Chapter 2. One-Dimensional Systems 2.1 Free Particles The task of this class is to solve time-dependent Schrödinger equation to reveal physics of target systems. For a time-independent Hamiltonian H , ψ(t) = exp(−iHt / ) ψ(0) = ∑ exp(−iEnt / )cn φn (2.1) n where ψ(0) = ∑ cn φn , and En and φn are eigenvalues and eigenstates of H : n H φn = En φn . (2.2) Therefore we need solve the time-independent Schrödinger equation ⎡ P2 ⎤ ⎢ ⎥ φ =E φ , +V(X ) n n ⎢ 2m ⎥ n ⎢⎣ ⎥⎦ (2.3) particularly, in real-space representation, ⎡ 2 d 2 ⎤ ⎢− ⎥ φ(x) = Eφ(x). +V(x) ⎢ 2m dx 2 ⎥ ⎢⎣ ⎥⎦ (2.4) The simplest case is a free particle with mass m, V(x) = 0 , then we solve − 2 d 2 d2 φ(x) = Eφ(x) ⇒ φ(x) + k 2φ(x) = 0, 2 2 2m dx dx where the wave number k = (2.5) 2mE . The solutions are Ae ikx and Be −ikx ; because they are degenerate, the eigenfunction for wave number k is φk (x) = Ae ikx + Be −ikx . (2.6) We then use the boundary and normalization conditions to determine A and B. (1) Since there is no bound, the probability density should be a constant, i.e., 2 2 2 const. = φ(x) = A + B + 2Re ⎡⎢AB *e 2ikx ⎤⎥ , ⎣ ⎦ thus, either A or B must be zero ⇒ φ(x) = Ae ikx (now we allow k be negative). (2) Normalization condition: 25 (2.7) δ(k ′ −k) = φk ′ φk = ∫ ∞ −∞ then A = 1 2π ⇒ φk (x) = e iϕ 2π 2 ′ A e −ik xe ikx dx , (2.8) e ikx . We can set the irrelevant phase ϕ = 0 , then we finally reach the familiar eigenfunction for a 1D free particle: 1 φk (x) = with eigenvalue (eigenenergy) Ek = 2π e ikx (2.9) 2k 2 . In addition, this is also an eigenfunction of 2m momentum, with eigenvalue pk = k . 2.2 Confined Particles Consider a particle with mass m in an infinite potential well with width L. We are solving exactly the same equation, Eq. (2.5), obtaining the same general solution, φk (x) = Ae ikx + Be −ikx 0 L/2 L . However, the boundary condition is different: φk (0) = 0, and φk (L) = 0 ⇒ A + B = 0; Ae ikL + Be −ikL = 0 ⇒ A = −B and sin(kL) = 0 ⇒ kL = nπ , with n = 1, 2, 3... (Question: can n be 0?) Then kn = nπ / L and φn (x) = An sin(knx) . Normalization condition: δn ′n = φn ′ φn = ∫ ∞ −∞ A*n ′An sin(kn ′x)sin(knx)dx ⇒ An = 2 / L . The eigenfunctions are φn (x) = ⎛ nπ ⎞ 2 sin ⎜⎜⎜ x ⎟⎟⎟ , L ⎝ L ⎟⎠ (2.10) with eigenvalues En = (kn )2 2m = n 2π 2 2 . 2mL2 Question: are these states of H still eigenstates of momentum? 