434 I 12. Identical Particles where the term in the first brackets is just the sum of two hydrogenic Hamiltonians Z,with charge whose expectation value is straightforward to determine for the state (12.52).6 Then (£ ) = imec2a2( - 2Z2 + 4Z(Z- Z) + ~Z) (12.54) = ~mec2a2(2Z2 - 4ZZ + ~Z) Setting ()(~) = 0 ('12.55 ) az we find the value of Z that minimizes the energy: * (12.56) Substituting this result into (£), we obtain (12.57) =which is much closer to the observed value Eexp -79.0 eV than our earlier first- order perturbative result (12.32). Notice that (12.56) has the simple interpretation that each electron is partially screened from the nucleus by the presence of the other electron in the atom. As before, we have put Z = 2 in (12.57) to obtain the numerical result. The variational method for the ground state of helium can be improved with more complicated trial wave functions. In fact, Pekeris has used a wave function involving 1075 parameters to obtain numerically an estimate for the ground-state energy that agrees with the experimental results to within the experimental errors? In general, the reason for working so hard to obtain such good agreement is that then one can use the wave function that has been deduced to calculate other quamities, such as the lifetimes of excited states. We can also estimate the energies of these excited states and obtain the corre- sponding wave functions using the variational method. Note in (12.48) that if we choose a trial state lo/) that does not have any of the true ground state in it, that is, if it is orthogonal to the ground state: (Eo lo/) = 0 ( 12.58) 6 Note: The Hamiltonian itself does not depend on Z. We have added and subtracted terms such as Ze2/If, ] to make the evaluation of (1/IIHil/1} as easy as possible. For the variational method to work, the variational parameter cannot be a parameter appearing in if. 7 C. L. Pekcris, Phys. Rev. 115, 1216 (1959). Page 451 (metric system)
12.2 The Helium Atom I 435 then the expectation value of the energy in this trial state satisfies (E} :::: £ 1, and we can obtain an estimate of the energy £ 1 by minimizing this expectation value. Sometimes it is easy to satisfy the condition (12.58). For example, if the excited state has nonzero orbital angular momentum, choosing a trial wave function involving the appropliate sphelical harmonic for the angular dependence automatically guarantees that this state is orthogonal to the ground state. Or, if necessary, we can determine an excited-energy trial state that is orthogonal to the trial state 11ft) that we obtained from estimating the ground state energy. We choose a new trial state llfl) and then explicitly construct a state that is orthogonal to the trial ground state 11ft): (12.59) However, since we are using trial state i·ifi) rather than the ~rue ground state in this superposition, we should not expect minimizing the expectation value of the energy in this state to yield an upper bound for £ 1. EXAMPLE 12.2 Use the variational method to estimate the ground-state energy of the hydrogen atom. l SOLUTION Of course, we already know the ground-state energy of hy- drogen. Our goal here is to show how to implement the variational method in a step-by-step approach. The Hamiltonian for the hydrogen atom is IAs a first step, we need to make a judicious choice of the trial wave function. Since we are looking for the ground state, we need to choose a spherical ly symmetric wave function, since we know the ground state has I = 0. More- over, for a sphe1ically symmetric potential energy, we knov.. that for small r the radial wave function R(r) behaves as r1• Finally, since our goal is to can-y out the calculation analytically, we want to choose a wa,·e function for which we can actually do the integrals needed to evaluate (£ }. There are two likely possibilities for the trial wave function, a simple exponential and a Gaussian. Here we will use the trial wave function R(r) = Ne- br where N is the normalization constant. First, we determine N: Page 452 (metric system)
436 I 12. Identical Particles Choosing N = 2b312 the full trial wave function is therefore 1ft = R (r)Yo 0 = ~1R(r) = b3/2 . \"\\( 4.7r '-e- br \"\\( .7r Consequently, (£ } = (1/ti HllJr> Minimizing the expectation value of the energy: leading to Substituting this value forb into the expression for (E), we see that £0 <(E) = -JLe-4 = - -1J.LC2a 2 - 2ti2 2 =where a e2f (fie) is the fine-structure constant. Since the exact ground-state energy of the Hamiltonian is -tJ-e2a2 / 2, the variational method in this case has yielded the exact energy. This is a consequence of our using a trial wave function that included the ground-state energy eigenfunction for a particular value of the parameter b. If we had used the same trial wave function for the three-dimensional harmonic oscillator, the upper limit on the ground-state energy would not be the exact eigenvalue. See Problem 12.8. -=.........: - Page 453 (metric system)
12.3 Multielectron Atoms and the Per iodic Table I 437 12.3 Multielectron Atoms and the Periodic Table Let's tum our attention to multielectron atoms. Even if we neglect the contribution of the kinetic energy of the nucleus (with charge Ze), then, as for helium, the energy eigenvalue equation for the Hamiltonian Pf~ ~ +~ ze2 \" ' e2 (12.60) - L._; - L._; H = L._; - . i =l 2me i=l li\\1 i<j li\\- rjl is too complicated to solve exactly. Of course, if we could neglect the contribution of the Coulomb repulsion between the individual electrons, the solution would be straightforward. The Hamiltonian would then just be the sum of Z individual Coulomb Hamiltonians (with charge Ze on the nucleus for ea't:h), and the allowed eigenstates could be formed from a direct product of these individual Coulomb eigenstates, provided we require that the total state, including spin, is antisymmetric under exchange of each pair of electrons. However, there is no reason to expect that we can treat the Coulomb repulsion of the electrons as a perturbation, as we attempted for helium. In particular, the typical distance separating the electrons in the atom should be of the same size as their distance from the nucleus, and although the size of the mutual interaction between each pair of electrons should be smaller than the interaction of the electrons with the nucleus because of the factor of Z in the nuclear charge, there are Z (Z - I)/ 2 different pairings of the Z electrons to take into account. Thus, even for modest Z we should expect the mutual interaction term of all the electrons in the atom to be comparable in size to the interaction of these electrons with the nucleus. We clearly need an altemative way of dealing with the mutual interaction of the electrons. One approach, first used by Hartree, is to treat each of the electrons as moving independently in a spherically symmetric potential energy V (r) due to the nucleus and the other Z - 1 electrons. Tbi·s potential energy should have the form IV(r) = ze2 r~ 0 ( 12.6 1) r ~ oo ; e r because for small r the electron sees only the nucleus while for large r the nucleus is shielded by the other Z - I electrons. How do we determine the form for this potential energy, since it depends on the charge distribution of the electrons, which is itself detemlined by solving the Schrl.'\\dinger equation? One approach is to guess a reasonable form for V(r ) and then solve the SchrOdinger equation (numerically) to detennine the wave functions. As in (12.25), we can use these wave functions to determine a charge distribution, which can then be used to determine the potential energy that each electron experiences. We can continue with this procedure until we Page 454 (metric system)
438 I 12. Identical Particles obtain a self-consistent solution-namely, the potential energy that we determine from charge distribution of the electrons is the same, to a given accuracy, as the potential energy that we used to determine the wave function s that yielded this charge distribution. We can label the energy eigenstates by the three quantum numbers 11. l, and m, just as we did for the hydrogen atom. However, since the potential energy does not have a simple 1/r dependence, states with a particular nand different l do not have the same energy. In particular, states with lower l should have lower energy, since the lowering of the centrifugal banicr with lower l permits these wave functions to penetrate more deeply inside the charge distribution formed by the other electrons and thus \"see,\" at least at small r. the full attractive potential energy of the nucleus with its charge Ze. Thus the accidental degeneracy of the hydrogen atom disappears. The independence of the energy on m persists, however, ~ince the potential energy is taken to be spherically symmetric. Let's discuss the ground-state electronic configuration of the elements, especially those of low z. Start with the first row of the periodic table, shown in Fig. 12.5. Hydrogen has n = I and l = 0 for the ground state. We call this a 1s electron configuration. For helium, as we saw in Section 12.2, we can put the two electrons both in the Is state, which we now denote by ls 2 as a shorthand notation for Is ls. Of course, the spin state must be a total-spin-singlet state in order to make the total state antisymmetric under exchange. These two electrons fill then = I shell. It takes more than 19 eV to excite one of the electrons to then= 2 level and 24.6 eV to ionize the atom by removing one of the electrons. Helium is exceptionally stable and, correspondingly, not chemically active.8 The next element in the periodic table, lithium, has three electrons, but we cannot add this third electron to the lv level, for then two of the electrons would be in the same state, since there are only two possible spin stales, spin up and spin down, for each of the electrons. Thus the total state could not be completely antisymmetric under exchange of any pair of the electrons. One of the electrons must therefore be in the next highest energy state, the 2s state, which has lower energy than the 2p state. We label this electron configuration by 1s22s . It takes only 5.4 eV to ionize lithium and thus lithium is chemically quite active. The ground state of the next element in the periodic table, beryllium, is ls22s2, while boron with five electrons is in the ls2 2s2 2p state. The next five elements-C. N, 0, F, and Ne-fill up the 2p level, which can accommodate 6 electrons, since for l = 1, we can have m = 1, 0, and -1 and two possible intrinsic spin states for each of these orbital states. Figure 12.6 shows the ionization energy for each of the elements. Notice that as Z increases within a shell, the electrons are pulled in toward the nucleus and the ionization energy 8 See also the discussion at the end of Section 12.4. Page 455 (metric system)
,•\"10' I I I II I c iil lS I Ri i.n ~12 ;l 3Li ,.Be ~ zs• zs2 .,\"C ~ ~ ~ [ ;:;· uNa 12Mg E! 3s • 3s2 0' 2S112 1So Ci\" .0..., t9K zuCa 21Sc zzTi 23V a& 24Cr zsMn z6Fc 27 4s1 4s2 Jd 1 3d 2 Jd3 4 s 1Jd5 Jd s u/6 '£ 0 6Ssn 50 ,1 4F 9 2S 112 1Sn 2D312 3Fz 4F312 7S3 4)fc •t4Ru 45 (I) 37Rb 3sSr 39Y 40Z r 4tNb 4zl\"lo ss 14d6 5,,·1 ::I ss1,td 5 6Ssn ss 1.ut' '1F ss1 sx2 4d t 4li2 ss1·td 4 '~ 7S3 1sRc ·5Fs 77 ::I 2Stt2 1So 2D312 3Fz 6Diiz sds s !2. 74W 6Ss12 1c.Os ~,: a·c: sd4 5<t<• ssCs s6Ba 57-7 1 nHf 7)fa 1078h w~ 0.. 51Jo 6ds 51)4 6s• (,~2 Rare sd2 sd3 61 (Jq 10&Sg 61Pm wHils 6d4 6d6 e13 & 2S 112 1So earths 3F2 4F312 4f5 0 60Nd 62Sm 4 4/4 1'11sn 4/•\" 8S ~7Fr xs R a 89-103 104Rf 10sDb 514 7s2 Acti- sjl46d2 6d3 93Np 7Fo 95A ~ 7s' •so ttides nU Mls_r s 6dlsj3 6L1112 94Pu 8S ae. 2s112 sj6 0 54, 7Fo 0 Rarc car.tbs 57La ssCe s9Pr 0 (Lantham des) ::I sdt 4f2 4f3 2Dm ;:1) 1G4 4/912 (JO c cr~ s9Ac 90Th 91Pa \"'::I Actinides 6<1' 6</2 6dlsj2 2D312 3Fz 4K1 112
[ II Ill I IV I v I VI I VII 1 Vlll 1 2 ~ sB ~ 7N sO 9F 10Ne 2pt 2pz ~ zp4 zpS 2P6 ~~~ ~~ ~ t&S 11CI tsAr Jr13AI t4Si tsP 3p• 3p2 3p3 3p4 3p6 2P tn 3Po 4Sm 3P2 2P:~n. 1So ~ 7Co 211Ni 29Cu 30Zn 31Ga 32Ge J:v\\S 34Se Jsll•· 36Kr w1\\J £1' 3d8 4s13d10 Jd 10 4p1 4p2 4p3 4p4 4p5 4p6 c3:: Fw2 3F4 2St/2 1So 2P ll2 3Po 4S312 3P2 2P Ji2 1So ;:;: Rh 46Pd 47Ag 4sCd 49ln soSn stSb szTe s3l s..Xe 14118 sso<Jd 10 ss 14<f 10 4c/10 sp 1 sp2 sp3 irioi' 2P1.12 3Po 4S3n sp 4 sp5 Sfl 6 Fwz 1So 2S 112 1So 2P312 (.r.T.), s3Bi 3Pz 1So 7ll' 7~ 1'1 79Au soHg stTI szPb 6(>3 Hs A t 0 sd' l,.~ • si/9 6s •s d ' o 6p t 6p2 4S312 s4Po c.p5 s6Rn :,)12 sd'o 3Po 6p4 2 1'~12 (>fJ6 :;:! .If)-' 2S 112 2P1 12 3P2 ~MI 1So 1So - ,)..>.. - 111Rg 111 1101>~ 7.126lt9 112Cn '' 0 1.,•1><1'1 7s26dl0 3 3Eu MGd 6sTb 66D v 67Ho 68Er 69Tm 10Y b 71 l~u 4f7 5£/•tj7 1>.11 4/\" 4/ t o· 4/ • -' 4/ 1'. 4/14 V> Sm ,1/h 1'H1~12 5/s 4/11 3· /16 2F·m st/14/14 !JJ 4/1 512 1So 21)3fl :;:! Q. Am 96Cm 978k 9sCf 99Es 10ofm 10 1M d 102No 103Lr sj1 6lfl<oj1 6</l<ojK <o/10 sjll sjl2 sj13 sfl4 sjl46dl r::Tr 51s 4/1512 3H6 2Fm S112 9lh 6H1512 Pa1Sgoe 4562P(1mn.etric system) If) \"0 I..f.). 6' Q. r;· 0} Q: If) .t> t.l I!)
