History and Some Aspects of The Lamb Shift: G. Jordan Maclay
History and Some Aspects of The Lamb Shift: G. Jordan Maclay
History and Some Aspects of The Lamb Shift: G. Jordan Maclay
Abstract: Radiation is a process common to classical and quantum systems with very different effects in
each regime. In a quantum system, the interaction of a bound electron with its own radiation field leads
to complex shifts in the energy levels of the electron, with the real part of the shift corresponding to a
shift in the energy level and the imaginary part to the width of the energy level. The most celebrated
radiative shift is the Lamb shift between the 2s1/2 and the 2p1/2 levels of the hydrogen atom. The
measurement of this shift in 1947 by Willis Lamb Jr. proved that the prediction by Dirac theory that
the energy levels were degenerate was incorrect. Hans Bethe’s calculation of the shift showed how to
deal with the divergences plaguing the existing theories and led to the understanding that interactions
with the zero-point vacuum field, the lowest energy state of the quantized electromagnetic field, have
measurable effects, not just resetting the zero of energy. This understanding led to the development
of modern quantum electrodynamics (QED). This historical pedagogic paper explores the history of
Bethe’s calculation and its significance. It explores radiative effects in classical and quantum systems
from different perspectives, with the emphasis on understanding the fundamental physical phenomena.
Illustrations are drawn from systems with central forces, the H atom, and the three-dimensional harmonic
oscillator. A first-order QED calculation of the complex radiative shift for a spinless electron is explored
using the equations of motion and the mass2 operator, describing the fundamental phenomena involved,
and relating the results to Feynman diagrams.
Keywords: Bethe; radiative shift; vacuum fluctuations; vacuum field; mass renormalization; Lamb shift;
QED; radiative reaction; radiative shift harmonic oscillator; zero point fluctuations
1. Introduction
1.1. Background
The shift of atomic energy levels from the levels given by the Dirac or Klein–Gordon equations with
the appropriate potentials results from effects that may be classified into four groups [1–13]: (1) The
interaction of the bound particle with its own radiation field, or equivalently with the zero-point (0 K
temperature) quantized vacuum electromagnetic field; (2) vacuum polarization effects; (3) finite nuclear
mass effects, including recoil corrections; and (4) nuclear structure effects, including finite size and
polarization corrections. The most frequently discussed and measured shift in energy levels is the
celebrated 2s1/2 − 2p1/2 Lamb shift in the hydrogen atom.
Although measurements of the shift were attempted in the 1930s, it was not measured accurately
until 1947 when Lamb and Retherford employed rf spectroscopy and exploited the metastability of the
2s1/2 level and determined the shift was about 1050 MHz, or 1 part in 106 of the 2s1/2 level [14–18]. Shortly
thereafter Bethe [19] published a nonrelativistic quantum theoretical calculation of the shift assuming it
was due to (1), the interaction of the electron with the vacuum field. This radiative shift accounted for
about 96% of the measured shift.
In this historical pedagogic paper, we discuss aspects of Bethe’s pivotal calculation, including its
history, its significance, and its impact on the development of quantum electrodynamics. We then consider
radiative shifts from different perspectives, classical, and QED, with the objective of highlighting the
connections between different aspects of the Lamb shift, and clarifying the physical processes involved.
As a pedagogic paper, the QED calculations in this paper are limited to the lowest-order shift for
spinless electrons, the same as Bethe’s calculation. To explore the connections between the physical
phenomena and the mathematics, we derive the complex first-order radiative shift in terms of the mass2
operator using the fundamental equations of motion, and then relate the results to Feynman diagrams. This
is a more difficult derivation than simply using second-order perturbation theory or Feynman diagrams.
Generally, textbook derivations only consider the real part of the shift. The radiative shifts are interpreted
as the difference in energy or mass renormalization between a free electron and a bound electron both
in the vacuum field, precisely as Bethe described it. The real part of the shift is the level shift and the
imaginary part the level width, and we derive a dispersion relation between these parts. Atomic level
shifts can be modeled as arising from transitions with the absorption and emission of virtual photons that
are causing the atom to be in different energy states some of the time. To offer two perspectives, we discuss
results for two central forces systems, a H atom and a three-dimensional isotropic simple harmonic oscillator.
The hydrogen atom is the fundamental two-body system and perhaps the most important tool of
atomic physics and the continual challenge is to calculate its properties to the highest accuracy possible.
The current QED theory is the most precise of any physical theory [20]:
The study of the hydrogen atom has been at the heart of the development of modern
physics...theoretical calculations reach precision up to the 12th decimal place...high resolution
laser spectroscopy experiments...reach to the 15th decimal place for the 1S–2S transition...The
Rydberg constant is known to 6 parts in 1012 [20–22]. Today the precision is so great that measurement
of the energy levels in the H atom has been used to determine the radius of the proton.
This remarkable precision began with the measurement and calculation of the first-order radiative
Lamb shift and that is why we are presenting a historical and pedagogic discussion of it. The derivation
of this shift is present, in one form or another, in virtually every book on quantum field theory [23–27]. The
derivation is often based on the Dirac equation for an electron with spin and second-order perturbation theory.
There are many excellent and comprehensive reviews of the Lamb shift and the computation of energy
levels to high precision in hydrogen-like atoms, including all the different effects [1–13]. As noted above,
the purpose of this paper is quite different from those reviews. No new physics is presented. Instead, we
offer some new perspectives on the old physics which began the new age of QED. We hope this exploration
will be of value, particularly to students and non-experts.
Bethe’s interpretation of the divergences he encountered. We derive an expression for the complex shift in
terms of matrix elements of the mass2 operator M2 , which corresponds to the total self energy squared of
the bound particle. Using the equations of motion for a relativistic scalar particle in a potential, we derive
an expression for M2 to order α in the radiation field, i.e., assuming that only one radiation field photon is
exchanged. We also consider the requirements for gauge invariance in our expressions for a physical shift.
In Section 5, we consider the radiative level shifts in the non-relativistic dipole approximation,
demonstrating that the shift is complex: the imaginary part corresponding to the width for decay by
dipole emission and the real part corresponding to the displacement of the energy level. This result is an
extension of Bethe’s second-order perturbation theory calculation of only the level shift. We show that
the real and imaginary parts satisfy a dispersion relation, which is fundamentally just an expression of
causality [28]. We interpret the radiative shift as due to the virtual transitions induced by the interaction of
the particle with its own radiation field. This interaction means that a given energy level has a finite width
and that the mean energy of the particle, averaged over time, is shifted. After developing the results for an
arbitrary central force potential , we illustrate with two particular cases: the harmonic oscillator potential
and the Coulomb potential.
In Section 6, we apply the methods developed in the calculation of the radiative shift to a
fully relativistic, spinless electron bound in a harmonic potential. In Section 7, we offer a conclusion.
The Appendix A includes brief biographies of Willis Lamb Jr. and Hans Bethe.
The theory thus leads to the false prediction that spectral lines will be infinitely displaced from
the values predicted by the Bohr frequency condition. . . As it stands the integral over ν diverges
absolutely. We have treated these difficulties in some detail because they show that the present
theory will not be applicable to any problem where relativistic effects are important, where
that is, we cannot be guided by the limiting case c → ∞ [c is the speed of light.] ... It appears
improbable that the difficulties discussed in this work will be soluble without an adequate theory
of the masses of the electron and the proton; nor it is certain that such a theory will be possible
on the basis of the special theory of relativity.
In 1938 Kramers had suggested the idea of renormalization of the mass due to interactions with the
vacuum field and its necessity in classical as well as in quantum theories, but had no clear idea how to do
it in practice [30]. As Bethe said in an interview in 1996 [31,32]:
Kramers had said [at the Shelter Island Conference] that we misunderstood the self energy of the
electron. The divergent self energy of the electron was already included in the physical mass. We
need to consider the difference in the self energy between a free electron and one bound in an atom.
It was believed that the divergence in the self energy of a electron due to its interaction with the
radiation field was linear in the cutoff frequency, until, in 1939, at Fermi’s suggestion, Weisskopf used the
relativistic Dirac theory and showed (after correcting a critical error in sign pointed out by Furry [33]) that
the electron self energy divergence was logarithmic [34]. He computed that the electron charge distribution
Physics 2020, 2 108
was spread over a Compton wavelength with a shape described by a Hankel function because of its
interaction with the vacuum field, a calculation that remains valid today [23].
The Dirac theory predicted that the 2s1/2 and 2p1/2 levels in the H atom were degenerate.
Measurements of the energy difference had been done but with mixed results. Then, in 1947, Willis Lamb
Jr. applied the expertise in microwave technology that he developed working with Prof. Isador Rabi [35] at
Columbia on radar research during WWII to the precise determination of the 2s1/2 − 2p1/2 energy difference
of 1050 MHz or 0.004 eV. Dyson who, as a graduate student working with Bethe at Cornell, recalled [36]:
And of course the people at Cornell were very closely in touch with the people in Columbia,
and in particular Willis Lamb talked to Hans Bethe who was the professor at Cornell, and Bethe
then sat down and gave the first more or less adequate theory of the Lamb shift, just from a
physical point of view. He understood that the reason why you had the Lamb shift was that
the electron in the hydrogen atom was interacting with the Maxwell electromagnetic field, in
addition to interacting with the proton, so that the effect of the fluctuations in the Maxwell field
were disturbing the electron while it was revolving around the proton, causing a slight change
in the position of the orbits. And so it was the back reaction of the electromagnetic field on
the electron that Lamb had been measuring. And so Bethe understood that from a physical
point of view. The problem was then, could you actually calculate it? And with the quantum
electrodynamics as it was then, it turned out you couldn’t; that if you just applied the rules of the
game as they were then understood and tried to calculate the Lamb shift, the answer came out
infinity, not a number of megacycles but an infinite number of megacycles. So that wasn’t very
useful and so it was clearly a real defect of the theory that it couldn’t grapple with this problem.
Lamb presented his results at the Conference on the Foundations of Quantum Mechanics held at
Shelter island during 1–3 June 1947, and published them 18 June 1947 in a three-page paper in Physical
Review [14]. Dyson later commented on the reaction to Lamb presenting his results at the conference [36]:
The hydrogen atom being the simplest and most deeply explored object in the whole universe,
in a way—I mean if you don’t understand the hydrogen atom, you don’t understand anything,
and to find that things were wrong even with a hydrogen atom was a big shock. So it became
the ambition of every theoretical physicist to understand this.
At the conference, many people, including Schwinger, Weisskopf, and Oppenheimer, suggested that
the deviation resulted from quantum fluctuations acting on the electron in the atom. However, the shift
from this interaction was infinite in all existing theories and therefore had been ignored. The consensus
was that the current theory was fundamentally flawed and that a radically new idea was needed to deal
with this. On the 75-mile train ride home to Schenectady, NY, Bethe did a non-relativistic calculation using
second-order perturbation theory, assuming an interaction with the vacuum field arising from minimal
coupling. The calculation predicted that the interaction of the electron with the vacuum field would lead
to a shift of 1040 MHz [19]. Bethe wrote a paper that was three pages long and sent it to the participants
on 9 June. The paper was received by the Physical Review and published on 15 August. As Bethe later
recalled in an interview [31,32]:
The combination of these two talks of Kramers and Lamb stimulated me greatly and I said to
myself: lets try to calculate that Lamb shift, lets try to calculate the difference between the self
energy of a free electron and that of an electron bound in the hydrogen in the N = 2 state. At the
conference I said to myself: I can do that. And indeed once the conference was over I traveled
to Schenectady to General Electric Research Labs. On the train I figured out how much that
difference might be. I had to remember the interaction of the electromagnetic quanta with the
Physics 2020, 2 109
electron. I wasn’t sure about a factor of two. So if I remembered correctly, I seem to get just
about the right energy separation of 1000 MHz, but I might be wrong by a factor of two. So the
first thing I did when I came to the library at General Electric was to look up Heitler’s book on
radiation theory. I found that indeed I had remembered the number correctly and that I got
1000 MHz. ...I was helped very much by a previous paper by Weisskopf who had show that
in Dirac pair theory that the energy of an electron only diverged logarithmically when you get
to high energy. So I said to myself once I take the difference between bound electron and free
electron the logarithmic divergence will probably disappear and it will converge. So lets just
calculate the effect of quanta up to the energy of the electron mass times c squared and lets hope
the relativistic correction won’t make any difference.
Dirac has called this result the “most important calculation in physics for decades.” Freeman Dyson
described it as “a turning point in the history of physics. . . It broke through a thicket of skepticism and
opened the way to the modern era of particle physics. It showed us all how to connect QED with the real
world” [36,37]. In his Nobel lecture, Feynman called Bethe’s calculation “the most important discovery in
history of quantum electrodynamics” [38,39]. The importance of this calculation cannot be understated. In
a major 2001 review article, Eides states: “Discovery of the Lamb shift, a subtle discrepancy between the
predictions of the Dirac equation and the experimental data, triggered development of modern relativistic
quantum electrodynamics and subsequently the Standard Model of physics” [7].
The key to Bethe’s success was in his interpretation of the infinities that arise in the calculation. He saw
that one infinite energy shift was independent of the Coulomb potential, and therefore, he reasoned, should
correspond to a mass renormalization of the free electron. He interpreted the infinity as a renormalization
of a bare electron resulting in an electron with the observed physical mass. This insight allowed him to
continue with the calculation and compute the finite energy shift due to the interaction of the electron with the
vacuum field for a specific atomic state. The resulting frequency integration led to another divergence, but only
logarithmic, thus he used an energy cutoff of mc2 to insure a finite result, reasoning that since the calculation
was non-relativistic a cutoff was justified. His insightful assumptions led to a result of surprising accuracy.
