Difference between revisions of "Interaction of an atom with an electromagnetic field"
imported>Mhirose |
imported>Wburton |
||
(46 intermediate revisions by 11 users not shown) | |||
Line 6: | Line 6: | ||
<span id="SEC_ERL"></span> | <span id="SEC_ERL"></span> | ||
− | Einstein's 1917 paper on the theory of radiation | + | Einstein's 1917 paper on the theory of radiation <ref> A. Einstein, Z. Phys. 18, 121 (1917), translated in Sources of Quantum Mechanics, B. L. Van der Waerden, Cover Publication, Inc., New York, 1967. This book is a gold mine for anyone interested in the development of quantum mechanics. </ref> provided seminal concepts for the quantum theory of radiation. It also anticipated devices such as the laser, and pointed the way to the field of laser-cooling of atoms. In it, he set out to answer two questions: |
1) How do the internal states of an atom that radiates and absorbs energy come into equilibrium with a thermal radiation field? (In answering this question Einstein invented the concept of spontaneous emission) | 1) How do the internal states of an atom that radiates and absorbs energy come into equilibrium with a thermal radiation field? (In answering this question Einstein invented the concept of spontaneous emission) | ||
Line 12: | Line 12: | ||
2) How do the translational states of an atom in thermal equilibrium (i.e. states obeying the Maxwell-Boltzmann Law for the distribution of velocities) come into thermal equilibrium with a radiation field? (In answering this question, Einstein introduced the concept of photon recoil. He also demonstrated that the field itself must obey the Planck radiation law.) | 2) How do the translational states of an atom in thermal equilibrium (i.e. states obeying the Maxwell-Boltzmann Law for the distribution of velocities) come into thermal equilibrium with a radiation field? (In answering this question, Einstein introduced the concept of photon recoil. He also demonstrated that the field itself must obey the Planck radiation law.) | ||
− | The first part of Einstein's paper, which addresses question 1), is well known, but the second part, which addresses question 2), is every bit as germane | + | The first part of Einstein's paper, which addresses question 1), is well known, but the second part, which addresses question 2), is every bit as germane to contemporary atom/optical physics. Because the paper preceded the creation of quantum mechanics there was no way for him to calculate transition rates. However, his arguments are based on general statistical principles and provide the foundation for interpreting the quantum mechanical results. |
Einstein considered a system of <math>N</math> atoms in thermal equilibrium with a radiation field. The system has two levels (an energy level consists of all of the states that have a given energy; the number of quantum states in a given level is its multiplicity.) with energies <math>E_ b</math> and <math>E_ a</math>, with <math>E_ b > E_ a</math>, and <math>E_ b - E_ a =\hbar \omega </math>. The numbers of atoms in the two levels are related by <math>N_ b + N_ a = N</math>. Einstein assumed the Planck radiation law for the spectral energy density temperature. For radiation in thermal equilibrium at temperature <math>T</math>, the energy per unit volume in wavelength range <math>d\omega </math> is: | Einstein considered a system of <math>N</math> atoms in thermal equilibrium with a radiation field. The system has two levels (an energy level consists of all of the states that have a given energy; the number of quantum states in a given level is its multiplicity.) with energies <math>E_ b</math> and <math>E_ a</math>, with <math>E_ b > E_ a</math>, and <math>E_ b - E_ a =\hbar \omega </math>. The numbers of atoms in the two levels are related by <math>N_ b + N_ a = N</math>. Einstein assumed the Planck radiation law for the spectral energy density temperature. For radiation in thermal equilibrium at temperature <math>T</math>, the energy per unit volume in wavelength range <math>d\omega </math> is: | ||
− | + | <equation id="erad1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \rho _ E (\omega )d\omega = \frac{\hbar \omega ^3}{\pi ^2 c^3} \frac{1}{{\rm exp} (\hbar \omega /kT) -1 }d\omega . \end{align}</math> | |
− | + | </equation> | |
− | |||
The mean occupation number of a harmonic oscillator at temperature <math>T</math>, which can be interpreted as the mean number of photons in one mode of the radiation field, is | The mean occupation number of a harmonic oscillator at temperature <math>T</math>, which can be interpreted as the mean number of photons in one mode of the radiation field, is | ||
− | + | <equation id="erad2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \bar{n} = \frac{1}{{\rm exp} (\hbar \omega /kT) -1}. \end{align}</math> | |
− | + | </equation> | |
− | |||
According to the Boltzmann Law of statistical mechanics, in thermal equilibrium the populations of the two levels are related by | According to the Boltzmann Law of statistical mechanics, in thermal equilibrium the populations of the two levels are related by | ||
− | + | <equation id="erad3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \frac{N_ b}{N_ a} = \frac{g_ b}{g_ a} e^{-(E_ b -E_ a)/kT} = \frac{g_ b}{g_ a} e^{-\hbar \omega /kT} . \end{align}</math> | |
− | + | </equation> | |
− | |||
Here <math>g_ b</math> and <math>g_ a</math> are the multiplicities of the two levels. The last step assumes the Bohr frequency condition, <math>\omega = (E_ b -E_ a)\ \hbar </math>. However, Einstein's paper actually derives this relation independently. | Here <math>g_ b</math> and <math>g_ a</math> are the multiplicities of the two levels. The last step assumes the Bohr frequency condition, <math>\omega = (E_ b -E_ a)\ \hbar </math>. However, Einstein's paper actually derives this relation independently. | ||
Line 39: | Line 36: | ||
According to classical theory, an oscillator can exchange energy with the radiation field at a rate that is proportional to the spectral density of radiation. The rates for absorption and emission are equal. The population transfer rate equation is thus predicted to be | According to classical theory, an oscillator can exchange energy with the radiation field at a rate that is proportional to the spectral density of radiation. The rates for absorption and emission are equal. The population transfer rate equation is thus predicted to be | ||
− | + | <equation id="erad4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \dot{N}_ b = - { \rho _ E (\omega ) B_{ba}} N_ b + \rho _ E (\omega ) B_{ab} N_ a = -\dot{N}_ a . \end{align}</math> | |
− | + | </equation> | |
− | |||
− | This equation is | + | This equation is compatible with <xr id="erad1"/>, <xr id="erad2"/>, <xr id="erad3"/> it follows that |
− | + | <equation id=" erl5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ g_ b B_{ba} & =& g_ a B_{ab} \\ \frac{\hbar \omega ^3}{\pi ^2 c^3} B_{ba} & =& A_{ba} \\ \rho _ E (\omega ) B_{ba} & =& \bar{n} A_{ba} \\ \end{align}</math> | |
− | + | </equation> | |
− | |||
− | |||
− | |||
− | |||
− | |||
− | |||
− | |||
− | |||
Consequently, the rate of transition <math>b\rightarrow a</math> is | Consequently, the rate of transition <math>b\rightarrow a</math> is | ||
− | + | <equation id=" erl6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ B_{ba} \rho _ E (\omega ) + A_{ba} = (\bar{n} +1 )A_{ba}, \end{align}</math> | |
− | + | </equation> | |
− | |||
while the rate of absorption is | while the rate of absorption is | ||
− | + | <equation id=" erl7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ B_{ab} \rho _ E (\omega ) = \frac{g_ b}{g_ a} \bar{n} A_{ba} \end{align}</math> | |
− | + | </equation> | |
− | |||
If we consider emission and absorption between single states by taking <math>g_ b = g_ a = 1</math>, then the ratio of rate of emission to rate of absorption is <math>(\bar{n} + 1) /\bar{n}</math>. | If we consider emission and absorption between single states by taking <math>g_ b = g_ a = 1</math>, then the ratio of rate of emission to rate of absorption is <math>(\bar{n} + 1) /\bar{n}</math>. | ||
Line 81: | Line 67: | ||
<br style="clear: both" /> | <br style="clear: both" /> | ||
+ | |||
== Quantum Theory of Absorption and Emission == | == Quantum Theory of Absorption and Emission == | ||
Line 94: | Line 81: | ||
Our starting point is Maxwell's equations (S.I. units): | Our starting point is Maxwell's equations (S.I. units): | ||