440 I 12. Ident ical Particles 25 20 c ·.:I2:':i: 10 .9 5 L - - -- -- ----2L0_ _ _ __ _ _ __ _v Z 4LQ--------~6L()----------8LQ----- Figure 12.6 The ionization energy of the elements. increases. Neon, like helium, has a completely filled shell and a high ionization energy of 21.6 eV. The charge dcnsi[)' for a closed shell I ( 12.62 ) L e1Rn,t(r ) I2 1Yt, 111 (e, ¢)12 m=- 1 is spherically symmetric since ceI ,~-.)1 2 - 214+n 1 L\" ''Y - IYl ,m (12.63) m=-1 Thus the electronic charge effectively shields the nuclear charge. Although exciting one of the electrons to an excited state would change this situation, the energy gap between then = 3 and n = 2levels is sufficiently large that the atom has little affinity for other electrons and is chemically quite inert . The third row of the periodic table starts with sodium. After filling then= 2 shell, the eleventh electron must go into the n = 3 level. Again, the 3s level lies lower than the 3p level and thus the electron configuration is ls2 2s2 2p6 3s. Since this last electron is in a new shell, it is primarily shielded from the nucleus by the inner ten electrons and thus, like lithium, has a low ionization energy (5.1 eV) and a high chemical activity. Both lithium and sodium are alkali metals, which are quite reactive chemically. As an example, iffluorine is present, sodium sees a natural home for its \"extra\" electron; it can donate it to fluorine completing then = 2 shell for that element. These ions then bind together through electrostatic attraction, fom1ing an ionic bond. In moving from sodi um to argon along the third row of the periodic table, the s and the p levels fill up just as they did in going from lithium to neon along the Page 457 (metric system)
12.4 Covalent Bonding I 441 second row, and thus the chemical properties of these elements are strikingly similar to the ones above them in the periodic table.9 In particular, chloline, like Jl.uorine, is a halogen that needs just one electron to complete a shell- the 3p shell in this case. Chlorine can take that electron from sodium, producing sodium chloride, or ordinary salt. Argon, like neon, has a closed shell and is relatively inert; it is one of the noble gases. The next row in the periodic table contains some surprises. Instead of fi lling the 3d level next, as you might expect based on the energy levels of hydrogen, the 4s electrons penetrate the electron cloud of the inner electrons and have their energy pul1ed down below that of the 3d level. In fact, there is very little separation between the 4s and 3d levels. By the time the 3d level is filled with four electrons, the interaction between the electrons raises the energy of the 4s level so that it is slightly above the 3d level. Thus chromium has an outer sh!Uelectron configuration of 4s 1 3d5 instead of 4s2 3d4. The next few elements- manganese, iron, cobalt, and nickel-have a filled 4s level with electronic configurations ranging from 4s2 3d5 to 4s2 3d8. The chemical properties of these elements are all similar. Since the average radius of the 3d electrons is somewhat less than the4s electrons, the outer 4s electrons tend to shield the inner 3d electrons from outside infiuences.10 At copper, which is 4s 1 3d10, the pattern shifts again, with one of the electrons from the 4s level shifting to the 3d level. However, the two configurations 4s2 3d9 and 4s 1 3d1o are so close together in energy that copper can behave as if it has one or two valence electrons depending on its chemical environment. Finally. after both the 4s and 3d levels are fi lled, the 4p level fills, repeating the pattern of the previous two rows. One of the triumphs of quantum mechanics is that it provides us with a detailed understanding of the physics responsible for the periodicity in the chemical proper- ties of the elements. 12.4 Covalent Bonding In our discussion of chemical activity of the elements, we have indicated that certain elements can bind ionically together through electrostatic attraction after they have transferred an electron from one of the elements to the other. There is another type of bond, the covalent bond, in which the elements actually share rather than exchange their electrons. This sott of bonding is pure quantum mechanics in action. To see how it arises, we first consider as a specific case the positively charged hydrogen 9 The ionization energies are slightly less, since n = 3 instead of2. 10 This effect is even more pronounced in the sixth and seventh rows of the periodic table. The 4f electrons do not fi ll up until after the 6s shell. Thus the rare earth elements with Z ranging from 57 to 71 have similar chemical prope11ies. Also, the Sf electrons do not fill until after the 7s shell, leading to very similar chemical properties for the actinides, Z = 89 to Z = 103. Page 458 (metric system)
442 I 12. Identical Particles molecule ion, where a single electron is shared by two protons; then we consider the more prototypical case ofthe hydrogen molecule, where two electrons are shared. We will see that the identical nature of the electrons plays a cmcial role in this covalent bonding. THE HYDROGEN MOLECULE ION Although molecules, even simple diatomic molecules, are complex systems with many degrees of freedom, fortunately there are approximations that we can make that make the problem of determining the bound states of molecules a reasonably tractable one. In particular, recall from Section 9.7 that the energy of vibration of the nuclei of a diatomic molecule is on the order of (melm N) l/ 2 smaller than the electronic energy of the molecule. Thus the typical period of the motion of electTons in the molecule is much sho1ter than that of the nuclei, and we can neglect the motion of the nuclei as a fin;t approximation, part of the Born-Oppenheimer approximation. For a diatomic molecule we then consider the behavior ofelectrons under the influence of two fixed nuclei separated by a distance R. In fact, we can take R as a parameter that also appears in the wave function and that we can adjust using the variational method to determine the value of R that minimizes the energy of the molecule, thus determining both the energy and the size of the molec ule. HiA natural trial wave function for the hydrogen molecule ion is determined by first considering the lowest energy state of the system when the two protons are widely separated. Then there are clearly two possible states: either the electron is attached to one of the protons, forming a hydrogen atom in the ground state, or the electron is attached to the other proton, again in the ground state of a hydrogen atom. These two states are indicated in Fig. 12.7. In terms of the coordinates shown in Fig. 12.8, the corresponding position-space wave functions are given by {r l l) = _l_ e-lr - R/2(/ao (12.64a) (12.64b) R {r !2) = _l_ e-lr+R/2(/ao R There are, ofcourse, many other possible states of the system that we are neglecting in which, for example, the electron is in an excited state of the hydrogen atom. What is the proper linear combination of states 11) and 12) to use for the variational method? Here, as was the case for the N atom in the ammonia molecule that we treated as a two-state system in Chapter 4, there is an amplitude for the electron attached to one of the protons to jump to the other proton. This amplitude means Page 459 (metric system)
12.4 Covalent Bonding I 443 12> l l } Figure 12.7 A sc.:hematic representation of the two states used as a basis for a. variational calculation of the ground state of the hydrogen molecule ion. ln 11} the electron combines with one of the protons to form a hydrogen atOm in its ground state, while in 12} it combines with the other proton, again forming hydrogen in the ground state. Figure 12.8 The coordinates of the two protons and p the electron used in the discussion of the hydrogen molecule ion. that the matrix representation of the Hamiltonian ( 12.65) using the states 11) and 12) as a basis will have off-diagonal matrix elements, similar to what we assumed in our treatment of the anm10nia molecule. Here. because of the relative simplicity of the H2 ion, we can actually calculate the matrix elements: (12.66) where E1 is the ground-state energy of the hydrogen atom: p 2 e2 ) e2 e2 H22 = (21 ( 2me - lr + R / 2! 12) - (21 lf _ R / 2ll2) + R (212) J= E1 -d3r Jr ~ l(r l2)1 2 +~- = H 11 ( 12.67) - R /21 R Page 460 (metric system)
444 12. Identical Partic les \\\\here the last step follows from the symmetry of the two configurations; p 2 e2 ) e2 e2 H12 = (11 ( 2me - lr + R /21 12) - (111r- R /2112) + R (112) I= d 3r ( l lr) e2 (rl2) ( £1+ -e2) (1 12 ) - lr- R/21 (12.68) R ''here the amplitude I{112) = d 3r {llr }(r l2) (12.69) is called the overlap integral. Note that the nonvanishing of the off-diagonal matrix element depends on the stares 11) and 12) overlapping in space. Since the wave functions are real, the off-diagonal matrix elements are equal: p2 e2 ) e2 e2 H 21 = (21 ( 2me- lr _ R / 21 11) - {2l lf + R / 2111) + /? {211 ) ( e2) I e2+ -= £ 1 (21 1 ) - d3r{21r ) {r l l) = H12 ( 12.70) R lr + R /21 =Because H11 = H22 as well as H 12 H21, the linear combinations of the states II) and 12) that diagonalize the Hamiltonian are (12.71) where the overall factor is needed to normalize the s tates because the basis states are not orthogonal (see Problem 12.10). Note from the form of the wave func- tions ( 12.64) that the wave function {r l+) has even parity, while the wave function {r l-) has odd parity. We could have selected the two states I+} and 1-} initially as the proper linear combinations from the inversion symmetry of the Hamiltonian (12.65). Figure 12.9 shows a sketch of the wave functions, and Fig. 12.10 shows the expec- tation values of the energies E= I ± l{112) ( H 11 ± H) (1 2.72) 12 ± plotted as a function of R. Only the even parity state hac; a minimum, which occurs for a ~eparation of 1.3 A for the protons, corresponding to a binding energy of 1.8 eV. Thus the state I+} is referred to as a bonding molecular orbital, while the state - is called an antibonding molecular orbital. These molecular orbitals are linear combinations of atomic orbitals. Note from Fig. 12.9 that only for the bonding orbital is the electron shared between the two protons. For the antibonding orbital, on the other hand, there is a node in the wave function midway between the two protons Page 461 (metric system)
12.4 Cova le nt Bonding I 445 (a) Bonding Antibonding (b) Figure 12.9 The wave functions for the bonding and antibonding orbital of the hydrogen molecule ion plotted (a) along the axis connecting the rwo protons and (b) in three dimensions. where the potential energy is quite negative, and thus the electron in this state doesn' t benefit from the full attraction of the two protons. The experimental separation between the protons for the hydrogen molecule ion is 1.06 A, with a binding energy of2.8 eV. The reason for the lack of better agreement between our variational results and experi ment resides in our choice of trial wave f unction. In particular, notice that as R ~ 0, the system reduces to He+, while our trial wave function remains the Is ground state of hydrogen. Thus we should not be surprised that we have underestimated the size of the binding energy and Page 462 (metric system)