To obtain the final numerical result required a calculation of the so-called Bethe log (which he credited
to GE workers Dr. Stehn and Miss Steward) which can be interpreted as the average excitation energy
for the radiative interaction. It equals the average energy difference between the level whose shift is
being computed and the other levels which are reached by virtual transitions due to interaction with the
quantum vacuum. The calculations showed that the average excitation energy for the N = 2 state was
about 17.8 Rydbergs or 240 eV (1 Rydberg = 13.6 eV, corresponding to the energy of the ground state of the
H atom), which Bethe thought was “an amazingly high value” that indicated scattering states dominated
the Bethe log, but the result was still clearly in the non-relativistic energy range since 240 eV << mc2
= 0.5 MeV. That value of the Bethe log was in error, and the currently accepted value for the 2s state is
16.6392 [7], which changes the calculated 2s1/2 − 2p1/2 shift from 1040 MHz, the value Bethe gave in his
paper, to 1052 MHz, compared to the currently accepted value of about 1057.8 MHz.
Some reflections of Freeman Dyson shed some light on Bethe’s personality and his work style that
may have led to his success [36]:
He had this intense love of doing physics collectively. I mean that it wasn’t really physics if you
did by yourself, it was something you did with a group of people. And so I just loved it from the
beginning and became very much a part of it right away. And then, of course, his way of work
was actually quite unique, I mean if you compare Bethe with anybody else I knew. First of all, he
had total command of the facts, that he absolutely just—you never needed to look up a number
in a table because he knew them all. He knew all the energy levels of hydrogen and he knew the
atomic weights of the different elements and the density of lead and gold and uranium, all these
Physics 2020, 2 110
just physical quantities, he knew them all. In addition of course, he had an extraordinary ability
to sit down and calculate and just simply go at it...And he was, of course, also just extraordinarily
reliable: if he said something, you could believe it. He was very careful about everything he
said. So just a thoroughly solid person. Very different from Feynman, because Feynman was
far more imaginative. I mean, one thing Bethe did not have was imagination; he never really
invented anything, he just used the theories that were there to explain the facts, and he knew
the facts and he knew the theories, so he just put them together; whereas Feynman was always
inventing things and he didn’t believe the theories that were taught in the textbooks, he had to
make them up for himself, so he had a much harder time; but still, of course, in the end you need
imagination too; I mean, both kinds of physicists are needed.
The lowest-order radiative shift which Bethe computed is of magnitude mα( Zα)2 , where α is the
fine-structure constant [40], Z is atomic number (the number of protons), and m is the mass of the electron. The
shift involves the emission and absorption of one virtual photon (so-called one-loop correction, thus α is raised to
the first power) and accounts for about 96 % of the difference in energy between the 2s1/2 and 2p1/2 states.
The other major effect of the same order contributing to the classic Lamb shift is vacuum polarization,
often called the Uehling contribution, which had been computed successfully before the Lamb shift
measurement and gives a shift of about −27 MHz [13,41,42]. Vacuum polarization arises from the presence
of a virtual electron positron cloud, approximately a Compton wavelength in radius, surrounding a charge,
essentially producing a dielectric constant in the vacuum region near a charge. For s states, the electron
goes very close to the proton, penetrating this cloud, and therefore effectively seeing a larger charge and
experiencing a stronger binding force, which lowers the energy level [7,33]. The fact that including the
effect of the vacuum polarization insured greater agreement with the experiment convinced physicists
that the vacuum polarization contribution was real and correct.
And as far as I know, this paper both disappointed and stimulated other people who were
who were more versed in relativistic theory, namely Schwinger and Feynman. . . and also
Weisskopf. Weisskopf pursued the theory in an old fashioned way and calculated the relativistic
part, together with some of his collaborators. And Schwinger was stimulated to produce a
completely new theory, relativistically invariant theory of quantum electrodynamics. But
essentially extending the old quantum electrodynamics, making it relativistically invariant
and so on. . . Feynman at Cornell used the completely novel and independent way of getting at
the same problem. He had his own way of doing quantum mechanics, his own way of putting
in the electric field. And it turned out that in the end that Feynman’s new way was very much
easier than Schwinger’s way.
Shortly after Bethe’s calculation, Dyson published, as a problem assigned by Bethe, a calculation of
the Lamb shift for a spinless electron [43]. Formal and rigorous relativistic calculations using perturbation
theory and including spin were done in 1949 by J. French and V. Weisskopf [44] and N. Kroll and W. Lamb [45].
Weisskopf later commented about these calculations that they “. . . resulted in good agreement with the experiment.
However, the methods used by those authors of subtracting two infinities were clumsy and unreliable [33].”
However, history has been kinder to these calculations which were not dependent on cutoffs, which were perhaps
clumsy and difficult, but produced excellent results that have stood the test of time [23,25].
Bethe’s breakthrough in understanding the role of the vacuum electromagnetic field and how to deal
with divergences led to intense theoretical work in quantum electrodynamics. It is most remarkable that
Physics 2020, 2 111
within a year three different approaches to quantum electrodynamics were independently developed that
were relativistic and could deal with divergences with some success. Schwinger, Tomonaga, and Feynman
each had proposed a manifestly covariant method, and shown its capability to address a broader range of
QED problems that just the energy levels of the H atom [38,46]. Although these methods all appeared to
be different, with his characteristic insight Freeman Dyson showed that they had essential similarities and
were mutually consistent [47]. He summarized: “The advantages of the Feynman theory are simplicity
and ease of application, while those of Tomonaga-Schwinger are generality and theoretical completeness.”
These new methods could be used to treat the radiative interaction as a perturbation to any desired order
of approximation. Dyson also compared the results to those from the S matrix theory [48]. Dyson observed
that Oppenheimer was particularly reluctant to accept Feynman’s approach [49].
Welton provided some physical insight into the radiative shift with an approximate calculation
based on a semi-classical model of the vacuum field which caused oscillation of the electron bound in the
Coulomb field, effectively increasing its size [50]. This motion meant that the electron saw a modified
Coulomb potential. Only for s states was the spread of the electron sufficient to modify the energy level, in
rough agreement with Bethe’s result. This calculation is discussed in more detail in Section 3.3.
In their comprehensive 2001 review [7], Eides et al. give a different perspective on the spread of the
electron: “According to QED an electron continuously emits and absorbs virtual photons and as a result its
electric charge is spread over a finite volume instead of being pointlike,” and then they use the expression
for the form factor, F (−k2 ) = 1 − (1/6) < r2 > k2 , to obtain the rms radius, obtaining a value of 1330
MHz for the Lamb shift. Their calculation differs from that of most authors [23,27], in that they assume the
bound electron is slightly off mass shell so the cutoff term becomes ln(1/Zα)2 rather than ln(1/Zα).
A period of intense theoretical development followed Bethe’s calculation, characterized by
calculations of the energy levels of the H atom, and QED in general, done with greater and greater
precision and complexity. Some of the key developments from 1950 to about 1970 are in the papers
[12,51–56]; from 1980 to 2000 are in [57–74]; and from 2000 to current are in [75–89]. Theorists applied
themselves to compute the numerous other effects leading to the total shift between the 2s1/2 and 2p1/2
levels, as well as for other levels, including relativistic corrections, center of mass effects, recoil corrections,
radiative recoil corrections, nuclear size and spin effects, and more rigorous, more precise and higher order
calculations of the radiative shifts (for reviews, see [1–10]).
One of the biggest challenges in the computation of the radiative shifts is the necessity to deal
with frequencies from the IR to relativistic values. For the low frequencies, the starting point is the
non-relativistic dipole approximation, and the Coulomb gauge is the most convenient. On the other hand,
for the high frequencies, relativistic dynamics is needed, the binding energy can be neglected, and the most
convenient gauge is the covariant Feynman gauge. Matching the contributions from both regions is a challenging
procedure. Commenting on these perennial matching issues in a 2001 review, Eides et al. observe [7]
It is a strange irony of history that due to these difficulties it became common wisdom in the
sixties that it was better to avoid separation of the contributions coming from different momenta
regions than to try to invent an accurate matching procedure... Bjorken and Drell wrote, having
in mind the separation procedure: ‘The reader may understandably be unhappy with this
procedure. . . we recommend the recent treatment of Erickson and Yennie which avoids the
division into soft and hard photons.’ Schwinger wrote ‘...there is a moral here for us. The artificial
separation of high and low frequencies, which are handled in different ways, must be avoided.’
All this advice was written even though it was understood that the separation of the large and
small distances was physically quite natural and the contributions coming from large and small
distances have a different physical nature.
..the explanation of the Lamb shift is a far more orderly affair it is is consistently carried through
within the framework of old-fashioned perturbation theory...the joining up of the low- and high-
energy contributions does not involve any new physics: it is a simple mathematical device to
enable the use of two distince approximation schemes [74].
In actual fact, the attitude has changed over the last decade and theorists have developed more
elaborate methods for dealing with matching contributions from high and low frequency regions, and are
now trying to embrace the split in order to clarify the physical nature of the corrections and to improve the
results of computations [7,87].
In Steven Weinberg’s 1995 classic “The Quantum Theory of Fields,” he uses an elegant method of
computing radiative shifts in which he introduces a photon mass in the photon propagators that ultimately
cancels when the low and high momenta regions are combined. As he says, his result is 1052.19 MHz,
“just the same as the old result of Kroll and Lamb [45] and French and Weisskopf [44] which they obtained
using the techniques of old-fashioned perturbation theory [25].” Lowell Brown in his book “Quantum
Field Theory” advocates using analytical continuation in the spatial dimensionality n of the field [26]. He
notes that in n > 4 dimensions there is no IR divergence and in n < 4 there is no UV divergence, thus, in
limit of n → 4, one can secure the correct results.
perturbation theory, which should rightfully deal with small perturbations, is dealing with infinite terms.
Three years before he died, Feynman wrote:
The shell game that we play..is technically called renormalization. But no matter how clever
the word is, it is what I would call a dippy process! Having to resort to such hocus-pocus has
prevented us from proving that the theory of quantum electrodynamics is mathematically self
consistent [94].
It is ironic that Bethe’s original calculation appears to have set this direction for the development of
QED. Had he not has such success with his original calculation, perhaps we would have a theory without
infinities today that provided a more satisfying intellectual and philosophical viewpoint. However, it is
hard to argue with success.
3. Radiative Shifts, Classical Physics, and the Zero Point Fluctuations of the Electromagnetic Field
s
2πh̄c2
A(r, t) = ∑ ( a ei(k·r−ωk t) + akλ
ωk V kλ
† −i (k·r−ωkλ )
e ) ek,λ , (1)
k,λ
where the raising and lowering operators obey the commutation rules
and the two polarization vectors (λ = 1, 2) are orthogonal to k, thus k · ek,λ = 0, and
The electric field is E(r, t) = −∂A(r, t)/∂t and B(r, t) = ∇ × A(r, t). The interaction Hamiltonian for
a particle of charge e and mass m in the vacuum field is
1 ~ )2 ,
HI = (~p − e A (4)
2m
where A ~ is the vector potential for the vacuum field. The radiative shift in energy levels, such as the Lamb
~ term.
shift, arises from the ~p · A
To summarize the properties of the vacuum field in QED: no real photons are present, only random
virtual photons of energy h̄ωk /2 and momentum hk/2c, with all possible momenta present consistent
with Equation (1). The expectation values of the electromagnetic fields vanish but the variances do not.
The fields are isotropic (invariant under rotations), invariant under space-time translations (homogeneous),
and under boosts (Lorentz invariant). The energy density spectrum which is proportional to ω 3 is also
Lorentz invariant. For temperatures above 0 K, there is an additional black body component to the vacuum
field, which we do not consider here.
In QED, we can model mass or charge renormalization with the process:
Physics 2020, 2 114
A similar process occurs for an atom, in which the atom undergoes allowed virtual (energy conserving)
transitions due to radiative reaction or the vacuum field. These transitions can be seen as shifting the
average energy of the atom. This mechanism responsible for the radiative part of the Lamb shift is
discussed in Section 5.2.3 from the QED viewpoint.
In QED, radiative shifts are often calculated using Feynman diagrams, in which the atom is depicted
as propagating in time, and it absorbs or emits a virtual photon changing its state correspondingly, then a
short time later (consistent with the time-energy uncertainty principle) emits or absorbs the same virtual
photon and returns to the initial state. This model in a sense describes the interaction of the electron
with its own radiation field. For QED radiative shifts, this process is equivalent to interacting with the
ubiquitous virtual fluctuating zero-point vacuum field.
d2 r Ze2 r 2e2 d3 r
m 2
= − 3 + 2 3. (5)
dt r 3c dt
The second term on the right is the Abraham–Lorentz force, the non-relativistic radiative reaction
force for an accelerating charged particle. The radiation field from the particle is essentially exerting a
force on itself, sometimes called a “self-field”, a phenomena which leads to renormalization and radiative
shifts. The classical equations of motion become sufficiently complicated so that they are usually solved
only in an approximation [96]. We illustrate the effects by considering the non-relativistic simple harmonic
oscillator and the non-relativistic classical hydrogen atom.
Including a damping force in the equations of motion produces a complex shift in the resonant
frequency [96]
i
ω0 → ω0 + ∆ω0 + Γ, (7)
2
where [40]
5 αh̄ 2 3 2 αh̄ 2
∆ω0 = − ( ) ω0 , Γ= ω . (8)
18 mc2 3 mc2 0
We display the factors of c and h̄ for clarity. The term αh̄/mc2 is the time it takes for light to travel a
distance equal to α times the reduced Compton wavelength, which also equals the time it takes for light to
Physics 2020, 2 115
travel a distance equal to the classical electron radius [97]. Only for accelerations that result in changes in
velocity for times less than αh̄/mc2 are radiative effects important. For the classical harmonic oscillator,
the shift ∆ω0 is a higher order effect than the width Γ.
