<<

The Physics of Spontaneous Emission and Mechanical Decay

G. Steele Department of Physics, MIT, MA 02139 (Dated: May 10, 2000) By using some very simple mathematical arguments, it is possible to prove that quantum me- chanics forbids a strictly exponential form for the decay. Examining more closely the derivation of the approximate exponential decay from Fermi’s Golden rule reveals a clear physical interpretation of this strange result. Closely linked to this field is the topic of the quantum Zeno effect and how the observation affects state in , and experimental confirmations of this.

INTRODUCTION the field, plus a coupling between them. The stationary states are then the eigenstates of the combined -field If one looks at the landmark successes of quantum me- Hamiltonian. chanics, one of the most important results that quantum Physically, what this means is that the eigenstates we mechanics predicted was that were stable. It is obtained by solving (1) are not true eigenstates. a famous classical result that electrons in orbit around From a practical point of view, this is a small and nearly a proton in continuously radiate en- always negligible effect: the shift in the atomic energy 5 ergy, and eventually fall into the nucleus: the stability of levels, the Lamb shift, are corrections of the order α . atoms could not be explained in the context of classical In contrast, from a philosophical point of view, these ef- mechanics. fects are indispensable. The effect of the coupling to the In the formulation of quantum mechanics, instead of quantised field can be considered as a perturbation of our solving the classical equations of motion, we take the Hamiltonian: Hamiltonian as an , and define a wavefunction H = H + V 0 (2) that this operator acts on: 0 where V 0 describes the change in the coupling of the atom p2 H = + V (r) (1) to the field due to quantisation. If the system is initially 0 2m prepared in an eigenstate of H0, we can the apply time where V (r) is the Coulomb potential. The electrons are dependent perturbation theory (see [1] page 475) to ob- then eigenstates of energy: their energy is stationary in tain a transition rate between the atomic states from the time, the electrons don’t radiate, and atoms are stable. golden rule: An important subtle difference in the quantum me- 2π chanical picture of the atom is that the everything is W = |hk|V 0|si|2ρ(E ) (3) ¯h s stationary in time. The electron is “moving” in the sense that it has kinetic energy, but every physical property where k is the final state, and s is the initial state, and ρ is of the atom is time independent. It is tempting to try the density of final states of the atom-field system, which to construct a physical picture in which the electron is we assume to form a continuum. Our excited states now rapidly “whizzing” around the nucleus and where what have a mechanism by which they can decay, consistent we are seeing is some kind of time average, but this pic- with the physics we expect. ture would be physically inaccurate: in non-relativistic From this, one is lead to conclude that the decay is quantum mechanics, all of the physics of the atom is sta- then exponential, the probability of finding the excited tionary on all timescales. state s after a time t given by: This poses a problem: quantum mechanics predicts −W t that all of the eigenstates of the atom are time indepen- PS(t) = e (4) dent: excited states are also stationary in time–but we know that excited states decay! and one finds the familiar Briet-Wigner form for the spec- This apparent paradox is solved by quantum field the- trum of the : ory. In our treatment above, we have described the 1 W/2 electric field of the nucleus as a fixed external poten- ρBW (ε) = 2 2 (5) π (ε − εs) + W /4 tial. tells us that the electric field is not simply an external variable we enter in to our This Lorentzian line shape is found to agree very well Hamiltonian as in (1), but that the electromagnetic field with the natural lineshape of atomic transitions. is itself also described by a quantum mechanical equation However, as we will see in the following sections, we of motion, and that the problem we should be solving is can prove on very general grounds that the decay of a for a Hamiltonian which includes both that atom and quantum mechanical system cannot be exponential! Of 2 course, the experiments are not wrong: what we will see is W dt, then the probability that it has not decayed is is that at both very short and very long timescales the (1 − W dt). Thus if the probability that it was in state s decay deviates from exponential, but that exponential was Ps(t), then: behaviour is seen interpolating between the two. Ps(t + dt) = Ps(t) × (Prob. of no decay in dt)