446 I 12. Identical Particles E(eV) r-+--r~--~2----~3----~4----~s-----76-----?~R(A) -13.6 ---- -- Figure 12.10 The energies of the bonding and antibonding orbitals as a function of the interproton separation R. overestimated the size of the molecule. One way of doing bencr is to use a trial wave function with an effective charge Z (as in Section 12.2) as well as the interproton distance R as parameters. Nonetheless, our solution demonstrales the key qualitative features of the binding. MUON-CATALYZED FUSION Note that the interproton separation in the hydrogen molecule ion is on the order of the Bohr radius a0 of the hydrogen atom. This length scale enters in the trial wave functions (12.64) that we used in our variational approach to the molecule. Since a0 = fijJLCa, the size of the molecule depends on the reduced mass JL of the corresponding atom. In particular, suppose we were to replace the electron in the hydrogen molecule ion with a muon. Then the reduced mass is given by + +mtl.mp/(m11 mp) instead of memp/(me mp). Since the reduced mass for the muonic atom is roughly a factor ofmtl./me larger than the reduced mass of hydrogen, the interproton separation in a muonic hydrogen molecule ion should be a factor of melmJ.L smaller than for the molecule with an electron generating the binding. Replacing the electron with a muon produces a much smaller molecule, just as replacing the electron with a muon in the hydrogen atom produces a much smaller atom. Suppose that the nuclei of this diatomic molecule consist of two deuterons instead of two protons. The small size of the muonic molecule means that the deuterons have a significantly greater probability of being close together than is the case when an electron is responsible for the binding of the molecule. ln fact, for this muonic molecule nuclear reactions such as d + d--+ t + p + 4.0 MeV (12.73) Page 463 (metric system)
12.4 Covalent Bonding I 447 have been observed to take place. This is a typical fusion reaction of the sott that is contemplated as a plentiful energy source (using the deuterium naturally present in sea water). However, whereas the attempts to generate fu sion commercially depend on thermonuclear reactions generated by heating the deuterons to sufficiently high temperatures that they have a significant chance of overcoming their Coulomb repulsion and being close enough together (on the order of a fermi) to pennit the reaction (12.73) to occur, the quantum mechanical binding in the muonic molecule does this at ordinary room temperature. Thus this muon catalysis can be described as a form of\"cold fusion .\" Unfortunately, muons themselves are not freely available (except in cosmic rays) and thus an accelerator is required to generate the energy required to produce muons, for example, in the fom1 of muon-antimuon pairs. The only way for reactions such as (12.73) to be a net generator of energy is for each muon to fom1 a large number of muonic molecules in which it catalyzes a nuclear reaction before the muon decays in roughly 2.2 tnicroseconds.11 So far this has not been feasible, but work is still in progress. THE HYDROGEN MOLECULE We are now ready to turn our attention to the hydrogen molecule, H2, where we must examine the effect of the identical nature of the two electrons on the binding of the molecule. Although we won't spell out the details here, we can use an approach similar to the one we used for the hydrogen molecule ion to understand the binding of the hydrogen molecule. For H2 , each hydrogen atom in the molecule supplies, of course, one electron. For each of the electrons to be in the same region of space between the two protons, the spatial state must be symmetric under exchange of the two patticles, and consequently the total-spin state must be a spin-0 state. With two electrons being shared instead of one, the molecule has a binding energy of H;,4.7 eV, as compared with 2.8 eV for despite the extra Coulomb repulsion of the electrons. The interproton separation for the molecule is 0.7 A, as compared with 1.3 Afor the ion. Binding occurs only for the total-spin-0 state; the total-spin- I state, corresponding to an antisymmetric spatial state in which the electrons. in general, do nol reside together in the region between the two protons, does not exhibit a minimum in Lhe total energy for any separation of the protons. Thus repulsion rather than binding occurs in this case. We can now also see why, for example, a hydrogen atom and a helium atom do not bind together to form a molecule.The two electrons in the ground state of heli um are both in the same spatial Is state, and therefore their total-spin state is the singlet spin-0 state. We say these two electrons in helium are paired together.The electron in the hydrogen atom cannot form a covalent bond and pair up with either one of these 11 L. Al varez et al., Phys. Rev. 105, I 127 (1957). For an cnt.e1tai.n.ing accounl o f Luis Alvarez's discovery of muon-catalyzed fusion, see Nobel Lectures- Physics, vol. II, Elsevier, New York, 1969 . Page 464 (metric system)
448 I 12. Identical Particles elecrrons, since this would either mean that three electrons are in the same spatial ~tate (which is forbidden) or else one of the elecrrons would have to be excited to Lhe 2s state of helium, which is energetically quite costly. On the other hand, as we have seen in the hydrogen molecule, if the electron from hydrogen interacts with an electron from helium in a total-spin-1 state, repulsion occurs between the two atoms. This is the reason helium is an inert element without affinity for other atoms, unlike an element with an unpaired outer electron that can pair up with an electron from another atom to form a covalent bond. Moreover, once two unpaired electrons from different atoms have paired up to form a covalent bond, an electron from a third atom cannot pair up with either one of them. Thus we see why the chemical forces saturate. 12.5 Conclusion Most of our attention in this chapter has been devoted to systems containing iden- tical fermions. The requirement that the state of the system be antisymmetric under the exchange of any two identical fermions means, in particular, that two identical fermions cannot occupy the same state (the Pauli principle). For identical bosons, on the other hand, the stale of the system must be symmetric under exchange of any two of the particles. Consequently, it is possible, and in fact preferred, to have many identical bosons in the same state. In Chapter 14 we will see an example when we discuss how a laser operates. Other examples in which many bosons condense to the ground state at sufficiently low temperatures include superconductivity and super- fluidityP Like the laser, these phenomena are interesting and exciting macroscopic manifestations of purely quantum behavior. Problems 12.1. Verify for an antisymmetric spatial state under exchange of two particles that 1/ . ( ' 1 )lo/A)= 2 d\\·1 d\\2 .J21rt, rz) - .J2 Ir2, r1) x (~ (r1, r2io/A) - ~ (r2, rJio/A)) J i= d3r1 3r2 lr 1, r2) (r1, r2 1o/A) 12 This condensation is often referred to as Bose- Einstein condensation, although this te1111 should probably be restJictcd to the condensation to the ground state that takes place when a dilute gas of atoms confined in a trap is cooled to temperatures in the microkelvin range. See J. S. Townsend. Quantum Physics: A Fundamental Approach to Modern Physics, University Science Books. Sausalito. CA, 20 I0, pp. 242- 247. Page 465 (metric system)
Problems I 449 12.2. Two identical, noninteracting spin-1 particles of mass m are in the one- dimensional ham1onic oscillator for which the Hamiltonian is (a) Determine the ground-state and first-excited-state kets and corresponding energies when the two particles are in a total-spin-0 state. What are the lowest energy states and corresponding kets for the particles if they are in a total- spin-! state? (b) Suppose the two particles interact with a potential energy of interaction V. (Jx1 - x21) = { - V0 Jx1 - x2 J<a 0 elsewhere ~ Argue what the effect will be on the energies that you determined in (a), that is, whether the energy of each state moves up, moves down, or re- mains unchanged. Suggestion: Examine which spatial wave functions for the total-spin-0 and total-spin-1 states tend to have the pa1iicles closer together. Consider, for example, the special case of x1 = x2. 12.3. Obtain an order-of-magnitude estimate for the singlet-triplet splitting of the energy levels of the two electrons in helium due to a direct spin-spin interaction. Suggestion: Compare with the magnitude of the hyperfine interaction in hydrogen as discussed in Chapter 5. 12.4. Use the variational principle to estimate the ground-state energy for the one- dimensional anharmonic oscillator A? fi = P; + bx4 2m Compare your result with the exact resull 12.5. For the delta function potential well 2m V(x) = A - - - o(x ) Ji2 b use a Gaussian wave function as a trial wave function to obtain an upper bound for the ground-state energy. Compare with the result of Problem 6. 19. 12.6. Consider the one-dimensional system of a particle of mass m in a uniform gravitational field above an impenetrable plane. Take the potential energy to be infinite at the plane and locate the plane at z = 0. Page 466 (metric system)
450 I 12. Identical Particles (a) Plot the potential energy of the particle. What is the Hamiltonian? Sketch roughly the ground-state wave function. Cb) Use an appropriate trial wave function to estimate the ground-state energy. (c) What is the average position (z) of the particle above the plane? 12.7. Use as a trial wave function the Gaussian (b)= ;1/f(x) l/4 .2 e-bx /2 to obtain an estimate of the ground-state energy for the linear potential energy V(x) = alxl. 12.8. Use the trial wave function R(r) = N e-br to estifnate the ground-state energy =of the three-dimensional isotropic hannonic oscillator, for which V (r) ~ ttw2r 2 . Compare your estimate with the exact value (see Section 10.5). 12.9. A muon and a proton are bound together in the ground state of a muonic \"hydrogen atom.\" As this atom diffuses around, it bumps into a deuteron. What is the incentive in terms of energy for the muon to jump from the proton to the deuteron, forming another muonic atom, but this time with the deuteron instead of the proton as the nucleus? Explain. 12.10. Show that the linear combinations of states that diagonalize the Hamiltonian of the hydrogen molecule ion are given by (12.71 ). Verify that these states are properly nonnalized and that the conesponding energy expectation values are given by (12.72). Note: If 11/f) is not normalized, then (E)= (1/lllht/1)/(1/1 11/f). Page 467 (metric system)
