Difference between revisions of "Derivation of the QED Hamiltonian"
imported>Ichuang (New page: We describe here a rigorous derivation of atom-photon interactions. Our overall goal is to arrive at a Hamiltonian description for the energy of a system of atoms, photons, and atoms inte...) |
imported>Yanzhou |
||
(9 intermediate revisions by 3 users not shown) | |||
Line 1: | Line 1: | ||
+ | <categorytree mode=pages style="float:right; clear:right; margin-left:1ex; border:1px solid gray; padding:0.7ex; background-color:white;" hideprefix=auto>8.422</categorytree> | ||
+ | |||
We describe here a rigorous derivation of atom-photon | We describe here a rigorous derivation of atom-photon | ||
interactions. Our | interactions. Our | ||
Line 15: | Line 17: | ||
field will be quantized, and described by operators <math>a</math> and | field will be quantized, and described by operators <math>a</math> and | ||
<math>a^\dagger</math>. | <math>a^\dagger</math>. | ||
+ | |||
+ | This result is simple and can be obtained with less rigorous derivations. Here, we want to at least mention all the steps of the rigorous derivation, starting from classical Maxwell's equations. | ||
== Quantum electrodynamics == | == Quantum electrodynamics == | ||
Line 39: | Line 43: | ||
\end{array}</math> | \end{array}</math> | ||
How many of the field components are true degrees of freedom? We can | How many of the field components are true degrees of freedom? We can | ||
− | understand this by taking a spatial Fourier transform. It turns out | + | understand this by taking a spatial Fourier transform. The component of the (vectorial) Fourier component, which is parallel to the k vector, is called the longitudinal part, the component orthogonal to the k vector is the transverse part. By transforming back to the spatial domain, we can distinguish the longitudinal and transverse parts. It turns out |
− | that the longitudinal electric field is not a free degree of freedom. | + | that the longitudinal electric field is not a free degree of freedom, i.e. it follows instantaneously the positions of the particles through the Coulomb potential. |
− | + | Introducing the vector potential reduces the number of field components to four. | |
:<math>\begin{array}{rcl} | :<math>\begin{array}{rcl} | ||
E(r,t) &=& \nabla U(r,t) - \partial_t A(r,t) | E(r,t) &=& \nabla U(r,t) - \partial_t A(r,t) | ||
Line 49: | Line 53: | ||
\,. | \,. | ||
\end{array}</math> | \end{array}</math> | ||
− | In the Coulomb gauge, <math> \nabla \cdot A =0</math>, | + | In the Coulomb gauge, <math> \nabla \cdot A =0</math>, i.e. the longitudinal component of A vanished. Furthermore, the scalar potential is |
− | is | ||
:<math> | :<math> | ||
U(r,t) = \frac{1}{4\pi\epsilon_0} \int d^3r' \frac{\rho(r',t)}{|r-r'|} | U(r,t) = \frac{1}{4\pi\epsilon_0} \int d^3r' \frac{\rho(r',t)}{|r-r'|} | ||
\,. | \,. | ||
</math> | </math> | ||
+ | Now we are down to two components, the transverse components of the vector poential A. | ||
+ | |||
Keep in mind that the equation of motion is now a second order | Keep in mind that the equation of motion is now a second order | ||
differential equation for the vector potential. All that is needed to | differential equation for the vector potential. All that is needed to | ||
specify the field evolution is thus the initial values of <math>A_\perp</math> | specify the field evolution is thus the initial values of <math>A_\perp</math> | ||
− | and <math>\dot{A}_\perp</math>, the transverse vector potential. | + | and <math>\dot{A}_\perp</math>, the transverse vector potential. The differential equation is identical to the equation for r and p for an harmonic oscillator potential. |
− | |||
− | potential. | ||
− | We can further | + | We can go further by decomposing the Fourier |