= Ps(t)(1 − W dt) (10) SHORT TIME BEHAVIOUR which leads to exponential decay. We start by giving a mathematical argument why The flaw in our argument can be traced to this last quantum mechanical decay cannot be exponential on step: we have assumed that Ps is changing entirely due short times scales, and then proceed to describe the to transitions out of state s, and have not accounted for physics in subsequent section. the transitions from other states back into s. In quan- tum mechanics, the probability Ps(t + dt) is given by the square modulus of the complex amplitude: Argument for Non-Exponential 2 Decay at Short Times Ps(t + dt) = |cs(t + dt)| dc = |c (t) + s dt|2 (11) We will consider a system that is in a state s at time s dt t = 0. We will make only one assumption: we will assume that the energy of the initial state, hs|H|si is finite. The where evolution of the system is given by: dc 1 X s iωkst = Vksckse (12) ∂ dt i¯h i¯h |ψ(t)i = H|ψ(t)i (6) k ∂t at first order in perturbation theory. In general, we have with |ψ(0)i = |si. Defining cs(t) = hs|ψ(t)i, the surviv- this equation to (10): ing amplitude, gives: dc |c (t) + s dt|2 6= |c (t)|2(1 + W dt) (13) i¯ha˙(t) = hs|H|ψi (7) s dt s

At t = 0 with a(0) = 1, we get: The above can give approximate equality in certain sit- uations. One such situation is the coupling of an initial i¯ha˙(t) = hs|H|si = Es (8) state to a continuum of states densely packed around the final state. In our expression (12), we the have a sum which is real and finite. Thusa ˙(0) is purely imaginary. over a large number of states. At t = 0, transitions from The probability of survival of the initial state is Ps(t) = s start feeding into the amplitudes of the final states k. ∗ a(t)a (t): taking the derivative and using a(0) = 1 gives: At the same time, the states k start to feed back into the amplitude of state s: however, each state feeds back dPs ∗ =a ˙(0) +a ˙ (0) = 0 (9) with a different phase eiωkst. For a continuous set of final dt , after some time T , the phase factors sincea ˙(0) is imaginary. Thus the slope of the decay must will all be uncorrelated. There will be a net cancella- be zero at t = 0, which is clearly violated by exponential tion of the feedback of the states k back into s, and the decay. Quantum mechanics predicts that the decay curve state s will then decay exponentially as we predicted with of an excited state is flat at small times. As we shall see classical statistics. later, this will lead to some interesting experimentally We see that for times less than the coherence time accessible effects. First, however, we will examine more of the continuum, the evolution of the unstable excited closely the steps that earlier had lead us to exponential state will be coherent. Treatments of coherent evolution decay. such as Rabi oscillations are earmarked by a initial slope of zero, and we see that our physical interpretation is consistent with the mathematics of Section (II.A). A Re-examination of Exponential Behaviour: What Went Wrong? Some Interesting Consequences: Exponential decay is a familiar result from classical The Quantum Zeno Effect statistics, which describes the decay of a system through an underlying stochastic process. The classical argument One may question the relevance of this short time be- is as follows[1]: IF we know that the system is in a state s haviour on the physics of the problem: it is more that at a time t and the probability of decay in a time interval likely that the decoherence time of the continuum states 3 is very fast, and it seems unlikely that one would be able to observe the evolution of the system on these timescales. Consider an experiment, however, where we make rapid successive measurements on the system. One of the postulates of quantum mechanics is that the pro- P(t) cess of making a measurement results is a collapse of the wavefunction onto one of the eigenstates: essentially, in performing a measurement, we are “preparing” the sys- tem again at t = 0. In the context of density matrices, the system is described by a diagonal . For a two level system, we would have: µ ¶ 1 0 ρ(t = 0) = 0 0 t1 t2 t3 As time evolves, two things happen: first, ρ22 increases at Time the expense of ρ11, but we also develop coherence terms in the off diagonal elements. If at a later time t = τ we FIG. 1: The short time behaviour of Ps(t) for repeated mea- surements at times t1, t2, and t3 made in the coherent evolu- make a measurement of the system (with an tion regime. whose eigenstates are the 1,2 states), we “collapse” the system onto one of the two states: µ ¶ 1 0 quantum Zeno effect was proposed. However, making ρ(t = 0) = ρ 11 0 0 successive observations on a system on timescales much faster than the decoherence time of the continuum is a or µ ¶ very difficult experiment, and has yet to be achieved, to 0 0 the author’s knowledge. ρ(t = 0) = ρ22 0 1 Note that the physics of the quantum Zeno effect is re- If we have collapsed onto the 1 state, we have effec- lated to the collapse of the wavefunction on observation, tively “reset” the system to the initial state with a re- and is not necessarily tied to a system that is decaying. duced initial amplitude. An important note is that the The effect can be seen in any system where one makes process of making the measurement sets the off-diagonal observations at a rate much faster than the decoherence terms to zero: ie. it destroys the coherences. Mak- time. The longer the decoherence time, the easier the ing repeated measurements of the system at and interval measurement. τ smaller than the decoherence time, the system would Perhaps it is not surprising then that the experimental evolve as shown in Figure (1). We can see that the decay system where the quantum Zeno effect was first convinc- of state 1 is slower than if we had not been making the ingly demonstrated was the inhibition of coherent tran- observations. sitions in a two level system with Rabi oscillations. After n such observations, the probability that were The experiment was performed by Itano et al. [2] at are in the 1 state is given by: NIST Boulder in 1989. In idealised picture, the system they were studying was a three level atom as shown in n P1(nτ) = [ρ11(τ)] Figure (2). Level 1 is the , level 2 is an ex- We note that for τ < T , we have inhibited the decay cited metastable state, and level 3 is an excited state with of the state. We would predict that for τ → 0 (contin- a strongly allowed transition to level 1. There are two ex- uous observation), that the system would be preserved ternal fields applied: there is a strong field on resonance indefinitely in the unstable excited state. The inhibition with the 1,2 transition that drives the two states into co- of quantum transitions through observation is referred to herent Rabi oscillations and there is a short pulsed field as the quantum Zeno effect, and has been confirmed by resonant on the 1,3 transition used the probe/observe the experiment. atom. The idea is that we start with the system in the 1 state at time t = 0 and apply a π pulse on the driving field. If Experimental Evidence of the Quantum Zeno Effect we make no intermediate observations, at the end of the driving pulse, the atom will be entirely in the 2 state. In the previous section, we had discussed the possi- We then start applying short probe pulses. If the atom bility of inhibiting spontaneous emission by frequent ob- is in the 1 state, if will be resonant on the 1,3 transi- servation. This was the original context in which the tion and it will scatter with it’s large resonant 4

