You are on page 1of 887
Definition of some units Angstrom 1A=10"!%m (order of magnitude of the atomic dimensions) Fermi IF = 107! m (order of magnitude of the nuclear dimensions) Barn 1b = 10-8 m? = (10-4 AY = (10 FP Electron Volt 1 eV = 1,602 189(5) x 107"? joule Useful orders of magnitude Electron rest energy : m,c? = 0.5 MeV [0.511 003(1) x 10% ev] Proton rest energy : M,c? = {000 MeV [938.280(3) x 10° eV] Neutron rest energy : Myc? = | 000 MeV [939.573(3) x 10° ev) One electron volt corresponds to : a frequency vy = 2.4 x 10" Hz through the relation E = Av [2.417 970(7) x 10% Hz] awavelength 2 = 12000A ‘through the relation 4 = c/v [12 398.52(4) A] a wave number i 8.000 em=! [8 065.48(2) cm-*] atemperature T~12000K through the relation E = kg’ (11 604.5(4)K] Ina | gauss magnetic field (10 * Tesla) : the electron cyclotron frequency v, = w,/2x = — gB/2xm, is v, = 2.8 MHz [2.799 225(8) x 106 Hz] the orbital Larmor frequency vy, = «w,/2n = — pipBfh = v2 isv, = L4 MHz [1.399 612(4) x 10 Hz] (this corresponds by definition to a g = 1 Landé factor) it Some general physical constants h= 6626 18(4) x 10-** joule second Planck’s constant he & = 1.054 589(6) x 10° joule second Speed of light (in vacuum) cc = 2.997 924 58(1) x 10° m/s Electron charge 4 = — 1,602 189(5) x 107'* coulomb Electron mass m, = 9.109 53(5) x 1079! kg Proton mass M, = 1.672 65(1) x 10727 kg P ‘Neutron mass M, = 1.674 95(1) x 10-77 kg M, ape = 1 836.1515(7) Ac = Wing: = 2426 309(8) = 107? A Elect Ce jel ‘on Compton wavelength {i = fifmgc = 3861 591(7) x 10-2 A Fine structure constant a a 1 (dimensionless) * = Tnedie he 1370360) Bohr radius ay = & = 0.529177 1(5)A Hydrogen atom ionization energy (without proton recoil — E,, = am,c?/2 = 13.605 80(5) eV effec) Rydberg’s constant Ra, = — Ey, /he = 1.097 373 18(8) x 105 emt ” Classical” electron radius = — 2 = 2817 938(7) fermi Aneym,c? Bohr magneton Hy = Gh/2m, = — 9.274 08(4) x 10-* joule/tesla Electron spin g factor 9, = 2 x 1.001 159 657(4) ‘Nuclear magneton My = — gh/2M, = 5.05082(2) x 10727 joule/testa Boltzmann's constant ky = 1.380.66(4) x 107? joule/K Avogadro's number N, = 6.022.05(3) x 1074 Useful Identities U : scalar field; A,B, ... : vector fields. Vx (VU) =0 v.(VU) = AU V.(Vx A)=0 Vx (Vx A)=V(V.A)—AA A x (B x C) = (A. C)B — (A. BIC Ax (Bx C)+Bx (Cx A) +C x (A x B)=0 (A x B).(C x D) = (A. C)(B.D) — (A. D.C) (A x B)x(C x D) = [(A x B). DJc — [(A x B).C]D = [(C x D). AJB - [(C x D).BJA vy) =uws+vw A (Uv) UAV + (VU). (VV) + VAU Vv. (UA) UV.A+A.VU vx (UA) UVxA+ (WU) A V.(Ax B) =B.(VxA)—A.(VxB) V (A.B) Ax (Vx B)+Bx (Vx A)+BV A+AV B Vx (Ax B)=A (V.B)—B(V.A)+B.V A-A.V B NB. : BVA vector field whose components are = pas Vax, (BV A) = BPA, = 58 i= J, (= PAG We = 2"'¥) + = e yaa] = Av xa) afoa"ve ~ evo 2) 2[02"¥e~ 4 a (ous si Lire vane us ~ 00%F2] = AV A) e"¥2— 22%) = Wea) ‘axpi¥e — 2H + 2 (ous siLaeeye — o2t"¥ 0 use] = IW * A) “re — #oo"¥0) = Wea) Mer 'H = QA 8 - 0 oy yas wy gE 2h apn Boy 248. gg ae ert aie) "i aera ee 25 i2et (42 \eod ve ee 2 eee (yee. fee 93 ae 4 i 3 (6.08 92:02) = aa) ne = Maa) "4salne) + 3 sileeinel = aa) “oein2] = aa) ‘acne + 5 deine = Koa) 2 ota 'y Foun — = ein a gus "y - 900% = 8 500'y +O .us"y ~ = *y (Oya 3 s00'y + ous'y = “r| us 'y + 6 500"7 =v (Cyt ay 7 y+ ey teV= Vv vee + v a ee 8 5 @ean= a Fon = a jeoueyds OPUNAD, uersen1e9 swiaysAs 9}eUIps00D Introduction Structure and level of this text It is hardly necessary to emphasize the importance of quantum mechanics in modern physics and chemistry. Current university programs naturally reflect this importance. In French universities, for example, an essentially qualitative introduction to fundamental quantum mechanical ideas is given in the second year. In the final year of the undergraduate physics program, basic quantum mechanics and its most important applications are studied in detail. This book is the direct result of several years of teaching quantum mechanics in the final year of the undergraduate program, first in two parallel courses at the Faculté des Sciences in Paris and then at the Universités Paris VI and Paris VII. We felt it to be important to mark a clear separation, in the structure of this book, between the two different but complementary aspects (lectures and recitations) of the courses given during this time. This is why we have divided this text into two distinct parts (see the “Directions for Use” at the beginning of the book). On the one hand, the chapters are based on the lectures given in the two courses, which we compared, discussed and expanded before writing the final version. On the other hand, the “complements” grew out of the recitations, exercises and problems given to the students, and reports that some of them prepared. Ideas also came from other courses given under other circumstances or at other levels (particularly in the graduate programs). As we pointed out in the “Directions for Use”, the chapters as a whole constitute, more or less, a course we would envisage teaching to fourth-year college students or those whose level is equivalent. However, the complements are not intended to be treated in a single year. The reader, teacher or student, must choose between them in accordance with his interests, tastes and goals. Throughout the writing of this book, our constant concern has been to address ourselves to students majoring in physics, like those we have taught over the past several years. Except in a few complements, we have not overstepped those limits. In addition, we have endeavored to take into account what we have seen of students’ difficulties in understanding and assimilating quantum 3 mechanics, as well as their questions. We hope, of course, that this book will also be of use to other readers such as graduate students, beginning research workers and secondary schoo! teachers. The reader is not required to be familiar with quantum physics : few of our students were, However, we do think that the quantum mechanics course we propose (see “General approach”, below) should be supplemented by a more descriptive and more experimentally oriented course, in atomic physics for example. General approach We feel that familiarity with quantum mechanics can best be acquired by using it to solve specific problems. We therefore introduce the postulates of quantum mechanics very early (in chapter III), so as to be able to apply thei in the rest of the book. Our teaching experience has shown it to be preferable to introduce all the postulates together in the beginning rather than presenting them in several stages. Similarly, we have chosen to use state spaces and Dirac notation from the very beginning. This avoids the useless repetition which results from presenting the more general bra-ket formalism only after having developed wave mechanics uniquely in terms of wave functions. In addition, a belated change in the notation runs the risk of confusing the student, and casting doubts on concepts which he has only just acquired and not yet completely assimilated. After a chapter of qualitative introduction to quantum mechanical ideas, which uses simple optical analogies to familiarize the reader with these new concepts, we present, in a systematic fashion, the mathematical tools (chapter [) and postulates of quantum mechanics as well as a discussion of their physical content (chapter III), This enables the reader, from the beginning, to have an overall view of the physical consequences of the new postulates. Starting with the complements of chapter III we take up applications, beginning with the simplest ones (two-level systems, the harmonic oscillator, etc.) and becoming gradually more complicated (the hydrogen atom, approximation methods, etc.). Our intention is to provide illustrations of quantum mechanics by taking many examples from different fields such as atomic physics, molecular physics and solid state physics. In these examples we concentrate on the quantum mechanical aspect of the phenomena, often neglecting specific details which are treated in more specialized texts. Whenever possible, the quantum mechanical results are compared with the classical ones in order to help the reader develop his intuition concerning quantum mechanical effects. This essentially deductive viewpoint has led us to avoid stressing the historical introduction of quantum mechanical ideas, that is, the presentation and discussion of experimental facts which force us to reject the classical ideas. We have thus had to forego the inductive approach, which is nevertheless needed if physics is to be faithfully portrayed as a science in continual evolution, provoked by constant confrontation with experimental facts. Such an approach seems to us to be better suited to an atomic physics