When we recall that in quantum mechanics the energy is proportional to the frequency E = h̄ω
and that the time dependence of an eigenstate of energy E is e−itE , it is no surprise that in quantum
electrodynamics radiative effects produce a complex shift in the bound state energies of a system, the real part
being the shift in the energy level and the imaginary part being the width of the state that determines its lifetime.
We can verify the Bohr Correspondence Principle for the three-dimensional isotropic harmonic
oscillator. This principle states that in the limit of large quantum numbers the classical power radiated in
the fundamental band is equal to the product of the photon energy and the quantum mechanical transition
probability (or the reciprocal of the lifetime). The power radiated from the classical isotropic oscillator is
all in the fundamental band and has the value
2 4 2
P= αω A , (9)
3 0
where A2 is the mean square amplitude of oscillation. The corresponding transition rate or line width Γ is
P 2
Γ= = αω03 A2 . (10)
ω0 3
For a quantum mechanical three-dimensional oscillator, the energy for a state N is EN = ( N + 32 )ω0 ≈
mω02 A2 and we find
2
3 1
A = N+ . (11)
2 mω0
Accordingly in the limit of large quantum numbers, it follows from the Bohr Correspondence Principle that
2α 2
ΓN = ω0 N. (12)
3 m
We show in Section 5.2.4 that this width Γ N equals the radiative level width computed in quantum
mechanics. The Correspondence Principle makes no statement about the level shift, the real part of the
radiative shift, and indeed the classical calculation yields a level shift of order (α)2 while the quantum
mechanical result is of order α.
α( Zα)
rc3 (t) = rc3 (0) − t, (13)
m2
with classical orbital frequency s
Zα
ωcl = . (14)
mrc3
Physics 2020, 2 116
2 α( Zα)2 1
P(t) = . (15)
3 m2 rc4
Applying the Correspondence Principle we obtain the transition probability
P(t) 2 ( Zα)3/2
Γ= = m . (16)
ωcl 3 (mrc )5/2
Substituting the quantum mechanical result for the radius for large principal quantum number N
N2
rc = , (17)
mZα
gives the transition rate or width for state N
2 α( Zα)4
ΓN = m . (18)
3 N5
This width is 2π times the energy lost classically by radiation in one revolution (about 2π 48h̄ MHz
assuming N = 2). We show that for large N this width equals the imaginary part of the radiative shift
calculated from quantum field theory.
3.3. The Relationship between Radiative Shifts and the Zero Point Field
In classical physics, the electromagnetic field in the vacuum vanishes. However, from quantum
electrodynamics, we know that we must consider the zero point vibrations of the electromagnetic field [101].
~ the non-relativistic
For a particle in an electromagnetic field with scalar and vector potentials φ and A,
Hamiltonian is
1 ~ )2 + eφ,
H= (~p − e A (19)
2m
and the relativistic Klein–Gordon equation is
~ )2 − ( E − φ)2 + m2 = 0.
(~p − e A (20)
The radiative shift for an energy level for a particle interacting with its own radiation field, like the
~ term [23]. The A
Lamb shift, is due to the ~p · A ~ 2 term contributes to the free particle mass renormalization
but does not contribute to the radiative shift of an atomic level since its expectation value does not depend
on the state of the atom.
Physics 2020, 2 117
To understand the radiative shift on a more intuitive basis, we investigate the link between the zero
point vibrations and the energy or mass shift of free and bound particles following an approach of Welton
and Weiskopf [50,102]. The zero point vibrations are incoherent and the mean field h~Ei vanishes but h~E2 i
does not. A free charged point particle is constantly being accelerated in the field, acquiring a mean kinetic
energy that increases its effective mass. Since the particle is oscillating, the effective volume occupied
by the particle increases and it can no longer be usefully regarded as a point particle. It cannot radiate
because the zero point vibrations represent the lowest energy state of the vacuum.
Now, consider the effect of the zero point vibrations on the same particle when bound in an external
central force potential, such as a Coulomb or harmonic potential. The external potential will modify the
motion of the particle in the zero point field. The difference between the effective energy for this particle
when bound and when free constitutes the finite measurable radiative shift. To estimate the radiative shift
from the zero point vibrations we can derive an expression for the real part of the radiative shift in terms of
the Laplacian of the potential and the mean square displacement ~ξ 2 of a charged particle in the zero point
field. If ~r is the location of the particle when unperturbed by the zero point field, then when perturbed the
particle effectively sees a potential V (~r + ~ξ ). For weak binding, ξ << r, and we make the expansion [103]
1 ~ ~ 2
V (~r + ~ξ ) = V (~r ) + ~ξ · ∇
~ V (~r ) + ξ · ∇ V (~r ). (21)
2
Since h~ξ i vanishes, the radiative shift is given approximately by the vacuum expectation value of the
last term:
h~ξ 2 i D 2 E
∆E = ∇ V (~r ) . (22)
6
where we assume the potential has spherical symmetry, thus hξ 12 i = hξ 22 i = hξ 32 i = h~ξ 2 /3i. Equation (22)
gives ∆E as the product of two factors, one depending on the nature of the fluctuations of the radiation
field and the other depending on the structure of the system. To estimate h~ξ 2 i for the vacuum field we
consider the Hamiltonian for a particle of mass m and charge e in the vacuum using the radiation gauge
~ = 0) :
(V = 0, ∇ · A
1 ~ (t, 0))2 .
H= (~p − e A (23)
2m
We use the value of the vector potential for the free vacuum field at the origin, which is equivalent to
the dipole approximation. The proton and the electron can be considered to become a point dipole [23].
Hamilton’s equations give the result
m d2~ξ/dt2 = e d A/dt.
~ (24)
Integrating gives
Z t
e
ξ~(t) = ~ t0 , 0 .
dt A (25)
m −∞
Squaring this and taking the vacuum expectation value gives:
e 2 Z t 0
Z t
00
D E
h~ξ · ~ξ i = dt0 e+et dt00 e+et ( A ~ t0 , 0 · A
~ t00 , 0 )+ ,
(26)
m −∞ −∞
the sign “+” at the bottom signifies that the operators in parenthesis are time ordered. The vacuum
expectation value on the right side is simply −igij Dij , where Dij is the radiation gauge propagator in
configuration space [104]:
1 1 1 −iω (t0 −t00 )
Z
0 00 4
Dij t − t = d k δij − k i k j · −
→2 e . (27)
(2π )4 k k2
Physics 2020, 2 118
Accordingly, we find
~ξ 2 = 2α ( h̄ )2 dω
D E Z ωc
. (28)
π mc E0 ω
where we display the factors of h̄ and c to stress that the term in parenthesis is λc the reduced Compton
wavelength of the particle, which we take to be the electron, thus λ is 3.86 × 10−11 cm. We take the upper
limit to correspond to approximately the mass of the particle. For greater frequencies, it is clear that
our semiclassical calculation is invalid because of relativistic kinematical effects and particle–antiparticle
pair creation, which will become possible. (Another justification for taking this limit is given when we
discuss this process from the viewpoint of the uncertainty principle). For the lower limit, we take some
characteristic energy of the bound state system, for example the magnitude of the ground state energy.
The final estimate for the shift in the energy of a particle bound in a potential V (r ) is
D
α h̄ 2 ωC E
∆E = ( ) ln ∇2 V (~r ) . (29)
3π mc E0
If we use a quantum mechanical average for the Laplacian, then this formula is precisely the same as
the first term in the quantum mechanical result for the real part of the radiative shift for a potential V (r )
(see Equation (134)), and gives a shift of 1340 MHz for a H atom with N = 2. However this formula does
not give a complex shift because of simplifications made in the treatment of the zero point vibrations. For
the Coulomb potential the Laplacian is proportional to δ3 (r ), so classically the shift vanishes since the
classical electron is never at the center, while quantum mechanically the shift is for S states only. For the H
atom, the logarithmic term is about 10.5 if we take E0 as the ground state and mc2 for the upper limit and
√ 2
hξ i is about 0.22λc . For the three-dimensional harmonic oscillator, the Laplacian is a constant, thus we
get the same constant shift whether we take a classical or a quantum mechanical average. The logarithmic
term is about 12.4 for an oscillator with ground state energy 2 eV.
field [109]. For example, there is a shift in the mass, charge, and magnetic moment of an electron or a shift
in the Lamb shift of an atom when we put it near a surface or between two surfaces [105,106,110].
A second observation we would like to mention is that radiative shifts can occur whenever we have
an interaction between a particle and a field, not necessarily just the electromagnetic field. For example
there are shifts for the gravitational field or for the meson field of a nucleus [107].
− ∂ 02 φ0 ( x 0 ) + m20 φ0 x 0 = 0,
(30)
where m0 is the unrenormalized mass [112]. The propagator for the bare meson G0 ( x 0 , x 00 ) satisfies
the equation
−∂ 02 + m20 G0 x 0 , x 00 = δ x 0 − x 00 .
(31)
1
G0 x 0 , x 00 = δ x 0 − x 00 ,
2
(32)
− ∂ 02 + m0
or in momentum space
1
G0 ( p) = . (33)
p2 + m20
The meson has a charge distribution and therefore interacts with its own electromagnetic field,
producing a change in the mass. The propagator for a free self-interacting meson becomes
1
GF ( p ) = , (34)
p2 + m20 + M2F ( p)
where M2F ( p) is the mass2 operator for a free, self-interacting or dressed meson. If m2 is the observed
(renormalized) physical mass, then the propagator GF ( p) must have a pole at p2 = −m2 . Thus,
m2 = m20 + M2F p2 = −m2 . (35)
A discussed in Section 2, the space-time methods of Feynman, which were developed right
after Bethe’s calculation, were helpful to provide a physical picture of the phenomena and facilitated
calculations [38]. In that spirit, we consider the diagrams in Figure 1 that show to the order e2 or α in the
meson’s radiation field (one radiation field photon present) the processes that represent the mass2 operator
Physics 2020, 2 120
M2F . By analyzing the mass2 operator in Section 4.3, we show these are indeed the appropriate Feynman
diagrams.
Figure 1. Feynman diagrams for mass renormalization. Time axis is horizontal. The diagram on the left
~ term and shows an electron emitting a virtual photon and then at a later time
corresponds to the ~p · A
~ 2 term.
reabsorbing the photon. The diagram on the right corresponds to the A
In configuration space, the equation of motion for the free self-interacting meson is
Z
p2 + m20 GF x 0 , x 00 + d4 x 000 M2F x 0 − x 000 GF x 000 , x 00 = δ x 0 − x 00 .
(36)
The presence of the convolution integral indicates we can view the meson as having a finite extent.
The “shape” of the meson “centered” at r 0 is proportional the Fourier transform of m20 + M2F ( p), namely
1
δ r 0 − r 00 + 2 M2F r 0 − r 00 + m20 − m2 .
(37)
m
The effective finite extent of the meson in the vacuum field is central to the interpretation of the Lamb
shift, as discussed in Sections 2 and 3.3. Evidently, we can still say we have a point particle but now it
is in a non-local potential. Although we need never explicitly mention the zero point vibrations in our
field theoretic calculation we could interpret the Feynman diagrams as corresponding to the zero point
fluctuations.
We can estimate the amplitude h~ξ 2 i of the zero point oscillations (or equivalently the emission and
absorption of virtual photons) by applying the uncertainty relations to the process depicted in Figure 1.
When the photon is emitted, the particle receives a momentum k i with uncertainty ∆k i . Accordingly, the
uncertainties in position ~ξ and velocity ~v of the particle satisfy the relations ∆ξ > 1/∆k i and ∆vi ≈ ∆k i /m.
Requiring that ∆vi ≈ 1 implies that ∆k i ≈ m and ∆ξ i > 1/m = Compton wavelength. To get the
effective h~ξ 2 i, we must multiply by the probability that the photon has been emitted. The diagram has two
vertices so the probability is proportional to α, which leads to the result α(∆ξ )2 = h~ξ 2 i ≈ 3α/m2 the mean
amplitude squared of the zero point vibrations, which is comparable to the result (Equation (28)) obtained
using the equations of motion for the vector potential.
When we put a bare meson in an external potentia, we assume it forms a bound state. The propagator
and therefore the equations of motion are as before except: (1) the free (mass)2 operator M2F is replaced
by a bound state mass operator M2 ; (2) the propagator GF for a free particle with radiative interaction is
replaced by the corresponding propagator for a bound particle G; and (3) pµ is replaced by the four-vector
by Πµ = pµ − Vµ , where Vµ is the external four-potential in accordance with minimal coupling [23]. The
energy of the state is shifted by a mechanism similar to that for a free bare meson. The Feynman diagrams
are shown in Figure 2.
The double line represents a meson propagating in the external potential. The difference between the
diagrams for the bound meson and the free meson is the radiative level shift (Figure 3). In other words,
the radiative shift in a bound state level is the change in the self-energy of a particle that occurs when it
becomes bound. As discussed in Section 2, this is exactly the way Bethe framed the problem of computing
the Lamb shift. The intermediate state of the atom, i.e., while the virtual radiation field photon has been
Physics 2020, 2 121
exchanged, is unknown. In his historic approach, the cumulative effect of these virtual transitions in given
by the Bethe Log term.
Figure 2. Feynman diagrams for the bound state mass renormalization. The double line represents a meson
bound in an external potential.
Figure 3. Feynman diagrams showing the level shift is the difference between the bound state mass
renormalization and the free particle mass renormalization. The double line represents a meson bound in
an external potential.