26 (2.11) A few crucial remarks: (1) Confinement leads to discrete energy levels. Only bound states are quantized, while the unbound states have continuous eigenvalues. (2) A bound state can’t be an eigenstate of momentum; therefore, it has no well-defined momentum. Momentum conservation is not required for transitions involved bound states. (3) In nanoscale systems (size < 10 nm), quantum confinement plays a major role. (4) When L shrinks, the energy level differences increase; L → 0 , V(x) → −λδ(x) with λ > 0 , there is only one bound state: mλ −|κ|x mλ 2κ 2 mλ 2 φ(x) = e with κ = 2 ; E = − =− 2 . 2m 2 (2.12) (5) A finite potential well has both bound and continuous states, depending on if E <V0 or E >V0 . (6) Finally, wavefunctions are either symmetric or anti-symmetric, while the potential well is symmetric (w.r.t. the well center). 2.3 Particles in a Symmetric Potential For a symmetric potential, V(x) =V(−x) , Question: are the eigenfunctions symmetric or antisymmetric or neither? 2 d 2 − φ(x) +V(x)φ(x) = Eφ(x) . 2m dx 2 (2.13) 2 d 2 φ(−x) +V(x)φ(−x) = Eφ(−x) . 2m dx 2 (2.14) Replacing x with −x : − Define the symmetric and antisymmetric functions: ψS (x) ≡ φ(x) + φ(−x) and ψA(x) ≡ φ(x)− φ(−x) , respectively. Adding and subtracting Eqs. (2.13) and (2.14), we find both ψS (x) and ψA(x) are solutions. However, if ψS (x) and ψA(x) are degenerate, then a linear combination of them, ψ(x) = c1ψS (x) +c2ψA(x) , which is neither symmetric nor anti-symmetric, is also an 27 eigenfunction for a symmetric potential! This is the famous spontaneous symmetry breaking, responsible for many fascinating phenomena, such as ferromagnetism and superconductivity. 2.4 Periodic Potential: the Bloch Wavefunction Consider a particle in a periodic potential, e.g., in a crystal as illustrated on the right, V(x + d) =V(x) ⇒ V(x + nd) =V(x) , (2.15) with n = 1, 2, 3 Question: does the wave function have the same translational symmetry, i.e., φ(x + d) = φ(x) ? The translation operator: T(d) x = x + d , T(d)φ(x) = φ(x −d) , and T(d) = exp(−iPd / ) . ⇒ exp(−iPd / )φ(x) = φ(x −d) . ⇒ φ(x) = exp(iPd / )φ(x −d) . V(x + d) =V(x) ⇒ H(x + d) = H(x), Hamiltonian is invariant under translation ⇒ T(d)H(x) = H(x)T(d), or [T(d),H ] = 0 Thus one could construct the common eigenstates φ(x) for Hamiltonian and translation operator T(d) : [Note: though it could fail under certain boundary conditions…] T(d)φ(x) = λφ(x), while T(d)φ(x) = φ(x −d) ⇒ λφ(x) = φ(x −d) ⇒ | λ |2 = 1, and λ is a function of d ⇒ λ = exp(−ikd), with k an arbitrary real number. Therefore, (2.16) φ(x) = exp(ikd)φ(x −d). Obviously φ(x + d) ≠ φ(x) , but the wave function must have certain periodic component. Define φ(x) ≡ F(x)u(x), and u(x) is periodic, i.e., u(x) = u(x −d) . Then φ(x −d) = F(x −d)u(x −d) ⇒ F(x)u(x) = exp(ikd)F(x −d)u(x −d) ⇒ F(x) = exp(ikd)F(x −d) , so that F(x) = exp(ikx) . We thus derived the Bloch theorem: φk (x) = exp(ikx)uk (x) , uk (x + nd) = uk (x) (2.17) Quantum phase in the term exp(ikx) is relevant now, which defines the crystal momentum k due to periodic crystal lattice. [Though an irrelevant arbitrary phase term eiδ can still be added.] 