2 1Z, N213

C

13C

14C

120Sn.40Ca,4He,

4He

4

He

12C.

12C?Z Z N

¢E12O.

12Be

238U.4He

UCoul102R = 11.07 fm2A1>3,UCoul102 = 1.5 UCoul1R2.UCoul1r2.

1Z - 12eUCoul

208Pb.12C

V

1r

2=

kQ

2R

 

¢3 -

r2

R2

≤ 

1r … R

216.38 •• (a) To estimate the importance of the surface term

in the binding-energy formula, consider a model of 

that consists of 27 identical cubical nucleonspacked into a cube. Each small cube hassix faces. What is the fraction of all these faces thatare exposed on the nuclear surface? (b) Repeat for

considered as a cube. (c) Commenton the difference in your answers.

16.39 •• The Coulomb energy of the charges in a nucleus ispositive and hence reduces the binding energy B. If we approximate the nucleus as a uniform sphere of charge then according to Eq. (16.50) of Prob-lem 16.45, the Coulomb correction to B should be

Putting show that this agrees with theform (16.26) of the Coulomb correction given inSection 16.7. Use your answer here to calculatein MeV, and compare with the empirical value

(You should not expect preciseagreement since the nucleus is certainly not a perfect-ly uniform sphere, but you should get the right orderof magnitude.)

16.40 •• (a) Use the binding-energy formula (16.30) to pre-dict the binding energy of  (b) Compare youranswer with the actual binding energy found from themasses listed in Appendix D.

16.41 •• Use the semiempirical binding-energy formula to

find for Make a table showingall five terms for each nucleus, and comment on theirrelative importance. Do your final answers follow thegeneral trend of Fig.16.14?

16.42 •• Use the binding-energy formula (16.30) to find the

total binding energies of andWhich terms are different in the four cases,and why?

16.43 •• If captures a neutron to form in its groundstate, the energy released is (a)Prove this statement. (b) Use the binding-energy for-mula (16.30) to estimate the energy released, andcompare with the observed value of 6.5 MeV. (Note:We have assumed here that is formed in itsground state, and the 6.5 MeV is carried away, by aphoton, for example. An important alternative is that

can be formed in an excited state, 6.5 MeVabove the ground state. This excitation energy canlead to oscillations that cause the nucleus to fission.)

16.44 •• Compare the surface area of a sphere to that of acube of the same volume.How much energy, in MeV,would be needed to distort the normally spherical

nucleus into a cubical shape if the volume andCoulomb energies changed by negligible amounts?

16.45 ••• In this problem you will calculate the electrostat-ic energy of a uniform sphere of charge Q with radiusR. The electric potential at any radius isgiven by (16.49) in Problem 16.28, and the potential

energy of a charge q at radius r  is (a) Writedown the total charge contained between radius r andqV1r2.

r 6 RV1r260Ni

236U

236U

B1236U2 - B1235U2.

236U235U

25

56

Mn.26

56

Fe27

56

Co,28

56

Ni,

259No.60Ni,20Ne,B>A40Ca.

aCoul = 0.711 MeV.

aCoul

R = R0 A1>3,BCoul = -UCoul = -  

3

5 k1Ze22

R

Q = Ze,

5 * 5 * 5125Te

3 * 3 * 327Al

TAYL16-001-035.I 1/22/03 12:46 PM Page 34

Page 34: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 34/35

34 Chapter 16 • The Structure of Atomic Nuclei

and hence find the potential energy of thatcharge. (b) If you integrate your answer to (a) from

to you will get twice the total potentialenergy of the whole sphere since you will have count-ed the energy of any two charge elements twice.Showthat the total Coulomb energy of the sphere is

(16.50)

16.46 ••• (a) Substitute the binding-energy formula (16.30)into the relation

to obtain the semiempirical mass formula. [To simpli-fy matters, take A to be odd, so that the pairing termin (16.30) is zero.] (b) Among any set of isobars, thenucleus with lowest mass is the most stable. To identi-fy this nucleus, first write your mass formula in termsof the variables A and Z (that is, replace N , whereverit appears, by ). With A fixed, differentiate mwith respect to Z . The minimum mass is determinedby the condition Show that this leads to arelation of the form

(16.51)

where and are related to the coefficients (16.31) asfollows:

(c) For 115, 185, find the values of Z that givethe most stable nuclei and compare with the observed

values from Appendix D.

SECTION 16.8 (The Shell Model★)

16.47 • Which of the following nuclei has a closed subshell for

Z ? Which for N ?

16.48 • Use the levels shown in Fig. 16.17 to predict for

the ground states of the nine nuclei

Compare with the data in Appendix D.

16.49 • Use the levels shown in Fig. 16.17 to predict thetotal angular momenta of the following nuclei:

and

16.50 • Use the shell model to predict for the groundstates of and

16.51 • The ground state of has and its first

excited state has Explain these observations in

terms of the shell model.

16.52 • (a) Use the levels shown in Fig. 16.17 to draw an oc-cupancy diagram like Fig. 16.10(b) for the groundstate of  (b) What is (c) What would you pre-dict for the first excited state? (Draw an occupancydiagram, and give ) jtot .

 jtot ?7Be.

 j =12.

 jtot =32

7Li

2141Sc20 .20

40Ca20 ,1939K20 ,

 jtot

71Ga.59Co,37Cl,33S,

29Si,

Á 48Ca.

40Ca, 41Ca,

 jtot

48Ca.42Ca,28Si,27Al,17O,11B,10B,

A = 37,

a =

1mn - mp2c2

4asym

  and  b =aCoul

4asym

ba

Z =A

1 + a

1 + bA2>30m

>0Z = 0.