CHAPTER13 Scattering How do we learn about the nature of the fundamental interactions on a microscopic level? Solving the Schrodinger equation for the hydrogen alOm yields an infinite set of energy levels that gives us proof that the Coulomb interaction is the predominant interaction between an electron and a proton on a distance scale on the order of angstroms. However, for the two-body problem in nuclear physics, as we discussed in Chapter 10, there is a single bound state, so we must resort to scattering techniques to learn about the nature of the nuclear force. After all, it was through a scattering experiment that Rutherford first discovered the very existence of the nucleus within the atom. Subsequently, scattering has played a major role in helping us learn about nuclear physics and particle physics, as well as more about atomic physics. After introducing the concept of the scattering cross section, we will use the Born approximation and the partial wave expansion to calculate the cross section in quantum mechanics. These two approaches are in a sense complementary to each other: the Born approximation works best at high energies and the partial wave expansion has its greatest utility at low energies. 13.1 The Asymptotic Wave Function and the Differential Cross Section In a typical scattering experiment, a beam of particles, often from an accelerator, is projected at a fixed target composed of other particles. In Ruthetford's experiment, the incident particles were a panicles, 4He nuclei, that were emitted in radioactive decay, while Lhc target consisted of gold atoms in the form of a thin gold foil. A schematic diagram of this experiment is given in Fig. 13.1. The angular distribution of the scattered a particles provided clear evidence of the existence of a relatively massive gold nucleus. More recently, experiments done at the SLAC National Ac- celerator Laboratory with high-energy electrons accelerated in the two-mile-long 451 Page 468 (metric system)
452 13. Scattering ZnS detector a-particle source Thin foil Figure 13.1 A schematic diagram of the Rutherford scattering experiment. accelerator to 20 GeY as the incident projectiles and target protons in the form of liq- uid hydrogen revealed that protons were actually compose'tl of fractionally charged constituents, which are called quarks. A convenient way to describe such experi- ments is to picture the incident particle, initially far from the target, at least on a microscopic scale, as an essentially free particle that interacts with the target only when it is within the range of the potential energy of interaction V (r), which we will take to be spherically symmetric. Just as a comet can be deflected by its gravita- tional interaction with the Sun, so too can the incident projectiles in these scattering experiments be deflected by their interaction with the target particle. Since the par- ticles interacting through this potential energy V (r) are not bound, the energy of the incident particle can take on a continuum of different values, just as it did when we analyzed the free particle in one dimension in Section 6.6. As we saw there, physical states in the form of a wave packet can be formed from the superposition of these continuum states. In practice, this wave packet generally has a sharp peak in momentum space about some incident momentum p0, and consequently the wave packet in position space is quite broad. In fact, we will assume that it is sufficiently broad that we can treat it as a plane wave for the purposes of analyzing the experiment. This is generally the way one-dimensional scattering is discussed within wave mechanics. Specifically, as we discussed in Section 6.10, one considers a potential energy function such as the potential barrier shown in Fig. 13.2. Outside the range a of the potential , we can express the wave function as a plane wave: X <0 (13 .1) x>a where k = JzmE jh2 . The time dependence of such an energy eigenfunction is the usual e- i Er /t! . This is of course a stationary state. However, by superposing these energy eigenstates together, we can produce a wave packet with time dependence such that the incident wave packet alone approaches the barrier, and after interaction with the banier there is an amplitude for the wave packet to be reflected and an amplitude for it to be transmitted, as depicted in Fig. 6. I2. Examining the stationary Page 469 (metric system)
13.1 The A symptotic Wave Function and the Differential Cross Section I 453 v --------------~L-----~L--------------- x 0a Figure 13.2 A potential barrier of width a in a one- dimensional scattering experiment. ~ states to dctennine these amplitudes is analogous to doing a scattering experiment with water waves in a pond in which the source of the waves is not a single stone thrown into the pond (which would generate a wave packet) but a harmonic source that continually beats up and down in the water at a steady frequency. What is the analogue of (13.1) in three dimensions? We take the incident wave to be traveling along the z axis, with the target located at the origin. Far from the target the asymptotic wave function should include this incident wave together with an outgoing wave produced by interaction with the potential: 1jr ~ Aeikz + (outgoing wave) (13.2) r--+oc (see Fig. 13.3). Outside the range of the potential, where the pmticle detectors are located, the outgoing spherical wave must be a solution to the Schrodinger equation Aeikz A.t.ce,¢) -e,i-k.r Figure 13.3 A schematic diagram of a tlu·ee-dimensional scattering experiment indicating tl1e incident plane wave and the outgoing spherical wave. Page 470 (metric system)
454 I 13. Scattering with V = 0. Thus, we start with the differential equation (10.4) for the radial wave functio n u = r R with V = 0, (13.3) For large r we can neglect the centrifugal barrier term and obtain, for any l, the equation /i2 cPu (13.4) - --=Eu 2J.L dr 2 which has the two solutions (13.5) <t = =u eikr and u e-ikr wi th (13.6) If we attach the time dependence e-iEtfli to each or these solutions, we see that only eikr corresponds to the outgoing wave, since as time increases, r must also increase to keep the phase constant. This outgoing wave is the type that we expect to be generated by the interaction of the incident wave with the potential. Since R = uj r, this suggests that we express the asymptotic wave function (13.2) in the form ikr =1/f---* Aeikz + Af(() , </>):._ +1/linc 1/fsc r (13.7) r-.+oo where the function f (e, </>) allows for angular dependence in the outgoing scattered wave. After all, we have no reason to expect that the outgoing scattered wave should have the same amplitude in the forward direction (() = 0) as at other angles. Also, the spherical harmonics that appear in the energy eigenfunctions (rl E, l, m) = R(r ) Y1,111 ((), ¢) can add up to produce, in principle, any angular dependence. How do we relate this asymptotic wave function (13.7) to what is measured in the laboratory? Let's return to the way the experiments are actually performed. As in Section 6. 10, the incident flux of particles can be related to a probability current, whose form follows from the Schrodinger equation in position space, (13.8) We start with the probability density (13.9) Page 471 (metric system)
13.1 The A symptotic Wave Function and the Diffe rential Cross Section I 455 By evaluating the time derivative of this probability density and taking advantage of the form of the Schrodinger equation (13.8), we find that we can write where (13. I 0) (13.11) The dimensions of the probability current j are probabili ty per unit area per unit time. Equation (13.1 0) expresses conservation of probability in the form of a local 'i conservation law, since it implies that (13.12) where in the last step we have used Gauss's theorem to convert the volume integral to a closed surface integral over the surface enclosing the volume. Note that if the integral ove r the surface of the dot product of the probability current j with outward normal n to the surface is positive, there is a net outflow of probability from the volume, and consequently the probability of finding the particle in the volume decreases. In Section 6.1 0 we saw that the one-dimensional probability c urrent for the wave function (13.1) is given by . -liIkCI-? x >a ( 13.13a) J= m (13.13b) Thus the reflection coefficient can be calculated from (13 .14) R = Jrcr = (likfm.) IBI2 = IBI2 line (likfm)IAI 2 IA12 while the transmission coefficient is given by (13.15) =ln one dimension, conservation of probability requires that R + T 1. In three dimensions, there is more than jus t reflection and transmission. The eexperimentalist counts the number of particles scattered through angles and ¢ Page 472 (metric system)
456 I 13. Scattering dQ Ur / <Detector Transmined beam Scattered beam Figure 13.4 An experimental setup characteristic of a three-dimensional scattering experiment. The detector subtends a solid angle dQ. ~ that enter a detector that subtends a certain solid angle, as indicated in Fig. 13.4. In particular, the differential cross section da is defined by da = da dQ = number of particles scattered into ctQ per second (1 3. 16) dQ number of particles incident per unit area per second Note that the di mcnsion of the cross section is area. We can think of it as an effective area that the target presents to the incident flux of particles. Also note from (13.16) that if we multiply this incident flux, which is the number of incident pa1ticles per unit area per unit time, by the differential cross sectional area da, we obtain the number of particles scattered into the solid angle dQ per unit time. The total cross section a is obtained by integrating over all angles: f da (l3.17) a= dQ dQ This total area may have a simple physical significance in some cases. For example, in a nuclear scattering experiment with neutrons as the projectiles and a nucleus as the target, the total cross section is on the order of the size of the nucleus, since the nuclear force is short range and neutrons that strike the nucleus are likely to interact with it. On the other hand, ir neutrinos are the projectiles and the target is a nucleus, the cross section is many orders of magnitude smaller. This is not because the nucleus has suddenly shrunk in size but because neutrinos interact so weakly with the nucleus that most of the neutrinos pass right through the nucleus without scattering at all. The flux of particles in a scattering experiment is proportional to the probability current. If we use the asymptotic wave function (13.7) to calculate the probability current, we find that the current does not simply break up into two pieces consisting of an incident current and a scattered current. Unlike the one-dimensional case (13.13), there is inteiference between the incident wave and the scattered wave. This interference in the forward direction (B = 0) is, in fact, responsible for the reduction of flux in the forward direction from its incident value and therefore is necessary Page 473 (metric system)