To indicate in more detail the process involved in the radiative shift for a Coulomb potential,
we expand the double line representation of the bound meson, indicating separate meson and proton lines
and the photons exchanged that represent the Coulomb force (Figure 4). The graphs giving the radiative
shifts are of the form shown in Figure 5. The lowest-order shift, to order α (first order) in the radiation
field and ( Zα)4 (second order) in the Coulomb field, is given simply by the vertex correction (Figure 6).
Rather than consider separately all the various graphs in the Coulomb field and obtain an answer in
a series with powers of Zα or ln( Zα) as is done with higher order calculations [7,51,67,72], we calculate
the radiative correction using the equations of motion for a meson (spinless electron) in a Coulomb field
and then make approximations to first order assuming that the proton or Coulomb source is an infinitely
heavy point charge. We are neglecting recoil effects, center of mass corrections, radiative corrections and
size effects for the proton. To include these effects we would use the Bethe–Salpeter equation [3,13,79]. On
the other hand, Weinberg (in 1995) did not think the Bethe–Salpeter equation was the correct equation for
relativistic interactions (it includes no crossed photon diagrams), and he concluded: "It must be said that
the theory of relativistic effects and radiative corrections in bound states is not yet in entirely satisfactory
shape [25]".
Figure 4. Feynman diagrams for the meson (top line) bound to the Coulomb field of a proton
(bottom line). The dots indicate that all possible configurations of Coulomb photons, including crossed
photon lines, are to be included.
In general, we are concerned with directly measurable quantities, namely the shift in the difference
between two energy levels of a bound meson. For example, we compute the change in the 2s − 2p
separation. Clearly, this shift is given by the difference in renormalization between a meson bound in a 2s
state and one bound in a 2p state. Thus, the renormalization of a free meson is never actually used.
Physics 2020, 2 122
Figure 5. Feynman diagrams for the meson (top line) bound to the Coulomb field of a proton (bottom line),
with the exchange of Coulomb photons and one radiative photon emitted and reabsorbed by the meson.
Figure 6. Feynman diagram for lowest order radiative correction to the bound meson.
4.2. Expressing the Radiative Shift in Terms of the Matrix Elements of the Mass2 Operator
From the equation for the propagator of a self- interacting meson in a potential V µ ( x ), we find the
equation obeyed by the corresponding meson wave functions. Taking mass renormalized wave functions
of the meson in the potential field as our unperturbed states, we apply first-order perturbation theory to
2
find the expression for the radiative shift in terms of matrix elements of the perturbation M . The Green’s
function or propagator for a meson field φ( x 0 ) that interacts with its own radiation field and the external
potential Vµ satisfies the equation:
2
Π 02 + m2 + M G x 0 , x 00 = δ x 0 − x 00 ,
(38)
where
1 0
Π0µ = ∂µ − Vµ x 0 ,
(39)
i
2
∂0µ = ∂/∂xµ0 for µ = 0, 1, 2, 3, and m is the physical mass. M2 is the mass2 operator and M is the
renormalized mass2 operator both for a meson in a Coulomb potential
2
M = M2 + m20 − m2 . (40)
In Equation (38), we use a shorthand notation for the integration as in Equation (36). We assume our
4-potential is such that we can work in a gauge with Vi = 0, V 0 = V (r ). Since we want an energy shift, we
take the Fourier transform of Equation (38) with respect to time
2 2
p 0 − ( Ē − V 0 )2 + m2 + M ( Ē) G Ē,~r 0 ,~r 00 = δ ~r 0 −~r 00 ,
(41)
where we define Z
2 e e ~r 0 ,~r 00 ≡ d3 r 000 M2 E,
e ~r 0 ,~r 000 G E,
e ~r 000 ,~r 00 ,
M (E ) G E, (42)
and Ee is the relativistic total energy. We can convert Equation (41) to an equation for the wave functions by
expressing the Green’s function as the vacuum expectation value of the time ordered product, signified by
a plus sign, of the meson field φ( x 0 ) and its adjoint φ† ( x”):
Physics 2020, 2 123
0 00
0
† 00
G x ,x =i φ x φ x . (43)
+
The Φm (~r ) are the relativistic bound state particle wave functions h0|φ(~r, 0)| E
em i with the renormalized
mass and a relativistic total energy Em . If Equation (41) is to be satisfied when we substitute this form for
e
G and let m = n, E e=E en , and r 0 6= r 00 , then it follows that
2 2
p0 2 + m2 − Ẽn − V 0 + M Ẽn Φn ~r 0 = 0.
(46)
2 e
We now use first order perturbation theory to calculate the radiative shift due to M ( E n ).
en (~r 0 ) for a meson
The unperturbed wave functions are the renormalized relativistic wave functions ψ
which satisfy the equation h i
p 02 − ( E
en0 − V 0 )2 + m2 ψ̃n (~r 0 ) = 0, (47)
en0 is the unperturbed relativistic energy eigenvalue. For our normalization, we choose
where E
ψ̃n , Ẽn0 − V ψ̂n = m, (48)
We take the scalar product of Equation (46) with ψ en and substitute Equations (48) and (49) to obtain,
in lowest order in the radiation field, the shift for the state N:
∆E
eN ≡ E e0 = 1 (ψ
eN − E eN , M2 ( E
eN )ψ
eN ), (50)
N
2m
which is shorthand for
1
Z Z
2
∆ ẼN = d3 r ψ̃∗N (~r 0 ) d3 r 0 0 M ẼN ,~r 0 ,~r 00 ψ̃N ~r 00 .
(51)
2m
1 D e 2 e E
∆E
eN = N M ( EN ) Ñ . (53)
2m
Physics 2020, 2 124
S Matrix Approach
As an alternative to our approach, we should mention that it is possible to use the S matrix formalism
to find the radiative shift. As mentioned in Section 2, Dyson showed the equivalence of the formulations
of QED of Schwinger and Feynman with the S matrix formalism [47,48]. For the interaction Lagrangian,
we use
µ
Lint = ejµ Arad , (54)
µ
where Arad is the meson’s radiation field and jµ is the meson current in the potential field. We calculate
the S matrix element between pure bound states with the usual harmonic time dependence. Since we have
e0
a perturbation to a bound state the matrix element must be expressible in the form hSi = e−iT ( EN − EN )
e
N
where T is the interaction time. To obtain the shift we perform the integrations and use the usual trick of
equating T and 2πδ(0).
4.3. Derivation of Mass2 Operator for Relativistic Meson (Spinless Electron) in an External Potential
2 e
We now outline the calculation of M ( E ) in a covariant gauge in which the meson’s radiation field
µ
Arad and the meson field φ obey the equations:
h i
Arad ( x 0 ) , Arad ( x 00 ) 0 00 = igµν δx 0 − x 00 ,
µ ν
t =t
h i
A0rad (~r 0 , t) , φ(~r 00 , t) = 0, (55)
h i
∂0 A0rad (~r 0 , t) , φ(~r 00 , t) = 0.
Since the results are gauge invariant, we can choose the Feynman gauge in order to simplify
the calculation. In the final answer, we simply replace the Feynman propagator with the radiation
gauge propagator. The derivation proceeds by converting the Klein–Gordon equation for a self-inter-
acting meson in an external potential into an equation for the corresponding Green’s function G ( x 0 , x 00 ).
An explicit form for M2 ( E
e) is then obtained by comparing this equation to the defining equation for G
2
which includes M (Equation (38)). If desired, one may skip to Section 4.3.2.
where n o
j x 0 = e Arad x 0 , Π 0µ φ x 0 − e2 Arad x 0 Aµ,rad x 0 φ x 0 .
µ
µ (58)
Physics 2020, 2 125
The anticommutator insures that the A · p term is Hermitean. To convert Equation (57) into an
equation for G ( x 0 , x 00 ), we make use of Equation (43). We multiply by φ† ( x 00 ), time order, and take the
vacuum expectation value. We use the equation
∂ 0 2 A x 0 B x 00 + = ∂ 0 2 A x 0 B x 00
+
0 0 00
δ t0 − t00
+ ∂0 A x , B x (59)
+ ∂0 0 A x 0 , B x 00 δ t0 − t00
+ A x 0 , B x 00 δ0 t0 − t00 ,
∂0 0 A x 0 B x 00 + = ∂0 A x 0 B x 00 + + A x 0 , B x 00 δ t0 − t00 ,
(60)
2
Since we are calculating M to order e2 in the radiation field the term
e2 h( Arad 0 µ 0 0 † 00 0 † 00
µ ( x )rad ( x ) φ ( x ) φ ( x ))+ i in h( j ( x ) φ ( x ))+ i may be calculated with a free photon field rather
than the radiation field. In essence this follows since the radiation field is equal to the free field plus terms
of higher order. To show the formal justification, consider the matrix element
σ = h Aµ ξ 0 Aν ξ 00 φ x 0 φ† x 00
i, (62)
+
∂ξ20 Aµ ξ 0 = ejµ ξ 0 ,
(63)
where thus
∂ξ 0 2 σ = e jµ ξ 0 Aν ξ 00 φ x 0 φ† x 00
+
(64)
+igµν δ ξ 0 − ξ 00 φ x 0 φ† x 00
.
+
To lowest order, we may drop the first term. Solving for σ gives
" #
1 0 00
G x 0 , x 00 .
σ= g µν
δ ξ −ξ (65)
∂ξ 02
Considering the boundary conditions, we realize the term in brackets is just the usual Feynman
propagator. Accordingly, we obtain
σ = − D µν ξ 0 − ξ 00 G x 0 , x 00 .
(66)
This result is to be expected since to lowest order the complete Hilbert space factors into two
independent spaces, one for φ( x 0 ) and one for A( x 0 ). Thus, we show that
Physics 2020, 2 126
D E D E
j x 0 φ x 00 + = ie2 Aµ x 0 Aµ x 0 + G x 0 , x 00
(67)
0
rad
0µ 0µ rad 0 0 00
Aµ x Π + Π Aµ x
+e φ x φ x .
+
We can rewrite the second term on the right side using the notation
−
→0 µ rad 0
Π Aµ ( x )φ x 0 ≡ Arad x 0 Π0µ + Π0µ Arad x0 φ x0
µ µ
1 µ 2 µ
(68)
0
Arad ξ 0 φ( x 0 )ξ 0 = x0 .
= ∂ 0 + ∂ x0 − 2V xµ
µ
i ξ i
−
→
† 00
Π02 + m20 G x 0 , x 00 = δ x 0 − x 00 + ie Π 0µ h Arad 0 0
µ x φ x φ x
+
D E (69)
2 0 0 0 00
−e µ
Aµ x A x + G x , x .
Using Equations (38) and (40) for the unrenormalized mass2 operator M2 shows the last two terms on
the right side of Equation (69) are equal to
−→ D E
− M2 G ( x 0 , x 00 ) = ie Π 0µ h Arad x 0 φ x 0 φ† x 00 i − e2 Aµ x 0 Aµ x 0 + G x 0 , x 00 .
µ (70)
+
where M2 G ( x 0 , x 00 ) represent a convolution integral as in Equation (42). To order e2 , we may replace the
full propagator G by the propagator G c for a particle in the potential with the physical mass:
Π0 2 + m2 GC x 0 , x 00 = δ x 0 − x 00 .
(71)
2−
→µ ←−
0 00 0
0 00 0 00
2
= −ie Π †
Πν x 00
ν
M x ,x x Aµ x A x φ x φ x
+
D (73)
2 0
0
E 0 00
+e Aµ x A µ
x +
δ x −x ,
Since our calculation is to order e2 , we again substitute G c for G ( x 0 , x 00 ). Now that we have derived
the equation for M2 ( x 0 , x 00 ), we return to the radiation gauge.
Physics 2020, 2 127
For our calculation of the radiative shift, we need the operator corresponding to the time Fourier
transform of M2 ( x 0 , x 00 ). To obtain this result, we use the expression for G c which follows from Equation (71)
and time translation invariance [113]:
∞
Z dE 1 00 −i Ee(t0 −t00 )
G c x 0 , x 00 = r 0 2
e
r e , (75)
−∞ 2π Π + m − ie
2
where
Πk = pk , Π0 = E
e − V (r ). (76)
d4 k ik(ξ 0 −ξ 00 )
Z
0 00
Dµν ξ − ξ = e Dµν (k), (77)
(2π )4
into our expression for M2 , Equation (74), and we note the derivative with respect to ξ µ0 brings down a
factor of k µ , we find, after some computation, the important result for the unrenormalized relativistic
mass2 operator
2 Z d4 k
e) = ie
M2 ( E Dµν (k ) T µν , (78)
2 (2π )4
where
1
T µν = (2Πµ − kµ ) (2Πν − kν )
( Π − k )2 + m2
(79)
1
+ (2Πν + kν ) (2Πµ + kµ ) − 2gµν .
( Π + k )2 + m2
We exploit the symmetry of the photon propagator under k → −k to write T µν in a form that manifests
crossing symmetry. From the Feynman rules we see that the diagrams corresponding to the T µν operator
are as shown in Figure 7.
Figure 7. Feynman diagrams for the Compton scattering amplitude T µν of a photon by a bound meson
(double line).
The double line in the figure refers to the meson propagating in an external potential. T µν is the
operator Compton scattering amplitude in the forward direction. The seagull term on the right in Figure 7
must be included to insure gauge invariance. At threshold, it gives the Thomson scattering amplitude. As
Equation (78) indicates, we obtain the diagrams for M2 by contracting the above diagrams for T µν with
the diagram for the photon propagator Dµν , giving the resulting Feynman diagrams for M2 in Figure 8.
The crossed diagram may be deformed into the uncrossed diagram, therefore both diagrams give equal
contributions to M2 . Note that, in a calculation of the shift between two levels, the bubble term gives no
contribution since its matrix elements are independent of the state.