3

Probe 2

Driving Field

1

FIG. 2: A schematic of the idealised level structure used in the quantum Zeno effect experiment of Itano et al. FIG. 3: Level structure of the Be+ atoms used in the quantum Zeno effect in the experiment by Itano et al. crossection. If the atom is in the 2 state when the probe pulse is sent, no photons will be scattered. If we observe time scales. To obtain these results, we required no de- scattered photons, we have measured the atom to be in tailed information about the nature of the system we the 1 state, while if we see no photons, we have mea- were studying. The only assumption we made was that sured the atom to be in the 2 state. Either way, we have the energy of the initial state was finite. successfully observed the state of the atom and have col- To obtain some general predictions about the be- lapsed it’s wavefunction. Note that we don’t even need haviour quantum mechanical decay at large times, we to detect the photons: the state of the atom has been must look at the problem from a slightly different angle. recorded in the EM field, regardless if we try to look for In the previous sections, we had determined the evolu- the photons or not. tion of the wavefunction by treating the problem in the The prediction is that as we increase the number n of basis of the atomic Hamiltonian H0 and using perturba- probe pulses we send during the π pulse, the probability tion theory. We will now take a different approach and of finding the atom in the 2 state at the end of the π expand our unstable excited atomic state in energy eigen- pulse will decrease, eventually going to zero for large n. functions of the total Hamiltonian, which will will denote For the actual experiment, Be+ ions placed in a mag- by ε: netic field, and the transitions were between the Zeeman Z X split 2s and 2p levels (see Figure 3). The results they ob- |ψ(t)i = hε|si|εie−iεt/h¯ (14) tained for P2 for different n are shown in Figure 4, and are consistent with the predictions of the quantum Zeno where we are integrating (summing) over the continuous effect. (discreet) eigenstates of H. We then supposed that our While these measurements do indeed confirm our phys- quasi- is orthogonal to the discreet states: ical picture of the wavefunction collapsing onto an eigen- this is in keeping with our physical intuition that decay is state on observation, these do not address the non- linked to the continuum states in the problem. We then exponential decay of spontaneous emission at small have: times. Some interesting theoretical proposals for study- Z ing the short time behaviour of spontaneous decay using −iεt/h¯ the quantum Zeno effect to study atom in resonant cav- cs(t) = hs|ψ(t)i = ρ(ε)e (15) ities and waveguides, where the continuum coupling can be more carefully controlled have been suggested by [3]. where we have defined ρ(ε) = | < ε|s > |2, the energy probability distribution of the initial state s. In this for- malism, instead of calculating a set of interaction matrix LONG TIME BEHAVIOUR elements, we calculate an energy distribution function, and instead of solving coupled equations, we integrate. In all of the previous sections, we have been discussing We can see that to get exponential decay, we would put the consequences of non-exponential behaviour at small in the Briet-Wigner form: 5