13.1 The A sy mp totic Wave Function and the Differential Cross Section I 457 Figure 13.5 If the incident \"plane\" wave extends over a distance b transverse to the z axis, ea distant detec tor located a distance r from the origin and at an angle ~ bj r will not be in the path of the incident beam. As r --+ oo, such a detec-4or would only detect the incident beam as e--+ 0. for scattedng to occur at all. 1 Nonetheless, in practice it is possible to calculate probability flux entering a detector that is not directly in the forward direction by using the scattered outgoing wave function alone. In the actual experiments, the incident beam is limited in the transverse direction by the sides of the beam tube, if by nothing else. If we call this transverse dimension b, then at a distance r from the target the incident wave is present only for angles e ~ bj r (see Fig. 13.5). Since the patticle detectors are located at large distances from the target (r ~ oo) , the scattered wave will be the only wave of interest unless the detector is placed essentially at e= 0, namely, in the incident beam. To detennine the probability current that flows into a detector that subtends a solid angle dQ, we calculate J•sc = - fi . (·'+\"'*scnv .'+t',sc - .V11sc\"v'·V'' s*c) (13. 18) 2 f.Ll . where the asymptotic form of the scattered wave function is given by eikr 1/lsc ~ Af(8, ¢) - (1 3.19) r (13.20) r-;.oo The scattered probability current is then given by (see Problem 13.2) . ~ nk lAIzIJ IzUr Jsc -? r--+oo J.Lr~ where the unit vector Ur shows that this current is radially directed. The probability flow (probability per unit time) into a solid angle dQ is detennined by taking the dot product of the scattered probability current (probability per unit area per unit time) 1 See also the optical theorem in Section 13.4. Page 474 (metric system)
458 13. Scattering \\\\ilh the area u,.r2 dQ covered by an infinitesimal detector: (13.21) Since the incident probability flux, which is directed along the z axis, is given by J·m. c -- hk1A 12 (I 3.22) IL the differential cross section that follows from the definition (13.16) is (13.23) Thus (13.24) It is this differential cross section that replaces the reflection coefficient R used e,to describe one-dimensional scattering. The function I ( <P) is referred to as the scattering amplitude. 13.2 The Born Approximation A patticularly good approach for calculating the scattering amplitude when the energy ofthe incident beam is large in compm·ison with the magnitude of the potential energy is the Born approximation. We begin by expressing the position-space energy eigenvalue equation 2(- r~2 \\72 + V ) l{r(r) = El{r(r) (13.25) in the form (13.26) with k given by (13.6). The incoming plane wave Aeikz is a solution to the equation (13.27) It is convenient to express the formal solution to (13.26) in the form J1{r(r) = Aeikz +dV G(r, r ') 2 V (r1)1{r (r') ( 13 .28) h_2M' Page 475 (metric system)
13.2 The Born Approximation I 459 The function G(r, r') is called the Green's function of the differential equation (13.26) and itself satisfies the di1Terential equation (13.29) That (13.28) is a solution to (13.26) can be verified by applying the differential operator \\12 + k2 to (13.28). It is only a formal solution since the wave function 1{1 appears on the right-hand side of this equation as well as on the left-hand side. Nonetheless, we will see that (13.28) provides a useful route for determining the wave function in an iterative procedure known as the Born approximation. We first detem1ine the Green's function, making sure that the solution (13.28) satisfies the appropriate boundary conditions. As indicated by the argument of the .;, delta function, the Green's function itself depends only on the difference of the vectors rand r'. With this in mind, let's first set r' to zero in (13.29) and detennine the solution to the equation (13.30) Given the spherical symmetry of this differential equation, we naturally search for a solution in spherical coordinates. Notice that, except at the origin, the Green's function is a solution to (13.27). We saw in Section 13.1 that a solution to the SchrOdinger equation for a free pruticle in the form of an outgoing wave is given by = eikr (13.31) G(r, 0) C- r where Cis some constant. In fact, (13.31) is actually a solution to (13.30) even at the origin, provided we choose the value of C properly. Recall from (10.9) that v?--1 = - 4n83(r ) ( 13.32) r Note that 1 ( 1)= = +\\12ik~r \\1. ( \\1 -ei-kr) V . (V eikr) - V. V - eikr rr rr 1n2 (n I) I=- v eikr +-?vDeikr · v - +eikr nv2- r rr 2'k)- - - + -_ 1 k2l eikr - 2l.k-eikr- 4nu~3(r ) eikr r( r r2 (13.33) Page 476 (metric system)
460 13. Scattering Detector Figure 13.6 The r ' integration is restricted to lie within the range of the potential energy, which is indicated by the shaded region, while the particle is detected at r. where the evaluation of the action of \\72 on eikr has bee~ carried out using or a\\72;.J2 = 2 +angular derivatives (13.34) - +-- 2 r ilr Thus (13.3 5) Comparing this result with ( 13.30), we see that eikr (13.36) G(r,0) = - - 4nr Therefore e i kl r - r' l (13.37) G(r, r') = - - -- 4n lr - r 'l and f'1/J(r ) = Aeikz- ~ dV e iklr-r'l 2n fi2 V(r')'f/l(r ') (13.38) jr - r 'l Generally, the range of the r' integral is limited by the range of the potential energy V (r ') to a microscopic distance (see Fig. 13.6). In order to determine the scattering amplintde, we need to examine the behavior of the wave function as »r--+ oo. In this case, since r lr'l, we can approximate r' r')12) 1/2 (+jr - r' I = (r.-, - 2r · r ' r12) 1I2 = r 1- 2ur · - + -r , -----+ r I - ur · - ( r r- r -+oo r ( 13.39) Page 477 (metric system)
13.2 The Born Approximation I 461 where we have neglected terms of order (r'j r)2. Thus in (13.38) we can make the replacement __I_------+ ~ (1 +Ur · ~)------+ ~ (13.40) lr - r'l r->00 r r r~ oo r since the terms that we m·e neglecting make a vanishing contribution to the integral relative to the one that we have retained as r ~ oo. However, within the exponent all that matters is the value of the phase kr modulo 2n, no matter how large r is. Thus we need to retain both the first two terms in the expansion (13.39), namely, eiklr-r'l ------+ eikr(!- u, ·r'! r) = eikre-ikrr' (]3.41) r->CO where (13.42) points in the direction of the outgoing scattered wave. Again, the terms of order P(r'I r can be safely neglected relative to the two terms that we have retained. With these approximations, (13.38) in the asymptotic limit becomes ('13.43) In retrospect, we can now see why we made the choices wedidinde1iving (13.43). Clearly, we chose the particular solution Aeikz to ( I3.27) to match up with the bound- ary condition that our wave function include the correct incident wave. Similarly, in deriving the Green's function (13.37), we discarded the incoming spherical wave ce-ikr; r because of the physical requirement that the potential generate only out- going waves. In practice it may be possible to do scatteting with incoming spherical waves,2 but most experiments are similar to the approach described earlier in which an incoming plane wave generates an outgoing spherical wave upon interaction with the target. We are now ready for the Born approximation. As we remarked earlier, (13.43) is an integral equation that involves the wave function 1/1 on the right-hand side within the integral, as well as on the left-hand side. If the potential energy V were set to zero, the solution for 1jJ would be simply 1jJ = Aeikz. This suggests that if the mag- nitude of the potential energy is small compared with the energy E , we can replace the wave function 1/l(r') within the integral with that of the incident wave. Then (13.44) 2 An example might be using laser light to implode a pellet of deuterium in an attempt to generate them1onuclear fusion reactions. Page 478 (metric system)
462 13. Scattering CLmparing this equation with general asymptotic expression (13.7) reveals that the ~anering amplitude is given in the Bom approximation by (13 .45) where we have introduced the vector ki with magnitude k directed along the z axis, the direction of the incident wave. Note that the vector hq = hki - fikf = Pi - pf is just the momentum transferred from the incident beanf to the target during the scattering process and that the scattering amplitude is, up to an overall constant, just the Fourier transform of the potential energy with respect to q. This Born approximation can be considered as the first in a series of approximations arising from <m iterative procedure in which the wave function determined by the previous iteration, such as ( 13.44), is then substituted for the exact wave function on the right-hand side of (13.43). A rough estimate of the range of validity of the Born approximation can be made «by noting that since we replace 1j1(r') by 1/linc in (13 .43), we want llfrsc/1/lincl 1 within the range of the potential (where V =f. 0), that is, in the vicinity of the origin. Comparing (13.7) with (I 3.38), we see that w.. f ik!r-r'l '~c ( r ) = _ ___£__ dV e V(r')ljl(r ') (13.46) 2rrh2 lr- r'l Thus since 1/linc(O) =A, flfrsc(O) 1 _ ___£__ cPr' eikr' V(r')eikz' (13.47) 1o/inc(0) = 2rr fi2 r' where we have replaced the exact wave function on the right-hand side of (13.46) with the incident wave function in accord with the Bom approximation. If the =potential energy is spherically symmetric, V (r') V (r'), we can carry out the angular integrals, and the condition for the validity of the Born approximation becomes 1/JscCO) I = ___£__ roo dr' {2;r; d¢' rrr de' vsine' r'eikr' (r')eikr' cos B' 2rrli2 Jo Jo Jo 1 1/JincCO) kr'llo2 «1 =1 : r:x:J dr' eikr' v (r') sin (13.48) hk Page 479 (metric system)
13.3 An Example of the Born Approximation: The Yukawa Potential I 463 At high energies (k--+ oo) , the exponential and the sine in (13.48) oscillate rapidly and cut off the integral for r' .< 1/k. The condition (13.48) becomes in this case (13.49a) or v. « 1 (13.49b) _Q E as we argued earlier. The Born approximation may also be valid at low energies, but under more restrictive conditions (see Problem 13.3). 13.3 An Example of the Born Approximation: The Yukawa Potential Let's evaluate the scattering amplitude for the potential energy e - m or (1 3.50) V(r) = g-- r known as the Yukawa potential. With the appropriate choice for the values of g and m 0 , this potential could represent the short-range potential energy between two nucleons, say a neutron and a proton. Or if we choose g = Z 1Z2e2 and m 0 = 0, the potential reduces to the Coulomb potential energy between a projectile with charge Z 1e and a nucleus with charge Z2e, as in Rutheiford scattering. However, the Coulomb potential does not vanish fast enough for larger to ensure that (13.7) is an asymptotic solution to the Schri:idinger equation in this case, and thus our formalism is not appropriate for a pure Coulomb interaction. Nonetheless, we can consider the factor e- mor as a mathematically convenient way to introduce the screening that actually occurs in Rutherford scattering, where the electrons within the atom shield the incident a particle from the nucleus until the a particle penetrates the electron cloud. Recall that the size of the atom is on the order of an angstrom. while the size of the nucleus is between 104 and 105 times smaller. From (13.45) ff(B, ¢) = _...!!:L o -mor' . (13.51 ) d ·'r' e_ _ _ ezq·r' 2rrti2 r' In order to carry out the integrals, it is convenient to choose our dummy integration e', e'variables so that the z' axis is parallel to q. Then q · r' = qz' = qr' cos where is the usual polar angle in spherical coordinates. Thus In Iaf= - J.J..g. 00 dr' 2n: d¢' 7t df3'sin B1 r1e- mor'eiqr'cos &' (13.52) In2rrh2 o o o Page 480 (metric system)