Physics 2020, 2 128
Figure 8. Feynman diagrams for M2 which give the radiative shift of a bound meson, which arise from the
Compton scattering amplitude (Figure 7) of virtual radiation photons by a bound meson (double line) .
induces no change in the observed shift. Under a gauge transformation, the radiative shift changes by an
amount
1 ie2 d4 k D e 0
Z E
0 0
δ ∆ EN =
e N n µ k ν T µν
+ µ n ν k µ T µν
+ + ν k µ k ν T µν e
N . (81)
2m 2 (2π )4
λ
k(2Π + k) = (Π + k)2 + m2 − Π2 + m2
, (82)
k(2Π − k) = −[(Π − k)2 + m2 ] + Π2 + m2
to obtain 1
k µ T µν = (2Πν + kν ) − Π2 + m2 (2Πν + kν )
( Π + k )2 + m2
1
2 2
(83)
− (2Πν − kν ) + (2Πν − kν ) Π + m
( Π − k )2 + m2
− 2kν .
For our unperturbed basis states, we have
Π2 + m2 | N
e i = 0. (84)
Consequently, h N e i = 0 and since T µν (k) = T νµ (−k ) it follows that h N
e k µ T µν N e i = 0.
e |k ν T µν | N
Accordingly, we see that T µν is gauge invariant between physical states and that δ(∆ Ef N ) vanishes.
e ~ λ − e∂t λ.
Λ=− ~p · ∇ (86)
m
To obtain gauge invariance, the matrix elements of between the initial and final states must vanish:
~
h f |Λ|i i = 0. If we let λ = eik·~r−iωt , then gauge invariance requires that
1 ~ ~
h f | ~p ·~k eik·~r − ω eik·~r |i i = 0. (87)
m
Following the customary prescription for the dipole approximation, we set exp(i~k ·~r ) equal to unity,
then, since h f |i i = 0, we conclude that the matrix element h f |~p ·~k |i i must vanish if we are to obtain gauge
invariance. Clearly, this is not generally the case and gauge invariance is violated. The difficulty lies in the
fact that setting the exponential equal to one resulted in approximating the change in the vector potential
to first order in k and the change in the scalar potential to zero order in k. If we approximate the change in
the scalar potential to one order higher, then we find that gauge invariance requires
1
h f | ~p ·~k − iω~k ·~r |i i = 0. (88)
m
This quantity does indeed vanish since
~p
f i = i h f |[ H,~r ]ii i = i E f − Ei h f |~r |i i
m (89)
= iω h f |~r |i i.
In the radiation gauge, the scalar potential vanishes, thus we circumvent these difficulties.
Alternatively, we may obtain the unrenormalized M2 operator in the nonrelativistic approximation
from a different perspective, by noting that the pole in the photon propagator in Equation (78) insures
that the integration over k0 leads to the result |~k| = k0 but since |~k | is a momentum it equals a frequency
over the speed of light |~k| = ω/c. As c increases the magnitude of the spatial momentum vanishes and
we obtain the dipole approximation. Seen in this way, the dipole approximation is not gauge dependent
but simply part of the nonrelativistic approximation. If we work in the radiation gauge, then this method
gives the result obtained from the usual proscription.
From dynamical considerations we can show that in a bound system characterized by a small
coupling constant the motion is nonrelativistic and |~k|, the approximate change in momentum for radiative
transitions between states, may be neglected with respect to the momentum p of the bound particle.
Consider a potential of the form
1
V (r ) = mgn+2 (mr )n , n > −2. (90)
n
The exponent of the mass m is chosen so that the coupling constant g is dimensionless; the exponent
of g and the overall coefficient are chosen so that V agrees with the conventional expressions for the
√
simple harmonic oscillator (n = 2, g = ω0 /m and the Coulomb potential (n = −1, g = Zα). The
Physics 2020, 2 130
total nonrelativistic energy of the atom is E = T + V. Employing the virial theorem for our potential
T̄ = −(n/2)V̄ and the uncertainty principle gives the results
1 n+2 2 2
p≈ ≈ gmc, E≈ g mc , (91)
r 2n
where c is the speed of light. These results justify the use of nonrelativistic dynamics for small g.
The contribution to the shift of a bound state energy level will be greatest for resonant virtual transitions,
that is, when the photon energy equals the difference between two energy levels. For these resonant
transitions E ≈ |~k|c and
~k n + 2
∼ g << 1, (92)
p 2n
for weak coupling. To insure that the nonrelativistic approximations remain valid during the integration
over frequency, it may be necessary to use a cut off which is proportional to the mass. The shift for greater
(and therefore nonresonant) frequencies for physically realistic situations can be calculated by neglecting
the bound state energy and keeping only the lowest order terms in the coupling constant.
To understand the physical meaning of the dipole approximation more clearly, we employ the
~
translation operator in momentum space eik·~r to show that for a function f ( p) we have the identity
−→ → − → → − → −
→ −→ → − −→ −
→− →
h N |(2−
→
p − k ) f (−
p − k )(2−
p − k )| N i = (h N |ei k ·~r )(2−
→
p + k ) f (−
p )(2 →
p + k )(e−i k · r | N i). (93)
Applying this result to the expressions for M2 ( E)and T µν (Equations (78) and (79)), we see that
−
→−→
the matrix elements for the shift are between translated atomic states (e−i k · r | N i) that have a center
of mass momentum −~k in order to conserve momentum when the virtual photon of momentum +~k is
emitted. In addition, from the Feynman rules for spinless mesons, we know that the ~k present in 2~p +~k
insures momentum conservation at the vertex. Accordingly dropping the ~k dependence means that we
are violating momentum conservation and neglecting the recoil of the particle, which is a reasonable
approximation since we are dealing with long wavelength photons whose momentum is much less than
the particle’s momentum. In more accurate calculations, we need to maintain center of mass momentum
conservation and include the corresponding recoil terms [3,7,51,67,72,87].
1
T ij = (2pi − k i )
−→ 2p j − k j
(−
→p − k )2 + 2mV − ( E − k0 ) 2m
(94)
1
(2pi + k i ) − 2gij ,
+ 2p j + k j − → −→ 2
( p + k ) + 2mV − ( E + k0 ) 2m
1
T 00 = 4m2 −→ 2
−
→
( p − k ) + 2mV − ( E − k0 ) 2m
(95)
1
+ 4m2 −→ −
→ 2 − 2g00 ,
0
( p + k ) + 2mV − ( E + k ) 2m
Physics 2020, 2 131
1
T i0 = 2m (2pi − k i ) −
→ 2
−
→
( p − k ) + 2mV − ( E − k0 ) 2m
(96)
1
+ 2m −
→ −→ 2 (2pi + k i ) ,
( p + k ) + 2mV − ( E + k0 ) 2m
e − m (which is negative for the hydrogen atom). As a check on
where E is the nonrelativistic energy E = E
the nonrelativistic limit, we can prove gauge invariance by noting
and remembering that for matrix elements between physical states we can use the Schrodinger equation
( H − E)| N i = 0, (98)
where
p2
H= − V. (99)
2m
The expression for the (mass)2 operator in the non-relativistic limit is given by
ie2 d4 k
Z
M2 ( E ) = Dµν (k ) T µν , (100)
2 (2π )4
where T µν is given by the nonrelativistic form in Equations (94)–(96). We use the photon propagator in the
radiation gauge:
g00 Pij
D00 = 2 Dij = 2 , (101)
k k
where
ki k j
Pij = (δij − ). (102)
~k2
(~p−~k)2
We perform the k0 integration first. There are poles in the complex k0 plane at k0 = E − V − 2m + ie
and ±(ω/c − ie) where ω = c|~k | and we display the speed of light c. Closing the contour in the lower half
plane enclosing the single pole at k0 = ω/c − ie gives the result
αc R R dΩk
M2 = − ωdω ×
2mπ 4π
1 ω
Pij (2pi − n̂i ωc ) ω 2 (2p j − n̂ j ) − 2mgij
~p − n̂ c
c + V − (E − ω) (103)
2m
2 2 1
+4m c ω − 2mg00 ,
(~p − n̂ )2
c + V − (E − ω)
2m
where n̂ = ~k/|~k| and we have combined cross terms since they give equal contributions to M2 . As we let
c → ∞, the terms in n̂ω/c vanish leaving us with the expression for M2 obtained by making the dipole
approximation in the usual manner (|~k| → 0).
The angular integration for the gij T ij term is
Physics 2020, 2 132
dΩk 2
Z
= δij , (104)
4π 3
corresponding to the two transverse polarization states of a photon. Using the identity,
ω H−E
= 1− , (105)
H − (E − ω) H − (E − ω)
we find
" #
8 p2 4 H−E H−E
Z
2αc 2 2
M =− dω − 3ω + 2mc − p p − 2mc . (106)
π 3 2m 3m i H − ( E − ω ) i H − (E − ω)
The expectation value of the last term, which comes from g00 T 00 , vanishes for physical states. The first
term can be interpreted as the change in the kinetic energy due to the mass renormalization in the
nonrelativistic limit [23]. The second and third terms compose the free particle mass renormalization. The
next to the last term is the only term that depends on the potential V, and gives a vanishing shift in the free
particle limit V → 0. Thus, the renormalized mass2 operator in the nonrelativistic limit is
4αc H−E
Z
2
M ( E) = dωpi p. (107)
3πm H − ( E − ω ) − ie i
En V
Z
d3 r 0 |hr 0 |n
ei|2 1 + 2
− 2 = 1, (108)
mc mc
where we make the factors of c explicit. Clearly, in the nonrelativistic limit, we obtain the usual Schrodinger
wave functions hr 0 |ni with the normalization
Z
d3 r 0 |hr 0 |ni|2 = 1, (109)
or
where n, l, m are the usual quantum numbers. The effective shift in the unperturbed level E0N due to the
radiative interaction is the matrix element of the renormalized (mass)2 operator with respect to | N >:
1 2
∆EN = EN − E0N = h N | M ( EN )| N i. (111)
2m
2
Substituting the expression for M (Equation (107)) and inserting a complete set of intermediate states
gives the result
2α s [ ( En − EN ) h N | pi | ni hn | pi | N i]
Z ωc
∆EN = ∑
3πm2 n 0
dω
En − EN + ω − ie
, (112)
Physics 2020, 2 133
where the s on the summation indicates we also include scattering states [114]. This is the same result as in
Bethe’s original paper and his book [13,19].
Equation (112) can be easily derived from second-order perturbation theory, as Bethe did, in which the
complete set of states |ni represent intermediate states [23] and this is often the approach in calculations of
the radiative Lamb shift in textbooks. We derive this equation for the shift using the fundamental equations
of motion. We now show that the term in brackets in this equation is proportional to the probability for a
transition between state N and state n by the emission or the absorption of dipole radiation, which leads
to a model for the radiative shift. The interaction Hamiltonian is
e ~ rad (~r (t), t),
Hint (t) = ~p(t) · A (113)
m
where A~ rad is the vector potential for the spinless electron’s or meson’s radiation field. The S matrix
operator is R
∞
S = ei −∞ :dtHint (t): , (114)
+
where the double dots mean the Hamiltonian is normally ordered, with creation operators to the left of the
annihilation operators. We want the matrix element ρ for a transition n → n0 , n0 < n by the emission of a
photon of momentum ~k and polarization ~e:
~
~ (~r (t), t)|0i = ~e e−ik·~r+iωt .
hke| A (116)
Accordingly, we find
e
ρ = −2πi δ ( En0 + ω − En ) hn0 |~e · ~p|ni, (118)
m
where we use the dipole approximation ~k ·~r ≈ 1. The decay rate for n → n0 by dipole emission is
total probability
Γen0 ,n = . (119)
interaction time
d3 k 1 | ρ |2
Z
Γen0 ,n = ∑ 3
(2π ) 2ω 2πrδ(0)
. (120)
pol
Recalling
ki k j
∑ eµi eµj = δij − k2
, (121)
µ
Physics 2020, 2 134
we obtain
4α
Γen0 ,n = ( En − En0 )hn0 | pi |nihn| pi |n0 i > 0, n0 < n, (122)
3m2
for the decay rate from n → n0 by dipole emission where En − En0 = ωnn0 . Similarly, the rate for the
transition n → n0 for n0 > n, by absorption of dipole radiation, is
4α
Γna 0 ,n = ( En0 − En ) n0 | pi | n n | pi | n0 > 0, n0 > n.