3D Bloch theorem ( R1 , R 2 and R 3 are crystal primitive lattice vectors): φk (r) = exp(ik ⋅ r)u k (r), with u k (r) = u k (r + n1R1 + n2R 2 + n3R 3 ). 28 (2.18) The Kronig-Penney Model: ⎪⎧0, (0 < x < a) , with V(x + d) =V(x) . V(x) = ⎪⎨ ⎪⎪V0, (a < x < d) ⎩ Step 1: Solving wave function in one periodic interval, 0 < x < d . 1. 0 < x < a . φ(x) = A cos(αx) + B sin(αx) , with α = 2mE / . Here A and B are two complex numbers, but could be simplified to real by neglecting e iϕ . 2. a < x < d . We only study bound states, whose energy E < V0 . φ (x) = C exp(− β x) + D exp( β x) , with β = 2m(V0 − E) / . Step 2: Extend to other intervals using Bloch’s theorem Eq. (2.17): u(x + d) = u(x) , φ(x) = exp(ikx)u(x) . Denote φ I (x) the wave function in the range of 0 < x < d , assuming φ I (x) is known. We want to find φ− I (x) in the range of −d < x < 0 . u(x + d) = exp[−ik(x + d)]φI (x + d) , and u(x) = exp(−ikx)φ−I (x) ⇒ exp[−ik(x + d)]φI (x + d) = exp(−ikx)φ−I (x) ⇒ φ−I (x) = exp(−ikd)φI (x + d) (2.19) For −nd < x < −(n −1)d, φ−N (x) = exp(−iknd)φI (x + nd); (2.20) Then For (n −1)d < x < nd, φN (x) = exp(iknd)φI (x −nd). (2.21) Boundary conditions: φ (x) and d φ (x) are continuous at x = a and x = 0 , therefore we obtain four equations: dx A cos(αa) + B sin(αa) = C exp(−βa) + D exp(βa) (2.22) −Aα sin(αa) + Bα cos(αa) = −C β exp(−βa) + Dβ exp(βa) (2.23) A = exp(−ikd)[C exp(−βd) + D exp(βd)] (2.24) Bα = exp(−ikd)[−C β exp(−βd) + Dβ exp(βd)] (2.25) 29 In order to find non-trivial solutions to Eqs. (2.22)-(2.25), the determinant must be zero: cos(αa) sin(αa) − exp(−βa) − exp(βa) −α sin(αa) α cos(αa) β exp(−βa) − βexp(βa) det =0 1 0 − exp(−βd −ikd) − exp(βd - ikd) 0 α β exp(−βd −ikd) − βexp(βd - ikd) ⇒ ⎛ β 2 − α2 ⎞⎟ ⎟⎟ sinh(βb)sin(αa) cos(kd) = cosh(βb)cos(αa) + ⎜⎜⎜ ⎜⎝ 2αβ ⎟⎠ (2.26) where b = d −a . Because α and β both depend on energy E , the right-hand-side of Eq. (2.26) is a function of E denoted by F(E) . Accordingly, cos(kd) = F(E) ⇒ F(E) ≤ 1 . (2.27) This leads to the most important feature of periodic potentials: continuous energy bands with forbidden energy gaps. Periodicity in crystal potential leads to electronic band structures, which determine electronic properties: Small band gap: semiconductors; Large gap: insulators; No