representation of the field and potential in terms of its normal modes, | representation of the field and potential in terms of its normal modes, | ||
:<math>\begin{array}{rcl} | :<math>\begin{array}{rcl} | ||
Line 80: | Line 83: | ||
:<math> | :<math> | ||
\vec{A}_\perp(r) = \int d^3k \sum_\epsilon A_\omega [ | \vec{A}_\perp(r) = \int d^3k \sum_\epsilon A_\omega [ | ||
− | \vec{\epsilon} \alpha_\epsilon(k) e^{ | + | \vec{\epsilon} \alpha_\epsilon(k) e^{ikr} + h.c. ] |
\,. | \,. | ||
</math> | </math> | ||
Line 88: | Line 91: | ||
Let us now quantize the field. We identify an equivalence between <math>A</math> | Let us now quantize the field. We identify an equivalence between <math>A</math> | ||
− | and <math>\dot{A}</math> with <math>r</math> and <math>p</math>, and quantize accordingly. The | + | and <math>\dot{A}</math> with <math>r</math> and <math>p</math>, and quantize accordingly. The normal mode coordinates are a linear combination of <math>A</math> and <math>\dot{A}</math> from which we obtain the commutator |
− | commutator | ||
:<math> | :<math> | ||
[a_ \epsilon (k),a_{ \epsilon '}^\dagger(k')] = \delta_{ \epsilon \epsilon '} \delta(k-k') | [a_ \epsilon (k),a_{ \epsilon '}^\dagger(k')] = \delta_{ \epsilon \epsilon '} \delta(k-k') | ||
\,. | \,. | ||
</math> | </math> | ||
+ | in full analogy to the raising and lowering operators for the simple harmonic oscillator. | ||
+ | |||
== Energy in the radiation field == | == Energy in the radiation field == | ||
It is helpful to go back to consider for a moment what the energy in | It is helpful to go back to consider for a moment what the energy in | ||
Line 108: | Line 112: | ||
\frac{\rho(r)\rho(r')}{|r-r'|} | \frac{\rho(r)\rho(r')}{|r-r'|} | ||
</math> | </math> | ||
− | as the energy of the Coulomb field due to the charge configuration. | + | as the energy of the instantaneous Coulomb field due to the charge configuration. |
− | + | The second term is the energy of the transverse component | |
− | |||
of the field, <math>H_{\rm trans}</math>, the radiation energy, which we can | of the field, <math>H_{\rm trans}</math>, the radiation energy, which we can | ||
understand by introducing again our expression for the vector | understand by introducing again our expression for the vector | ||
Line 123: | Line 126: | ||
Hamiltonian. Now introduce normal modes, using | Hamiltonian. Now introduce normal modes, using | ||
:<math> | :<math> | ||
− | \alpha \sim [ -\dot{A}_\perp + i\omega | + | \alpha \sim [ -\dot{A}_\perp + i\omega A_\perp ] |
\,. | \,. | ||
</math> | </math> | ||
Line 133: | Line 136: | ||
\,, | \,, | ||
\end{array}</math> | \end{array}</math> | ||
− | a purely classical expression for energy in the radiation | + | a purely classical expression for the energy in the radiation |
− | field. | + | field. But it looks quantum. Where does <math>\hbar</math> come from? It enters in the constant |
relating the normal modes with <math>\alpha</math>. It is just a unit used in | relating the normal modes with <math>\alpha</math>. It is just a unit used in | ||
the definitions at this moment, which is convenient to use because | the definitions at this moment, which is convenient to use because | ||
Line 155: | Line 158: | ||
When quantized, this becomes <math>H = \hbar\omega ( a^\dagger a + 1/2)</math>. | When quantized, this becomes <math>H = \hbar\omega ( a^\dagger a + 1/2)</math>. | ||
We conclude that the radiation field is just a bunch of oscillators, | We conclude that the radiation field is just a bunch of oscillators, | ||
− | with one per <math>k</math> vector, and each one is described in its quantized | + | with one per <math>k</math> vector and polarization, and each one is described in its quantized |