text or an introductory quantum physics course on a more elementary level. Similarly, we have deliberately avoided any discussion of the philisophical intropucTion implications of quantum mechanics and of alternative interpretations that have been proposed. Such discussions, while very interesting (see section 5 of the biblio- graphy), scem to us to belong on another level. We feel that these questions can be fruitfully considered only after one has mastered the “orthodox” quantum theory whose impressive successes in all fields of physics and chemistry compelled its acceptance. Acknowledgements ‘The teaching experiences out of which this text grew were group efforts, pursued over several years. We want to thank all the members of the various ‘groups and particularly, Jacques Dupont-Roc and Serge Haroche, for their friendly collaboration, for the fruitful discussions we have had in our weekly meetings and for the ideas for problems and exercises which they have suggested. Without their enthusiasm and their valuable help, we would never have been able to undertake and carry out the writing of this book. Nor can we forget what we owe to the physicists who introduced us to research, Alfred Kastler and Jean Brossel for two of us and Maurice Lévy for the third. It was in the context of their laboratories that we discovered the beauty and power of quantum mechanics. Neither have we forgotten the importance to us of the modern physics taught at the C.E.A. by Albert Messiah, Claude Bloch and Anatole Abragam, at a time when graduate studies were not yet incorporated into French university programs. We wish to express our gratitude to Ms. Aucher, Baudrit, Boy, Brodschi, Emo, Heyvaerts, Lemirre, Touzeau for preparation of the manuscript. Foreword This book is essentially a translation of the French edition which appeared at the end of 1973, ‘The text has undergone a certain number of modifications. The most important one is an addition of a detailed bibliography, with suggestions concerning its use appearing at the end of cach chapter or complemeats. This book was originally conceived for French students finishing their undergraduate studies or beginning their research work. It seems to us however that the structure of this book (Separation into chapters and complements — see the “Directions for use”) should make it suitable for other groups of readers. For example, for an undergraduate elementary Quantum Mechanics course, we would recommend using the most important chapters with their simplest complements, For a more advanced course, one could add the remaining chapters and use more 1/4k, and the function to be integrated oscillates several tots within the interval AK. In figure (b,x is fixed such that [x — xo] < I/Ak, and the function tobe Integrated hardly otllats, eo that ite integral over fakes ona relatively large vale. Consequently, the center of the wave packet [point where |y(x, 0)| is maximum) is situated at When x moves away from the value xp, |y(x,0)| decreases. This decrease becomes appreciable if ¢!~*"*-*) oscillates approximately once when & traverses the domain 4k, that is, when: Ak. (x = x9) = 1 (c-17) If Ax is the approximate width of the wave packet, we therefore have : Ak. Ax 21 (c-18) We are thus brought back to a classical relation between the widths of two functions which are Fourier transforms of each other. The important fact is that the product 4x . 4k has a lower bound; the exact value of this bound clearly depends on the precise definition of the widths Ax and Ak A wave packet such as (C-7) thus represents the state of a particle whose probability of presence, at the time ¢ = 0, is practically zero outside an interval of approximate width Ax centered at the value xy. COMMENT: The preceding argument could lead one to believe that the product Ax. Ak is always of the order of I [¢f. (C-17)]. Let us stress the fact that this is a lower limit. Although it is impossible to construct wave packets for which the product Ax . Ak is negligible compared to I, it is perfectly possible to construct packets for which this product is as large as desired [see, for example, complement G,, especially comment (ii) of §3-c]. This is why (C-18) is written in the form of an inequality. C._WAVE PacKETs 3. Heisenberg uncertainty relation In quantum mechanics, inequality (C-18) has extremely important