464 13. Scattering =Figure 13.7 The incident wave vector k1, the scattered wave vector kf, and the vector q k; - kf . Carry ing out the angular integrals first, we obtain (13.53) Note that q2 = (ki - k /)2 += (k2 - 2 k i . k f k2) = =2k2(1 - cos 0) 4k2 s.m2 (-} (13. 54) 2 =and therefore f j(B) only. The angle f) is shown in Fig. 13.7. Thus = =~d 4w?g. ? dQ - lf(O)i2 fi4 fm ~ + 2 /2)f- (13.55) 4k2 sin (0 The differential cross section depends only on the angle (} and not on the angle ¢ because of azimuthal symmetry under rotations about the z axis for a spherical potential. We now specialize to the case of Coulomb scattering. At energies and/or an- m5,»gles such that 4k2 sin2(8/2) the expression ( 13.55) for the differential cross section reduces to do J.L2(z tZ2e2)2 -= dQ fi44k4 sin4(8 / 2) = _(_Zt:_Z___2:::e...,2._)_2 _ ( 13.56) 16£2 sin4(0/2) the famous result for Rutherford scaltering. Interestingly, the dependence on Planck's con...tanL has disappeared entirely. This differential cross section (13.56) is in com- plete agreement with that obtained fi·om a classical analysis of Coulomb scattering, a-. \\\\ ell as with that obtained from an exact solution using quantum mechanics. Without this rather fortunate agreement between the classical and quantum results, Page 481 (metric system)
13.4 The Partial Wave Expansion I 465 the hist01ical development of quantum mechanics would almost certainly have been quite different. It was the agreement between (13.56) and experiment thatled Ruther- ford, before the advent of quantum mechanics, to the nuclear model of the atom. In classical physics, this model could not be stable as the electrons in their classical orbits radiated away their energy and spiraled into the nucleus. Bolu first addressed these problems with his own model, with its stationary states and discrete energies, in the crucial transition between classical and quantum mechanics. 13.4 The Partial Wave Expansion In general, the Born approximation is a high-energy approxi mation for calculating the differential cross section. There is another approach, k.now'il as the partial wave expansion, that is most useful at low energies and is therefore somewhat comple- mentary to the Bom approximation. As we have seen in Rutherford scattering, if the potential energy is spherically symmetric, the scattering amplitude is a function of eonly: J(e. ¢) = f(e) ( 13.57) We begin by writing 00 (13.58) (13.59) f (e) = L (2l + l )az(k) Pt(COS e) 1=0 as a superposition of the partial waves, where v2i+l .~P1(cos e) =Yz o is a Legendre polynomial (see Problem 9.17). In a sense, all we are saying in (13.58) is that we can write any function of e as a superposition of Legendre polynomials. After all, these functions form a complete set. The rationale for expressing the coefficients in the expansion in the form (13.58) will become clear shottly. For now, let us note that the value of the coefficient a1(k) will, in general, depend on the value of the energy, a dependence exhibited explicitly by indicating that the pattial wave is a function of k. Although in principle we have traded j(e) for an infinite set of partial waves, the utility of (13.58) comes from the fact that at low energies only a few of the a1(k) are significat1tly different from zero. To see why, we fi rst give a heuristic semiclassical argument. In general, a beam of particles incident on a target in a scattering experiment with a broad spectrum of impact parameters consists of many different orbital <mgular momenta, as can be seen by evaluating r x p for each impact parameter (see Fig. 13.8). However, if the potential energy has a finite range a, scattering will occur only for those impact parameters that ru·e less than a. Thus for Page 482 (metric system)
466 13. Scattering ----- ----- ----·z T b ...L Figure 13.8 A classical representation of the incident ftux in tenllS of particles follow ing well-defined trajectories. Those particles with impact parameter b possess orbital angular momentum ILl= lr x pi = btik. Only particles with impact parameters less than or equal to the range a of the potential energy would interact with the target. interaction there is a maximum angular momentum, whose value is roughly given by tilmax = ap = a(hk), or !max= ak. The lower the energy and the smaller the value of k, the fewer angular momentum states can interact with the target. To start our partial wave analysis in quantum mechanics, we express the incident plane wave in the fonn (see Problem 13.8) 00 ( 13.60) eikz =eikrcosB= L i ' (2l + l )h(kr)Pt(cos8) 1=0 which can be considered as a special case of the more general expansion L1/1(r) = c1R1(r )Y1,0(e) ('13.61) I Only the Y1,0's enter because the plane wave (I 3.60) is independent of <f>. Since the plane wave is the wave function of a free particle, the radial functions must be spherical Bessel functions . We must discard the spherical Neumann functions in the expansion because they blow up at the oligin, as we discussed in Chapter 10. Clearly, the plane wave is finite at the origin. Note that the appearance of all angular momenta in this expansion is consistent with a picture of a plane wave, which is infinite in extent, as having all impact parameters. What can we say about the asymptotic form of the full wave function (13.7), including the scattered wave, in terms of partial waves? Since we are searching for a olution of the Schri:idinger equation for r --+ oo, a region in which V = 0 but one rhat excludes the origin, we must include both spherical Bessel and Neumann functions in our general expression for the function R1(r) in (13.61 ): (13.62) Page 483 (metric system)
13.4 The Partial Wave Expansion I 467 The asymptotic expressions for the Bessel functions themselves are given by l.t (kr ) -----+ -s-i=n(-k--r---=lr-c-/-2')- rJI (kr ) -----+ - cos(kr - lrc /2) (13.63) r->oo kr r -+ oo kr Substituting these forms into (13.60) and ( 13.62), we obtain ~eikz -----+ i'(2/ +I) sin(kr- Lrcj2) PI(cos()) (13.64) r-.oo L.. kr l=O for the incident wave and [Az B, J1/r(r) -----+ '\"\"\" sin(kr - lrc /2) _ cos(kr - lrc/ 2) P,(cos e) kr kr r -> CO L_. l t= Ct sin[kr - lrc/2 + o,(k)] P (cos e) 1 (13 .65) 1=0 kr for the complete wave function, where in the last step we have combined the sine and cosine into a sine function with its phase shifted by Ot(k). Comparing (13.64) and (13.65), we see that the effect ofthe potential is to introduce a phase shift in the asymptotic wave function. Figure 13.9 shows qualitatively how this happens. We can express (13.65) in the fonn co ei(kr- ln/2+0{) - e- i(kr - /;r j2+0J) 1/r -----+ 'L\" \".\". C t . P1(cos e) (13.66) 2t kr r-+00 1= 0 which contains both incoming and outgoing spherical waves. What is the source of these incoming spherical waves? They must be due to the presence of the incident uu (a) (b) Figure 13.9 A depiction of how the potential energy affects the phase of a wave. (a) A =potential well (an attractive potential) produces a positive phase shift (80 > 0) for the radial function u r R while in (b) a potential barrier (a repulsive potential) generates a negative phase shift (80 < 0). The dashed curve shows u when V = 0 in each case. Page 484 (metric system)
468 13. Scattering plane wave in the full asymptotic wave function. u· we rewrite (13.64) as (13. 67) '' e see these incoming spherical waves explicitly. Since3 ikr 1/1 - eikz---+ f(B):._ r (13.68) r->00 which is an outgoing spherical wave only, the incoming spherical waves must cancel if we sublract ( 13.67) from (13 .66), which implies that (13.69) With this result, we see that (13.70) Thus, comparing (13.58) and ( 13.70), we find (13.71 ) Determining the scattering amplitude through a decomposition into partial waves is equivalent to detennining the phase shift for each of these partial waves. In order to determine the total cross section ( 13.72) I I=a dQ ddQa = dQ lf(B)I?- we take advantage of (13.59) and the orthogonality of the spherical harmonics, namely, (13.73) to do the integral over the solid angle: 4 00 (13.74) a= : L(2/ + 1) sin2 81 k 1=0 3 In this section we have set the amplitude A of the incident plane wave equal to unity for mathematical simplicity. Note that the differential cross section ( 13.24) is independent of A. Page 485 (metric system)
13.5 Examples of Phase-Shift Analysis I 469 Co mparing this result with the expression (13.70) for the scatte ring amplitude, we see that cr = -47Il'm f(O) ( 13.75) k where we have taken advantage of the fact that the Legendre polynomials satisfy P1(1) = 1. Equation ( 13.75) is known as the optical theorem and is a reflection of the fact that, as we discussed earlier, the very existence of scattering requires scattering in the forward direction in order to interfere with the incident wave and reduce the probability current in the forward direction. Finally, it is common to express (13.74) as ( 13.76) where ( 13.77) the lth partial cross section, is the contribution to the total cross section by the /th partial wave. Note that the maximum value for the /th partial cross section occurs when the phase shift 81 = 1r/2. 13.5 Examples of Phase-Shift Analysis HARD-SPHERE SCATTERING We first analyze the scattering from the repulsive potential V (r ) = { oo r< a (13.78) 0 r >a which characterizes a very hard (impenetrable) sphere. Our earlier discussion sug- gests that at sufficiently low energy the l = 0 partial wave dominates the expansion (13.70). Determining the phase shift is plliiicularly easy for S-wave cattering, since =when I 0 the radial equation simplifies considerably with the elimination of the +centrifugal barrier I (l l)li2j2J,ir 2. Outside the sphere, the function u = r R satisfies the free-particle equation -2liJ-2,i dd-2ru2 =Eu r> a (13.79) Rather than write the solution to (13.79) in the form u = B cos kr + C sin kr (or in the form u = Beikr + ce- ikr), we are guided by asymptotic form for the radial wave function in (13.65) to write u = C sin(kr + 80) r> a (13.80) Page 486 (metric system)
470 I 13. Scattering u I ,,- ,' i ~ , // I ', / 1 I I' 1 I I I I I I I I I I 1 I I I I II I \\ / I I\\ I ~----~L-----~\\------~------~------~r \\ \\ \\ \\ \\ \\ \\ \\ \\ __' ' .... ,.,\"'' =Figure 13.10 A plot of the wave function u r R for S- wave scaUering from a hard sphere showing the phase shift 80 = - ka. The dashed curve shows u when V = ~. =where, as usual, k )2t.tE j li.2. Figure 13.10 shows a plot of u . The boundary =condition u(a) 0 determines the S-wave phase shift: C sin(ka + 80) = 0 or oo = -ka (13.81) Thus the S-wave total cross section is given by 4rr .2 ~ 4rr .2 k o0 a Sll1 Stn k2 k.2=0'[= 0 = (13.82) Problem 13.1 0 shows that the higher partial waves, such as the P-wave, can be neglected relative to the S-wave for hard-sphere scattering at low energy. Thus (1.3.83) Notice that the cross section is indeed an area, but in this case the area is four times the classical cross section rra2 that the sphere presents in the form of a disk that blocks the incident plane wave. Of course, low-energy scattering corresponds to a very long wavelength for the incident wave, and thus we should not expect to obtain the classical result. However, even at high energies and very short wavelengths we cannot completely avoid diffraction effects. We give a heuristic argument. At high energies, many partial waves, up to lmax = ka, should contribute to the scattering. Therefore L : +a ~ ka 4 (2l 1) sin2 81 (13.84) k- ka» l 1= 0 Page 487 (metric system)