− | + | <equation id="Maxwell" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \nabla \cdot {\bf E} & = & \rho /\epsilon _0 \\ \nabla \cdot {\bf B} & = & 0 \\ \nabla \times {\bf E} & = & - \frac{\partial {\bf B}}{\partial t} \\ \nabla \times {\bf B} & = & \frac{1}{c^2} \frac{\partial \bf { E}}{\partial t} + \mu _0 \bf {J} \end{align}</math> | |
− | + | </equation> | |
− | |||
The charge density <math>\rho </math> and current density '''J''' obey the continuity equation | The charge density <math>\rho </math> and current density '''J''' obey the continuity equation | ||
− | + | <equation id=" wd2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \nabla \cdot {\bf J} + \frac{\partial \rho }{\partial t} = 0 \end{align}</math> | |
− | + | </equation> | |
− | |||
Introducing the vector potential '''A''' and the scalar potential <math>\psi </math>, we have | Introducing the vector potential '''A''' and the scalar potential <math>\psi </math>, we have | ||
− | + | <equation id=" wd3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf E} & = & - \nabla \psi - \frac{\partial {\bf A}}{\partial t} \\ {\bf B} & = & \nabla \times {\bf A} \end{align}</math> | |
− | + | </equation> | |
− | |||
We are free to change the potentials by a gauge transformation: | We are free to change the potentials by a gauge transformation: | ||
− | + | <equation id=" wd4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf A}^\prime = {\bf A} + \nabla \Lambda , ~ ~ ~ ~ ~ \psi ^\prime = \psi - \frac{\partial \Lambda }{\partial t} \end{align}</math> | |
− | + | </equation> | |
− | |||
where <math>\Lambda </math> is a scalar function. This transformation leaves the fields invariant, but changes the form of the dynamical equation. We shall work in the <i> | where <math>\Lambda </math> is a scalar function. This transformation leaves the fields invariant, but changes the form of the dynamical equation. We shall work in the <i> | ||
Line 124: | Line 107: | ||
</i> (often called the radiation gauge), defined by | </i> (often called the radiation gauge), defined by | ||
− | + | <equation id=" wd5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \nabla \cdot {\bf A} = 0 \end{align}</math> | |
− | + | </equation> | |
− | |||
In free space, '''A''' obeys the wave equation | In free space, '''A''' obeys the wave equation | ||
− | + | <equation id=" wd6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \nabla ^2 {\bf A} = \frac{1}{c^2} \frac{\partial ^2 {\bf A}}{\partial t^2} \end{align}</math> | |
− | + | </equation> | |
− | |||
Because <math>\nabla \cdot {\bf A}= 0</math>, '''A''' is transverse. We take a propagating plane wave solution of the form | Because <math>\nabla \cdot {\bf A}= 0</math>, '''A''' is transverse. We take a propagating plane wave solution of the form | ||
− | + | <equation id="A-field" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf A}(r, t) = A{\bf \hat{e}} \cos ({\bf k}\cdot {\bf r} -\omega t) = A{\bf \hat{e}} \frac{1}{2} \left[ e^{i({\bf k}\cdot {\bf r} -\omega t)} + e^{-i({\bf k}\cdot {\bf r} -\omega t)} \right], \end{align}</math> | |
− | + | </equation> | |
− | |||
where <math>k^2 =\omega ^2 / c^2</math> and <math>{\bf \hat{e}}\cdot {\bf k}= 0</math>. For a linearly polarized field, the polarization vector <math>{\bf \hat{e}}</math> is real. For an elliptically polarized field it is complex, and for a circularly polarized field it is given by <math>{\bf \hat{e}} = ({\bf \hat{ x}} \pm i {\bf \hat{ y}} ) /\sqrt {2}</math> , where the + and <math>-</math> signs correspond to positive and negative helicity, respectively. (Alternatively, they correspond to left and right hand circular polarization, respectively, the sign convention being a tradition from optics.) The electric and magnetic fields are then given by | where <math>k^2 =\omega ^2 / c^2</math> and <math>{\bf \hat{e}}\cdot {\bf k}= 0</math>. For a linearly polarized field, the polarization vector <math>{\bf \hat{e}}</math> is real. For an elliptically polarized field it is complex, and for a circularly polarized field it is given by <math>{\bf \hat{e}} = ({\bf \hat{ x}} \pm i {\bf \hat{ y}} ) /\sqrt {2}</math> , where the + and <math>-</math> signs correspond to positive and negative helicity, respectively. (Alternatively, they correspond to left and right hand circular polarization, respectively, the sign convention being a tradition from optics.) The electric and magnetic fields are then given by | ||
− | + | <equation id="E-field" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf E}(r, t) = \omega A{\bf \hat{e}} \sin ({\bf k}\cdot {\bf r} -\omega t) = - i \omega A {\bf \hat{e}} \frac{1}{2} \left[ e^{i({\bf k}\cdot {\bf r} -\omega t)} - e^{-i({\bf k}\cdot {\bf r} -\omega t)} \right]. \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | <equation id="B-field" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf B}(r, t) = k ({\bf \hat{k}} \times {\bf \hat{ e}}) \sin ({\bf k}\cdot {\bf r} -\omega t) = - i k A ({\bf \hat{k}} \times {\bf \hat{ e}}) \frac{1}{2} \left[ e^{i({\bf k}\cdot {\bf r} -\omega t)} - e^{-i({\bf k}\cdot {\bf r} -\omega t)} \right]. \end{align}</math> | |
− | + | </equation> | |
− | |||
The time average Poynting vector is | The time average Poynting vector is | ||
− | + | <equation id=" wd9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf S} = \frac{ \epsilon _0 c^2}{2} ( {{\bf E} \times {\bf B}^* )} = \frac{\epsilon _0 c}{2} \omega ^2 A^2 {\bf \hat{k}} . \end{align}</math> | |
− | + | </equation> | |
− | |||
The average energy density in the wave is given by | The average energy density in the wave is given by | ||
− | + | <equation id="energy-density" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ u = \omega ^2 \frac{\epsilon _0 }{2} A^2 {\bf \hat{k}} . \end{align}</math> | |
− | + | </equation> | |
− | |||
<br style="clear: both" /> | <br style="clear: both" /> | ||
Line 177: | Line 153: | ||
The behavior of charged particles in an electromagnetic field is correctly described by Hamilton's equations provided that the canonical momentum is redefined: | The behavior of charged particles in an electromagnetic field is correctly described by Hamilton's equations provided that the canonical momentum is redefined: | ||
− | + | <equation id=" int1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf p}_{\rm can} = {\bf p}_{\rm kin} + q {\bf A} \end{align}</math> | |
− | + | </equation> | |
− | |||
The kinetic energy is <math>{\bf p}_{\rm kin}^2 /2 m</math>. Taking <math>q = - e</math>, the Hamiltonian for an atom in an electromagnetic field in free space is | The kinetic energy is <math>{\bf p}_{\rm kin}^2 /2 m</math>. Taking <math>q = - e</math>, the Hamiltonian for an atom in an electromagnetic field in free space is | ||
− | + | <equation id=" int2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H = \frac{1}{2m} \sum _{j=1}^{N} {\left( {\bf p}_ j + e {\bf A} (r_ j )\right)^2} + \sum _{j=1}^{N} V ({\bf r}_ j ), \end{align}</math> | |
− | + | </equation> | |
− | |||
where <math>V ( {\bf r}_ j )</math> describes the potential energy due to internal interactions. We are neglecting spin interactions. | where <math>V ( {\bf r}_ j )</math> describes the potential energy due to internal interactions. We are neglecting spin interactions. | ||
Line 193: | Line 167: | ||
Expanding and rearranging, we have | Expanding and rearranging, we have | ||
− | + | <equation id=" int3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H & =& \sum _{j=1}^{N} \frac{{\bf p}_ j^2}{2m} + V ({\bf r}_ j ) + \frac{e}{2m} \sum _{j=1}^{N} {\left({\bf p}_ j \cdot {\bf A} ( {\bf r}_ j) + {\bf A} ({\bf r}_ j ) \cdot {\bf p}_ j \right)} + \frac{e^2}{2m} \sum _{j=1}^{N} A_ j^2 ({\bf r} ) \\ & = & H_0 + H_{\rm int} + H^{(2)} . \end{align}</math> | |
− | + | </equation> | |
− | |||
Here, <math>{\bf p}_ j = - i\hbar \nabla _ j </math>. Consequently, <math>H_0</math> describes the unperturbed atom. <math>H_{\rm int}</math> describes the atom's interaction with the field. <math>H^{(2)}</math>, which is second order in '''A''', plays a role only at very high intensities. (In a static magnetic field, however, <math>H^{(2)}</math> gives rise to diamagnetism.) | Here, <math>{\bf p}_ j = - i\hbar \nabla _ j </math>. Consequently, <math>H_0</math> describes the unperturbed atom. <math>H_{\rm int}</math> describes the atom's interaction with the field. <math>H^{(2)}</math>, which is second order in '''A''', plays a role only at very high intensities. (In a static magnetic field, however, <math>H^{(2)}</math> gives rise to diamagnetism.) | ||