form by the Hamiltonian <math>H = \hbar\omega ( a^\dagger _k a_k + 1/2)</math>. | form by the Hamiltonian <math>H = \hbar\omega ( a^\dagger _k a_k + 1/2)</math>. | ||
− | |||
− | |||
− | |||
− | |||
− | |||
− | |||
− | |||
− | + | The most important aspect of our derivation was the separation of the fields into longitudinal and transverse components. For example, a naive approach using Cartesian coordinates, | |
− | |||
without eliminating the longitudinal field, would fail miserably, | without eliminating the longitudinal field, would fail miserably, | ||
because | because | ||
Line 173: | Line 168: | ||
\,. | \,. | ||
</math> | </math> | ||
− | The treatment above, | + | The treatment above, eliminating the dependent components, is thus essential. |
− | + | ||
+ | Also note that quantizing the electromagnetic field is only possible if we have an expression for the energy of the radiation field that also corresponds to a valid Lagrangian for the system. | ||
+ | Finally, our approach is not relativistically covariant. There are covariant formulations of the quantized radiation field, but they are more complicated. | ||
+ | |||
+ | == Quantum description of the radiation field == | ||
+ | |||
+ | The inverse Fourier transform provides us with the field components in terms of the quantum operators, | ||
:<math>\begin{array}{rcl} | :<math>\begin{array}{rcl} | ||
A_\perp(r) &=& \int d^3k \sum_ \epsilon A_\omega [ \epsilon a_ \epsilon (k) | A_\perp(r) &=& \int d^3k \sum_ \epsilon A_\omega [ \epsilon a_ \epsilon (k) | ||
Line 209: | Line 210: | ||
momentum <math>p</math> with <math>p-qA</math>. | momentum <math>p</math> with <math>p-qA</math>. | ||
The Coulomb interaction energy is standard. | The Coulomb interaction energy is standard. | ||
− | The radiation field energy is <math>H_R \sum_i \hbar\omega_i ( a^\dagger _i a_i + | + | The radiation field energy is <math>H_R = \sum_i \hbar\omega_i ( a^\dagger _i a_i + |
1/2)</math>. | 1/2)</math>. | ||
− | + | ||
+ | We may write the total Hamiltonian as a sum of parts | ||
:<math> | :<math> | ||
H = H_P + H_R + H_I | H = H_P + H_R + H_I | ||
\,, | \,, | ||
</math> | </math> | ||
− | where <math>H_P</math> is the particle Hamiltonian | + | where <math>H_P</math> is the particle Hamiltonian including the Coulomb field |
:<math> | :<math> | ||
H_P = \sum_\alpha \frac{p^2_\alpha}{2m_\alpha} + V_{\rm Coul} | H_P = \sum_\alpha \frac{p^2_\alpha}{2m_\alpha} + V_{\rm Coul} | ||
Line 230: | Line 232: | ||
H_{I1}^S &=& -\sum_\alpha g_\alpha \frac{q_\alpha}{2m_\alpha} | H_{I1}^S &=& -\sum_\alpha g_\alpha \frac{q_\alpha}{2m_\alpha} | ||
S_\alpha \cdot B(r_\alpha) | S_\alpha \cdot B(r_\alpha) | ||
− | |||
\\ | \\ | ||
H_{I2} &=& \sum_\alpha \frac{q_\alpha}{2m_\alpha} A^2_\perp(r_\alpha) | H_{I2} &=& \sum_\alpha \frac{q_\alpha}{2m_\alpha} A^2_\perp(r_\alpha) | ||
Line 236: | Line 237: | ||
\end{array}</math> | \end{array}</math> | ||
For atoms, <math>H_{I1}\gg H_{I1}^S, H_{I2}</math>, typically. | For atoms, <math>H_{I1}\gg H_{I1}^S, H_{I2}</math>, typically. | ||
+ | We will often perform perturbation theory in <math>H_I</math>. Note that this is NOT a perturbation theory in charge or field strength, since the Coulomb field has been separated off and is fully included in the particle Hamiltonian. | ||
== The dipole approximation == | == The dipole approximation == | ||
Line 267: | Line 269: | ||
\,, | \,, | ||
</math> | </math> | ||
− | which in the limit of a resonant interaction, becomes | + | which in the limit of a near-resonant interaction, becomes |
:<math> | :<math> | ||
H_{int} = q\vec{E}_0 \langle 2|\vec{r}|1 \rangle \sim -\vec{d}\cdot\vec{E} | H_{int} = q\vec{E}_0 \langle 2|\vec{r}|1 \rangle \sim -\vec{d}\cdot\vec{E} | ||
Line 277: | Line 279: | ||
A rigorous way to obtain the full solution, which is essentially the | A rigorous way to obtain the full solution, which is essentially the | ||