gap: metals. But there are exceptions… Silicon: a semiconductor Aluminum: a metal 30 2.5 Simple Harmonic Oscillators The simple harmonic oscillator (SHO) is one of the most important models in quantum mechanics: (1) Any realistic potential well near the bottom can be approximated by a SHO. (2) In particular, molecular vibrations, lattice vibrations in crystals, nuclear structures, etc. (3) Theoretically, the foundation of field quantization, leading to quantum field theory. One example is the potential energy for a diatomic molecule, V(r −r0 ) =V(r0 ) + 1 d 2V(r) (r −r0 )2 +O(r −r0 )3 . 2 2 dr r=r (2.28) 0 Then the basic Hamiltonian for a one-dimensional SHO is H= P2 1 2 P2 1 + kX = + mω 2X 2 . 2m 2 2m 2 (2.29) Classically, the Hamiltonian equation for SHO: dx ∂H p dp ∂H = = , =− = −mω 2x . dt ∂p m dt ∂x (2.30) dx 2 + ω 2x = 0 . 2 dt (2.31) 1 x(t) = A sin(ωt + ϕ); E = mω 2A2 . 2 (2.32) Therefore, Solution: However, quantum mechanically the solutions are very different. We solve the timeindependent Schrödinger equation to obtain eigenfunctions and eigenenergies: ⎛ 2 d 2 ⎞ 1 2 2⎟ ⎜⎜− ⎟⎟ ψ(x) = E ψ(x) . + mω x ⎜⎜ 2m dx 2 2 ⎟⎠ ⎝ (2.33) We can solve Eq. (2.33) using lowering (annihilation) and raising (creation) operators defined as a≡ mω ⎛⎜ P ⎞⎟ mω ⎛⎜ P ⎞⎟ ⎟⎟, then a † = ⎟⎟ , ⎜⎜X + i ⎜⎜X −i 2 ⎝ mω ⎟⎠ 2 ⎝ mω ⎟⎠ 31 (2.34) then a †a = H i H 1 + [X,P ] = − . ω 2 ω 2 (2.35) aa † = H i H 1 − [X,P ] = + . ω 2 ω 2 (2.36) Similarly, Eqs. (2.35) and (2.36) ⇒ ⎡a,a † ⎤ = 1 , ⎢⎣ ⎥⎦ (2.37) 1 1 H = ω(a †a + ) = ω(N + ) , 2 2 (2.38) where the number operator N ≡ a †a . The equation we want to solve is 1 ⇒ ω(N + ) n = En n . 2 H n = En n (2.39) Obviously, H and N have the same eigenvectors n : N n = λn n with λn the eigenvalue. Then a n and a † n are eigenvectors of N as well, with eigenvalues of λn −1 and λn + 1 , respectively. This is because ( a †a a n ) = (aa † )( ) = aa a n −a n = aλn n −a n = (λn −1) a n , i.e., ( ) = (λ ( −1 a n N an † n ( ) ( −1) a n ; N a † n ) = (λ n ( ) ) + 1) a † n . (2.40) Summary: a † increases the eigenvalue, while a decreases the eigenvalue. Question: are n eigenvectors of a † or a or both? Consider the ground state 0 , N 0 = λ0 0 , and then no other eigenstates have eigenvalue ( lower than λ0 . But a 0 has an eigenvalue of λ0 −1 , thus the state a 0 ) simply doesn’t exist, i.e., a 0 = 0 . This is a very important conclusion! Now we can easily derive En for a SHO. a 0 = 0 ⇒ a †a 0 = 0 ⇒ λ0 = 0 ( ) ( ) ( a †a a † 0 = (λ0 + 1) a † 0 = 1 a † 0 ) ⇒ a † 0 = C 0 1 and λ1 = 1. 