Line 202: | Line 175: | ||
Because we are working in the Coulomb gauge, <math>\nabla \cdot {\bf A} =0</math> so that '''A''' and '''p''' commute. We have | Because we are working in the Coulomb gauge, <math>\nabla \cdot {\bf A} =0</math> so that '''A''' and '''p''' commute. We have | ||
− | + | <equation id=" int4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H_{\rm int} = \frac{eA}{m} \hat{\bf {e}} \cdot {\bf p} \cos ({\bf k}\cdot {\bf r} -\omega t) . \end{align}</math> | |
− | + | </equation> | |
− | |||
It is convenient to write the matrix element between states <math> | a \rangle </math> and <math> | b \rangle </math> in the form | It is convenient to write the matrix element between states <math> | a \rangle </math> and <math> | b \rangle </math> in the form | ||
− | + | <equation id=" int5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \langle b | H_{\rm int} | a \rangle = \frac{1}{2} H_{ba} e^{-i\omega t} + \frac{1}{2} H_{ba} e^{+i\omega t}, \end{align}</math> | |
− | + | </equation> | |
− | |||
where | where | ||
− | + | <equation id=" int6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H_{ba} = \frac{eA}{m} {\bf \hat{e}} \, \langle b |{\bf p} \, e^{i {\bf k} \cdot {\bf r}} | a \rangle . \end{align}</math> | |
− | + | </equation> | |
− | |||
− | Atomic dimensions are small compared to the wavelength of radiation involved in optical transitions. The scale of the ratio is set by <math>\alpha \approx 1/137</math>. Consequently, when the matrix element in | + | Atomic dimensions are small compared to the wavelength of radiation involved in optical transitions. The scale of the ratio is set by <math>\alpha \approx 1/137</math>. Consequently, when the matrix element in <xr id="int6"/> is evaluated, the wave function vanishes except in the region where <math>{\bf k}\cdot {\bf r} = 2 \pi r /\lambda \ll 1</math>. It is therefore appropriate to expand the exponential: |
− | + | <equation id=" int7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H_{ba} = \frac{eA}{m} {\bf \hat{e}} \cdot \langle b | {\bf p} (1 + i{\bf k} \cdot {\bf r} - 1/2 ({\bf k}\cdot {\bf r} )^2 + \cdots ) | a \rangle \end{align}</math> | |
− | + | </equation> | |
− | |||
Unless <math>\langle b | {\bf p} | a \rangle </math> vanishes, for instance due to parity considerations, the leading term dominates and we can neglect the others. For reasons that will become clear, this is called the dipole approximation. This is by far the most important situation, and we shall defer consideration of the higher order terms. In the dipole approximation we have | Unless <math>\langle b | {\bf p} | a \rangle </math> vanishes, for instance due to parity considerations, the leading term dominates and we can neglect the others. For reasons that will become clear, this is called the dipole approximation. This is by far the most important situation, and we shall defer consideration of the higher order terms. In the dipole approximation we have | ||
− | + | <equation id=" int8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H_{ba} = \frac{eA}{m} {\bf \hat{e}} \cdot \langle b | {\bf p} | a \rangle = \frac{-ieE}{m\omega } {\bf \hat{e}} \cdot \langle b | {\bf p} | a \rangle \end{align}</math> | |
− | + | </equation> | |
− | |||
− | where we have used, from | + | where we have used, from <xr id="E-field"/>, <math>A = -iE/\omega </math>. It can be shown (i.e. left as exercise) that the matrix element of '''p''' can be transfomred into a matrix element for <math>{\bf r}</math>: |
− | + | <equation id=" int9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \langle b | {\bf p} | a \rangle = - i m \omega _{ab} \langle b | {\bf r} | a \rangle = + i m \omega _{ba} \langle b | {\bf r} | a \rangle \end{align}</math> | |
− | + | </equation> | |
− | |||
This results in | This results in | ||
− | + | <equation id=" int10" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H_{ba} = \frac{e E \omega _{ba}}{\omega } {\bf \hat{e}} \cdot \langle b | {\bf r} | a \rangle \end{align}</math> | |
− | + | </equation> | |
− | |||
We will be interested in resonance phenomena in which <math>\omega \approx \omega _{ba}</math>. Consequently, | We will be interested in resonance phenomena in which <math>\omega \approx \omega _{ba}</math>. Consequently, | ||
− | + | <equation id=" int11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H_{ba} = + e {\bf E}_0 \cdot \langle b | {\bf r} | a \rangle = - {\bf d}_{ba} \cdot {\bf E} \end{align}</math> | |
− | + | </equation> | |
− | |||
where '''d ''' is the dipole operator, <math>{\bf d} = - e {\bf r}</math>. Displaying the time dependence explictlty, we have | where '''d ''' is the dipole operator, <math>{\bf d} = - e {\bf r}</math>. Displaying the time dependence explictlty, we have | ||
− | + | <equation id=" int12" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H_{ba}^\prime = - {\bf d}_{ba}\cdot {\bf E}_0 e^{-i\omega t}. \end{align}</math> | |
− | + | </equation> | |
− | |||
However, it is important to bear in mind that this is only the first term in a series, and that if it vanishes the higher order terms will contribute a perturbation at the driving frequency. | However, it is important to bear in mind that this is only the first term in a series, and that if it vanishes the higher order terms will contribute a perturbation at the driving frequency. | ||
− | <math>H_{ba}</math> appears as a matrix element of the momentum operator '''p''' in | + | <math>H_{ba}</math> appears as a matrix element of the momentum operator '''p''' in <xr id="int8"/>, and of the dipole operator '''r''' in <xr id="int11"/>. These matrix elements look different and depend on different parts of the wave function. The momentum operator emphasizes the curvature of the wave function, which is largest at small distances, whereas the dipole operator evaluates the moment of the charge distribution, i.e. the long range behavior. In practice, the accuracy of a calculation can depend significantly on which operator is used. |
<br style="clear: both" /> | <br style="clear: both" /> | ||
+ | |||
== Quantization of the radiation field == | == Quantization of the radiation field == | ||
<span id="SEC_QRD"></span> | <span id="SEC_QRD"></span> | ||
− | We shall consider a single mode of the radiation field. This means a single value of the wave vector '''k''', and one of the two orthogonal transverse polarization vectors <math>{\bf \hat{e}}</math>. The radiation field is described by a plane wave vector potential of the form | + | We shall consider a single mode of the radiation field. This means a single value of the wave vector '''k''', and one of the two orthogonal transverse polarization vectors <math>{\bf \hat{e}}</math>. The radiation field is described by a plane wave vector potential of the form <xr id="A-field"/>. We assume that '''k''' obeys a periodic boundary or condition, <math>k_ x L_ x = 2\pi n_ x</math>, etc. (For any '''k''', we can choose boundaries <math>L_ x , L_ y , L_ z</math> to satisfy this.) The time averaged energy density is given by <xr id="energy-density"/>, and the total energy in the volume V defined by these boundaries is |
− | + | <equation id="energy-total" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ U = \frac{\epsilon _0 }{2}\omega ^2 A^2 V, \end{align}</math> | |
− | + | </equation> | |
− | |||
where <math>A^2</math> is the mean squared value of <math>A</math> averaged over the spatial mode. We now make a formal connection between the radiation field and a harmonic oscillator. We define variables Q and P by | where <math>A^2</math> is the mean squared value of <math>A</math> averaged over the spatial mode. We now make a formal connection between the radiation field and a harmonic oscillator. We define variables Q and P by | ||
− | + | <equation id=" qrd5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ A = \frac{1}{\omega } \sqrt {\frac{1}{\epsilon _ o V}} (\omega Q + iP ), ~ ~ A^* =\frac{1}{\omega }\sqrt {\frac{1}{\epsilon _ o V}} (\omega Q - iP ). \end{align}</math> | |
− | + | </equation> | |
− | |||
− | Then, from | + | Then, from <xr id="energy-total"/>, we find |
− | + | <equation id=" qrd6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ U = \frac{1}{2} (\omega ^2 Q^2 + P^2 ). \end{align}</math> | |
− | + | </equation> | |
− | |||