13.5 Examples of Phase-Shift Analysis I 471 With so many l values contributing, we assume we can replace sin2 81 by its average value, ~, in the sum.4 Then La --+ ka -21( (2l +I) --+ 2rra?- k2 (13.85) ka» I 1=0 ka» I Why do we obtain twice the classical result, even at high energy? For hard-sphere scattering, partial waves with impact parameters less than a must be deflected. How- ever, in order to produce a \"shadow\" behind the sphere, there must be scattering in the forward direction (recall the optical theorem) to produce destructive interference with the incident plane wave. In fact, the interference is not completely destructive and the shadow has a bright spot (known in optics as Poisson's bright spot5) in the forward direction. S-WAVE SCATTERING FOR THE FINITE POTENTIA L WELL As another example of the determination of the phase shift at low energy, we examine scattering from an attractive potential, namely, the spherical well V(r) = - Vo r< a (13 .86) 0 r> a { In terms of the function u = r R, the energy eigenvalue equation becomes (13.87a) (l3 .87b) Equation (13.87a) can be written as (l3 .88a) 4 For a detailed analysis, see J. J. Sakurai and J. Napolitano, Modem Quantum Mechanics, 211d edition, Addison- Wesley, San Francisco, CA, 2011, p. 421. 5 Ironically, Poisson, who supported a corpuscular theory for light, refused to believe Fresnel's prediction that a bright spot would occur in the shadow of an illuminated disk, until it was experimentally verified. Page 488 (metric system)
672 13. Scattering \\\\ hde equation ( 13.87b) is the usual (13.88b) The solution to (13.88a) that satis{ies the boundary condition u(O) = 0 is u = A sin k0r r < a (13.89a) As before, we write the solution outside the well in the form u = C sin(kr + 80) r >a (1 3.89b) ~ allowing explicitly for the appearance of a phase shift. The !inite spherical well is especially nice because we can determine analytically the wave fu nction everywhere, both inside and outside the weJJ.6 Making sure that u is continuous and bas a continuous first derivative at r = a, we obtain A sin koa = C sin(ka + oo) (13.90a) +Ak0 cos k0a = Ck cos(ka 80) ( 13.90b) Dividing these two equations, we obtain tan (ka = -+ k ka (13.91 ) 80) = - tan k0a k0a tan k0a k0 This equation for the S-wave phase shift simplifies considerably at sufficiently low energy, that is, ka --+ 0. Since kaI k{Ja « I, as long as tan k0a is not too large, the right-hand side of (13.91) is much less than one and we can replace the tangent of a small quantity with the quantity itself. Thus +ka 80 ~ ka tan k0a (13.92a) - k0a or (13.92b) 6 For other potential energies for which an analytic solution is not so easy to determine, we can still solve the energy eigenvalue equation by integrating the Schrodinger equation numerically outwards from the origin. Comparison of the numerical solution with the asymptotic fom1 of the radial wave function that appears in ( 13.65) permits a determination of the phase shift. Page 489 (metric system)
13.5 Examples of Phase-Shift Analysis I 473 From (13.77) 2 Since 4rr= = -<Yf= O st.n 2 80 \"' 4rr [ ka (t-an-k0-a - 1) ] k2 2k ~a 2 (13 .93) = 4rra2 (t-an-k0-a - 1) k0a (13.94) then for sufficiently small ka (13 .95) and the total (S-wave) cross section is independent of energy. RESONANCES There is a significant exception to this independence of the cross section on energy. Suppose that the quantity .j2f.lV0a2j fi2 is slightly less than rrj2. Then as the energy increases, k0a, as given in ( I3.94), can reach the value of rr/2. In this case, tan k0a is infinite, and therefore we can no longer assume that the right-hand side of (13.91) is small, even for small ka. In fact, at the value of the energy when k0a = rr/2, t.an(ka + 80) = oo and, consequently, ka + 80 = rr/2, which implies 80 ~ rr/2 since «we are presuming ka 1_7 Thus <Yt=O = 4rr s.m. 2 80 \"=' 4rr = 4na2 ( ?l2) (13.96) 2k 2k k-a Here we see a pronounced dependence of the total cross section on energy. Also notice that the magnitude of the total cross section is much larger than that given in (13.93). Instead of being a small quantity, the phase shift 80 = rr/ 2. What is causing this unusual behavior? If you return to our discussion of the bound states of the finite spherical well in Chapter 10, you will recognize the condition (13.97) 7 Since the phase shift 80 starts at zero, we are assuming that the condition 80 = rr/2 is the first resonance condition that we reach as this phase sh.ift grows. Page 490 (metric system)
474 I 13. Scattering as the condition thatthe well has a bound state at zero energy. Thus for a potential well atisfying (I 3.97), the energy of the scattedng system is essentially the same as the energy of the bound state. In this case, the incident particle in a scattering experiment would like to form a bound state in the well. Since the system has a small but positive energy, the bound state isn't stable. However, pulling in the wave function in an arrempt to form such a bound system dramatically changes the scattering behavior. Although it is more difficult to determine the phase shifts and cross sections in such a nice analytic form as (13.92) for higher partial waves, it is easy to see physically why these higher partial waves may exhibit resonant behavior, with large \"bumps\" in the partial cross sections. These bumps arise when the phase shift goes through odd multiples of rr / 2. Figure 13.11 shows a plot of the effective potential energy Veff (l ·) = V(r ) + L_(;l_+--l-)':t-i-2 (13.98) 2p,r2 for the spherical well. A particle with energy E greater than zero but less than the height of the barrier can tmmel through the banier and form a metastable bound state in the well. This state is metastable (and not stable) because a particle \"trapped\" inside the well can also tunnel out. Thus if the energy of the beam in a scattering experiment is tuned to one of the energies of these metastable states, there is an enhanced tendency for the particle to get stuck in the well . The system then loses track of the mechanism by which the bound state was formed, that is, in particular it may lose track of the direction of the incident beam. When this metastable state decays, it emits the pmticle with the characteristic angular distribution for that particular decay mode. A convenient way to parametrize the behavior of the phase shift in the vicinity of a resonance leads to the famous Breit- Wigner formula. We assume the phase shift 81 of the Lth partial wave goes through rr /2 at an energy £ 0: (13 .99) We next make a Taylor series expansion of cot 81 in the vicinity of the resonant energy: o ocot +1(E0) dcot81) E0) + ··· 1(E) =cot ( dE (E - E=Eo =- ( - . - 12 - -do-1. ) (E - Eo) + ... (13.100) sm 81 dE (13 . 101) E=Eo Defining do1(E)) 2 ( dE E= Eo r Page 491 (metric system)
13.5 Examples of Phase-Shift Analysis I 475 v Veff \\ \"'--.; 'I I EJ ~--~~-----~~~~=-------------------------------- 0~-----+\\------~--~==~~==~~-~- ~-~- -~-~-~-~- ~--~-~-~-~---------- I I I I I \\ \\ \\ \\ \\ \\ ', ', - Vo ! - - - - - - - . . . 1 Figure 13.11 The centrifugal banier combines with the potential well to produce an effective potential that can produce a metastable state, as indicated. lf the energy of the incident beam coincides with the energy of one of these metastable states, a resonance in the scattering cross section can occur. we obtain ocot 1(E) = - -r2 (E - E0) + · ·· (13 . 102) Finally, we can express the function a1(k) [see (13.71)] as eio1 at(k) = -· - sin 81 k 1------ k cot 81 - i \"' 1 1 1 f /2 (13. 103) k - (2/ r) (E - E0) - i k (E - E0) +if'/2 Repeating the steps leading to (13.74), we find that the total cross section for the lth partial wave in the vicinity of a resonance is thus given by r_2 (13.104) (J ::: _4rr_. (2/ .L I) _ _ _ ./:.._4_ __ 1 - k2 ' (E - Eo) 2 +f2j4 As an example, Fig. 1.3.12a shows a strong resonance in rr+-p scattering with a peak at roughly 190 MeV of incident pion kinetic energy. This resonance, known as the !:.(1232) because lhe center-of-mass energy of the resonance at the peak is 1232 MeV, has a full width at half maximum r of 110-120 MeV. Figure 13.12b shows the P-wave phase shift, which reaches rr/2 at the resonance peak. The reso- nance is thus formed in the l = 1 channel. In fact, the intrinsic spin of the !:.(1232) Page 492 (metric system)
476 13. Scattering 150 125 160 \"\"75 ........ 140 <1) 120 \"Y ~ ~ 100 '0 '-\"' :a¢::: 80 60tl) 0 :\"e' ~ 0.. I00 200 300 400 I00 200 300 400 500 600 700 Lab kinetic energy (MeV) Lab kinetic energy (MeV) (a) (b) Figure 13.12 (a) The total cross section for n+- p scattering with pion kinetic energies up to 400 MeV. Adapted from W. R. Frazer, Elementary Particles, Prentice-Hall, Englewood Cliffs, NJ, 1966. (b) The P-wave phase shift for n +-p scattering. Adapted from L. D. Roper, R. M. Wright, and B. T. Fcld, Phys. Rev.l38, Bl90, 1965. f·is j = From (13.1 04), we expect a1=1 = 12n f k2 when E = £0 . However, when we ~dd the orbital angular momentum L= 1 of the pion-proton system to an intrinsic 1 1spins = of the proton, we can from a total angular momentum j = (two states) as well as .i = ~ (four states). Thus of the six total angular momentum states that are generated (presumably incoherently) in the collision, only the four j = ~ states can produce the resonance. If we assume the scattering cross section for the j = ~ states is negligible in comparison with the j = ~ resonance scattering in the vicinity of the peak, then +- p total cross section at the resonance should be ~ of the P-wave peak cross section, that is 8nf k2 , which is about 190 mb, in good agreement with the observed value. Finally, we return to our discussion of S-wave scattering for the finite potential well and ask what happens if we increase the energy of the beam so that the phase shift 80 --+ TC, as illustrated in Fig. 13.13. Then the wave function outside the well, u = C sin(kr + ;r) = -C sin kr (13.105) is the same as the wave function outside the well with zero phase shift [see (13.89b)] up Lo an overall phase. Moreover, si nce sin 80 = 0, the S-wave partial cross section Page 493 (metric system)
13.6 Summary I 477 u Figure 13.13 The wave function u(r) for S-wave scattering for a potential well at an energy such that the phase shift 80 = rr. The dashed curve shows u when V = 0. vanishes (CJz=o = 0). This effect, known as the Ramsauer-Townsend effect,8 is clearly seen in the very low scattering cross section from noble gases at about 0.7 eV. These rare-gas atoms have a potential well with a sharply defined range, and it is possible at low energies to have 80 = rr, with all other phase shifts negligible, leading to essentially petfect transmission of the incident wave. 13.6 Summary The differential scattering cross section is given by (13.106) where the scattering amplitude f(e, ¢ ) determines the angular dependence in the asymptotic form for the solution to the Schrodinger equation: 1/1 ~ Aeikz + Af(e, ikr (1 3.107) ¢):_ r r ~ oo At \"high\" energies, the scattering amplitude can be determined through the Born approximation Jj (e , ¢) = - 2:fi2 d\\' V(r')eiq·r' (1.3.108) 8 Unfortunately, a different Townsend. Page 494 (metric system)