− | same as that sketched above, is given in API. It involves using a | + | same as that sketched above, is given in API. It shows that the simple <math> \vec{d}\cdot\vec{E} </math> is exact (in the long wavelength approximation) and includes the <math>H_{I2}</math> term. |
+ | |||
+ | The derivation involves using a | ||
canonical transformation with the operator | canonical transformation with the operator | ||
:<math> | :<math> | ||
Line 311: | Line 315: | ||
Note that this formulation already takes into account the | Note that this formulation already takes into account the | ||
polarizability of matter, and the relation between <math>E</math> and <math>D </math>. | polarizability of matter, and the relation between <math>E</math> and <math>D </math>. | ||
+ | |||
+ | Conclusion: What you should remember from this lengthy derivation is the separation into longitudinal and transverse fields, the fact that the Coulomb interaction is included in the particle Hamiltonian, and that the electric dipole approximation includes the A^2 term. | ||
+ | |||
+ | |||
+ | === References === | ||
+ | |||
+ | [[Category:Photon-atom interactions]] | ||
+ | |||
+ | * [https://cua-admin.mit.edu:8443/wiki/images/8/88/API_appendix_pp_621-639.pdf API Appendix] |
Latest revision as of 21:19, 8 February 2011
We describe here a rigorous derivation of atom-photon interactions. Our overall goal is to arrive at a Hamiltonian description for the energy of a system of atoms, photons, and atoms interacting with photons through the radiation field.
The basic interaction we will obtain is the dipole interaction Hamiltonian,
where is the atom's dipole moment, and is the electric field at the position of the atom. In the end, the electric field will be quantized, and described by operators and .
This result is simple and can be obtained with less rigorous derivations. Here, we want to at least mention all the steps of the rigorous derivation, starting from classical Maxwell's equations.
Contents
Quantum electrodynamics
The classical Hamiltonian which describes one particle
and this is transformed into the quantum picture by enforcing the commutation relation . We will do the same for an atom interacting with light.
Let us begin with Maxwell's equations.
How many of the field components are true degrees of freedom? We can understand this by taking a spatial Fourier transform. The component of the (vectorial) Fourier component, which is parallel to the k vector, is called the longitudinal part, the component orthogonal to the k vector is the transverse part. By transforming back to the spatial domain, we can distinguish the longitudinal and transverse parts. It turns out that the longitudinal electric field is not a free degree of freedom, i.e. it follows instantaneously the positions of the particles through the Coulomb potential.
Introducing the vector potential reduces the number of field components to four.
In the Coulomb gauge, , i.e. the longitudinal component of A vanished. Furthermore, the scalar potential is
Now we are down to two components, the transverse components of the vector poential A.
Keep in mind that the equation of motion is now a second order differential equation for the vector potential. All that is needed to specify the field evolution is thus the initial values of and , the transverse vector potential. The differential equation is identical to the equation for r and p for an harmonic oscillator potential.
We can go further by decomposing the Fourier representation of the field and potential in terms of its normal modes,
Using this, we can represent the equations of motion for the field as
Note
in the normal mode decomposition, where we now use as the photon polarization, which carries the vector direction of the potential.