This describes the energy of a harmonic oscillator having unit mass. We quantize the oscillator in the usual fashion by treating Q and P as operators, with | This describes the energy of a harmonic oscillator having unit mass. We quantize the oscillator in the usual fashion by treating Q and P as operators, with | ||
− | + | <equation id=" qrd7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ P = - i\hbar \frac{\partial }{\partial Q}, ~ ~ ~ [Q,P] = i\hbar . \end{align}</math> | |
− | + | </equation> | |
− | |||
We introduce the operators <math>a</math> and <math>a^\dagger </math> defined by | We introduce the operators <math>a</math> and <math>a^\dagger </math> defined by | ||
− | + | <equation id=" qrd8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ a = \frac{1}{\sqrt {2\hbar \omega }} (\omega Q + iP ) \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | <equation id=" qrd9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ a^\dagger = \frac{1}{\sqrt {2\hbar \omega }} (\omega Q - iP ) \end{align}</math> | |
− | + | </equation> | |
− | |||
The fundamental commutation rule is | The fundamental commutation rule is | ||
− | + | <equation id=" qrd10" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ [a, a^\dagger ] = 1 \end{align}</math> | |
− | + | </equation> | |
− | |||
from which the following can be deduced: | from which the following can be deduced: | ||
− | + | <equation id=" qrd11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H = \frac{1}{2} \hbar \omega [a^\dagger a + a a^\dagger ] = \hbar \omega \left[a^\dagger a + \frac{1}{2} \right] = \hbar \omega \left[N+ \frac{1}{2} \right] \end{align}</math> | |
− | + | </equation> | |
− | |||
where the number operator <math>N = a^\dagger a </math> obeys | where the number operator <math>N = a^\dagger a </math> obeys | ||
− | + | <equation id=" qrd12" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ N| n \rangle = n| n \rangle \end{align}</math> | |
− | + | </equation> | |
− | |||
We also have | We also have | ||
− | + | <equation id=" qrd13" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \langle n-1| a | n \rangle & =& \sqrt {n} \\ \langle n+1| a^\dagger | n \rangle & =& \sqrt {n +1} \\ \langle n| a^\dagger a | n \rangle & =& n \\ \langle n |a a^\dagger | n \rangle & =& n+1 \\ \langle n| H | n \rangle & =& \hbar \omega \left(n+ \frac{1}{2} \right) \\ \ \langle n| a | n \rangle & =& \langle n | a^\dagger | n \rangle = 0 \end{align}</math> | |
− | + | </equation> | |
− | |||
The operators <math>a </math> and <math>a^\dagger </math> are called the annihilation and creation operators, respectively. We can express the vector potential and electric field in terms of <math>a</math> and <math>a^\dagger </math> as follows | The operators <math>a </math> and <math>a^\dagger </math> are called the annihilation and creation operators, respectively. We can express the vector potential and electric field in terms of <math>a</math> and <math>a^\dagger </math> as follows | ||
− | + | <equation id=" part1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ A = \frac{1}{ \omega \sqrt {\epsilon _ o V}} (\omega Q + iP) = \sqrt {\frac{2 \hbar }{ \omega \epsilon _ o V}} a \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | <equation id=" part2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ A^\dagger = \frac{1}{ \omega \sqrt {\epsilon _ o V}} (\omega Q - iP) = \sqrt {\frac{2 \hbar }{ \omega \epsilon _ o V}} a^\dagger \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | <equation id=" part3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf E} = - i \sqrt {\frac{ \hbar \omega }{2 \epsilon _ o V} } {\left[ a{\bf \hat{e}} e^{i({\bf k}\cdot {\bf r} - \omega t)} - a^\dagger {\bf \hat{e}}^* e^{-i({\bf k}\cdot {\bf r} -\omega t)}\right]} \end{align}</math> | |
− | + | </equation> | |
− | |||
In the dipole limit we can take <math>e^{i {\bf k}\cdot {\bf r}} = 1</math>. Then | In the dipole limit we can take <math>e^{i {\bf k}\cdot {\bf r}} = 1</math>. Then | ||
− | + | <equation id=" part3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf E} = - i \sqrt {\frac{ \hbar \omega }{2 \epsilon _ o V} } \left[ a {\bf \hat e} e^{-i \omega t}- a^\dagger {\bf {\hat e}}^* e^{i \omega t}\right] \end{align}</math> | |
− | + | </equation> | |
− | |||
The interaction Hamiltonian is, | The interaction Hamiltonian is, | ||
− | + | <equation id=" qrd16" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ H_{\rm int}= -ie \sqrt {\frac{\hbar \omega }{2\epsilon _ o V}}{\bf r}\cdot {\left[ a{\bf \hat{e}} e^{-i\omega t} - a^\dagger {\bf \hat{e}}^* e^{+i\omega t}\right]}, \end{align}</math> | |
− | + | </equation> | |
− | |||
where we have written the dipole operator as <math>{\bf d} = - e {\bf r}</math>. | where we have written the dipole operator as <math>{\bf d} = - e {\bf r}</math>. | ||
<br style="clear: both" /> | <br style="clear: both" /> | ||
+ | |||
== Interaction of a two-level system and a single mode of the radiation field == | == Interaction of a two-level system and a single mode of the radiation field == | ||
Line 380: | Line 331: | ||
We consider a two-state atomic system <math> | a \rangle </math>, <math>| b \rangle </math> and a radiation field described by <math>| n \rangle ,\ n = 0,1,2 \dots </math> The states of the total system can be taken to be | We consider a two-state atomic system <math> | a \rangle </math>, <math>| b \rangle </math> and a radiation field described by <math>| n \rangle ,\ n = 0,1,2 \dots </math> The states of the total system can be taken to be | ||
− | + | <equation id=" vac1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ | I \rangle = | a,\ n \rangle = | a \rangle \ | n \rangle , ~ ~ ~ | F \rangle = | b,\ n^\prime \rangle = |b \rangle \ |n^\prime \rangle . \end{align}</math> | |
− | + | </equation> | |
− | |||
We shall take <math>{\bf \hat{e}} = {\bf \hat{ z}} </math>. Then | We shall take <math>{\bf \hat{e}} = {\bf \hat{ z}} </math>. Then | ||
− | + | <equation id=" vac2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \langle F |H_{\rm int} | I \rangle = -i e z_{ab} \sqrt {\frac{2\pi \hbar \omega }{V}} \langle n^\prime | a e^{-i\omega t} - a^\dagger e^{i\omega t} | n \rangle e^{-i\omega _{ab} t} \end{align}</math> | |
− | + | </equation> | |
− | |||
− | The first term in the bracket obeys the selection rule <math>n^\prime = n - 1</math>. This corresponds to loss of one photon from the field and absorption of one photon by the atom. The second term obeys <math>n^\prime = n + 1</math>. This corresponds to emission of a photon by the atom. Using | + | The first term in the bracket obeys the selection rule <math>n^\prime = n - 1</math>. This corresponds to loss of one photon from the field and absorption of one photon by the atom. The second term obeys <math>n^\prime = n + 1</math>. This corresponds to emission of a photon by the atom. Using <xr id="qrd13"/>, we have |
− | + | <equation id=" vac3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \langle F | H_{\rm int} | I \rangle = -i e z_{ab} \sqrt {\frac{2\pi \hbar \omega }{V}} {\left( \sqrt {n}\, \delta _{n\prime ,n-1} \ e^{-i \omega t} - \sqrt {n+1}\, \delta _{n\prime ,n+1} e^{+i\omega t} \right)} \ e^{-i\omega _{ab} t} \end{align}</math> | |
− | + | </equation> | |
− | |||
Transitions occur when the total time dependence is zero, or near zero. Thus absorption occurs when <math>\omega =- \omega _{ab}</math>, or <math>E_ a + \hbar \omega = E_ b</math>. As we expect, energy is conserved. Similarly, emission occurs when <math>\omega = + \omega _{ab}</math>, or <math>E_ a - \hbar \omega = E_ b</math>. | Transitions occur when the total time dependence is zero, or near zero. Thus absorption occurs when <math>\omega =- \omega _{ab}</math>, or <math>E_ a + \hbar \omega = E_ b</math>. As we expect, energy is conserved. Similarly, emission occurs when <math>\omega = + \omega _{ab}</math>, or <math>E_ a - \hbar \omega = E_ b</math>. | ||
Line 403: | Line 351: | ||
A particularly interesting case occurs when <math>n = 0</math>, i.e. the field is initially in the vacuum state, and <math>\omega = \omega _{ab}</math>. Then | A particularly interesting case occurs when <math>n = 0</math>, i.e. the field is initially in the vacuum state, and <math>\omega = \omega _{ab}</math>. Then | ||
− | + | <equation id=" vac4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \langle F | H_{\rm int} | I \rangle = i e z_{ab} \sqrt {\frac{2\pi \hbar \omega }{V}} \equiv H_{FI}^0 \end{align}</math> | |
− | + | </equation> | |
− | |||