FIG. 4: Graph of the experimental and calculated 1 → 2 transition probabilities as a function of the number of measurement pulses. The decrease of the transition probability with increasing n demonstrates the quantum Zeno effect. From Itano et al.

This contradicts the common and well accepted exper- imental fact that exponential decay is observed in a num- −W t ρ = ρBW → cs(t) = e ber of quantum systems. In going back and looking more carefully at a derivation of exponential decay in quantum As we already know, although it may provide a good mechanics, we have gained a deeper understanding of the fit to the data, the Breit-Wigner distribution not a phys- way in which exponential decay arises through cancelling ically correct energy distribution function since it does phase contributions from continuum states. not reproduce coherent evolution at small times. Fur- thermore, it has an infinite variance, which is also some- what unphysical. Possible ways to observe this very short time scale be- It will not be covered in detail here (see reference [4] for haviour of decaying systems lead us consider the quan- an excellent discussion), we can predict deviations from tum Zeno effect. While the quantum Zeno effect has exponential decay at asymptotically large times based on been confirmed by experiments, it has not yet been ap- very general considerations on the energy distribution. plied convincingly in experiment for detecting the early For example, if the continuum has a ground state it can coherent evolution of a state that would otherwise display be shown that the probability at large times will decay exponential decay. −p as a power law, Ps(t) ∼ t . If the continuum has no ground state (for example, the We concluded with a brief discussion of another way coupling of an atom to a uniform electric field), the prob- in which quantum mechanics disobeys exponential de- −t p ability will decay faster that exponential, Ps(t) ∼ e /t . cay, this time as asymptotically long times, by imposing Again, any interested readers would be recommended to a different and independent set of physically motivated examine reference [4]. constraints on the properties of the quantum mechanical Unfortunately, experimental observation of these ef- system. It would seem, however, that this decay regime fects is much more difficult than with the short time be- is outside the realm of experiment. haviour. For example, considering the decay of a charge pion, it is estimated in reference [5] that by the time this behaviour sets in, the surviving pion probability is less that 10−80. [1] E. Merzbacher Quantum Mechanics, vol. I, (W. A. Ben- jamin, New York, 1966), Ch. 18 CONCLUDING REMARKS [2] W. M. Itano, D. J. Heinzen, J. J. Bollinger, and D. J. Wineland, Phys. Rev. A 41, 2295 (1990) [3] A. G. Kofman and G. Krurizki, Phys. Rev. A 54, 3750 By very general arguments, we have shown that the (1996) evolution of a quantum mechanical system must be at [4] K. Uniikrishnan, Phys. Rev. A 40, 41 (1998) some small time scale and will always begin with zero [5] C. B. Chiu, E. C. G. Sudarshan, B. Misra, Phys. Rev. D, slope. 16, 520 (1977)