32 (2.41) Similarly, a † 1 = C 1 2 , since state a † 1 has an eigenvalue of λ2 = 1 + λ1 = 2. Thus a † n −1 = C n−1 n ; and λn = n . (2.42) Eq. (2.42) leads to ⇒ H n = ω (n + 1 / 2) n , N n =n n (2.43) i.e., the eigenenergies of a SHO is En = (n + 1 / 2)ω . (2.44) A summary of these extremely useful second-quantization operators: The number operator: N = a †a , N n = n n , and H = (N + 1 / 2)ω . Raising (creation) operator: a † n = C n n + 1 , ⇒ n a = n + 1 C n* , then n aa † n = C n 2 2 n + 1 n + 1 = Cn . (2.45) On the other hand, aa † = a †a + 1 ⇒ aa † n = (n + 1) n . (2.46) Combining Eqs. (2.45) and (2.46), we obtain a† n = n + 1 n + 1 ; C n = n + 1; n a = n +1 n +1 . (2.47) Lowering (annihilation) operator: a n = Dn n −1 , similarly, we can obtain Dn = n; a n = n n −1 ; n a † = n −1 n . (2.48) Based on Eqs. (2.47) and (2.48), one can easily verify that a †a n = n n ; aa † n = (n + 1) n ; [a,a † ] = aa † −aa † = 1 . In addition, n = 1 n a † n −1 = 0 = 1 1 a1 = 1 † n(n −1) 1 1⋅2 (a ) 2 (a ) 33 2 1 n −2 == 2 == 1 n! n! n (a ) (a ) n . † n 0 . (2.49) (2.50) The (energy-level) number representation: ⎡ 0⎤ ⎡0 ⎢ ⎢ ⎥ ⎢ ⎢ 0⎥ ⎢0 ⎢ ⎥ ⎢ n = ⎢⎢ ⎥⎥ , a = ⎢0 ⎢ ⎢ ⎥ ⎢0 ⎢1 ⎥ ⎢ ⎢ ⎥ ⎢ ⎢⎣ ⎥⎦ ⎢ ⎢ ⎣ 1 0 0 2 0 0 0 0 ⎡0 0 ⎤⎥ ⎢ ⎥ ⎢ 0 ⎥ ⎢ 1 0 ⎥, ⎢ † ⎥ 2 3 ⎥ a = ⎢⎢0 ⎥ ⎢ 0 ⎥ 0 ⎢0 ⎥ ⎢ ⎥ ⎢ ⎥⎦ ⎣ 0 0 0 0 0 0 0 3 0 ⎡0 ⎤⎥ ⎢ ⎥ ⎢0 ⎥ ⎢ ⎥, ⎢ ⎥ N = ⎢0 ⎥ ⎢ ⎢0 ⎥⎥ ⎢ ⎥ ⎢⎣ ⎥ ⎦ 0 1 0 0 0 0 2 0 0 0 0 3 ⎤ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥⎦ Next, we want to work out the eigenfunctions for a SHO, ψn (x) = x n . We use the real , and p0 ≡ mω , then Eq. (2.34) ⇒ mω space presentation. Define x 0 ≡ a= 1 ⎛⎜ X P⎞ 1 ⎛⎜ x d ⎞⎟⎟ ⎜⎜ + i ⎟⎟⎟ = ⎜⎜ + ⎟. p0 ⎟⎟⎠ 2 ⎜⎝ x 0 2 ⎜⎝ x 0 p0 dx ⎟⎟⎠ It is always a good idea to define a dimensionless variable y ≡ a= (2.51) d d x = x0 ⇒ ⇒ dy dx x0 1 ⎛⎜ d ⎞⎟ 1 ⎛⎜ d ⎞⎟ † ⎜⎜y + ⎟⎟; a = ⎜⎜y − ⎟⎟ . dy ⎟⎠ dy ⎟⎠ 2⎝ 2⎝ (2.52) The strategy is finding ψ0(x) first, then obtaining ψn (x) iteratively. a 0 =0 ⇒ x a 0 = dψ (y) 1 ⎛⎜ d ⎞⎟ ⎜⎜y + ⎟⎟ ψ0(y) = 0 ⇒ 0 + yψ0(y) = 0 ⇒ dy dy ⎟⎠ 2⎝ dψ0(y) ψ0(y) = −ydy (2.53) Solving Eq. (2.53), wavefunction for the ground state is ψ0(y) = A0 exp(−y 2 / 2) , (2.54) 1/4 ⎛ mω ⎞⎟ ⎟ ψ0(x) = A0 exp −x / 2x , where A0 = ⎜⎜ ⎜⎝ π ⎟⎟⎠ ( 2 2 0 ) 1/4 ⎛1⎞ = ⎜⎜ ⎟⎟⎟ x 0−1/2 . ⎜⎝ π ⎟⎠ (a ) = † Then, wavefunction for any state of a SHO is readily computed using n (2.55) n n! 0 : n 1 ⎡⎢ 1 ⎛⎜ d ⎞⎟⎤⎥ ⎟⎟ A0 exp(−y 2 / 2) . ψn (x) = x n = y − ⎜ ⎢ ⎜⎝ dy ⎟⎠⎥⎥⎦ n ! ⎢⎣ 2 34 (2.56) Eq. (2.56) involves Hermite Polynomials: n ⎡ 1 ⎛ d ⎞⎤ H n (y) = exp(y / 2) ⎢⎢ ⎜⎜⎜y − ⎟⎟⎟⎥⎥ exp(−y 2 / 2) . dy ⎟⎠⎥⎦ ⎢⎣ 2 ⎝ 2 (2.57) Using Hermite polynomials, wavefunction for a SHO is ψn (x) = AnH n (y)exp(−y 2 / 2) = AnH n (x / x 0 )exp(−x 2 / 2x 02 ) , (2.58) where the normalization constant 1/4 1/4 ⎛ 1 mω ⎞⎟ ⎟ An = ⎜⎜⎜ 2n ⎜⎝ π2 (n !)2 ⎟⎟⎠ ⎛ ⎞⎟ 1 ⎟ x −1/2 . = ⎜⎜⎜ 2n ⎜⎝ π2 (n !)2 ⎟⎟⎠ 0 Question: How to evaluate 4 X 2 2 ? Answer: 4 X 2 2 = 3x 02 = (2.59) 3 . mω One can easily verify that for the ground state 0 , (Δx)2 (Δp)2 = 1 2 4 1 ⇒ Δx ⋅ Δp = . 