The situation describes a constant perturbation <math>H_{FI}^0</math> coupling the two states <math>I = | a , n = 0 \rangle </math> and <math>F = | b, n^\prime = 1 \rangle </math>. The states are degenerate because <math>E_ a = E_ b + \hbar \omega </math>. Consequently, <math>E_ a</math> is the upper of the two atomic energy levels. | The situation describes a constant perturbation <math>H_{FI}^0</math> coupling the two states <math>I = | a , n = 0 \rangle </math> and <math>F = | b, n^\prime = 1 \rangle </math>. The states are degenerate because <math>E_ a = E_ b + \hbar \omega </math>. Consequently, <math>E_ a</math> is the upper of the two atomic energy levels. | ||
Line 412: | Line 359: | ||
The system is composed of two degenerate eigenstates, but due to the coupling of the field, the degeneracy is split. The eigenstates are symmetric and antisymmetric combinations of the initial states, and we can label them as | The system is composed of two degenerate eigenstates, but due to the coupling of the field, the degeneracy is split. The eigenstates are symmetric and antisymmetric combinations of the initial states, and we can label them as | ||
− | + | <equation id=" vac5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ | \pm \rangle = \frac{1}{\sqrt {2}} (|I \rangle \pm | F \rangle ) = \frac{1}{\sqrt {2}} ( | a , 0 \rangle \pm | b, 1 \rangle ). \end{align}</math> | |
− | + | </equation> | |
− | |||
The energies of these states are | The energies of these states are | ||
− | + | <equation id=" vac6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ E_{\pm } = \pm | H_{FI}^0 | \end{align}</math> | |
− | + | </equation> | |
− | |||
If at <math>t = 0</math>, the atom is in state <math>| a \rangle </math> which means that the radiation field is in state <math>| 0 \rangle </math> then the system is in a superposition state: | If at <math>t = 0</math>, the atom is in state <math>| a \rangle </math> which means that the radiation field is in state <math>| 0 \rangle </math> then the system is in a superposition state: | ||
− | + | <equation id=" vac7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \psi (0) = \frac{1}{\sqrt {2}} ( | + \rangle + | - \rangle ) . \end{align}</math> | |
− | + | </equation> | |
− | |||
The time evolution of this superposition is given by | The time evolution of this superposition is given by | ||
− | + | <equation id=" vac8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \psi (t) = \frac{1}{\sqrt {2}} \left(| + \rangle e^{i\Omega /2t} + | - \rangle e^{-i\Omega /2t} \right) \end{align}</math> | |
− | + | </equation> | |
− | |||
where <math>\Omega / 2 = | H_{FI}^0 | / \hbar = e z_{ab}\sqrt {\omega / (e \epsilon _ o V \hbar )}</math>. The probability that the atom is in state <math> | b \rangle </math> at a later time is | where <math>\Omega / 2 = | H_{FI}^0 | / \hbar = e z_{ab}\sqrt {\omega / (e \epsilon _ o V \hbar )}</math>. The probability that the atom is in state <math> | b \rangle </math> at a later time is | ||
− | + | <equation id=" vac9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ P_ b = \frac{1}{2} (1 + \cos \Omega t ). \end{align}</math> | |
− | + | </equation> | |
− | |||
The frequency <math>\Omega </math> is called the vacuum Rabi frequency. | The frequency <math>\Omega </math> is called the vacuum Rabi frequency. | ||
− | The dynamics of a 2-level atom interacting with a single mode of the vacuum were first analyzed in | + | The dynamics of a 2-level atom interacting with a single mode of the vacuum were first analyzed in <ref> E.T. Jaynes and F.W. Cummings, Proc. IEEE, 51, 89 (1963). </ref> and the oscillations are sometimes called <i> |
Jaynes-Cummings | Jaynes-Cummings | ||
</i> oscillations. | </i> oscillations. | ||
− | The atom-vacuum interaction <math>H_{FI}^0</math>, | + | The atom-vacuum interaction <math>H_{FI}^0</math>, <xr id="vac4"/>, has a simple physical interpretation. The electric field amplitude associated with the zero point energy in the cavity is given by |
− | + | <equation id=" vac10" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \epsilon _ o E^2 V = \frac{1}{2} \hbar \omega \end{align}</math> | |
− | + | </equation> | |
− | |||
Consequently, <math>| H_{FI}^0 | = E d_{ab}= ez_{ab} E</math>. The interaction frequency <math>| H_{FI}^0 | / \hbar </math> is sometimes referred to as the vacuum Rabi frequency, although, as we have seen, the actual oscillation frequency is <math>2 \times H_{FI}^0 /\hbar </math>. | Consequently, <math>| H_{FI}^0 | = E d_{ab}= ez_{ab} E</math>. The interaction frequency <math>| H_{FI}^0 | / \hbar </math> is sometimes referred to as the vacuum Rabi frequency, although, as we have seen, the actual oscillation frequency is <math>2 \times H_{FI}^0 /\hbar </math>. | ||
− | Absorption and emission are closely related. Because the rates are proportional to <math>| \langle F | H_{\rm int} | I \rangle |^2</math>, it is evident from | + | Absorption and emission are closely related. Because the rates are proportional to <math>| \langle F | H_{\rm int} | I \rangle |^2</math>, it is evident from <xr id="vac3"/> that |
− | + | <equation id=" vac11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \frac{\rm Rate~ of~ emission}{\rm Rate~ of~ absorption} = \frac{n+1}{n} \end{align}</math> | |
− | + | </equation> | |
− | |||
This result, which applies to radiative transitions between any two states of a system, is general. In the absence of spontaneous emission, the absorption and emission rates are identical. | This result, which applies to radiative transitions between any two states of a system, is general. In the absence of spontaneous emission, the absorption and emission rates are identical. | ||
− | The oscillatory behavior described by | + | The oscillatory behavior described by <xr id="vac8"/> is exactly the opposite of free space behavior in which an excited atom irreversibly decays to the lowest available state by spontaneous emission. The distinction is that in free space there are an infinite number of final states available to the photon, since it can go off in any direction, but in the cavity there is only one state. The natural way to regard the atom-cavity system is not in terms of the atom and cavity separately, as in <xr id="vac1"/>, but in terms of the coupled states <math>| + \rangle </math> and <math>| - \rangle </math> (<xr id="vac5"/>). Such states, called <i> |
dressed atom | dressed atom | ||
</i> states, are the true eigenstates of the atom-cavity system. | </i> states, are the true eigenstates of the atom-cavity system. | ||
Line 481: | Line 421: | ||
In Chapter 6, first-order perturbation theory was applied to find the response of a system initially in state <math>|a\rangle </math> to a perturbation of the form <math>( H_{ba}/2 ) e^{-i\omega t}</math>. The result is that the amplitude for state <math>|b \rangle </math> is given by | In Chapter 6, first-order perturbation theory was applied to find the response of a system initially in state <math>|a\rangle </math> to a perturbation of the form <math>( H_{ba}/2 ) e^{-i\omega t}</math>. The result is that the amplitude for state <math>|b \rangle </math> is given by | ||
− | + | <equation id=" abem1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ a_ b (t) = \frac{1}{2 i\hbar } \int _0^ t H_{ba} e^{-i(\omega - \omega _{ba} )t^\prime } dt^\prime = \frac{H_{ba}}{2\hbar } {\left[ \frac{e^{-i(\omega - \omega _{ba} )t} -1}{\omega - \omega _{ba}} \right]} \end{align}</math> | |
− | + | </equation> | |
− | |||
There will be a similar expression involving the time-dependence <math>e^{+ i \omega t}</math>. The <math>- i \omega </math> term gives rise to resonance at <math>\omega = \omega _{ba}</math>; the <math>+ i \omega </math> term gives rise to resonance at <math>\omega = \omega _{ab}</math>. One term is responsible for absorption, the other is responsible for emission. | There will be a similar expression involving the time-dependence <math>e^{+ i \omega t}</math>. The <math>- i \omega </math> term gives rise to resonance at <math>\omega = \omega _{ba}</math>; the <math>+ i \omega </math> term gives rise to resonance at <math>\omega = \omega _{ab}</math>. One term is responsible for absorption, the other is responsible for emission. | ||
Line 490: | Line 429: | ||
The probability that the system has made a transition to state <math>| b \rangle </math> at time <math>t</math> is | The probability that the system has made a transition to state <math>| b \rangle </math> at time <math>t</math> is | ||