A - Z

mnuc

= Zmp

+ Nmn

-B

c2

UCoul =3

5 kQ2

R

r = R,r = 0

r + dr , 16.53 • In applying the shell model, we usually use the levelordering of Fig. 16.17, which is based on an assumed

spherical potential well. For certain nonspherical nu-clei, this assumption is incorrect and the ordering of 

the levels is different. Examples are andcheck this by using the rule (16.41) and Fig. 16.17 topredict the angular momenta of these two nuclei andcomparing with the observed values in Appendix D.

16.54 •• (a) Use the energy levels of Fig. 16.17 to draw anoccupancy diagram like Fig.16.10(b) for (b) Whatis for the ground state? (c) The first excited state isformed by lifting one neutron from toDraw an occupancy diagram for this state, and

explain what you would predict for of this state.(Does your answer agree with the observed value,

) (d) Now predict for the ground and

first excited states of (In this case the excitationinvolves a proton.)

16.55 •• For each choice of n and l (with ), a nucleonhas two different possible energy levels with

Prove that the sum of the degeneracies of these two levels is equal to the total degeneracy thatthe level nl  would have had in the absence of any

splitting.

16.56 ••• The first magic number for both nuclei and atomsis 2. The second is 8 for nuclei and 10 for atoms. Thedifference occurs because what we call the andlevels in nuclei are widely separated, whereas the cor-responding levels in atoms (called and by atom-ic physicists) are very close together. To explain thisdifference, compare the shape of an atomic potentialenergy with that of the nuclear potential well.Noting that states with higher l have probability dis-tributions that are pushed out to larger radii, qualita-tively explain the difference in the positions of these

two levels.

16.57 ••• The spin-orbit energy (16.34) is proportional toTo evaluate note that since

You can solve this for and then put in the valuesfor and [For instance,

where ] Show that if then

and if then

Use these answers to prove that the spin-orbit split-

ting increases with increasing l , as indicated inFig. 16.17.

SECTION 16.9 (Mass Spectrometers★)

16.58 • Where does the methionine molecule in Fig. 16.21break to account for the two peaks at 74 u and 75 u inits mass spectrum?

16.59 • Repeat Problem 16.58 for the peaks at 45 u and 104 u.

16.60 •• (a) Find the radius R of the curved path for aonce-ionized nitrogen atom in a mass spectrometerwith and using (16.44).V0 = 10 kV,B = 0.05 T

S # L = -

1l + 1

2 U

2

>2. j = l -

12 ,l U2

>2,

S # L = j = l +12

 j = l ;12.

J2= j1 j + 12 U

2L2.S2,J2,S # L

J2= 1S + L2 # 1S + L2 = S2

+ L2+ 2S # L

J = S + L,S # L,S # L.

1r1>r22p2s

2s1p

S # L

 j = l ;12.

l Z 0

13N.

 jtot jtot =32

 ?

 jtot

1p1>2 .1p3>2 jtot

13C.

121Sb;19F

TAYL16-001-035.I 1/22/03 12:46 PM Page 35

Page 35: TAYL16-001-035.I

8/14/2019 TAYL16-001-035.I

http://slidepdf.com/reader/full/tayl16-001-035i 35/35

Problems for Chapter 16 35

(b) Equation (16.44) was derived nonrelativistically.What would R be if calculated relativistically? How

many significant figures must you include to see thedifference?

16.61 •• Conditions in the ion source of a mass spectrometercan be adjusted to allow formation of multiply ionizedatoms. (a) Rewrite Eq.(16.44) for the case of an n timesionized atom, and solve for R. (b) Taking the masses of 

and to be 4 u, 12 u, and 16 u, respectively,

write down expressions for the radii, R, forand (c) Using the actual masses of the atoms, findthe fractional differences in these three radii (He versusC and C versus O). The tiny differences in these radii

can be measured very accurately and allow precise com-parisons of the three masses involved (one of which,

defines the atomic mass scale).

16.62 •• The chemical composition of a molecule can oftenbe deduced from a very accurate measurement of itsmass. As a simple example consider the following.Gas from a car’s exhaust is analyzed in a mass spec-trometer.Two of the resulting peaks are very close to28 u. (a) Can you suggest what gases produced thesetwo peaks? (b) Careful measurements indicate thatthe masses of the neutral molecules concerned are28.006 u and 27.995 u. What are the two molecules?

12C,

O4+.C3+,He+,

16O12C,4He,

COMPUTER PROBLEMS

16.63 •• (Section 16.2) Same as Problem 16.15, but inpart (c) use graph-plotting software to plot for

from to given the following

parameters:

and where Plot inunits of that is,plot

16.64 •• (Section 16.2) Same as Problem 16.63, but makeplots for and drawing both curves on thesame plot.

16.65•• (Section 16.2) Same as Problem 16.63, but usesuitable plotting software to make plots for and

16.66 ••• (Section 16.7) (a) According to the semiempiricalbinding-energy formula, the most stable value of Z forany fixed A is given by (16.51) in Problem 16.46. (Thisis for A odd.) Using suitable plotting software,plot thecurve of stability, showing the most stable values of 

for as predicted by (16.51).Thesimplest way to do this is to make a parametric plot of 

as a function of the parameter  A. Include onthe same plot the positions of the following odd- A

nuclei: and 235U.187Au,107Ag,63Cu,7Li,1N, Z

2

0 6 A 6 240,1N, Z2

238U.

4He

244Pu,12C

r 1r2>r 0 .r 0 ;r 1r2R0 = 1.07 fm.R = R0  A1>3, t = 0.55 fm,r 0 = 3.25 * 1017 kg>m3,

r = 15 fm,r = 0208Pb

r 1r2