2
(123)
3m
In accordance with the principle of detailed balance, we see
From our definition, Γen,n0 is defined only for n0 > n and then is always positive or zero. We see
formally that Γen,n0 = −Γen0 ,n . Accordingly, if n > n0 , we interpret Γen,n0 as −Γen0 ,n . Using this convention
with our expression for Γen,n0 , we find that, after changing variables, the expression in Equation (112) for
the shift may be written in the simpler form:
1 s Z En +ωc − 12 Γen,N
∆EN =
π ∑ En
dω
ω − EN − ie
. (125)
n
From Equations (94)–(96) for T µν , it is clear that ∆EN is an analytic function f ( N, EN ) of the energy
EN , which is in the denominator. We define
s
∆EN = f ( N; EN ) = ∑ f n ( N; EN ) . (126)
n
The partial shift f n ( N; EN ) represents the contribution to the shift in level N from virtual transitions
from level N to level n. We replace EN by the complex variable z and investigate the structure of the partial
shift as a function of z:
1 En +ωc
Z − 12 Γen,N
f n ( N; z) = dω . (127)
π En ω − z − ie
We extend the lower limit of integration to E1 and the upper limit to ∞ and multiply by the appropriate
theta functions (θ (t) = 0 if t < 0, = 1 if t > 0) so that the value of the integral is unchanged. After
summing over all states, we find that the complex radiative shift obeys the dispersion relation [28]
Z ∞
1 Im f ( N; ω )
f ( N; z) = dω , (128)
π E1 ω − z − ie
where
s
1
Im f ( N; z) = ∑ − 2 Γen,N θ (ω − En )θ (ωc + En − ω ). (129)
n
We can separate the integral into its real and imaginary parts
Z ∞
1 Im f ( N; ω )
f ( N; z) = P dω + i Im f ( N; z). (130)
π E1 ω−z
Figure 9. Cut Structure of f ( N; ω ) in the complex ω plane. At each value of En which is less than EN , there
is a cut with a discontinuity of − 12 Γen,N ; at EN , there is no cut. At each value of En which is greater than EN ,
there is a cut with a discontinuity of 12 ΓeN,n .
where Γ N is the total width for decay of state N by dipole radiation. The imaginary part of the shift
equals the half-width in magnitude and is always negative as it must be to insure that the probability
density decreases exponentially:|e−it( EN +∆EN ) |2 = e−Γ N t . Only states to which the state N can decay by
0
−Γen,N ωc − EN + En
n<N 1
Re f n ( N; EN ) = × ln . (132)
ΓeN,n n>N 2π | En − EN |
where E0 is an arbitrary energy parameter, which we shall take to be some characteristic energy of the
bound system, for example, the ground state energy. The first term is the same expression for the shift that
we obtained by considering the motion of the particle in the zero-point field (Equation (29)). Note that we
only assume the spinless electron is in a central force potential V (r ).
(1) Re f m ( N; EN ) + Re f N (m; Em ) = 0.
(2) Re f m ( N; EN ) < 0.
Im f m ( N; EN ) = Re f m ( N; EN )[ π1 ln ωc −1
(3) EN − Em ] .
The first relation shows that the average energy of two levels that shift each other is unchanged.
Together, the first two relations show that virtual transitions to lower states cause downward shifts and
transitions to upper states cause upward shifts. The third statement shows that a lower level’s contribution
to the width is less than its contribution to the shift by the factor π1 ln( ENω−cEm ). We can deduce relations (1)
and (2) for the level shifts exactly and relation (3) for the level width in an approximation by assuming that
the observed energy corresponds to a time-weighted average of the original energy and the energy of the
state to which the system made a virtual transition. To make this interpretation quantitative, we consider a
state N with a partial width Γ = ΓeN,m = Γm,N a for m < N. The system makes Γ transitions from N to m in
one second and remains in the state m for a time allowed by the time-energy uncertainty principle [115]
1
δt ≈ . (135)
EN − Em
Therefore, for a system in which Γ << EN − Em (e.g., atomic systems), the average energy ENave of
level N is shifted and is approximately
Γ Γ
ENave = Em + 1 − EN = EN − Γ. (136)
EN − Em EN − Em
The level shift for state N due to a transition from a state N to a lower state m is ENave − EN or
Re f m ( N; EN ) = −Γem,N . Similarly we find that for a transition from a state m to a higher state N the level
shift is Re f N (m, EN ) = Γ aN,m which is positive. From these two expressions, relations (1) and (2) follow.
Corresponding to the third relation we find using Equation (129) and the results directly above that the
model predicts a level width
1 1
Im f m ( N; EN ) = − Γ = Re f m ( N; EN ), (137)
2 2
This result agrees with relation (3) only if we replace π1 ln ENω−cEm by unity [116]. If we use the dipole
sum rule, then in our model we find that the total level shift is 4/3 of the result obtained in the discussion
after Equation (37) where we obtained the shift by applying the uncertainty principle to determine the
effects of the zero point field on a bound particle.
Physics 2020, 2 137
Figure 10. The energy level E0N is shifted to EN by intermediate virtual transitions to Em 0 , which also
increases the width of the level to Γ. The level Em is shifted to Em by virtual transitions to E0N . The latter
0
5.2.4. Two Examples: The Harmonic Oscillator and the Coulomb Potential
In our discussion thus far we only assume we have a spinless particle of mass m and charge e in a
central force potential V (r ) interacting with its own radiation field. Now, we can apply the results to these
two specific potentials.
(1) The Isotropic 3-D Harmonic Oscillator
Consider a simple isotropic harmonic oscillator in three dimensions for which
1
V (r ) = mω02 r2 , (138)
2
with energy levels
3
E N = ( N + ) ω0 , N = n 1 + n 2 + n 3 . (139)
2
The fact that V (r ) increases formally with r without bound does not introduce difficulties since
transitions are possible only between adjacent energy levels. Employing the matrix elements of the
momentum operator
r
mω0 p √
ni0 p j ni =
ni + 1δn0 ni +1 − ni δn0 ni −1 , (140)
2 i i
we can easily compute the real and imaginary parts of the radiative shift using Equations (122), (123), (131)
and (132). For the complex radiative shift of level EN , we find
α ω
∆EN = ω02 3 ln C − i2πN , (141)
3πm ω0
Since the matrix elements vanish except for S states, we may isolate the L dependence of the shift
by defining the Bethe log γ( N, L) for a state characterized by principal quantum number N and angular
momentum L [13]:
γ( N, L) ∑sn ( En − EN ) h N 0 | pi | ni hn | pi | N 0i
| En − EN | . (145)
= ∑sn ( En − EN ) h NL | pi | ni hn | pi | NLi ln 1 m ( Zα )2
2
Using Equation (134), setting the frequency cutoff to ωc = m, and substituting the Schrodinger wave
function
1 Zαm 3
2
|ψN (0)| = δL0 , (146)
π N
where the Kronecker delta function δL0 vanishes if L 6= 0. We find for the shift for level NL
4m 1 2
Re ∆ENL = α( Zα)4 δ L0 ln − γ ( N, L ) . (147)
3π N3 ( Zα)2
where γ( N, L) must still be numerically evaluated [117]. The result is the same result Bethe obtained in his
original calculation. The Bethe log is tabulated for a few energy levels in the original work in which it is was
introduced [13] and in various articles for additional levels and at a higher precision, for example [7,83].
To provide a scale of magnitude for the shift, we note that the term in square brackets is the energy
radiated in one revolution of the electron in the ground state according to the laws of classical physics and
equals Planck’s constant times 1090 MHz. For N = 2, the 2s shift is 1051.84 MHz. The currently accepted
value for the Lamb shift is about 1057.87 MHz. We can estimate
( Zα)2
∆EN /EN ≈ α , (148)
N
which is about 1 part in 1.3 × 106 for N = 2. The width for low-lying states may be obtained by computing
the sum in Equation (129) explicitly.
In the limit of very large quantum numbers for any central force field for circular orbits, we can
simplify the expression for the width Γ N by assuming that the most important transitions are those for
which ∆n << N. The strongest transitions in the classical limit are between wave packets corresponding
to the circular orbits n = N, l = N − 1 and n = N − 1, l = N − 2. This is equivalent to saying that the
classical radiation is primarily in the fundamental band. Accordingly, our sum collapses to
4α
Γ N = ωcl h N | pi | N − 1i h N − 1 | pi | N i , (149)
3m2
where ωcl is the classical frequency of rotation. This matrix element can be obtained without direct
computation by noting that
D E
N | p2 | N ∼= h N | p i | N + 1i h N + 1 | p i | N i + h N | p i | N − 1i h N − 1 | p i | N i , (150)
Physics 2020, 2 139
which follows from our assumption that the only significant transitions are those for which ∆N = ±1 and
from the fact that h N | pi | N i = 0 for a bound state. We assume that the matrix elements do not change
rapidly with N, thus
h N | p i | N − 1i h N − 1 | p i | N i ∼
= h N | p i | N + 1i h N + 1 | p i | N i . (151)
2α D E
ΓN = N p2 N . (152)
ω
3m2 cl
For the Coulomb potential ωcl = m( Zα)2 /N 3 (see Equations (14) and (17)), we find
2 ( Zα)4
ΓN = mα , (153)
3 N5
which is in accordance with the result obtained through the correspondence principle Equation (18).
Note that nowhere in our derivation of Equation (152) do we specify the detailed nature of the
central force. We only assume that the radiation was in the fundamental band, which is always true for
classical circular orbits. Indeed, this equation agrees with the expression for Γ obtained by applying the
correspondence principle to the classical expression for the radiated power Pc for any circular orbit of a
charged particle
2 α 2 2
Pc = p ωcl , (154)
3 m2
namely
2 α 2
Γ= p ωcl . (155)
3 m2
For both examples, the relative shifts go approximately as α(Bound State Energy Level)/(Rest Mass
Energy), reflecting the fundamental nature of radiative shifts (and that we are considering radiative shifts
in lowest order).
For exact nonrelativistic calculations, the sum over states for the real part of the energy shift was
trivial to compute for the oscillator since only two intermediate states contribute. Alternatively, if we
compute the shift from Equation (107) without inserting intermediate states, then from the equations of
motion we can easily compute the contraction over pi . We will follow this procedure in our calculations of
the level shift for the relativistic harmonic oscillator in Section 6. Unfortunately, to secure exact results for
the Coulomb potential is more difficult. If we use Equation (112), we must include an infinite number of
intermediate states in our sum. If we do not use intermediate states but use Equation (107) directly, then
we find that the equations of motion are intractable unless we use group theoretical techniques, which we
will publish elsewhere [118].
6. Radiative Shift of a Relativistic Meson (Spinless Electron) with a Harmonic Interaction Lagrangian
6.1. Introduction
We compute the radiative shift for a spinless, relativistic meson with a charge e with a harmonic
interaction Lagrangian Lint = Vψ2 where ψ is the meson field and V = C2 r2 , and C is a real constant.
From consideration of the equations of motion, we compute the the radiative shift of the energy levels that
corresponds to the difference of the contribution to the mass renormalization from a mass m bound by
the harmonic interaction and a free meson [119,120]. We derive an integral expression for the complex
radiative shift to order α in the radiation field and to all orders in the binding field. In Section 6.2, we
Physics 2020, 2 140
perform the computations after making the simplifying assumption that the virtual photon is spinless. In
Section 6.3 we include the effects of spin.
We assume the unperturbed meson state | N i obeys the Klein–Gordon equation with the interaction
term
( p2 − p20 + C2 r2 + m2 )| N i = 0. (156)
( H − E0N )| N i = 0, (157)
where
p2 m C
H= + ( )2 r 2 , (158)
2m 2 m
and
p20 − m2
E0N = . (159)
2m
This form shows that the equations of motion are the same as those of a simple harmonic oscillator
with frequency
ω = C/m. (160)
3
E0N = ( N + )ω, (161)
2
and
3
p20 = 2( N + )C + m2 . (162)
2
where g = 2me2 and D (~k ) is the inverse momentum space propagator for the bound meson:
D ~k, k0 = D (k) = (~p −~k)2 − ( p0 − k0 )2 + C2 r2 + m2 . (164)
~ ~ ~
e−itD(k) = eik·~r e−itD(0, k0 ) e−ik·~r , (166)
where
D ~0, k0 = 2mH + m2 − ( p0 − k0 )2 . (167)
Physics 2020, 2 141
By applying the equations of motion for the canonical variables for an elapsed time equal to mt,
we can compute the translations in Equation (166) explicitly with the result
R∞ ~
1
D (k)−ie0
= i 0 dte−itmH ei2k·~pν e+itmH
2 2 0
(169)
×e−iµk e−it(m −( p0 −k0 ) ) e−e t ,
2
where
sin(Ct) cos(Ct)
µ= C ,
sin(Ct) (170)
ν= C .
The integration over the scalar photon momentum can be performed by completing the square, and
employing the general formula
Z ∞
2 −2ibx ) π ∓ b2 ± i π
dxe±(iax = e a e 4. (171)
−∞ a
where we have used the product representation for the three-dimensional harmonic oscillator states
| N1 N2 N3 i = | N1 i| N2 i| N3 i and N = N1 + N2 + N3 . The quantities σN and Ω are
t2
g 2
− 12 −ip0 t+λ
σN = − ( λ + t ) e , (173)
16π 2
i~p2 ( ν2 )
− 23
Ω N1 N2 N3 = (λ + µ) N1 N2 N3 e λ + µ
N1 N2 N3 .
(174)
We can calculate the matrix elements directly and express the results in terms of the quantity
j
iν2 C
Ω( j) = . (175)
j+ 23
(λ + µ − iν2 C )
We find
Ω000 = Ω (0),
Ω100 = Ω (0) + Ω (1),
(176)
Ω200 = Ω(0) + 2Ω(1) + 32 Ω(2),
Ω110 = Ω(0) + 2Ω(1) + Ω(2).
The radiative shifts lift the degeneracy for some levels and this parameterization simplifies the
calculation of shifts between degenerate levels. The free particle mass shift is contained in Ω(0) to all
orders. This follows by noting that, for j > 0, as C → 0
lim Ω( j) → 0. (177)
C →0
Physics 2020, 2 142
For calculations of the shift between the nondegenerate energy levels, we would use a different
formulation, subtracting the free particle shift in the beginning. To check our equations, we consider the
limit C → 0, which should yield the free particle renormalization. In this limit we have µ → t and ν → t
so the only nonvanishing Ω is
3
Ω (0) → ( λ + t ) − 2 . (178)
Substituting these quantities into the expression for the shift, we find
Z 1 Z ∞
g dt −im2 yt
∆Efree = − dy e , (179)
16π 2 0 0 t
we find
g
2
∆E f ree = ln em + γ − 1, (182)
16π 2
where e in the infinitesimal cutoff for the t integration and γ is Euler’ constant. This result has the same
structure as the conventional result with respect to the divergences. The finite parts depend on the values
of the cutoffs and on the particular procedures used to evaluate the integrals. The infinite terms cancel in
the calculation of measurable shifts and consequently have no direct physical significance.
The expression for the bound state shifts can be rewritten in terms of y and τ = 2Ct:
y
λ + µ − iν2 C = e−iτ + iτ − 1 ,
τ
2Cy 1− iτ
iν2 C = 1
eiτ + e−iτ − 2 ,
4iC
(183)
1
eiτ − e−iτ .
µ = 4iC
which equals
j
dτ y j eiτ + e−iτ − 2 e−iyητ
Z 1 Z ∞
g 1
∆EN ( j) = − dy , (185)
16π 2 (2i ) j 0 2Ce τ j+1 1 − y e−iτ + iτ − 1 j+ 32
iτ
where the degree of coupling to the harmonic oscillator is given by the dimensionless parameter
p20
η= . (186)
2C
The shift can be expressed as a single integral of a confluent hypergeometric function with two
arguments. The structure is similar to that for H atom where the shift can also be expressed in terms of an
integral over a confluent hypergeometric function [11].
Physics 2020, 2 143
d4 k 1
Z D E
∆EN = EN − E0N
µ
= ig N Tµ N , (187)
(2π )4 k2 − ie
where
µ 1
Tµ = (2p − k)µ (2p − k)µ . (188)
D (k) − ie
Executing the trace gives
−→ −
→2
Tµ = 4 pi D1(k) pi − p20 D1(k) −2 −→
h i n o
1
− 4p0 k0 D1(k) − 1
µ
p · k, D (k)
k − k20 D (k)
. (189)
We can derive expressions for each of these quantities in terms of our previous results by employing
the Heisenberg equations of motion for pi and qi (Equation (168)) and also our form of the Klein–Gordon
equation (Equation (156)). Our final result is
Z ∞ Z ∞
3
∆EN1 N2 N3 = 4 dλ dt − p20 − C2 ν2 2C N + − 3iµC2
0 0 2
t λ 1 ∂
− p20 + 2µνC2 +
t+λ ν 2i ∂ν
1 (190)
∂
+C2 ν2 2C2 ν2 − 1 σ Ω
−i ∂(λ + µ) N N1 N2 N3
1
Z
+ dtσN Ω N1 N2 N3 |λ=0 ,
i
where σN and Ω N1 N2 N3 have the same meaning as before (Equations (173) and (174)).
7. Conclusions
We discuss the history of Lamb shift and Bethe’s pivotal calculation, and how it influenced the
direction of theoretical physics for over half a century.
We discuss the general nature of radiative shifts of bound state energy levels, from the classical and
the quantum perspectives, examining in some detail results for the harmonic oscillator and the hydrogen
atom. The radiative shifts are complex, the real part being the level shift and the imaginary part being
the level width. The shifts arise because of the emission and absorption of virtual photons which occurs
due to interaction of the charged particle with its own radiation field, or, equivalently, with the vacuum
zero-point fluctuations. We know vacuum fluctuations are affected by geometry and therefore radiative
shifts differ from free space values for atoms in a cavity, for example, or near a surface [105–108,110,121].
Lamb shifts have even been used to model gravitational energy in black holes [122].
Today, the computation of radiative shifts and atomic energy levels can be done very precisely, from 1
part in 1012 to 1 part in 1015 for certain energy levels, the most precise computations for any physical system
[20]. Today, the corresponding experiments demonstrate comparable precision. Some see the opportunity
for developing metrology [123–127]. This favorable situation allows atomic systems to be a platform for
the discovery of new physics beyond the standard model. Theoreticians are already calculating the effect
on energy levels due to the quantization of space, the non-commutativity of space-time coordinates and
space-time fluctuations for H atoms, muonic atoms and Rydberg states [128–135]. Measurements are being
done on collaborative Lamb shifts for mesoscopic arrays [136,137].
Physics 2020, 2 144
Because of this high precision, measurements of radiative shifts and atomic energy levels reveal
detailed information about phenomena causing shifts aside from radiative effects. This precision has
led to a new understanding low Z two body systems, including muonium, positronium, and tritium,
revealing nuclear structure effects and other higher order effects. We can expect that atomic energy level
measurements and computations will continue to contribute significantly to the development of quantum
physics in the future.
Appendix
We thought a few selected comments about the lives of Lamb and Bethe might help frame their
activities during the years in which they played such key roles in the development of QED.
federal position forced Bethe to leave. In 1935, he joined the physics faculty at Cornell, and enjoyed the
atmosphere very much, and remained there for most of his career. During WWII, he served as head of the
Theoretical Division at Los Alamos, under Oppenheimer. Bethe won the Nobel Prize in physics in 1967 for
“for his contributions to the theory of nuclear reactions, especially his discoveries concerning the energy
production in stars.” He explained why the sun keeps shining, and did not win it for his contributions to
QED. In later years, he advocated for peaceful use of nuclear energy and nuclear disarmament. He died in
2005 at age 98.
26. Brown, L. Quantum Field Theory; Cambridge Univ. Press: Cambridge, UK, 1992.
27. Bjorken, J.; Drell, S. Relativistic Quantum Mechanics; McGraw Hill: Neywork, NY, USA, 1964; Classic discussion
of remormalization in terms of Feynman diagrams.
28. Low, F. Natural Line Shape. Phys. Rev. 1952, 88, 53. [CrossRef]
29. Oppenheimer, J.R. Note on the Theory of the Interaction of Field and Matter. Phys. Rev. 1930, 35, 461. [CrossRef]
30. Kramers, H.A. Subtraction of infinities. Nuovo Cim. 1938, 15, 108. [CrossRef]
31. Calculating the Lamb Shift, Videos 104-107 of Hans Bethe Scientist. Recorded by Sam Schweber. December 1996.
Available online: webofstories.com (accessed on 15 March 2020).
32. Salam, A.; Bethe, H.; Dirac, P.; Heisenberg, W.; Wigner, E.; Klein O.; Lifshitz, E. Energy on Earth and in the Stars.
In From a Life in Physics; World Scientific: Singapore, 1989.
33. Weisskopf, V. The Development of Field Theory in the Last 50 Years. Phys. Today 1981, 34, 69–85. [CrossRef]
34. Weisskopf, V. On the Self-Energy and the Electromagnetic Field of the Electron. Phys. Rev. 1939, 56, 72. [CrossRef]
35. Isidor Isaac Rabi was awarded the 1944 Nobel Prize in physics for his development of the atomic and molecular
beam magnetic resonance method of observing atomic spectra.
36. Videos of Freeman Dyson 64-65 discussing Hans Bethe. Recorded by Sam Schweber. June 1998. Available online:
webofstories.com (accessed on 15 March 2020).
37. Dyson, F. Hans Bethe and Quantum Electrodynamics. Phys. Today 2005, 58, 48–51. [CrossRef]
38. Feynman, R. The Development of the Space-Time View of Quantum Electrodynamics. Phys. Today 1966, 19(8), 31. [CrossRef]
39. Schweber, S. QED and the Men Who Made It: Dyson, Feynman, Schwinger, and Tomonaga; Princeton Univ. Press:
Princeton, NJ, USA, 1994.
40. We generally employ natural Gaussian units so h̄ = 1, c = 1, and α = (e2 /h̄c) ≈ 1/137. The notation for indices
is µ, ν, .. = 0, 1, 2, 3; i, j, . = 1, 2, 3; pµ pµ = − p20 + ~p2 . Occasionally we will show factors of h̄ and c for clarity.
41. Serber, R. Linear Modifications in the Maxwell Field Equations. Phys. Rev. 1935, 48, 49. [CrossRef]
42. Uehling, E. Polarization Effects in the Positron Theory. Phys. Rev. 1935, 48, 55. [CrossRef]
43. Dyson, F. The Electromagnetic Shift of Energy Levels. Phy. Rev. 1948, 73, 617. [CrossRef]
44. French, J.; Weisskopf, V. The Electromagnetic Shift of Energy Levels. Phys.Rev. 1949, 75, 1240. [CrossRef]
45. Kroll, N.; Lamb, W.E., Jr. On the Self-Energy of a Bound Electron. Phys. Rev. 1949, 75, 388. [CrossRef]
46. Schwinger, J. Quantum electrodynamics. I. A covariant formulation. Phy. Rev. 1948, 74, 1439. [CrossRef]
47. Dyson, F. The Radiation Theories of Tomonaga, Schwinger, and Feynman. Phys. Rev. 1949, 75, 486. [CrossRef]
48. Dyson, F. The S-Matrix in Quantum Mechanics. Phy. Rev. 1949, 75, 1736. [CrossRef]
49. Dyson, F. Makers of Patterns, An Autobiography through Letters; Liveright Publishing: New York, NY, USA, 2018.
50. Welton, T. Some observable effects of the quantum-mechanical fluctuations of the electromagnetic field. Phys. Rev.
1948, 74, 1157. [CrossRef]
51. Baranger, M.; Bethe, H.A.; Feynman, R.P. Relativistic Correction to the Lamb Shift. Phys. Rev. 1953, 92, 482. [CrossRef]
52. Karplus, R.; Klein, A.; Schwinger, J. Electrodynamic Displacement of Atomic Energy Levels. II. Lamb Shift. Phys.
Rev. 1952, 86, 288. [CrossRef]
53. Wichmann, E.H.; Kroll, N.M. Vacuum Polarization in a Strong Coulomb Field. Phys. Rev. 1956, 101, 843. [CrossRef]
54. Erickson, W. Improved Lamb-Shift Calculation for All Values of Zα. Phys. Rev. Lett. 1971, 27, 780. [CrossRef]
55. Power, E.A. Zero-Point Energy and the Lamb Shift. Am. J. Phys. 1966, 34, 516. [CrossRef]
56. Weisskopf, V. The Privilege of Being a Physicist; W.H Freeman and Co.: New York, NY, USA, 1989.
57. Mohr, P.J.; Plunien, G.; Soff, G. QED corrections in heavy atoms. Phys. Rep. 1998, 293, 227. [CrossRef]
58. Mohr, P.H. Self-energy correction to one-electron energy levels in a strong Coulomb field. Phys. Rev. A 1992, 46, 4421.
59. Karshenboim, S.G. The Lamb shift of excited S-levels in hydrogen and deuterium atoms. Z. Phys. D 1997, 39, 109. [CrossRef]
60. Karshenboim, S.G. Two-loop logarithmic corrections in the hydrogen Lamb shift. J. Phys. B At. Mol. Opt. Phys.
1996, 29, L29–L31. [CrossRef]
61. Berkeland, D.; Hind, E.; Boshierm, M. Precision optical measurement of lamb shifts in atomic hydrogen. Phys.
Rev. Lett. 1995, 75, 2470. [CrossRef]
62. Karshenboim, S.G. Lamb Shift in the Hydrogen-Atom—Leading Logarithmic Corrections. Phys. At. Nucl. 1995, 58, 649–653.
63. Karshenboim, S.G. Lamb Shift in the Hydrogen-Atom—Lifetime of the 2P(1/2) Level. Phys. At. Nucl. 1995, 58, 835–838.
Physics 2020, 2 147
64. Karshenboim, S.G. The Lamb Shift in the Hydrogen-Atom—Shift. Phys. At. Nucl. 1995, 58, 262–266.
65. Eides, M.I.; Grotch, H. Corrections of order α2 ( Zα)4 and α2 ( Zα)6 to the Lamb shift. Phys. Rev. A 1995, 52, 3360.
[CrossRef] [PubMed]
66. Karshenboim, S.G. Lamb Shift in Hydrogen-Atom. JETP 1994, 79, 230–236; Available online: http://www.jetp.ac.
ru/cgi-bin/dn/e_079_02_0230.pdf (accessed on 15 March 2020).
67. Pachucki, K. Higher-Order Binding Corrections to the Lamb Shift. Ann. Phys. (NY) 1993, 226, 1–87. [CrossRef]
68. Palchikov, V.G.; Sokolov, Y.L.; Yakovlev, V.P. Lifetime of the 2P State and Lamb Shift in the Hydrogen-Atom. JETP
Lett. 1983, 38, 418–420; Available online: http://www.jetpletters.ac.ru/ps/1483/article_22631.pdf (accessed on
15 March 2020).
69. Lundeen, S.R.; Pipkin, F.M. Measurement of the Lamb Shift in Hydrogen, n=2. Phys. Rev. Lett. 1981, 46, 232.
70. Drake, G. Quantum electrodynamic Effects in Few-Electron Atomic Systems. Adv. At. Mol. Phys. 1982, 18, 399.
71. Grotch, H. Lamb Shift in Nonrelativistic Quantum Electrodynamics. Am. J. Phys. 1981, 49, 48–51. [CrossRef]
72. Sapirstein, J. Higher-Order Binding Corrections to the Lamb Shift. Phys. Rev. Lett. 1981, 47, 1723. [CrossRef]
73. Schwebel, S. Interaction Theory-Relativistic Hydrogen-Atom and the Lamb Shift. Int. J. Theor. Phys. 1978, 17,
931–939. [CrossRef]
74. Davies, B. Note on the Lamb shift. Am. J. Phys. 1982, 50, 331. [CrossRef]
75. Karshenboim, S.G.; Ozawa, A.; Ivanov, V.G. Higher-order logarithmic corrections and the two-loop self-energy
of a 1s electron in hydrogen. Phys. Rev. A 2019, 100, 032515. [CrossRef]
76. Karshenboim, S.G.; Shelyuto, V.A. Three-loop radiative corrections to the 1s Lamb shift in hydrogen. Phys. Rev. A
2019, 100, 032513. [CrossRef]
77. Eides, M.I.; Shelyuto, V.A. Hard three-loop corrections to hyperfine splitting in positronium and muonium. Phys.
Rev. D 2015, 92, 013010. [CrossRef]
78. Noble, J.H.; Jentschura, U.D. Dirac equations with confining potentials. Int. J. Mod. Phys. A 2015, 30, 1550002. [CrossRef]
79. Eides, M.I.; Shelyuto, V.A. Polarization operator contributions to the Lamb shift and hyperfine splitting. Phys.
Rev. A 2003, 68, 042106. [CrossRef]
80. Szafron, R.; Korzinin, E.Y.; Shelyuto, V.A.; Ivanov, V.G.; Karshenboim, S.G. Virtual Delbruck scattering and the
Lamb shift in light hydrogenlike atoms. Phys. Rev. A 2019, 100, 032507; Preprint in arxiv:1909.04116. Many
papers of Prof. Karshenboim and collaborators are available on Arxiv. [CrossRef]
81. Zamastil, J.; Patkos, V. Self-energy of an electron bound in a Coulomb field. Phys. Rev. A 2013, 88, 032501. [CrossRef]
82. Zamastil, J. Approximate numerical calculation of the self-energy of a bound electron. Ann. Phys. 2012, 327, 297–328.
83. Jentschura, U.D.; Mohr, P.J. Calculation of hydrogenic Bethe logarithms for Rydberg states. Phys. Rev. A 2005, 72,
012110. [CrossRef]
84. Jentschura, U.D. Techniques in analytic Lamb shift calculations. Mod. Phys. Lett. A 2005, 20, 2261–2276. [CrossRef]
85. Jentschura, U.D.; Mohr, P.J.; Soff, G. Electron self-energy for the K and L shells at low nuclear charge. Phys. Rev.
A 2001, 63, 042512. [CrossRef]
86. Jentschura, U.D.; Mohr, P.J. Electron self-energy for higher excited S levels. Phys. Rev. A 2004, 69, 064103. [CrossRef]
87. Eides, M.I.; Grotch, H.; Shelyuto, V.A. Radiative-recoil corrections of order α( Zα)5 (m/M)m to the Lamb shift
revisited. Phys. Rev. A 2002, 63, 052509. [CrossRef]
88. Holstein, B.R. Effective interactions and the hydrogen atom. Am. J. Phys. 2004, 72, 333–344. [CrossRef]
89. Jentschura, U.D.; Bigot, E.O.L.; Evers, J.; Mohr, P.J.; Keitel, C.H. Relativistic and radiative energy shifts for
Rydberg states. J. Phys. B At. Mol. Opt. Phys. 2005, 38, S97–S105. [CrossRef]
90. Agafonov, A.I. Hydrogen energy-level shifts induced by the atom motion: Crossover from the Lamb shifts to the
motion-induced shifts. Mod. Phys. Lett. B 2018, 32, 1850273. [CrossRef]
91. Kelkar, G.; Mart, T.; Nowakowski, M. Extraction of the Proton Charge Radius from Experiments. Makara J. Sci.
2016, 20, 119–126. [CrossRef]
92. Martynenko, A.P. Proton-polarizability effect in the lamb shift for the hydrogen atom. Phys. At. Nucl. 2006, 69,
1309–1316. [CrossRef]
93. The Alpha Collaboration. Investigation of the fine structure of antihydrogen, Nature 2020, 578, 375–380. [CrossRef]
94. Feynman, R. QED The strange theory of light and matter; Princeton University Press: Princeton, NJ, USA, 1988; p. 128.
Physics 2020, 2 148
95. The quantization volume V is an artifice to avoid infinite volumes. In this box normalization k x = 2πn x /L x ,
k y = 2πny /Ly , and k z = 2πnz /Lz , with V = L x Ly Lz , and the integers n x , ny , and nz go from −∞ to +∞.
96. Jackson, J. Classical Electrodynamics; Wiley and Sons: New York, NY, USA, 1962; Chapter 17.
97. The classical radius of the electron is rcl = e2 /(mc2 ) = 2.8 × 10−13 cm, which can be written at αh̄/mc = αλ where
λ is the reduced Compton wavelength of the electron 3.8 × 10−11 cm (Compton wavelength divided by 2π).
98. Cole, D.; Zou, Y. Quantum mechanical ground state of hydrogen obtained from classical electrodynamics. Phys.
Lett. A 2003, 317, 14–20. [CrossRef]
99. Boyer, T. Classical Zero-Point Radiation and Relativity:The Problem of Atomic Collapse Revisited. Found. Phys.
2016, 46, 880-890. [CrossRef]
100. An alternative theory to quantum mechanics, Stochastic Electrodynamics (SED) posits that the vacuum
fluctuations are a real, not a virtual, electromagnetic field, and that this field provides the energy lost by
radiation and is responsible for the stability of atoms. See [98,99] for information on SED and atomic stability.
For a critical evaluation for the H atom see Maclay, J. The Role of Vacuum Fluctuations and Symmetry in the
Hydrogen Atom in Quantum Mechanics and Stochastic Electrodynamics. Atoms 2019, 7, 39. [CrossRef]
101. We also mention the vacuum fluctuations of the charge density, characterized by virtual electron-positron pairs, which
leads to the renormalization of the electron charge. Since this charge renormalization contributes much less to the shift
between states than the mass renormalization from the zero point vibrations of the EM field, we shall not consider it
here. In mesic atoms, in which the meson orbit is largely within the nucleus, the conversé situation obtains.
102. Weisskopf, V. Recent Developments in the Theory of the Electron. Rev. Mod. Phys. 1949, 21, 305. [CrossRef]
103. This expansion is essentially the dipole approximation.
104. The metric is (−1, 1, 1, 1) for µ = 0, 1, 2, 3.
105. Barton, G. New Aspects of the Casimir Effect: Fluctuations and Radiative Reaction. In Cavity Quantum Electrodynamics;
Berman, P., Ed.; Academic Press: San Diego, CA, USA, 1994; pp. 425–455. This gives a clear discussion of how changes in
the vacuum field due to surfaces affect charge, magnetic moment, mass and energy levels.
106. Bordag, M.; Klimchitskaya, G.; Mohideen, U.; Mostepanenko, V. Advances in the Casimir Effect; Oxford Univ.
Press: New York, NY, USA, 2009. This book gives a very complete discussion of how surfaces affect vacuum
energy and can lead to Casimir forces between surfaces.
107. Milton, K. The Casimir Effect, Physical Manifestations of Zero-Point Energy; World Scientific: Singapore, 2001.
108. Milton, K.; Bordag, M. (Eds.) Quantum Field Theory Under the Influence of External Conditions; World Scientific:
Singapore, 2010; This book discusses a broad variety of systems, including gravitational and nuclear.
109. The shift is also temperature dependent since the vacuum field has a temperature dependent component.
110. Billaud, B.; Truong, T.-T. Lamb shift of non-degenerate energy level systems placed between two infinite parallel
conducting plates. J. Phys. A Math. Theor. 2013, 46, 025306. [CrossRef]
111. Maclay, J. The Symmetry of the Energy Levels of the Hydrogen Atom and the Application of the Symmetry to
the Calculation of Radiative Level Shifts. Ph.D. Thesis, Yale University, New Haven, CN, USA, 1972.
112. The primes indicate eigenvalues of operators, and unprimed quantities indicate abstract operators. The quantity
x 0 means the four-vector (t0 , ~r 0 ) and the volume element is d4 x 0 = dt0 dx 0 dy0 dz0 . The partial derivative is
defined as ∂0 = ∂/∂x 0 , and ∂02 = −∂2 /∂2 t0 + ∂2 /∂2 x 0 + ∂2 /∂2 y0 + ∂2 /∂2 z0 . We also may write this operator as
∂2ξ 0 = −∂2 /∂2 ξ 00 + ∂2 /∂2 ξ 10 + ∂2 /∂2 ξ 30 + ∂2 /∂2 ξ 40 for an arbitrary four-vector ξ.
113. To validate this expression for G c we operate on the integral with Π02 + m2 . We observe Π0k < r 0 | =< r 0 |Πk ,
Π00 < r 0 | =< r 0 |Π0 so (Π02 + m2 ) < r 0 | =< r 0 |(Π2 + m2 ). With the normalization < r 0 |r 00 >= δ(r 0 − r 00 ),
it follows the integral obeys the defining equation for G c .
114. Note that the sign of the energy shift is positive. This seems to contradict the rule that a perturbation must lower
the ground state energy. The rule holds, however, if we consider the total perturbation to be the unrenormalized
(mass)2 operator not the renormalized operator.
115. The time-energy relationship is not an uncertainly principle in the same sense as the position-momentum
uncertainty principle, which follows because the corresponding operators do not commute. The time-energy
relationship arises from the properties of Fourier transforms and is consistent with the position-momentum
uncertainty principle.
Physics 2020, 2 149
116. For the H atom, the value of (1/π ) ln(( En − Em )/ωc ) is roughly −3 assuming the cutoff is at hωc = mc2 and n
and m are adjacent bound state energy levels.
117. The Bethe log is commonly written as ln (K0 /Z2 Ry ) where K0 refers to the average excitation energy [ EN − Em ] ave ,
and Ry is the Rydberg constant.
118. Maclay, J., Revisiting the Symmetry of the H Atom: SO(4) to SO(4,2), and Its Use to Calculate Radiative Shifts, to
be submitted for publication in 2020 in the upcoming special issue on “Symmetries in Quantum Mechanics” of
the open access journal Symmetry.
119. Došlić, N and Danko, S Harmonic oscillator with the radiation reaction interaction Phys. Rev. A 1995 51, 3485.
120. Daeimohamad, M.; Mohammadi, M. Quantum Dynamics of a Harmonic Oscillator in a Deformed Bath in the
Presence of Lamb Shift. Int. J. Theor. Phys. 2012, 51, 3052–3061. [CrossRef]
121. Koshino, K.; Nakamura, Y. Control of the radiative level shift and linewidth of a superconducting artificial atom
through a variable boundary condition. New J. Phys. 2012, 14, 043005. [CrossRef]
122. Porto, R.A. Lamb shift and the gravitational binding energy for binary black holes. Phys. Rev. D 2017, 96, 024063.
123. Cagnac, B. Hydrogen metrology: Up to what limit? Phys. Scr. 1997, T70, 24–33. [CrossRef]
124. Cagnac, B. Progress on the Rydberg Constant—THE Hydeogen-Atom as a Freauency Standard. IEEE Trans.
Instrum. Meas. 1993, 42, 206–213. [CrossRef]
125. Cagnac, B.; Plimmer, M.D.; Julien, L.; Biraben, F. The Hydrogen-Atom, a Tool for Metrology. Rep. Prog. Phys.
1994, 57, 853–893. [CrossRef]
126. Hagel, G.; Schwob, C.; Jozefowski, L.; de Beauvoir, B.; Hilico, L.; Nez, F.; Julien, L.; Biraben, F.; Acef, O.; Clairon, A.
Metrology of hydrogen atom: Determination of the Rydberg constant and Lamb shifts. Laser Phys. 2001, 11, 1076–1082.
127. Mohr, P.; Taylor, B. Quantum electrodynamics and the fundamental constants. Adv. Quantum Chem. 1998, 30, 77–97.
128. Jones, M.; Potvliege, R.M.; Spannowsky, M. Probing new physics using Rydberg states of atomic hydrogen. Phys.
Rev. Res. 2020, 2, 013244. [CrossRef]
129. Alavi, A.; Rezaei, N. Dirac equation, hydrogen atom spectrum and the Lamb shift in dynamical non-commutative
spaces. Pramana J. Phys. 2017, 88, 77. [CrossRef]
130. Haghighat, M.; Khorsandi, M. Hydrogen and muonic hydrogen atomic spectra in non-commutative space-time.
Eur. Phys. J. C 2015, 75, 4. [CrossRef]
131. Gnatenko, K.P.; Krynytskyi, Y.S.; Tkachuk, V.M. Perturbation of the ns levels of the hydrogen atom in rotationally
invariant noncommutative space. Mod. Phys. Lett. A 2015, 30, 1550033. [CrossRef]
132. Rivas, I.; Camacho, A.; Goklu, E. Quantum spacetime fluctuations: Lamb shift and hyperfine structure of the
hydrogen atom. Phys. Rev. D 2011, 84, 055024. [CrossRef]
133. Zaim, S.; Khodja, L.; Delenda, Y. Second-Order Corrections to the Noncommutative Klein-Gordon Equation with
a Coulomb Potential. Int. J. Mod. Phys. A 2011, 26, 4133–4144. [CrossRef]
134. Bouaziz, D.; Ferkous, N. Hydrogen atom in momentum space with a minimal length. Phys. Rev. A 2010, 82, 022105.
135. Chaichian, M.; Sheikh-Jabbari, M.M.; Tureanu, A. Hydrogen atom spectrum and the Lamb shift in
noncommutative QED. Phys. Rev. Lett. 2001, 86, 2716–2719. [CrossRef] [PubMed]
136. Scully, M.O. Collective Lamb shift in single photon Dicke superradiance. Phys. Rev. Lett. 2009, 102, 143601. [CrossRef]
137. Meir, Z.; Schwartz, O.; Shahmoon, E.; Oron, D.; Ozeri, R. Cooperative Lamb Shift in a Mesocopic Atomic Array.
Phy. Rev. Let. 2014, 113, 193002. [CrossRef]
138. Milonni, P. Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA. Personal
communication, 2020.
139. Lamb, W. Super classical quantum mechanics: The best interpretation of nonrelativistic quantum mechanics. Am.
J. Phys. 2001, 69, 413–422. [CrossRef]
140. Schweber, S. The Happy Thirties. Phys. Today 2005, 58, 38. [CrossRef]
c 2020 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article
distributed under the terms and conditions of the Creative Commons Attribution (CC BY)
license (http://creativecommons.org/licenses/by/4.0/).