− | + | <equation id=" abem2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{a\rightarrow b} = | a_ b (t)|^2 = \frac{| H_{ba}|^2}{4 \hbar ^2} \frac{\sin ^2 [(\omega - \omega _{ba} )t/2]}{((\omega - \omega _{ba} )t/2)^2}t^2 \end{align}</math> | |
− | + | </equation> | |
− | |||
In the limit <math>\omega \rightarrow \omega _{ba}</math>, we have | In the limit <math>\omega \rightarrow \omega _{ba}</math>, we have | ||
− | + | <equation id=" abem3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{a\rightarrow b} \approx \frac{| H_{ba}|^2}{4 \hbar ^2} t^2 . \end{align}</math> | |
− | + | </equation> | |
− | |||
So, for short time, <math>W_{a\rightarrow b}</math> increases quadratically. This is reminiscent of a Rabi resonance in a 2-level system in the limit of short time. | So, for short time, <math>W_{a\rightarrow b}</math> increases quadratically. This is reminiscent of a Rabi resonance in a 2-level system in the limit of short time. | ||
− | However, | + | However, <xr id="abem2"/> is only valid provided <math>W_{a\rightarrow b} \ll 1</math>, or for time <math>T \ll \hbar /H_{ba}</math>. For such a short time, the incident radiation will have a spectral width <math>\Delta \omega \sim 1/T</math>. In this case, we must integrate <xr id="abem2"/> over the spectrum. In doing this, we shall make use of the relation |
− | + | <equation id=" abem4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \int _{-\infty }^{+\infty } \frac{\sin ^2 (\omega - \omega _{ba})t/2}{[(\omega - \omega _{ba})/2]^2} d \omega = 2t \int _{-\infty }^{+\infty } \frac{\sin ^2 (u - u_ o)}{(u - u_ o)^2} d u \rightarrow 2 \pi t \int _{-\infty }^{+\infty } \delta (\omega - \omega _{ba} ) d \omega . \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | <xr id="abem2"/> becomes | |
− | + | <equation id=" abem5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{a\rightarrow b} = \frac{| H_{ba}|^2}{\hbar ^2} 2\pi t \delta (\omega - \omega _{ba} ) \end{align}</math> | |
− | + | </equation> | |
− | |||
− | The <math>\delta </math>-function requires that eventually <math>W_{a\rightarrow b}</math> be integrated over a spectral distribution function. <math>W_{a\rightarrow b}</math> can | + | The <math>\delta </math>-function requires that eventually <math>W_{a\rightarrow b}</math> be integrated over a spectral distribution function. Absorbing an <math> \hbar </math> into the delta function, <math>W_{a\rightarrow b}</math> can be written |
− | + | <equation id=" abem6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{a\rightarrow b} = \frac{| H_{ba}|^2}{\hbar} 2\pi t \delta (E_ b - E_ a - \hbar \omega ). \end{align}</math> | |
− | + | </equation> | |
− | |||
Because the transition probability is proportional to the time, we can define the transition rate | Because the transition probability is proportional to the time, we can define the transition rate | ||
− | + | <equation id=" abem7a" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = \frac{d}{dt} W_{a\rightarrow b} = 2\pi \frac{| H_{ba}|^2}{\hbar} \delta (\omega - \omega _{ba}) \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | <equation id=" abem7b" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ = 2\pi \frac{| H_{ba}|^2}{\hbar } \delta (E_ b - E_ a - \hbar \omega ) \end{align}</math> | |
− | + | </equation> | |
− | |||
The <math>\delta </math>-function arises because of the assumption in first order perturbation theory that the amplitude of the initial state is not affected significantly. This will not be the case, for instance, if a monochromatic radiation field couples the two states, in which case the amplitudes oscillate between 0 and 1. However, the assumption of perfectly monochromatic radiation is in itself unrealistic. | The <math>\delta </math>-function arises because of the assumption in first order perturbation theory that the amplitude of the initial state is not affected significantly. This will not be the case, for instance, if a monochromatic radiation field couples the two states, in which case the amplitudes oscillate between 0 and 1. However, the assumption of perfectly monochromatic radiation is in itself unrealistic. | ||
Line 545: | Line 477: | ||
where <math>S_0</math> is the incident Poynting vector, and f(<math>\omega ^\prime </math>) is a normalized line shape function centered at the frequency <math>\omega ^\prime </math> which obeys <math>\int f (\omega ^\prime ) d\omega ^\prime = 1</math>. We can define a characteristic spectral width of <math>f(\omega ^\prime )</math> by | where <math>S_0</math> is the incident Poynting vector, and f(<math>\omega ^\prime </math>) is a normalized line shape function centered at the frequency <math>\omega ^\prime </math> which obeys <math>\int f (\omega ^\prime ) d\omega ^\prime = 1</math>. We can define a characteristic spectral width of <math>f(\omega ^\prime )</math> by | ||
− | + | <equation id=" abem8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Delta \omega = \frac{1}{f(\omega _{ab} )} \end{align}</math> | |
− | + | </equation> | |
− | |||
− | Integrating | + | Integrating <xr id="abem7b"/> over the spectrum of the radiation gives |
− | + | <equation id=" abem9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = \frac{2\pi | H_{ba}|^2}{\hbar ^2} f(\omega _{ab} ) \end{align}</math> | |
− | + | </equation> | |
− | |||
If we define the effective Rabi frequency by | If we define the effective Rabi frequency by | ||
− | + | <equation id=" abem10" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Omega _ R = \frac{| H_{ba}| }{\hbar } \end{align}</math> | |
− | + | </equation> | |
− | |||
then | then | ||
− | + | <equation id=" abem11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = {2 \pi } \frac{\Omega _ R^2}{\Delta \omega } \end{align}</math> | |
− | + | </equation> | |
− | |||
Another situation that often occurs is when the radiation is monochromatic, but the final state is actually composed of many states spaced close to each other in energy so as to form a continuum. If such is the case, the density of final states can be described by | Another situation that often occurs is when the radiation is monochromatic, but the final state is actually composed of many states spaced close to each other in energy so as to form a continuum. If such is the case, the density of final states can be described by | ||
− | + | <equation id=" abem12" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ dN= \rho (E) dE \end{align}</math> | |
− | + | </equation> | |
− | |||
− | where <math>dN</math> is the number of states in range <math>dE</math>. Taking <math>\hbar \omega = E_ b - E_ a</math> in | + | where <math>dN</math> is the number of states in range <math>dE</math>. Taking <math>\hbar \omega = E_ b - E_ a</math> in <xr id="abem7b"/>, and integrating gives |
− | + | <equation id=" abem13" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = 2\pi \frac{| H_{ba}|^2}{\hbar^2 } \rho (E_ b ) \end{align}</math> | |
− | + | </equation> | |
− | |||
This result remains valid in the limit <math>E_ b\rightarrow E_ a</math>, where <math>\omega \rightarrow 0</math>. In this static situation, the result is known as <i> | This result remains valid in the limit <math>E_ b\rightarrow E_ a</math>, where <math>\omega \rightarrow 0</math>. In this static situation, the result is known as <i> | ||
Line 589: | Line 515: | ||
</i>. | </i>. | ||
− | Note that | + | Note that <xr id="abem9"/> and <xr id="abem13"/> both describe a uniform rate process in which the population of the initial state decreases exponentially in time. If the population of the initial state is <math>P(0)</math>, then |
− | + | <equation id=" abem14" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ P(t) = P(0) e^{-\Gamma _{ba} t} \end{align}</math> | |
− | + | </equation> | |
− | |||
− | Applying this to the dipole transition described in | + | Applying this to the dipole transition described in <xr id="int11"/>, we have |
− | + | <equation id=" abem15" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = 2\pi \frac{E^2 d_{ba}^2}{\hbar ^2} f(\omega ) \end{align}</math> | |
− | + | </equation> | |
− | |||
The arguments here do not distinguish whether <math>E_ a < E_ b</math> or <math>E_ a > E_ b</math> (though the sign of <math>\omega = ( E_ b - E_ a )/\hbar </math> obviously does). In the former case the process is absorption, in the latter case it is emission. | The arguments here do not distinguish whether <math>E_ a < E_ b</math> or <math>E_ a > E_ b</math> (though the sign of <math>\omega = ( E_ b - E_ a )/\hbar </math> obviously does). In the former case the process is absorption, in the latter case it is emission. | ||