physical consequences. We intend to discuss these now (we shall stay, for simplicity, within the framework of a one-dimensional model). We have seen that a plane wave e'®*—°%" corresponds to a constant proba- bility density for the particle’s presence along the Ox axis, for all values of 1. This result can be roughly expressed by saying that the corresponding value of 4x is infinite. On the other hand, only one angular frequency w, and one wave vector k, are involved. According to the de Broglie relations, this means that the energy and momentum of the particle are well-defined : E = fia, and p = hky. Such a plane wave can, moreover, be considered to be a special case of (C-7), for which g(k) is a « delta function » (appendix 11): ak) = 5(k — ko) (c-19) The corresponding value of Ak is then zero. But this property can also be interpreted in the following manner, using the principle of spectral decomposition (cf. § A-3 and B-2). To say that a particle, described at = 0 by the wave function y(x,0) = 4c, has a well-determined momentum, is to say that a measurement of the momentum at this time will definitely yield p = hk. From this we deduce that e" characterizes the eigenstate corresponding to p = hk. Since there exists a plane wave for every real value of k, the eigenvalues which one can expect to find in a measurement of the momentum on an arbitrary state include all real values. In this case, there is no quantization of the possible results : as in classical mechanics, all values of the momentum are allowed. Now consider formula (C-8). In this formula, (x, 0) appears as a linear superposition of the momentum eigenfunctions in which the coefficient of e!* is a(k). We are thus led to interpret |a(k)|? (to within a constant factor) as the proba- bility of finding p = hk if one measures, at = 0, the momentum of a particle whose state is described by Y(x, £). In reality, the possible values of p, like those of x, form a continuous set, and |g(k)|? is proportional to a probability density: the probability d(x) of obtaining a value between hk and h(k + dk) is, to within a constant factor, |a(k)|? dk. More precisely, if we rewrite formula (C-8) in the form: Wx, 0) fro oh dp (C-20) Vinh we know that ¥(p) and y(x, 0) satisfy the Bessel-Parseval relation (appendix I): Jes oF a [Wor (ca) If the common value of these integrals is C, d%(x) 4 |W(x, 0)}? dx is the probability of the particle being found, at ¢ = 0, between x and x + dx. In the same way : (C22) 27 CHAPTER |_WAVES AND PARTICLES is the probability that the measurement of the momentum will yield a result included between p and p + dp [relation (C-21) then insures that the total probability of finding any value is indeed equal to 1]. Now let us go back to the inequality (C-18) We can write it as: Ax. dp zh (C23) (4p = hak is the width of the curve representing |/(p)|). Consider a particle whose state is defined by the wave packet (C-20), We know that its position probability at ¢ = 0, is appreciable only within a region of width 4x about x9: its position is known within an uncertainty 4x. If one measures the momentum of this particle at the same time, one will find a value between py + 42 and py — 4 since [¥(p)|? 2 is practically zero outside this interval : the uncertainty in the momentum is therefore 4p. The interpretation of relation (C-23) is then the following : it is impossible to define at a given time both the position of the particle and its momen- tum to an arbitrary degree of accuracy. When the lower limit imposed by (C-23) is reached, increasing the accuracy in the position (decreasing 4x) implies that the accuracy in the momentum diminishes (4p increases), and vice versa, This relation is called the Heisenberg uncertainty relation. We know of nothing like this in classical mechanics. The limitation expressed by (C-23) arises from the fact that h is not zero. It is the very small value of h on the macroscopic scale which renders this limitation totally negligible in classical ‘mechanics (an example is discussed in detail in complement B,). COMMENT The inequality (C-18) with which we started is not an inherently quantum ‘mechanical principle. It merely expresses a general property of Fourier transforms, ‘numerous applications of which can be found in classical physics, For example, it is well known from electromagnetic theory that there exists no train of electromagnetic waves for which one can define the position and the wavelength with infinite accuracy at the same time. Quantum mechanics enters in when one associates a wave with a ‘material particle and requires that the wavelength and the momentum satisfy de Broglie’s relation. 4. Time evolution of a wave packet Until now, we have been concerned only with the form of a wave packet at a given instant; in this paragraph, we are going to study its time evolution. Let us return, therefore, to the case of a free particle whose state is described by the one-dimensional wave packet (C-7). A given plane wave e"* ~ “” propagates along the Ox axis with the velocity : Yig=z (C24) since it depends on x and ¢ only through (« =r): lk) is called the phase velocity of the plane wave. 28 We know that in the case of an electromagnetic wave propagating in a vacuum, ¥, is independent of k and equal to the speed of light c, All the waves which make up a'wave packet move at the same velocity, so that the packet as a whole also moves with the same velocity, without changing in shape. On the other hand, we know that this is not true in a dispersive medium, where the phase velocity is given by : Vk) = (C-25) MN) = TH (C25) n(k) being the index of the medium, which varies with the wavelength. The case that we are considering here corresponds to a dispersive medium, since the a a is equal to [¢f. equation (C-3)]: Vfk) = pas (C-26) We shall see that when the different waves thus have unequal phase velocities, the velocity of the maximum x,, of the wave packet is not the average phase velocity 22 Bho, contrary to what one might expect. 0 As we did before, we shall begin by trying to understand qualitatively what happens, before taking a more general point of view. Therefore, let us return to the superposition of three waves considered in § C-2. For arbitrary 1, Y(x, ¢) is given by: SS We see, therefore, that the maximum of |¥(x, ¢)|, which was at x = 0 at ¢ = 0, is now at the point : 42 (C-28) and not at the point x = Re The physical origin of this result appears in figure 6. Part a) of this figure represents the positon at time 1 = 0 of three adjacent maxima (1), (2), (3) for the real parts of each of the three waves. Since the maxima denoted by the index (2) coincide at x = 0, there is constructive interference at this point, which thus corresponds to the position of the maximum of |Y(x, 0)|. Since the phase velocity increases with k [formula (C-26)], the maximum (3) of the wave + 4) will gradually catch up with that of the wave (kq), which will in turn eatch up with that of the wave(b - 4) After a certain time, we shall thus have the situation shown in figure 6-0; it will be the maxima (3) which coincide and 2» thus determine the position of the maximum xy(/) of (x, 1)|. We clearly see in the figure that xy(t) is not equal to 227, and a simple calculation again Ko yields (C-28). jk ko +f lo ly by ly ly le ke ky le ley lo le ley Ak b- Thy le ley lo lo ley es 1 fl : 0) i” alt) nouns 6 Positions of the maxima of the three waves of figure 4 at time 1 = 0 (fig. a) and at a subsequent 1 (ig. b). At time 1 = 0, ie isthe maxima (2), situated at x = 0, which interfere constructively : the position of the center of the wave packet is xy( {At time 1, the three waves have advanced ‘with diferent phase velocities ¥,. It ts then the maxima (3) whleh interfere constructively and the center of the wave packet is situated at x = xy(t), We thas see thatthe velocity ofthe center ofthe ‘wave packet (group velocity) is different frou the pise velocitis of the three waves. The shift of the center of the wave packet (C-7) can be found in an analogous fashion, by applying the “stationary phase” method. It can be seen from the form (C-7) of the free wave packet that, in order to go from W(x, 0) to v(x, ) all we need to do is change g(k) to g(k)e~". The reasoning of § C-2 thus remains valid, on the condition that we replace the argument a(k) of g(k) by: alk) = okt (C-29) The condition (C-16) then gives : éo-[42] «-[8] a We are thus brought back to result (C-28) : the velocity of the maximum of the wave packet is: vats = [$2] (cn Ve (ko) is called the group velocity of the wave packet. With the dispersion relation given in (C-3), we obtain: Velo) = To = 2g thy) (ca 30 This result is important, for it enables us to retrieve the classical description of the free particle, for the cases where this description is valid. For example, when one is dealing with a macroscopic particle (and the example of the dust particle discussed in complement B, shows how small it can be), the uncertainty relation does not introduce an observable limit on the accuracy with which its position and momen- tum are known. This means that we can construct, in order to describe such a particle in a quantum mechanical way, a wave packet whose characteristic widths 4x and dp are negligible. We would then speak, in classical terms, of the position xy(t) and the momentum p, of the particle, But then its velocity must be v = 22, This is indeed what is implied by formula (C-32), obtained in the quantum description : in the cases where 4x and 4p can both be made negligible, the maximum of the wave packet moves like a particle which obeys the laws of classical mechanics. COMMENT We have stressed here the motion of the center of the free wave packet. Ivis also possible to study the way in which its form evolves in time. TL is then easy to show that, if the width 4p is a constant of the motion, 4x varies over time and, for sufficiently long times, increases without limit (spreading of the wave packet). The discussion of this phenomenon is given in complement G,, where the special case of a Gaussian wave packet is treated, D. PARTICLE IN A TIME-INDEPENDENT SCALAR POTENTIAL We have seen, in § C, how the quantum mechanical description of a particle reduces to the classical description when Planck's constant h can be considered to be negligible. In the classical approximation, the wavelike character does not appear because the wavelengih i =" assuciated with the particle is much smaller than the characteristic lengths of its motion. This situation is analogous to the one encoun- tered in optics. Geometrical optics, which ignores the wavelike properties of light, constitutes a good approximation when the corresponding wavelength can be neglected compared to the lengths with which one is concerned. Classical mechanics thus plays, with respect to quantum mechanics, the same role played by geometrical optics with respect to wave optics. In this paragraph, we are going to be concerned with a particle in a time- independent potential, What we have just said implies that typically quantum effects (that is, those of wave origin) should arise when the potential varies appreciably over distances shorter than the wavelength, which cannot then be neglected. This is why we are going to study the behavior of a quantum particle placed in various “square potentials”, that is, “step potentials”, as shown in figure 7-a. Such a potential, which is discontinuous. clearly varies considerably over, intervals of the order of the wavelength, however small it is: quantum effects must therefore always appear. Before beginning this investigation, we shall discuss some important properties of the Schrddinger equation when the potential is not time- dependent. 3 1. Sepa jon of variables. Stationary states The wave function of a particle whose potential energy V(r) is not time- dependent must satisfy the Schrédinger equation : 2 yey = —© ayn + VO We) 1 ae 2m * ° 9, EXISTENCE OF STATIONARY STATES Let us see if there exist solutions of this equation of the form: Yee.) = of) 2) 2) Substituting (D-2) into (D-1), we obtain a ot 2 ~ x0] - F aot] + xOv 919 03 If we divide both sides by the product y(r)x(0), we find : ih dy 1 a Fe = sta - Fao] + v0 4) This equation equates a function of ¢ only (left-hand side) and a function of r only (right-hand side). This equality is only possible if each of these functions 1s in fact a constant, which we shall set equal to fiw, where « has the dimensions of an angular frequency. Setting the left-hand side equal to Aw, we obtain for y(t) a differential equation which can easily be integrated to give: wi) = Ae 5) In the same way, g(r) must satisfy the equation : —E soe + vie oh ho o(r) (D-6) If we set 4 = 1 in equation (D-5) [which is possible if we incorporate, for example, the constant 4 in g(r)], we achieve the following result: the function Wr, D = giryen'" (0-7) is a solution of the Schrédinger equation, on the condition that g(r) is a solution of (D-6). The time and space variables are said to have been separated. A wave function of the form (D-7) is called a stafionary solution of the Schrédinger equation: it leads to a time-independent probability den- sity Wr, )? = jo(r)|’. In a stationary function, only one angular frequency w appears; according to the Planck-Einstein relations, a stationary state is a state with a well-defined energy E — hw (energy eigenstate). In classical mechanics, when the potential energy is time-independent, the total energy is a constant of the motion ; in quantum mechanics, there exist well-determined energy states. 32 D._TIME-INDEPENDENT SCALAK POTENTIAL Equation (D-6) can therefore be written : [-F 4+ ve foe Eg) (D-8) or: Hol) ~ Ee (@» where H is the differential operator : Ps 1 py tvn (D-10) H H isa linear operator since, if 4, and A, are constants, we have: HAP) + 4,910] = 4.1 9@) + AHO) (Dy Equation (D-9) is thus the eigenvalue equation of the linear operator H: the application of H to the « eigenfunction » g(r) yields the same function, multiplied by the corresponding «eigenvalue» E. The allowed energies are therefore the eigenvalues of the operator H. We skall see later that equation (D-9) has square- integrable solutions g(r) only for certain values of E (ef. § D-2-c and § 2-c of complement H,): this is the origin of energy quantization, COMMENT Equation (D-8) [or (D-9)] is sometimes called the “time-independent Schrodinger equation”, as opposed to the “time-dependent Schrdinger equation” (D-1), We stress their essential difference : equation (D-1) is a general equation which gives the evotution of the wave function, whatever the state of the particle; on the other hand, the eigenvalue equation (D-9) enables us to find, amongst all the possible states of the particle, those which are stationary. b. — SUPERPOSITION OF STATIONARY STATES In order to distinguish between the various possible values of the energy E (and the corresponding eigenfunctions (r)), we label them with an index n Thus we have : He,t) = E,9,(t) (D-12) and the stationary states of the particle have as wave functions : Walls 1) = Gyr) em" (D-13) ¥,(r, 1) is a solution of the Schrédinger equation (D-1), Since this equation is linear, it has a whole series of other solutions of the form Wr, t) = Yc, gale) ea (D-14) 3B CHAPTER | WAVES AND PARTICLES where the coefficients c, are arbitrary complex constants. In particular, we have : VE, 0) = Yc, a(t) (D5) Inversely, assume that we know (r,0), that is, the state of the particle at 1 — 0, We shall see later that any function ¥(r, 0) can always be decomposed in terms of eigenfunctions of H, as in (D-15). The coefficients c, are therefore determined by ¥(F,0). The corresponding solution Yr, 1) of le Schrédinger equation is then given by (D-14). All we need to do to obtain it is to multiply each term of (D-15) by the factor e~"", where E, is the eigenvalue associated with 9,(r). We stress the fact that these phase factors differ from one term to another. It is only in the case of stationary states that the ‘-dependence involves only one exponential [formula (D-13)]. 2. One-dimensional “square” potentials. Qualitative study We said at the beginning of § D that in order to display quantum effects we were going to consider potentials which varied considerably over small distances. We shall limit ourselves here to a qualitative study, so as to concentrate on the simple physical ideas. A more detailed study is presented in the complements of this chapter (complement H,). To simplify the problem, we shall consider a one-dimensional model, in which the potential energy depends only on x (the justification for such a model is given in complement Fy) 2. PHYSICAL MEANING OF A SQUARE POTENTIAL We shall consider a one-dimensional problem with a potential of the type shown in figure 7-a, The Ox axis is divided into a certain number of constant- potential regions. At the border of two adjacent regions the potential makes an abrupt jump (discontinuity). Actually, such a function cannot really represent @ physical potential, which must be continuous. We shall use it to represent schema- tically a potential energy V(x) which actually has the shape shown in figure 7-b : there are no discontinuities, but V(x) varies very rapidly in the neighborhood of certain values of x. When the intervals over which these variations occur are much smaller than all other distances involved in the problem (in particular, the wave- length associated with the particle), we can replace the true potential by the square potential of figure 7-0. This is an approximation, which would cease to be valid, for example, for a particle having too high an energy, whose wavelength would be very short. The predictions of classical mechanics concerning the behavior of a particle in a potential such as that of figure 7 are easy to determine. For example, imagine that V(x) is the gravitational potential energy. Figure 7-b then represents the real profile of the terrain on which the particle moves : the corresponding disconti- nuities are sharp slopes, separated by horizontal platcaus. Notice that, if we fix the total energy F of the particle, the domains of the Ox axis where V > E are forbidden to it (its kinetic energy £, = E — V must be positive). 34 “Square” —— potential . — Real potential » | ° , = rovne 7 _ Stuare potatial (Bp. 2) which Senta repre 2 real : x ental (By. b) for which he Sarasa fie COMMENT: a) ang The force exerted on the particle is F(x) = - 440. tn figure 7, we have depicted this foree, obtained from the potential V(x) of figure 7-b. It ean be seen that this particle, in all the regions where the potential is constant, is not subject to any force. Its velocity is then constant. It is only in the frontier zones between these plateaus that a force acts on the particle and, depending on the case, accelerates it or slows it down, b. OPTICAL ANALOGY We are going to consider the stationary states (§ D-1) of a particle in a one- dimensional “square” potential. In a region where the potential has a constant value V. the eigenvalue equation (D-9) is written: wd 7 i [- rey v |) = Eolx) (D416) [es +E 7 | glx) = 0 (D-17) Now, in optics, there exists a completely analogous equation. Consider a transparent medium whose index n depends neither on © nor on time. In this medium, there can be electromagnetic waves whose electric field E(r, r) is inde- pendent of y and z and has the form: D = ee ™ (D-18) 35 CHAPTER | WAVES AND PARTICLES where e is a unit vector perpendicular to Ox, E(x) must then satisfy : @ nQ? s+ (x) = 0 (D-19) (ss c }a ! We see that equations (D-17) and (D-19) become identical if we set : 292 Img — yy = 28 (D-20) Moreover, at a point x where the potential energy [and, consequently, the index n given by (D-20)] is discontinuous, the boundary conditions for (x) and E(x) are the same : these two functions, as well as their first derivatives, must remain continuous (cf. complement H,, § I-h). The structural analogy hetween the two equations (D-17) and (D-19) thus enables us to associate with a quantum mechanical problem, corresponding to the potential of figure 7-a, an optical problem : the propagation of an electromagnetic wave of angular frequency @ in ‘a medium whose index n has discontinuities of the same type. According to (D-20), the relation between the optical and mechanical parameters is: MQ) = FeV ImcX{E ~ V) (B21) For the light wave, a region where E > V corresponds to a transparent medium whose index is real. The wave is then of the form e*, ‘What happens when V > £? Formula (D-20) gives a pure imaginary index. In (D-19), n? is negative and the solution is of the form e~**: it is the analogue of an “evanescent wave”. Certain aspects of the situation recall the propagation of an electromagnetic wave in a metallic medium. Thus we can transpose the well-known results of wave optics to the problems which we are studying here. It is important, however, to realize that this is merely an analogy. The interpretation that we give for the wave function is fundamentally different from that which classical wave optics attributes to the electromagnetic wave. 2. EXAMPLES. & — Potential step and barrier Consider a particle of energy E which, coming from the region of negative x. arrives at the potential “step” of height Yo shown in figure 8. If E > Vo, (the case in which the classical particle clears the potential step and continues towards the right with a smaller velocity), the optical analogy is the following : a light wave propagates from left to right in a medium of index n, : ¢ Vines (D-22) * This analogy should not be pushed too far, since the index n of a metallic medium has both a real and a complex part (in a metal, an optical wave continues to oscillate as i damps out). 36

You might also like