2 (2.60) Question: for an arbitrary energy eigenstate n , what is Δx ⋅ Δp ? (Your homework) Now let’s evaluate the optical transition matrix elements m exp(ikx) n between two eigenstates of a charged SHO. Light is an electromagnetic wave; consider a monochromatic laser with angular frequency ω , whose electric field is E = E0 exp[i(k ⋅ r − ωt)] (2.61) The interaction between the charged SHO and the field is H ′ = −qr ⋅ E (2.62) Based on perturbation theory (We will study it later this semester), the optical transition rate between two eigenstates is proportional to m H ′ n 2 , which can be decomposed to a couple of matrix elements in the form of m exp(ikx) n for 1D SHO. Method 1: in the x-representation (Eigenstates of X as basis) m exp(ikx) n = ∫ ∞ −∞ ψ*m (x)e ikx ψn (x)dx . Method 2: in the number-representation (Eigenstates of Hamiltonian as basis) 35 (2.63) (ik)j m Xj n . j ! j =0 ∞ m exp(ikx) n = ∑ (2.64) Method 3: Feynman’s approach First, we write e ikx in terms of lowering and raising operators e ikx = exp(iλa † + iλa) , where λ = kx 0 / 2 = k (2.65) . We use the Feynman-Glauber theorem to simplify Eq. (2.65). 2mω The Feynman-Glauber theorem (You have proved it in HW#3) states that for two operators A and B, if both of them commute with the commutator C ≡ [A,B] : [A,C ] = [B,C ] = 0 , then e A+B = e Ae Be −C /2 (2.66) Here A = iλa † , B = iλa , and C = (iλ)2[a †,a] = λ 2 , obviously [A,C ] = [B,C ] = 0 . Then exp(iλa + iλa † ) = exp(iλa † )exp(iλa)exp(−λ 2 / 2) 2 † (2.67) † Thus m exp(ikx) n = e −λ /2 m e iλa e iλa n , and we need to evaluate e iλa n and m e iλa . Using Eq. (2.48), we can derive n (iλ)j j (iλ)j a n =∑ j! j! j =0 j =0 n! n−j , (n − j)! ∞ e iλa n = ∑ me iλa † (iλ)l † l = m ∑ (a ) = m −l l! l=0 ∞ (iλ)l ∑ l! l=0 m m! . (m −l)! (2.68) (2.69) Combining Eqs. (2.68) and (2.69): m m e iλa e iλa n = ∑ † l=0 (iλ)l l! n m! (iλ)j ∑ (m −l)! j =0 j ! n! m −l n − j (n − j)! (2.70) Where m −l n − j = δm−l,n−j , thus for the non-zero terms, m −l = n − j ⇒ l = m + j −n . Without losing generality, we assume m ≥ n : (iλ)m+j−n m ! (iλ)j j! j =0 (m + j −n)! (n − j)! n m exp(ikx) n = e −λ /2 ∑ 2 n = e −λ /2(iλ)m−n m !n ! ∑ 2 j =0 n! (n − j)! (iλ)2j . j !(m + j −n)!(n − j)! 36 (2.71)