Line 611: | Line 535: | ||
<span id="SEC_sem"></span> | <span id="SEC_sem"></span> | ||
− | The rate of absorption for the transition <math>a \rightarrow b</math>, where <math>E_ b > E_ a</math>, is, from | + | The rate of absorption, in CGS units, for the transition <math>a \rightarrow b</math>, where <math>E_ b > E_ a</math>, is, from <xr id="qrd16"/> and <xr id="abem7b"/>, |
− | + | <equation id=" sem1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = \frac{4\pi ^2}{\hbar V} | {\bf \hat{e}} \cdot {\bf d}_{ba} |^2 n\omega \delta (\omega _0 -\omega ) . \end{align}</math> | |
− | + | </equation> | |
− | |||
where <math>\omega _0 = ( E_ b - E_ a ) /\hbar </math>. To evaluate this we need to let <math>n \rightarrow n (\omega )</math>, where <math>n (\omega ) d\omega </math> is the number of photons in the frequency interval <math>d\omega </math>, and integrate over the spectrum. The result is | where <math>\omega _0 = ( E_ b - E_ a ) /\hbar </math>. To evaluate this we need to let <math>n \rightarrow n (\omega )</math>, where <math>n (\omega ) d\omega </math> is the number of photons in the frequency interval <math>d\omega </math>, and integrate over the spectrum. The result is | ||
− | + | <equation id=" sem2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = \frac{4\pi ^2}{\hbar V} | {\bf \hat{e}}\cdot {\bf d}_{ba} |^2 \omega _0 n(\omega _0 ) \end{align}</math> | |
− | + | </equation> | |
− | |||
To calculate <math>n (\omega )</math>, we first calculate the mode density in space by applying the usual periodic boundary condition | To calculate <math>n (\omega )</math>, we first calculate the mode density in space by applying the usual periodic boundary condition | ||
− | + | <equation id=" sem3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ k_ j L = 2\pi n_ j , ~ ~ ~ j = x,y,z. \end{align}</math> | |
− | + | </equation> | |
− | |||
The number of modes in the range <math>d^3 k = dk_ x dk_ y dk_ z</math> is | The number of modes in the range <math>d^3 k = dk_ x dk_ y dk_ z</math> is | ||
− | + | <equation id=" sem4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ dN = dn_ x dn_ y dn_ z = \frac{V}{{\left(2 \pi \right)^3} } d^3 k=\frac{V}{{\left(2 \pi \right)^3} }k^2 dk \ d\Omega = \frac{V}{{\left(2 \pi \right)^3} } \frac{\omega ^2\, d\omega \ d\Omega }{c^3} \end{align}</math> | |
− | + | </equation> | |
− | |||
Letting <math>\bar{n} = \bar{n (\omega ) }</math> be the average number of photons per mode, then | Letting <math>\bar{n} = \bar{n (\omega ) }</math> be the average number of photons per mode, then | ||
− | + | <equation id=" sem5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ n (\omega ) = \bar{n} \frac{dN}{d\omega } = \frac{\bar{n} V\omega ^2 d\Omega }{(2\pi )^3 c^3} \end{align}</math> | |
− | + | </equation> | |
− | |||
− | Introducing this into | + | Introducing this into <xr id="sem2"/> gives |
− | + | <equation id=" sem6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = \frac{\bar{n}\omega ^3}{2\pi \hbar c^3} | {\bf \hat{e}} \cdot {\bf d}_{ba} |^2 d\Omega \end{align}</math> | |
− | + | </equation> | |
− | |||
− | We wish to apply this to the case of isotropic radiation in free space, as, for instance, in a thermal radiation field. We can take <math>{\bf d}_{ba}</math> to lie along the <math>z</math> axis and describe '''k''' in spherical coordinates about this axis. Since the wave is transverse, <math>{\bf \hat{e}} \cdot {\bf \hat{D}} = \sin \theta </math> | + | We wish to apply this to the case of isotropic radiation in free space, as, for instance, in a thermal radiation field. We can take <math>{\bf d}_{ba}</math> to lie along the <math>z</math> axis and describe '''k''' in spherical coordinates about this axis. Since the wave is transverse, <math>{\bf \hat{e}} \cdot {\bf \hat{D}} = \sin \theta </math> for one polarization, and zero for the other one. Consequently, |
− | + | <equation id=" sem7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \int | {\bf \hat{e}} \cdot {\bf d}_{ba} |^2 d\Omega = | {\bf d}_{ba} |^2 \int \sin ^2 \theta d\Omega = \frac{8\pi }{3} | {\bf d}_{ba}|^2 \end{align}</math> | |
− | + | </equation> | |
− | |||
− | Introducing this into | + | Introducing this into <xr id="sem6"/> yields the absorption rates |
− | + | <equation id=" sem8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ab} = \frac{4}{3} \frac{\omega ^3}{\hbar c^3} | {\bf d}_{ba} |^2 \bar{n} \end{align}</math> | |
− | + | </equation> | |
− | |||
It follows that the emission rate for the transition <math>b\rightarrow a</math> is | It follows that the emission rate for the transition <math>b\rightarrow a</math> is | ||
− | + | <equation id=" sem9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ba} = \frac{4}{3} \frac{\omega ^3}{\hbar c^3} | {\bf d}_{ba} |^2 (\bar{n} + 1) \end{align}</math> | |
− | + | </equation> | |
− | |||
If there are no photons present, the emission rate—called the rate of spontaneous emission—is | If there are no photons present, the emission rate—called the rate of spontaneous emission—is | ||
− | + | <equation id=" sem10" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ba}^0 = \frac{4}{3} \frac{ \omega ^3}{\hbar c^3} | {\bf d}_{ba}|^2 = \frac{4}{3} \frac{e^2\omega ^3}{\hbar c^3} | \langle b| {\bf r} | a \rangle |^2 \end{align}</math> | |
− | + | </equation> | |
− | |||
In atomic units, in which <math>c = 1 / \alpha </math>, we have | In atomic units, in which <math>c = 1 / \alpha </math>, we have | ||
− | + | <equation id=" sem11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Gamma _{ba}^0 = \frac{4}{3} \alpha ^3 \omega ^3 | {\bf r}_{ba} |^2 . \end{align}</math> | |
− | + | </equation> | |
− | |||
Taking, typically, <math>\omega = 1</math>, and <math>r_{ba}= 1</math>, we have <math>\Gamma ^0 \approx \alpha ^3</math>. The “<math>Q</math>'' of a radiative transition is <math>Q =\omega /\Gamma \approx \alpha ^{-3}\approx </math> <math>3 \times 10^6</math>. The <math>\alpha ^3</math> dependence of <math>\Gamma </math> indicates that radiation is fundamentally a weak process: hence the high <math>Q</math> and the relatively long radiative lifetime of a state, <math>\tau = 1 /\Gamma </math>. For example, for the <math>2P\rightarrow 1S</math> transition in hydrogen (the <math>L_{\alpha }</math> transition), we have <math>\omega = 3/8</math>, and taking <math>r_{2p,1s} \approx 1</math>, we find <math>\tau = 3.6\times 10^7</math> atomic units, or 0.8 ns. The actual lifetime is 1.6 ns. | Taking, typically, <math>\omega = 1</math>, and <math>r_{ba}= 1</math>, we have <math>\Gamma ^0 \approx \alpha ^3</math>. The “<math>Q</math>'' of a radiative transition is <math>Q =\omega /\Gamma \approx \alpha ^{-3}\approx </math> <math>3 \times 10^6</math>. The <math>\alpha ^3</math> dependence of <math>\Gamma </math> indicates that radiation is fundamentally a weak process: hence the high <math>Q</math> and the relatively long radiative lifetime of a state, <math>\tau = 1 /\Gamma </math>. For example, for the <math>2P\rightarrow 1S</math> transition in hydrogen (the <math>L_{\alpha }</math> transition), we have <math>\omega = 3/8</math>, and taking <math>r_{2p,1s} \approx 1</math>, we find <math>\tau = 3.6\times 10^7</math> atomic units, or 0.8 ns. The actual lifetime is 1.6 ns. | ||
Line 693: | Line 606: | ||
<br style="clear: both" /> | <br style="clear: both" /> | ||
+ | |||
== Line Strength == | == Line Strength == | ||
Line 699: | Line 613: | ||
Because the absorption and stimulated emission rates are proportional to the spontaneous emission rate, we shall focus our attention on the Einstein A coefficient: | Because the absorption and stimulated emission rates are proportional to the spontaneous emission rate, we shall focus our attention on the Einstein A coefficient: | ||
− | + | <equation id=" lines1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ A_{ba} = \frac{4}{3} \frac{e^2\omega ^3}{\hbar c^3} | \langle b | {\bf r} | a \rangle |^2 \end{align}</math> | |
− | + | </equation> | |
− | |||
where | where | ||
− | + | <equation id=" lines2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ | \langle b | {\bf r} | a \rangle |^2 = | \langle b | x | a \rangle |^2 + | \langle b | y | a \rangle |^2 + | \langle b | z | a \rangle |^2 \end{align}</math> | |
− | + | </equation> | |
− | |||