478 I 13. Scattering which is (up to constants) the Fourier transform of the potential energy with respect to the vector q = k; - kf · At \"low\" energies, on the other hand, the scattering amplitude for scattering from a central potential can be determined from the partial wave expansion (13.109) leading to a total cross section o+4 00 a= : L (2l I) sin2 1 (13.110) k 1=0 The phase shifts 81, which are generated when the partjcular partial waves interact with the potential, enter in the asymptotic expression for the wave function ljJ(r) ~f. Ct sin[kr - !;:2 + o1(k)l P,(cos $) (13.111) 1=0 A useful way to determine the phase shifts is to solve the Sehrodinger equation. =either numerically or analytically, for the radial wave function u r R, which obeys the one-dimensional SchrOdinger equation l(l + l)/i2 ?u 2w- =/i2 d2u + + V Eu (13.112) -- - 2 (r)u 2J.L dr and compare the asymptotic form of the solution with (13.111). Problems 13.1. Use the three-dimensional time-dependent Schrodinger equation at- -Ti\\2121j1 + Vljf = t. aljf 2J.L h- to establish that the probability density 1/J*(r, t)1f;(r, t) obeys the local conservation law where . =J .2.J!.!L:_i (.Ifl/r* \\ l w - w\\J·Ir'') . . If/ What would happen to your derivatjon if the potential e nergy V were imaginary? Is probability conserved? Explain. In nonrelativistic quantum mechanics, such an Page 495 (metric system)
Problems I 479 imaginary potential energy can be used, for example, to account for particle absorp- tion in interactions with the nucleus. 13.2. Evaluate the probability current for the scattered wave and show that where u,. is a unit vector in the direction of the radius. 13.3. Show that at low energies (ka ~ 0) the requirement (13.48) for the validity of the Born approximation becomes where V0 is the order of magnitude of the potential energy, a is the range of the potential, and we have neglected constants of order unity. By comparing this result with ( 13.49), argue that if the Born approximation is valid at low energies, it works at high energies as well. 13.4. Using the Born approximation to determine the one-dimensional reflection coefficient R for a potential energy V (x) that vanishes everywhere except in the vicinity of the origin: (a) Show that we can write the solution to the one-dimensional Schrodinger equation in the form Jljt(x) = Aeikx + dx1 G(x, x1) 2tmi V(x ,)1/l(x1) 2 where ax- +32 x1) k?-G(x, xI) = 8(x- xI) - 1 G(x, (b) Since G satisfies a second-order diiTerential equation. G must be a continuous function and, in particular. it must be continuous at x = x 1 By integrating the • differential equation for C from just below to just above x = x 1 show that the , =first derivative of G is discontinuous at x x 1 and that it satisfies (aG) _(ac)OX X=X~ =1 OX X=X~ Page 496 (metric system)
480 I 13. Scattering Then show that one solution for G is given by G= _ 1_ eik(x - x') x > x' X < x' 2ik 1_ 1_ e-ik(x-x') 2ik (c) Substitute this expression for G into the equation for 1/t in (a). Show that in the Born approximation . + ..1/f Aetkx------*. ~oo e2ikx' 2m Ae- tkx dx' ----V(x' ) - 00 2ik ti,2 x-+-oo and that consequently =I· r=1-ooR ~kmh·?- dx' e2ikx'v(x')l2 (d) For the potential bar1ier V(x) = V0 0 < x <a {0 elsewhere the exact reflection coefficient is given by R = 1 - T with =1 T - --:;------ --:;--; : : : : : =: : ; ; = = ==- 1+ [Vcff4E(E - V0)]sin2 j(2mjh2)(E - V0) a «Show that the exact result for R in the limit V0j E 1 agrees with the result of the Born approximation. 13.5. In the initial Rutherford scattering experiment Geiger and Marsden used a particles with an energy of 5 MeV. Choose a reasonable value for m0 and determine the range of angles for which (13.56) should be valid. Suggestion: Write m 0 = lja , where a is a characteristic screening length. How large should a be? Note: Interaction with the electrons in the atom produces a deflection on the order of w-4 radians. 13.6. Use the Born approximation to detemJ.ine the differential cross section for the potential energy \\\\here C is a constant, corresponding to a 1/r3 force. Note: The result depends on h . so it is not the same as the classical .result. 13.7. Use the Born approximation to detennine the differential cross section for the potential energy Page 497 (metric system)
Problems I 481 13.8. Our goal is to establish (13.60), which, with the aid of ( 13.59), can be put in thefonn = LeikrcosB 00 i1j4n:(21 + I) j 1(kr)Y1,0(8) 1=0 (a) Explain why the expansion of the plane wave must be of the form L00 eikr cost! = Ctit(kr)Y,,o(8) l=O (b) Use the fact that )lI -L~n-_ Il,l) (1L. )O- -J(-2I)-! (see Problem 9.18) to show J j[ jc1].1(kr) = --J(21-I)! 1 a a )]-1e1.cf> ( i - - cote- dQ Y* ae a¢i e1.krcosO /,1 f [· 1 = _I_ dQ y* (- 1 ) 1 et t</> sinI e d e t·kr coso] -J(2I)! u d ccos e)1 where the last step follows from the explicit form (9.142) for the raising operator in position space. (c) Use the explicit form (9.146) for Y1,1(e, ¢) to express this result in the form =c Il.l.( k r ) (' k r )I 2/l! f ce+4n: dQ IY ¢)12eikrcos e yf7(2i)l\")i\\!i t . (2/ 1)! I,/ ' (d) Finally, isolate c1 by evaluating this expression as r---+ 0. Hint: For small r, .it(kr) behaves as (kr)1j(2l + 1) !!, where (2l + l)!! = (21 + 1)(2/- 1) · · · 5 · 3·1. 13.9. A pa11icle is scattered by a spherically symmetric potential at sufficiently low energy that the phase shifts 81 = 0 for l > 1(that is, only 80 and 81are nonzero). Show that the differential cross section has the fonn da = A + B cos e+ C ? e - cos- dD. and determine A, B , and C in terms of the phase shifts. Determine the total cross section a in terms of A, B, and C. Page 498 (metric system)
482 I 13. Scattering 13.10. Evaluate the P-wave phase shift 81 for scattering from a hard sphere, for which the potential energy is given by oo r<a V(r) = { 0 r >a Express your result in terms of ) 1(ka) and 711(ka). Use the leading behavior of Jt(p) a nd 17 1(p) for small p to show that8 1 --?- -(ka) 3/ 3 as ka--?- 0, and thus 81 can indeed be neglected in comparison to 80 [sec (13 .81 )] at sufficiently low energy. 13.11. Compare the Born approximation result for the total cross section for scatter- ing from the potential well V(r) = - V0 r <a {0 r >a ~ with that obtained by using S-wave phase shift analysis. Using the condition for the validity of the Born approximation a t low energy (see Problem 13.3), show that the two approaches are in agreement when the Born approximation is valid. 13.12. Consider the spherically symmetric potential energy n- )2~V,(r) = yo~cr - a where y is a constant and o(r - a) is a Dirac delta function that vanishes everywhere except on the spherical surface specified by r = a. (a) Show that the S-wave phase shi ft 80 for scattering from this potential satisfies the equation tan(ka + 80 ) = - +-(yta-fnk)k-ata-nk-a 1 (b) Evaluate the phase shift in the low-energy limit and show that the total cross section for S-wave scattering is 13.13. (a) Determine the differential cross section daf dQ in the Born approximation =for scattering from the potential energy 2~V (r)jn2 yo(r - a) (see Prob- lem 13.12). Show the explicit dependence of daj dQ on(). (b) Evaluate daj dQ in the low-energy limit. Show that the differential cross section is isotropic. What is the total cross section? (c) Use the condition for the validity of the Bom approximation at low energy (see Problem 13.3) to establish that your result in (b) for the total cross section agrees with that given in Problem 13.12 in the appropriate limit. Page 499 (metric system)
Search
Read the Text Version
- 1
- 2
- 3
- 4
- 5
- 6
- 7
- 8
- 9
- 10
- 11
- 12
- 13
- 14
- 15
- 16
- 17
- 18
- 19
- 20
- 21
- 22
- 23
- 24
- 25
- 26
- 27
- 28
- 29
- 30
- 31
- 32
- 33
- 34
- 35
- 36
- 37
- 38
- 39
- 40
- 41
- 42
- 43
- 44
- 45
- 46
- 47
- 48
- 49
- 50
- 51
- 52
- 53
- 54
- 55
- 56
- 57
- 58
- 59
- 60
- 61
- 62
- 63
- 64
- 65
- 66
- 67
- 68
- 69
- 70
- 71
- 72
- 73
- 74
- 75
- 76
- 77
- 78
- 79
- 80
- 81
- 82
- 83
- 84
- 85
- 86
- 87
- 88
- 89
- 90
- 91
- 92
- 93
- 94
- 95
- 96
- 97
- 98
- 99
- 100
- 101
- 102
- 103
- 104
- 105
- 106
- 107
- 108
- 109
- 110
- 111
- 112
- 113
- 114
- 115
- 116
- 117
- 118
- 119
- 120
- 121
- 122
- 123
- 124
- 125
- 126
- 127
- 128
- 129
- 130
- 131
- 132
- 133
- 134
- 135
- 136
- 137
- 138
- 139
- 140
- 141
- 142
- 143
- 144
- 145
- 146
- 147
- 148
- 149
- 150
- 151
- 152
- 153
- 154
- 155
- 156
- 157
- 158
- 159
- 160
- 161
- 162
- 163
- 164
- 165
- 166
- 167
- 168
- 169
- 170
- 171
- 172
- 173
- 174
- 175
- 176
- 177
- 178
- 179
- 180
- 181
- 182
- 183
- 184
- 185
- 186
- 187
- 188
- 189
- 190
- 191
- 192
- 193
- 194
- 195
- 196
- 197
- 198
- 199
- 200
- 201
- 202
- 203
- 204
- 205
- 206
- 207
- 208
- 209
- 210
- 211
- 212
- 213
- 214
- 215
- 216
- 217
- 218
- 219
- 220
- 221
- 222
- 223
- 224
- 225
- 226
- 227
- 228
- 229
- 230
- 231
- 232
- 233
- 234
- 235
- 236
- 237
- 238
- 239
- 240
- 241
- 242
- 243
- 244
- 245
- 246
- 247
- 248
- 249
- 250
- 251
- 252
- 253
- 254
- 255
- 256
- 257
- 258
- 259
- 260
- 261
- 262
- 263
- 264
- 265
- 266
- 267
- 268
- 269
- 270
- 271
- 272
- 273
- 274
- 275
- 276
- 277
- 278
- 279
- 280
- 281
- 282
- 283
- 284
- 285
- 286
- 287
- 288
- 289
- 290
- 291
- 292
- 293
- 294
- 295
- 296
- 297
- 298
- 299
- 300
- 301
- 302
- 303
- 304
- 305
- 306
- 307
- 308
- 309
- 310
- 311
- 312
- 313
- 314
- 315
- 316
- 317
- 318
- 319
- 320
- 321
- 322
- 323
- 324
- 325
- 326
- 327
- 328
- 329
- 330
- 331
- 332
- 333
- 334
- 335
- 336
- 337
- 338
- 339
- 340
- 341
- 342
- 343
- 344
- 345
- 346
- 347
- 348
- 349
- 350
- 351
- 352
- 353
- 354
- 355
- 356
- 357
- 358
- 359
- 360
- 361
- 362
- 363
- 364
- 365
- 366
- 367
- 368
- 369
- 370
- 371
- 372
- 373
- 374
- 375
- 376
- 377
- 378
- 379
- 380
- 381
- 382
- 383
- 384
- 385
- 386
- 387
- 388
- 389
- 390
- 391
- 392
- 393
- 394
- 395
- 396
- 397
- 398
- 399
- 400
- 401
- 402
- 403
- 404
- 405
- 406
- 407
- 408
- 409
- 410
- 411
- 412
- 413
- 414
- 415
- 416
- 417
- 418
- 419
- 420
- 421
- 422
- 423
- 424
- 425
- 426
- 427
- 428
- 429
- 430
- 431
- 432
- 433
- 434
- 435
- 436
- 437
- 438
- 439
- 440
- 441
- 442
- 443
- 444
- 445
- 446
- 447
- 448
- 449
- 450
- 451
- 452
- 453
- 454
- 455
- 456
- 457
- 458
- 459
- 460
- 461
- 462
- 463
- 464
- 465
- 466
- 467
- 468
- 469
- 470
- 471
- 472
- 473
- 474
- 475
- 476
- 477
- 478
- 479
- 480
- 481
- 482
- 483
- 484
- 485
- 486
- 487
- 488
- 489
- 490
- 491
- 492
- 493
- 494
- 495
- 496
- 497
- 498
- 499
- 500
- 501
- 502
- 503
- 504
- 505
- 506
- 507
- 508
- 509
- 510
- 511
- 512
- 513
- 514
- 515
- 516
- 517
- 518
- 519
- 520
- 521
- 522
- 523
- 524
- 525
- 526
- 527
- 528
- 529
- 530
- 531
- 532
- 533
- 534
- 535
- 536
- 537
- 538
- 539
- 540
- 541
- 542
- 543
- 544
- 545
- 546
- 547
- 548
- 549
- 550
- 551
- 552
- 553
- 554
- 555
- 556
- 557
- 558
- 559
- 560
- 561
- 562
- 563
- 564
- 565
- 566
- 567
- 568
- 569
- 570
- 571
- 572
- 573
- 574
- 575
- 576
- 577
- 578
- 579
- 580
- 581
- 582
- 583
- 584
- 585
- 586
- 1 - 50
- 51 - 100
- 101 - 150
- 151 - 200
- 201 - 250
- 251 - 300
- 301 - 350
- 351 - 400
- 401 - 450
- 451 - 500
- 501 - 550
- 551 - 586
Pages: