Chapter 12

The Spin

12.1  INTRODUCTION

In Chapter 10, we solved Schrodinger equation for an electron in hydrogen atom and found discrete values of energy and orbital angular momentum. When an electron makes transition from higher energy state with energy Em to a lower energy state with energy En, allowed by selection rules (to be discussed in a later chapter), the emitted radiation has wavelength λ = hc /(Em − En). When the predicted spectrum was compared with the observed spectral lines, there was agreement in almost whole spectrum except in finer details of (some) lines. In fact, the theoretical spectrum cannot account for a few additional spectral lines seen experimentally. Most important of these is the fact that many of the observed spectral lines, instead of single lines predicted by Schrodinger equation, consist of two closely spaced lines. For example, this so-called fine-structure is seen in the first line of Balmer series of hydrogen atom (corresponding to the transition from n = 3 to n = 2 level). For this particular case, whereas the theoretical prediction is a single line of wavelength λ = 6563Å, the experimental spectrum, instead, shows two lines separated by ∆λ = 1.4Å. Though it is a small effect, it cries for the explanation.

We may cite another example of the discrepancy between the theoretical results and the corresponding experimental observations. This is seen in the Zeeman Effect. We shall discuss the details of Zeeman Effect in Chapter 15. Suffice here to mention that when an atom is put in a constant homogeneous magnetic field, the interaction of orbital magnetic moment and external magnetic field results in splitting of energy levels and hence in splitting of a spectral line into three lines, termed as ‘Normal Zeeman Effect’. On the other hand, what is generally seen is a single line splitting into four, six, or even more lines in the presence of magnetic field, termed as ‘Anomalous Zeeman Effect’.

Both the above mentioned discrepancies may be sorted perfectly well, if we assume, along with Goudsmit and Uhlenbeck, that the electron has some internal degrees of freedom and as a result has additional angular momentum and magnetic moment along with its orbital angular momentum and orbital magnetic moment. Goudsmit and Uhlenbeck imagined the electron as a classical charged spherical ball spinning around its axis. Though this picture of spinning electron may not be true, as it leads the electron to have angular speed so high that the equatorial velocity of the imagined spherical ball is many times larger than the velocity of light, still it is termed as spin angular momentum and spin magnetic moments, only for historical reasons. We really do not know what type of internal motion in electron causes the spin angular momentum and spin magnetic moment. Not only the experimental results cried for the presence of spin angular momentum and spin magnetic moment, the consistency of relativistic quantum mechanics, developed by Dirac, confirmed the existence of electron spin as a fundamental requirement. In this chapter, we shall have preliminary discussions about the quantum mechanical description of spin, spin operators, and spin dynamics, and so on. Firstly, we start with orbital angular momentum and magnetic moment.

12.2  ORBITAL ANGULAR MOMENTUM AND MAGNETIC MOMENT

Let us consider an electron, as a classical particle, moving in a circular orbit of radius r, with speed v. Let us take the orbit in x−y plane (at z = 0) with its centre at the origin 0, as shown in Figure 12.1.

Figure 12.1

Figure 12.1 Electron moving in a circular orbit of radius r in x-y plane

The angular momentum vector L points in z-direction (Lx = Ly = 0) and has magnitude

 

|L| =Lz = mevr        (12.1)

 

Now the rotating electron is equivalent to an electric current I which is given by

 

I = charge × number of revolutions in a unit time
equation

in the direction of electron motion. We know that a current I in a small loop of area A is equivalent to a magnet of dipole moment μ, perpendicular to the plane of the loop (in a sense given by the right handed screw rule with respect to sense of current) given by

 

|μ| = μz = IA
equation
μx = μy= 0

For the general case, we may write

equation

Here the subscript l signifies that the magnetic moment is present due to orbital motion. It is clear from Eq. (12.4) that as the electron carries negative charge (–e), the orbital magnetic moment vector μl points in opposite direction to that of the orbital angular momentum vector L.

Now switching over to quantum mechanical case, it is interesting to note that the same relation [Eq. (12.4)] holds even when the electron is described quantum mechanically. But we know from Chapter 9 that in quantum mechanical description of angular momentum, not only the magnitude of angular momentum vector L is quantized equation, the z-component of L is also quantized with values Lx = ml ħ [m1 having (2l + 1) values –l, –l + 1,... l]. Consequently, the z-component may have any one of the (2l + 1) values given by

equation
= − gl μBml        (12.5b)

where μB = (eħ/2me) (= 9.27 × 10–24JT–1) is known as Bohr magneton and gl is known as Lande g-factor which for orbital case is unity.

12.3  THE ELECTRON SPIN: SPIN OPERATORS AND SPIN EIGENSTATES

As mentioned in the Introduction, there are experimental evidences (e.g., fine structure and Zeeman effect) and theoretical compulsions (in Dirac’s relativities quantum mechanical theory of electrons) to assume that electrons, in addition to their orbital angular momentum and orbital magnetic moment, have intrinsic spin angular momentum and spin magnetic moment. Further, the results of Stern–Gerlach experiment (which shall be discussed in next section) reinforce in the same direction and help in drawing conclusions about spin angular momentum and spin magnetic moment.

So, based on various facts, for example, (i) results of Section 11.3 with eigenvalues of Jz = mjħ (mj = –j, –j + 1, ...,j) and allowing j to have values equation – odd integer alongwith integer values, (ii) Uhlenbeck and Goudsmit suggestion of equation explaining experimental results, (iii) Dirac’s relativistic quantum mechanical theory of electrons requiring the presence of spin angular momentum equation for consistency, and (iv) Stern–Gerlach experimental results [explainable only if we include spin magnetic moment μz= − gs(e/2me)Sz], we have the following expressions of spin angular momentum and spin magnetic moment.

 

Sz = msħ        (12.6a)

 

ms = −s, s        (12.6b)
equation
equation

where qs (≈ 2.0) is known as Lande g-factor for spin.

In fact, we had already obtained the operator form of angular momentum components for the case of quantum number j = equation (in Section 11.7). This case really corresponds to spin angular momentum of electron. We start writing s and ms in place of j and mj (s is called as spin quantum number and ms as spin magnetic quantum number). The spin operators Ŝx, Ŝy, Ŝz, are written as

equation
equation
equation

where

equation
equation
equation

are the famous Pauli spin matrices. The eigenstates and eigenvalues of the spin angular momentum component operator Ŝz are easily obtained as

equation
equation

So the operator Ŝz represented by a 2 × 2 matrix has eigenstates equation and equation with corresponding eigenvalues equation and − equation. Let us call these states as equation and equation, respectively.

equation
equation

Thus equation represents a state with ‘spin up’ and equation represents a state with ‘spin down’. Let us write a general spin state as equation

equation

Hence the matrix elements C1 and C2 are the complex amplitudes of ‘spin up’ and ‘spin down’ states, respectively. The column matrix in Eq. (12.11) is generally referred to as a spinor and, in fact, represents a state vector in an abstract two-dimensional complex linear vector space. The state vectors equation and equation are the two basis vectors, which form the complete orthonormal set, in this two-dimensional linear vector space. If the spin state of an electron is given by the spinor equation of Eq. (12.11), then the probability of finding the electron with spin up is |C1|2 and that of finding it with spin down is |C2|2. Obviously the normalization of state equation means

equation

Let us rewrite the spin operator components,

equation
equation
equation
equation
equation
equation
equation

Using notations of Eq(12.10) for up-spin state equation and down-spin state equation, we write below, how the components of Ŝ act on equation and equation

equation
equation
equation
equation
equation
equation
equation
equation
equation
equation

where equation is a general spin state. Above equations may be re-written in the notations of equation and equation for spin-up and spin-down states as

equation
equation
equation
equation
equation
equation
equation
equation
equation
equation

Let us now explore the algebra of spin equation operators in details. From Eqs (12.8) we have

equation
equation

from these, we get

equation

and

equation

Using Eq. (12.7), we get from Eqs (12.18)(12.20)

 

Ŝx Ŝy = Ŝz        (12.21)

 

Ŝx Ŝy + Ŝy Ŝx = 0        (12.22)

 

x, Ŝy] = 2 Ŝz        (12.23)

 

These equations are, in fact, true for any components of spin operator, so we may write

 

Ŝi Ŝj = Ŝk        (12.24)

 

Ŝi Ŝj + Ŝj Ŝi = 0        (12.25)

 

i, Ŝj] = 2 Ŝk        (12.26)

 

where i, j, k = x, y, or z, in cyclic order. These relations are similar to those obtained for the orbital angular momentum operator components.

12.3.1  Physical Significance of above Results

We have seen results of eigenstates and eigenvalues of various spin angular momentum operators. For example, the spin operator Ŝz has eigenvalues msħ, that is, ħ/2 and – ħ/2 with corresponding spin-up and spin-down states equation and equation [that means, if an experiment (say Stern–Gerlach experiment, discussed in Section 12.6) is performed to measure z-component of spin of an electron, the result would be either ħ/2 or –ħ/2]. Also we have eigenvalues of Ŝ2 as (3/4)ħ2 in both the above mentioned states. So the magnitude of the spin angular momentum vector S is equation. We show schematically in Figure 12.2a, the z-component Sz and magnitude S of spin angular momentum vector. These two quantities are related through

 

Sz = S cosθ        (12.27)

 

giving the value of cosθ as

equation

So, when we say an electron is in spin-up state, its spin vector is making an angle equation with z-axis. To maintain this angle θ with z-axis, the vector S may point in any direction along the surface of the cone shown in Figure 12.2a. Similarly in spin-down state, the vector S makes same angle θ with negative z-axis and may lie in any direction along the surface of the lower cone in Figure 12.2a. This simply means that as and when Sz is exactly known, the Sx and Sy components are totally uncertain. This is consistent with the uncertainty relation which states that two quantities corresponding to two non-commuting operators cannot be determined accurately simultaneously. And we have seen that no two components of S commute. Figure 12.2b shows a similar situation for spin-up and spin-down states with respect to y-axis.

Figure 12.2

Figure 12.2 (a) The electron spin in spin-up and spin-down states with respect to z-direction points along the surface of upper (lower) cone (b) Spin-up and spin-down states with respect to y-direction

12.4  TOTAL WAVE FUNCTION OF AN ELECTRON

We know that the wave function of an electron, in a three-dimensional space, is characterized by three quantum numbers nx, ny, nz (or n, l, ml) and is a function of three continuous space coordinates x, y, z (or r, θ, ϕ). The wave function is generally denoted by ψnx, ny, nz (x, y, z) or ψn, l, ml (r, θ, ϕ) This is all about the space part of the wave function. Now the electron, whether it is moving in an atom, in a solid or in field free space, has spin angular momentum. We have learnt in this chapter that the z-component of spin angular momentum. Sz,. has two values equation and − equation, corresponding to two different values of spin quantum number ms (equation and −equation). So, if we want to write the spin wave function of the electron at the same footing as its space wave function, the spin wave function should be a function of spin coordinates (say ξ) and be characterized by the spin quantum number ms. With this notion, let us write the spin wave function as χms (ξ).

Now, for two different values Of ms, the two wave functions χ1/2(ξ) and χ(1/2)(ξ) correspond, respectively to spin-up and spin-down states. Let us denote χ1/2(ξ) by α(ξ) and χ(1/2)(ξ) by β(ξ)· So, when an electron is in up spin state α(ξ), the spin operator Ŝz has eigenvalues equation while in down spin state β(ξ), Ŝz has eigenvalue − equation. At this stage, we really do not know the forms of α(ξ) and β(ξ) and also do not know what values the spin coordinate ξ may take in spin-space. And, therefore, if we want to apply orthonormality conditions on α(ξ) and β(ξ), that is,

equation

and

equation

we do not know, how to integrate over ξ. To overcome the difficulty we take help from the eigen-states of spin operator Ŝz. The operator Ŝz has eigenvalue equation in eigenstate equation and equation in eigenstate equation. Therefore, we conclude that the state equation is nothing but the up spin state α(ξ) and equation is same as β(ξ). If we assign only two values +1 and –1 to the spin coordinate ξ and if we permit the spin functions to have only two values 0 and 1, such that

equation

we see the above mentioned correspondence is fulfilled. Also the integrals (12.29) for orthonormality conditions may be evaluated.

equation

and

equation

The various notations for (orthonormal) spin–eigenstates along z-direction are summarized in Table 12.1.

Table 12.1 Various Notations of Spin Table 12.1 Eigenstate

Spin-up state Spin-down state
equation
equation
χ+
χ
α(ξ)
β(ξ)
equation
equation
equation
equation
equation
equation
equation
equation

Table 12.2 Spinor and Ket Notations

Description of state Spinor notation Ket notation

Spin-up state along z-direction

equation
equation

Spin down (z-direction)

equation
equation

Spin up (x-direction)

equation
equation

Spin down (x-direction)

equation
equation

Spin up (y-direction)

equation
equation

Spin down (y-direction)

equation
equation

Also in Table 12.2, we show various notations for spin-up and spin-down states along x-, y-, and z-directions.

Now, we turn to writing the total wave function of an electron including its space part as well as spin part. If we assume that the Hamiltonian of the electron does not contain a term, involving spin angular momentum operator Ŝ (which is the case when, for example, spin-orbit interaction is neglected in atomic case), the total wave function of electron may be written as the product of the space—as well as the spin wave functions,

 

ψnx, ny, nz, ms (x, y, z, ξ) = ψnx, ny, nz (x, y, z) χms(ξ)        (12.32)

As discussed above, the spin wave function χms, for two internal degrees of freedom (i.e., for ms = + equation and − equation) is represented by the spinors equation and equation, respectively. So we need two wave functions to describe the total state of an electron at a space-point (x, y, z). We may write these two wave functions in column matrix form as

equation

This is known as the Pauli wave function (designated by P). The spinor wave function ψΡ may also be written as a linear combination of basis spinors equation and equation.

equation

Now, normalization of wave function ψΡ gives

equation
equation

As we have

equation

So

equation

The first term in above equation represents the probability of the electron in spin-up state equation i.e., in state equation equation and the second term represents the probability for spin-down state.

If Ĥ represents the Hamiltonian of the electron, its equation of motion may be written as

equation

This is generally called as Pauli equation of motion. Now, just like the spin-dependent (total) wave function of an electron has been written in spinor (i.e., column matrix) form [Eq. (12.33)], the spin-dependent operator may be written in a 2 × 2 matrix form. For example, let us consider spin-orbit interaction part of an electron in a hydrogen-like atom (see Section 15.6) which is given as

equation

where

equation

So

equation

This is the Pauli operator form of Hamiltonian equation.

12.5  THE STERN-GERLACH EXPERIMENT

In 1922 Stern and Gerlach demonstrated in their experiment that the magnetic moment (and the angular momentum) of an electron in an atom has certain discrete orientations. Thus the results of Schrodinger equation of an electron in an atom that the angular momentum component (say) Lz has only certain discrete values ml ħ (or magnetic moment component μz has values –μΒ ml) were verified.

Before discussing the Stern–Gerlach experiment, let us consider a magnet in a constant magnetic field B. The potential energy of interaction is

 

V =μ · B        (12.42)

 

The magnet experiences a torque T, given by

 

Τ = μ × Β        (12.43)

 

The magnet will also experience a force F given by

 

F = −∇V        (12.44)

 

Combining Eqs (12.42) and (12.44), we get

 

F =∇(μ · B)        (12.45)

 

Now for a constant uniform field ∇B = 0 and, therefore, there is no net force on the magnetic moment. This case is shown in Figure 12.3a for a (classical) bar magnet, where equal and oppose forces work on the two poles of the magnet. So there is a finite torque but not net force on the magnet. For a non-uniform field

 

B = B ez
Figure 12.3

Figure 12.3 A classical magnet: (a) in a uniform magnetic field, (b) in a non-uniform magnetic field

B is non-zero and therefore there is a net force working on the magnetic moment. The case is shown in Figure 12.3b for a bar magnet put in a non-uniform field. The forces at the two magnetic poles of the magnet are of different magnitudes and, therefore, there is a net force on the magnet alongwith a finite torque.

The set up used in Stern–Gerlach experiment is shown in schematic form in Figure 12.4. The first experiment was performed using silver atoms. A piece of silver metal is vapourized in an oven. Silver atoms emerging from a small hole, and collimated by slits enter the magnetic field region in the form of narrow atomic beam. This beam passes between the two poles of a magnet which produce non-uniform magnetic field, as shown in Figure 12.4. The beam is finally detected on the plate P.

From Eq. (12.45) we have

 

F = μxΒx + μyBy + μzΒz        (12.46)

 

With the shape of the poles as in Figure 12.4, we have By = 0. Moreover the magnetic field is predominantly in z-direction, so we neglect the component Bx also. Therefore

 

F = μzBz        (12.47)

 

Now the magnet is symmetrical about the y–z plane and the atomic beam is confined in this plane, therefore, (∂Bz/∂x) = 0. Also, if we neglect edge effects, (∂Bz/∂y) = 0. Therefore, Eq. (12.47) may be written as

equation

As shown in Figure 12.4 the atoms travel a distance l along y-axis while a force F is working on these in z-direction. As a result, the displacement in z-direction, ∆z is given by

equation

Atoms of mass M are taken to enter the field with velocity v in y-direction. After crossing the field and with displacement ∆z in z-direction, the atoms move in straight line and reach the plate.

Figure 12.4

Figure 12.4 The Stern–Gerlach apparatus

If atomic magnets were behaving like classical magnets, as μzcosθ, the angle θ between the magnetic moment vector and the direction of magnetic field (z-direction) may have any value between 0 and π. So the magnetic moments of different atoms in the beam might have all possible orientations and after passing through the non-uniform magnetic field, should have spread along z-direction. Instead, to their surprise, Stern and Gerlach found silver atoms striking the plate only in two regions, symmetrically situated about the point of no deflection. The very fact that the silver beam had split into only two components on passing through the inhomogeneous magnetic field, dictates that the magnetic moment vector of silver atoms must have only two orientations with respect to z-axis.

The quantization of the component of the magnetic moment along the direction of the magnetic field is termed as space quantization. We now know that of the 47 electrons in silver atom, 46 electrons form a state with total orbital angular momentum L = 0 and total spin angular momentum S = 0. So the total angular momentum and total magnetic moment of the silver atom in its ground state is due to the spin of last electron.

12.6  SPIN AND ROTATION (SPINOR TRANSFORMATION)

Let us recollect, we studied the transformation of state vector ψ in a coordinate system obtained by rotating it about an axis along unit vector e through an angle ϕ. In this section, we are concerned with the transformation properties of spin wave function under similar rotation. Following Eq. (9.61b), a spin wave function χs of a particle of spin s transforms into a new spin wave function equation in the following way

equation

For a spin equation particle, we have equation, so

equation

Here in Eq. (12.50b), the definition of the exponential of an operator is,

equation

Now using the identity (see Exercise 12.3)

equation

We have

equation

and

equation

So collecting even and odd powers in the expansion (12.51), we may write the rotation operator as

equation

or

equation

Here I is 2 × 2 unit matrix. It may be checked from Eq. (12.54) that for ϕ = 2π, we get

 

Re(2π) = –I        (12.55)

 

Therefore, for ϕ = 2π, we have

 

χ= Re() χ = –χ        (12.56)

 

So a full rotation by 360° about a fixed axis, which, in fact, is equivalent to zero rotation or simply identity operator, as seen in Eq. (12.56) changes the sign of every spinor component. Let us mention here that the usual vectors and tensors in three-dimensional co-ordinate space behave differently under the said rotation operator: these simply return to their original value upon a rotation of 2π. However, it may be mentioned here that for spinors getting their sign changed under 2π-rotation does not de-value their usefulness. It may be remembered that all physical quantities (expectation values) evaluated using these sign changed spinors are same as those obtained by using original spinors.

Let us write the spin wave function χ as

equation

and consider a rotation, about z-axis, through an angle ϕ. Under this rotation, let χ be transformed to χ′, written as

equation

Using Eq. (12.54), we have

equation

or

equation

As a second example, let us consider rotation about x-axis through an angle ϕ. In that case we have,

equation

or

equation
12.7  A MAGNETIC MOMENT IN A UNIFORM MAGNETIC FIELD: THE LARMOR PRECESSION

12.7.1  Classical Treatment

Let us firstly consider a classical system. For example, we consider a classical particle of charge q and mass m moving in a circular orbit and having angular momentum L and magnetic moment μ. L and μ are collinear, related through the equation

equation

Here γ, the ratio of the magnetic moment and the angular momentum is called the gyromagnetic ratio. Let the classical system be put in a constant homogeneous magnetic field.

 

B = B0ez        (12.61)

 

As a result, a torque is exerted on the system. The classical equation of motion of L is given by

equation

or

equation

or

 

Δμ = γ μ Β0sinθΔt        (12.63)

 

Here θ is the angle that μ makes with Β. Δμ will be at right angles to μ and, therefore, magnetic moment μ, without changing angle θ, precesses about the direction of B, as shown in Figure 12.5. Let at time t = 0, μ be along OA and at t = Δt, along OB. Then

 

ΟΆ = μ sinθ

 

and

 

Δμ = O′A · α

 

= μsinθ · α = μsin θω0 Δt        (12.64)

 

From Eqs (12.63) and (12.64), we have

equation

ω0 is the frequency of precession of μ around direction of B and is called Larmor frequency. It may be noted that the Larmor frequency is only one-half of the cyclotron frequency, that is, the frequency with which a free charged particle goes around in a circular path in presence of a constant homogeneous magnetic field.

Figure 12.5

Figure 12.5 Larmor precession of a classical magnetic moment

12.7.2  Quantum Mechanical Treatment: Precession of an Electron Spin in a Magnetic Field

Let us consider an electron at rest in a constant uniform magnetic field pointing in z-direction.

 

B = B0ez        (12.66)

 

The Hamiltonian is the interaction energy

equation

In the matrix representation, where Ŝz is diagonal, Ĥ becomes

equation

The energy eigenvalues of the electron are given as (see Figure 12.6),

equation
equation

where

equation

with corresponding spin states

equation
equation
Figure 12.6

Figure 12.6 Splitting of spin energy level in the presence of constant magnetic field B = B0 ez

Let the electron be initially (t = 0) in state

equation

To find the state at time t, we start with the time-dependent Schrodinger equation

equation

where

equation

and we have to find quantities C1(t) and C2(t).

Substituting Eq. (12.73) into Eq. (12.72) and using Eq. (12.68), we get

equation

or

equation

or

equation
equation

Equations (12.75) have got solutions

 

C1(t) = C1(0) e(–iΩ/2) t        (12.76a)

 

C2(t) = C2(0) e(iΩ/2) t        (12.76b)

Therefore, the normalized state at time t is

equation

Let us consider time development of some particular initial states.

Case I: Let us take the initial state corresponding to C1(0) = C2(0) = 1, that is,

Let

equation

Then from Eq. (12.77), we get

equation

So, if T denotes the period, T = (2π/Ω) then

equation
equation
equation
equation

Here we have used notations mentioned in Table 12.2. From Eqs (12.80) we conclude the following:

  1. At t = 0, the electron is in an eigenstate equation of Ŝx with corresponding eigenvalue ħ/2.
  2. At t = T/4 it is in an eigenstate equation of Ŝy with eigenvalue ħ/2.
  3. At t = T/2 in eigenstate equation of Ŝx with eigenvalue –ħ/2.
  4. At t = 3T/4 in eigenstate equation of Ŝy with eigenvalue –ħ/2.

Let us remember the discussion of Section 12.3 and the connected Figure 12.2 where it was mentioned explicitly that when an electron is found in spin-up state equation of Ŝx, the spin is not really pointing along +x direction but lies on the surface of the cone with its axis coinciding with +x-axis. So at t = 0, the spin is lying on the cone surface as shown in Figure 12.7. In the presence of magnetic field B = B0 ez, the axis of the cone rotates in x–y plane such that at t = T/4, the axis of the cone is along y-direction. At time t = T/2 the axis of the cone is along negative x-axis. At t = 3T/4 along negative y-axis and at t = T again at starting point, that is, along x-axis. So the cone and, therefore, the spin is precessing around z-axis, with frequency Ω = (eB0/me). This is termed as Larmar’s precession.

Figure 12.7

Figure 12.7 Precession of a spinning electron in a constant uniform magnetic field B = B0ez. Initially the electron is in state equation

We may reach the same conclusions, as above, about the precession of electron spin (about z-axis), by finding the expectation values equation and equation. For example,

equation
equation

Putting equation from Eq. (12.79), we have

equation

or

equation

Similarly, we get

equation

and

equation

Equations (12.82) imply that the spin angular momentum vector and, therefore, spin magnetic moment vector rotates about z-axis, in presence of magnetic field B = B0ez, with angular frequency Ω = (eB0/me).

It may also be checked here that, though the initial (spin) eigenstate of the electron equation changes as a function of time to the sate equation in presence of magnetic field B and thus the spin precesses about z-axis with angular frequency Ω, its energy eigenvalue does not change with time. We can find its energy in state equation as

equation

Case II: Let the initial state correspond to C1(0) = 1, C2(0) = 0, that is,

equation

From Eq. (12.77), we may write the state at time t as

equation

It can be easily seen that the energy eigenvalue in state equation is

equation

Let us now find equation

equation

So, if an electron is put in magnetic field B = B0 ez in spin-up state equation at time t = 0, its energy eigenvalue is E1 = (ħΩ/2) and it remains in spin-up state for all time (with energy E1). Similarly if the electron is in spin-down state equation with eigenvalue E2 = –(ħΩ/2), it will remain in spin-down state (with energy E2) for all time.

12.8  ELECTRON SPIN RESONANCE

We studied in Section 12.7 the effect of constant homogeneous magnetic field B = B0 ez on the spin states and energy eigenvalues of an electron. We found that irrespective of the initial spin state of the electron, its energy eigenvalue does not change with time. If initially the electron is in state equation, with the passage of time the state changes to equation to equation to equation to equation; that is, the spin angular momentum vector precesses in x–y plane about z-axis with angular frequency Ω. However, if initially the electron is in state equation with energy eigenvalue equation, its state at time t remains equation (except a multiplicative phase factor ei(Ω/2)t). Similarly if initially it is in state equation with energy eigenvalue equation, its state remains equation with a phase factor ei(Ω / 2)t.

In this section, we shall consider, in addition to the uniform magnetic field B = B0 ez, a rotating radio frequency (r.f.) field B1, in x–y plane, given as

 

B1 = exB1 cos ωt + eyB1 sin ωt        (12.87)

 

We shall see that the r.f. field would cause transitions between spin-up and spin-down states. The total Hamiltonian of electron in presence of both the fields B and B1, is

 

Ĥ = Ĥ0 + Ĥ′        (12.88a)

 

where

equation

and

equation

We shall now solve the time-dependent Schrodinger equation

equation

Let us express the spin state equation in terms of basis state equation and equation (of Hamiltonian Ĥ0), as

equation

where, as mentioned above equation and equation. Substituting Eq. (12.90) in Eq. (12.89), we get

equation

First two terms on L.H.S. are same as first two terms on R.H.S. So, we get

equation

Taking inner product of Eq. (12.91) with eigen-bra (0 1) (i.e., by pre-multiplying Eq. (12.91) by equation and noting that equation and equation, we get

equation

or

equation

or

equation

where

equation

and

equation

Similarly by pre-multiplying Eq. (12.91) by eigen-bra (1 0), we have

equation

Now differentiating Eq. (12.94) and using Eq. (12.92) for (dC2/dt), we get

equation

where

equation

If we take a solution of the form

 

C1 (t) = eiβt        (12.97)

 

We get

 

β = α ± Γ        (12.98a)

 

with

equation

Therefore, the general solution of Eq. (12.95) may be written as

 

C1(t) = A1ei(α + Γ)t + A2ei(α – Γ)t        (12.99)

 

Let us take the initial spin state to be

equation

i.e.,

 

C1(0) = 0, C2(0) = 1        (12.100b)

 

Then Eq. (12.99) gives

 

C1(t) = Aeiαt sinΓt        (12.101)

 

To find C2(t), we use Eq. (12.94)

equation
equation

so

equation

and we get

equation

From Eq. (12.101) and Eq. (12.102), we have

equation

and

equation

giving the required condition

equation

Now the initial spin state (at t = 0) is [see Eq. (12.100b)]

equation

Therefore, the quantity | C1(t) |2 is nothing but the probability that the state equation has changed to equation at time t; alternatively, P(t) = | C1(t) |2 gives the probability of spin–flip. | C1(t) |2 may be re-written as

equation

We can see from Eq. (12.106) that the transition probability P(t) is small for | ω – Ω | >> ω1. For ω = Ω, the spin–flip probability is maximum. In fact, spin–flip probability is unity at times t = (π / 2T), (3π / 2Τ), ··· for ω = Ω. This is called the resonance condition. The resonance condition may be achieved either by varying the frequency ω of the r.f. field or by varying Ω (by varying the strength B0 of the homogeneous field in z-direction).

Figure 12.8 shows the variation of probability P(t) as a function ω/Ω. We get resonance at ω/Ω = 1. We show variation of transition probability P(t) as a function of time in Figure 12.9 for fixed values of ω – Ω; Figure 12.9a for ω − Ω = 6ω1 and 12.9b for ω = Ω.

Figure 12.8

Figure 12.8 Variation of transition probability P(t) = |C1(t)|2 as a function of (ω/Ω), for two different values of (ω1/Ω) = (B1/B0)

Figure 12.9

Figure 12.9 Variation of the transition probability P(t) between equation and equation states under the effect of rotating magnetic field B1(t) (a) shows the case for outside resonance, for example, ω – Ω = 6ω1, (b) shows the resonance case ω = Ω, where even for small value of B1 the transition probability P(t) may become 1

EXERCISES

Exercise 12.1

In a Stern–Gerlach experiment, a beam of hydrogen atoms with velocity 3 × 103 m/s, passes through an inhomogeneous magnetic field of length 50 cm and having a gradient of 200 T/m perpendicular to the direction of the incident beam. Find out the transverse deflection of the atoms of the beam at the point where the beam leaves the field.

Exercise 12.2

Show that Pauli spin matrices satisfy

  1. equation
  2. equation
  3. equation

(∊ijk is the Levi–Civita symbol:

= +1, if ijk = 123, 231, or 312
= –1, if ijk = 132, 213, or 321
= 0, otherwise)

Exercise 12.3

Show that

equation

where A and B are arbitrary vectors.

Exercise 12.4

Consider an electron at rest in a constant uniform magnetic field in x-direction B = B0ex. The corresponding Hamiltonian is

equation

What are the eigenvalues and eigenkets of Ĥ0?

Exercise 12.5

Spin eigenket of an electron is represented by the spinor equation. Express this spinor in a co-ordinate system obtained by rotation about y-axis through an angle ϕ.

Exercise 12.6

Express the spinor equation in a co-ordinate system obtained by

  1. rotation about x-axis through an angle 60°
  2. rotation about y-axis through an angle 60°
  3. rotation about z-axis through an angle 60°

Exercise 12.7

Show that any spinor equation may be expressed as a linear combination of any of the three pairs of eigenkets.

equation

or

equation

or

equation

Exercise 12.8

In a measurement, the z-component of spin of an electron is found to have value ħ/2.

  1. What is the spin state of the electron after the measurement?
  2. Show that in this state
equation

Exercise 12.9

Show that it is impossible for a spin equation particle to be in a state equation such that

equation

Exercise 12.10

A polarized beam of electrons in state equation is sent through Stern–Gerlach analyzer which measures Sx. What values (and with what probabilities) will be found?

Exercise 12.11

An assembly of electrons has an isotropic distribution of spin values. If you choose an electron at random, what is the probability of finding this electron with the following spin components?

  1. equation
  2. equation
  3. equation

Exercise 12.12

A neutron is in spin state equation. What is the probability that a measurement finds each of the following?

  1. equation
  2. equation
  3. equation
  4. equation
  5. equation
  6. equation

Exercise 12.13

Consider a spin equation particle in the state

equation

What is the probability that a measurement finds the following?

  1. equation
  2. equation
  3. equation
  4. equation
  5. equation
  6. equation

Exercise 12.14

Find the expectation values of Ŝx, Ŝy, Ŝz of spin equation particle in spinor state given in Exercise 12.13.

Exercise 12.15

Consider an electron in spin state

equation
  1. Determine the normalization constant.
  2. Find the expectation values of Ŝx, Ŝy and Ŝz.
  3. Find the expectation values of equation, equation and equation.
  4. Find the uncertainties ∆Sx, ∆Sy and ∆Sz. Here, for example,
equation

Exercise 12.16

For a normalized spinor

equation

Compute

  1. equation
  2. equation

Exercise 12.17

Suppose a 2 × 2 matrix equation is written as

equation

where b0, bk(k = 1, 2, 3) are numbers.

  1. Find tr(X) and equation
  2. Find b0 and bk in terms of matrix elements Xij(i, j = 1, 2). This shows that any 2 × 2 matrix X may be expressed a linear combination of Pauli spin matrices equation (written in this exercise as equation) and a 2 × 2 unit matrix I.

Exercise 12.18

Consider a two level system whose Hamiltonian operator is given as

equation

Here number a is having dimension of energy. Find the energy eigenvalues and the corresponding eigen-kets (in terms of states equation and equation).

Exercise 12.19

An electron is in spin state equation Compute

equation

where the expectation value is taken for the state equation. Check the generalized uncertainty relation.

equation

where

equation

Exercise 12.20

A beam of spin equation atoms goes through three Stern–Gerlach measurements 1, 2, and 3 in succession in following manner:

  1. Measurement 1 accepts Sz = ħ/2 atoms and rejects Sz = –ħ/2.
  2. Measurement 2 accepts Sn = ħ/2 atoms and rejects Sn = –ħ/2 atoms. Here Sn is the eigenvalue of the operator Ŝ · en; unit vector en makes an angle ϕ in the xz-plane with respect to z-axis.
  3. Measurement 3 accepts Sz = ħ/2 and rejects Sz = -ħ/2.

What is the intensity of the final Sz = ħ/2 beam if the Sz = ħ/2 beam surviving after measurement 1 is normalized to unity? For what orientation ϕ (of measurement 2) shall we get the maximum intensity in measurement 3?

SOLUTIONS

Solution 12.1

equation
μz = μΒ = 9.27 × 10–24 J/T
M = 1.67 × 10–24 kg

 

so

equation

Solution 12.2

  1. equation

    and so on.

  2. We can easily see
    equation

    so result follows,

  3. From (a) and (b), we can easily find result (c).

Solution 12.3

equation

Using relations of Solution 12.2, we get

equation

Solution 12.4

The Hamiltonian Ĥ0 may be written in matrix form as

equation

We can easily see, Ĥo has

eigenvalue = µBB0, corresponding to eigenstate equation,

eigenvalue = −µBB0, corresponding to eigenstate equation.

Solution 12.5

Let equation denote the spinor in the rotated coordinate system. Using Eq. (12.54), we have

equation
equation
equation

Solution 12.6

  1. Using relation (12.59b), we have
    equation
  2. Using relation (12.107), we have
    equation
  3. Using relation (12.58b), we have
    equation

Solution 12.7

Let us write

equation

so

equation

giving

equation

Therefore,

equation

Similarly, we may get

equation

and

equation

Solution 12.8

  1. After the measurement the electron remains in state equation.
  2. It is straight forward to evaluate these expectation values in state equation.

Solution 12.9

Let us start with operator Ŝz and find the state in which

equation

or

equation

Let us set | a | = 1. Then equation may be written, without loss of generality, as

equation

where θ1 and θ2 are the phases of a and b, respectively. Now let us find equation for this state

equation
= ħ cos(θ1θ2) = 0

Also, we may get

equation

Now there is no value of (θ1θ2), which gives

 

cos(θ1θ2) = sin(θ1θ2) = 0

 

Therefore, there is no state for which expectation values of all the three spin operator components are zero.

Solution 12.10

The state equation may be written as

equation

So, if Sx is measured on electrons in state equation, we will get value ħ/2 with probability equation and –ħ/2 with probability equation.

Solution 12.11

Suppose the spin component Sx is measured. Then obviously one of two values ħ/2 or –ħ/2 will be found. The spin is isotropic, so these two values are equally likely. Hence, probability is equation for both values Sx = ħ/2 and Sx = ħ/2. Similar is the case for Sy and Sz.

Solution 12.12

  1. 1
  2. 0
  3. equation
  4. equation
  5. equation
  6. equation

Solution 12.13

We have seen in Exercise 12.7 that a state equation may be written in following forms:

equation
equation

so

equation
equation
equation

From Eq. (12.110a)

  1. Probability of equation
  2. Probability of equation

    From Eq. (12.110b)

  3. Probability of equation
  4. Probability of equation

    From Eq. (12.110c)

  5. Probability of equation
  6. Probability of equation

Solution 12.14

First Method

Expectation value of equation

Expectation value of equation

Expectation value of equation

Second Method

equation
equation
equation

Solution 12.15

  1. equation

    so equation

  2. equation
    equation
    equation
  3. equation
  4. Values of standard deviations ∆Sx, and so on, may be easily calculated from the values in (a), (b), and (c) above.

Solution 12.17

equation
  1. equation
  2. We have
    X11 = b0 + bz
    X22 = b0 + bz
    X12 = bx – iby
    X21 = bx + iby

    So

    equation

Solution 12.18

Ĥ may be written in matrix form as

equation in the basis equation and equation

where equation and equation

It may be easily checked that matrix Ĥ has got eigenstate

eigenstate equation with eigenvalue equation

and

eigenstate equation with eigenvalue equation

Here normalization constants are

equation

Eigenstates equation and equation may be expressed in terms of basis states equation and equation as

equation

and

equation

Solution 12.19

equation

so

equation

Similarly, we may get

equation

Now

x, Ŝy] = Ŝz

So

equation

So, we get

equation

Solution 12.20

After passing through measurement 1, the beam has atoms in state

equation

In the co-ordinate system obtained by rotating it about y-axis through an angle ϕ, the state may be written (with the help of the result of Solution 12.5) as

equation

Measurement accepts only those atoms having Sn = ħ/2.

So after second measurement the eigenstate is

equation

Now after second measurement, the spin component is measured along z-direction of the original co-ordinate system, which is obtained by rotating the changed coordinate system about y-axis through an angle –ϕ. So we get

equation

So, the number of atoms having Sz = ħ/2 are having probability cos4ϕ/2. This probability is maximum (unity) for ϕ = 0.

REFERENCES
  1. Shankar, R. 1980. Principles of Quantum Mechanics. New York, NY: Plenum.
  2. Baym, G. 1969. Lectures on Quantum Mechanics. New York, NY: W. A. Benjamin.
  3. Ziman, J.M.1969. Elements of Advanced Quantum Mechanics. London, UK: Cambridge University Press.
  4. Merzbacher, E. 1999. Quantum Mechanics, 3rd edn., New York, NY: John Wiley & Sons.
  5. Tinkham, M. 1964. Group Theory and Quantum Mechanics. New York, NY: McGraw-Hill.
  6. Cohen-Tannoudji, C., Diu, B. and Laloe, F. 1977. Quantum Mechanics, Vol. I and II, New York, NY: John Wiley & Sons.
  7. Feynman, R.P., Lighton, R.B. and Sands, M. 1965. The Feynman Lectures on Physics, Vol. III, Reading, MA: Addison-Wesley.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.117.185.169