Let us now quantize the field. We identify an equivalence between and with and , and quantize accordingly. The normal mode coordinates are a linear combination of and from which we obtain the commutator
in full analogy to the raising and lowering operators for the simple harmonic oscillator.
Energy in the radiation field
It is helpful to go back to consider for a moment what the energy in the field is. Recall that
where we can identify
as the energy of the instantaneous Coulomb field due to the charge configuration. The second term is the energy of the transverse component of the field, , the radiation energy, which we can understand by introducing again our expression for the vector potential.
where is identified as the conjugate momentum. This looks much like a simple harmonic oscillator Hamiltonian. Now introduce normal modes, using
This gives
a purely classical expression for the energy in the radiation field. But it looks quantum. Where does come from? It enters in the constant relating the normal modes with . It is just a unit used in the definitions at this moment, which is convenient to use because later on appears in the quantum expression of the energy.
This expression for energy is really identical to that for the classical simple harmonic oscillator,
where . For this classical oscillator, it is helpful to introduce a variable describing superpositions of , giving
When quantized, this becomes . We conclude that the radiation field is just a bunch of oscillators, with one per vector and polarization, and each one is described in its quantized form by the Hamiltonian .
The most important aspect of our derivation was the separation of the fields into longitudinal and transverse components. For example, a naive approach using Cartesian coordinates, without eliminating the longitudinal field, would fail miserably, because
The treatment above, eliminating the dependent components, is thus essential.
Also note that quantizing the electromagnetic field is only possible if we have an expression for the energy of the radiation field that also corresponds to a valid Lagrangian for the system. Finally, our approach is not relativistically covariant. There are covariant formulations of the quantized radiation field, but they are more complicated.
Quantum description of the radiation field
The inverse Fourier transform provides us with the field components in terms of the quantum operators,
The particle operators are
Coupling of atoms and the radiation field
The total Hamiltonian for the radiation field and charges is
where the second term, with has been added by hand, and describes spin interacting with the magnetic field, which will be discussed later. It can be derived from first principles by starting with the Dirac equation, expressed in the non-relativistic limit. The important new term, compared with standard nonrelativistic quantum mechanics, is the replacement of momentum with . The Coulomb interaction energy is standard. The radiation field energy is .
We may write the total Hamiltonian as a sum of parts
where is the particle Hamiltonian including the Coulomb field
is the Hamiltonian for the radiation field, and can be written as the sum of three parts, , and
For atoms, , typically. We will often perform perturbation theory in . Note that this is NOT a perturbation theory in charge or field strength, since the Coulomb field has been separated off and is fully included in the particle Hamiltonian.
The dipole approximation
Typically, for atomic physics, the wavelength of radiation is much much larger than the size of the atom, so that we may write the main interaction between atoms and the radiation field as
where . This simplifies to
in the limit that the field wavelength is much larger than the atom, so we can take . Since , and
then using gives
The interaction energy is thus
which in the limit of a near-resonant interaction, becomes
Questions that arise in this loose derivation include: what happens for off-resonant interactions? And what happens with the other interaction term we derived above, ?
A rigorous way to obtain the full solution, which is essentially the same as that sketched above, is given in API. It shows that the simple is exact (in the long wavelength approximation) and includes the term.
The derivation involves using a canonical transformation with the operator
which is just a displacement operator acting on momentum, that transforms into , in a new frame of reference. We approximate with , which is valid in the dipole approximation, in which . The Hamiltonian in this frame of reference is
In this frame, the dipole interaction energy appears explicitly. The transformed electric field is
The interaction Hamiltonian is
Note that this formulation already takes into account the polarizability of matter, and the relation between and .
Conclusion: What you should remember from this lengthy derivation is the separation into longitudinal and transverse fields, the fact that the Coulomb interaction is included in the particle Hamiltonian, and that the electric dipole approximation includes the A^2 term.