For an isolated atom, the initial and final states will be eigenstates of total angular momentum. (If there is an accidental degeneracy, as in hydrogen, it is still possible to select angular momentum eigenstates.) If the final angular momentum is <math>J_ a</math>, then the atom can decay into each of the <math>2 J_ a + 1</math> final states, characterized by the azimuthal quantum number <math>m_ a = -J_ a , -J_ a + 1,\dots , +J_ a</math>. Consequently, | For an isolated atom, the initial and final states will be eigenstates of total angular momentum. (If there is an accidental degeneracy, as in hydrogen, it is still possible to select angular momentum eigenstates.) If the final angular momentum is <math>J_ a</math>, then the atom can decay into each of the <math>2 J_ a + 1</math> final states, characterized by the azimuthal quantum number <math>m_ a = -J_ a , -J_ a + 1,\dots , +J_ a</math>. Consequently, | ||
− | + | <equation id=" lines3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ A_{ba} = \frac{4}{3} \frac{e^2\omega ^3}{\hbar c^3}\sum _{m_ a} | \langle b, J_ b | {\bf r} |a, J_ a, m_ a \rangle |^2 \end{align}</math> | |
− | + | </equation> | |
− | |||
The upper level, however, is also degenerate, with a (<math>2 J_ b + 1</math>)–fold degeneracy. The lifetime cannot depend on which state the atom happens to be in. This follows from the isotropy of space: <math>m_ b</math> depends on the orientation of <math>{\bf J}_ b</math> with respect to some direction in space, but the decay rate for an isolated atom can't depend on how the atom happens to be oriented. Consequently, it is convenient to define the <i> | The upper level, however, is also degenerate, with a (<math>2 J_ b + 1</math>)–fold degeneracy. The lifetime cannot depend on which state the atom happens to be in. This follows from the isotropy of space: <math>m_ b</math> depends on the orientation of <math>{\bf J}_ b</math> with respect to some direction in space, but the decay rate for an isolated atom can't depend on how the atom happens to be oriented. Consequently, it is convenient to define the <i> | ||
Line 722: | Line 633: | ||
</i> <math>S_{ba}</math>, given by | </i> <math>S_{ba}</math>, given by | ||
− | + | <equation id=" lines4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ S_{ba} = S_{ab} = \sum _{m_ b} \sum _{m_ a} | \langle b, J_ b, m_ b | {\bf r} | a, J_ a, m_ a \rangle |^2 \end{align}</math> | |
− | + | </equation> | |
− | |||
Then, | Then, | ||
− | + | <equation id=" lines5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ A_{ba} = \frac{4}{3} \frac{e^2\omega ^3}{\hbar c^3} \frac{S_{ba}}{g_ b} = \frac{4}{3} \frac{e^2\omega ^3}{\hbar c^3} \frac{S_{ba}}{2J_ b +1} \end{align}</math> | |
− | + | </equation> | |
− | |||
The line strength is closely related to the average oscillator strength <math>\bar{f}_{ab}</math>. <math>\bar{f}_{ab}</math> is obtained by averaging <math>f_{ab}</math> over the initial state <math>|b\rangle </math>, and summing over the values of <math>m</math> in the final state, <math>|a\rangle </math>. For absorption, <math>\omega _{ab} > 0</math>, and | The line strength is closely related to the average oscillator strength <math>\bar{f}_{ab}</math>. <math>\bar{f}_{ab}</math> is obtained by averaging <math>f_{ab}</math> over the initial state <math>|b\rangle </math>, and summing over the values of <math>m</math> in the final state, <math>|a\rangle </math>. For absorption, <math>\omega _{ab} > 0</math>, and | ||
− | + | <equation id=" line11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \bar{f}_{ab} = \frac{2m}{3\hbar } \omega _{ab} \frac{1}{2J_ b + 1} \sum _{m_ b} \sum _{m_ a} |\langle b, J_ b, m_ b |{\bf r} | a, J_ a, m_ a \rangle |^2 \end{align}</math> | |
− | + | </equation> | |
− | |||
It follows that | It follows that | ||
− | + | <equation id=" line12" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \bar{f}_{ba} = - \frac{2J_ b + 1}{2J_ a +1} \bar{f}_{ab} . \end{align}</math> | |
− | + | </equation> | |
− | |||
In terms of the oscillator strength, we have | In terms of the oscillator strength, we have | ||
− | + | <equation id=" line13" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \bar{f}_{ab} = \frac{2m}{3\hbar }\omega _{ab} \frac{1}{2J_ b + 1} {S}_{ab} . \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | <equation id=" line14" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \bar{f}_{ba} = - \frac{2m}{3\hbar } | \omega _{ab} | \frac{1}{2J_ a + 1} {S}_{ab} . \end{align}</math> | |
− | + | </equation> | |
− | |||
<br style="clear: both" /> | <br style="clear: both" /> | ||
Line 766: | Line 671: | ||
<span id="SEC_BROAD"></span> | <span id="SEC_BROAD"></span> | ||
− | We have calculated the rate of absorption and emission of an atom in a thermal field, but a more common situation involves interaction with a light beam, either monochromatic or broad band. Here | + | We have calculated the rate of absorption and emission of an atom in a thermal field, but a more common situation involves interaction with a light beam, either monochromatic or broad band. Here ''broad band'' means having a spectral width that is broad compared to the natural line width of the system—the spontaneous decay rate. |
For an electric dipole transition, the radiation interaction is | For an electric dipole transition, the radiation interaction is | ||
− | + | <equation id=" broad1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ | H_{ba} | = e | {\bf r}_{ba} |\cdot {\bf \hat{e}} E/2, \end{align}</math> | |
− | + | </equation> | |
− | |||
− | where <math>E </math> is the amplitude of the field. The transition rate, from | + | where <math>E </math> is the amplitude of the field. The transition rate, from <xr id="sem7"/>, is |
− | + | <equation id=" broad2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{ab} = \frac{\pi }{2} \frac{e^2 | {\bf \hat{e}} \cdot {\bf r}_{ba} |^2 E^2}{\hbar ^2} f (\omega _0 ) = \frac{\pi }{2} \frac{e^2 | {\bf \hat{e}} \cdot {\bf r}_{ba} |^2 E^2}{\hbar } f(E_ b - E_ a ) \end{align}</math> | |
− | + | </equation> | |
− | |||
where <math>\omega _0 = ( E_ b - E_ a )/\hbar </math> and <math>f (\omega )</math> is the normalized line shape function, or alternatively, the normalized density of states, expressed in frequency units. The transition rate is proportional to the intensity <math>I_0</math> of a monochromatic radiation source. <math>I_0</math> is given by the Poynting vector, and can be expressed by the electric field as <math>E^2 = 8 \pi I_0 / c</math>. Consequently, | where <math>\omega _0 = ( E_ b - E_ a )/\hbar </math> and <math>f (\omega )</math> is the normalized line shape function, or alternatively, the normalized density of states, expressed in frequency units. The transition rate is proportional to the intensity <math>I_0</math> of a monochromatic radiation source. <math>I_0</math> is given by the Poynting vector, and can be expressed by the electric field as <math>E^2 = 8 \pi I_0 / c</math>. Consequently, | ||
− | + | <equation id=" broad3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{ab} = \frac{4\pi ^2}{c} \frac{e^2 | {\bf \hat{e}} \cdot {\bf r}_{ba} |^2}{\hbar ^2} I_0 f (\omega _0 ) \end{align}</math> | |
− | + | </equation> | |
− | |||
In the case of a Lorentzian line having a FWHM of <math>\Gamma _0</math> centered on frequency <math>\omega _0</math>, | In the case of a Lorentzian line having a FWHM of <math>\Gamma _0</math> centered on frequency <math>\omega _0</math>, | ||
− | + | <equation id=" broad4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ f(\omega ) = \frac{1}{\pi } \frac{(\Gamma _0 /2)}{(\omega - \omega _0 )^2 + (\Gamma _0 /2)^2} \end{align}</math> | |
− | + | </equation> | |
− | |||
In this case, | In this case, | ||
− | + | <equation id=" broad5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{ab} = \frac{8\pi e^2}{c\hbar ^2 \Gamma _0} | \langle b | {\bf \hat{e}} \cdot {\bf r} | a \rangle |^2 I_0 \end{align}</math> | |
− | + | </equation> | |
− | |||
Note that <math>W_{ab}</math> is the rate of transition between two particular quantum states, not the total rate between energy levels. Naturally, we also have <math> W_{ab} = W_{ba}</math>.\ | Note that <math>W_{ab}</math> is the rate of transition between two particular quantum states, not the total rate between energy levels. Naturally, we also have <math> W_{ab} = W_{ba}</math>.\ | ||
− | An alternative way to express | + | An alternative way to express <xr id="broad2"/> is to introduce the Rabi frequency, |
− | + | <equation id=" broad6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \Omega _ R = \frac{2 H_{ba}}{\hbar } = \frac{e |{\bf \hat{e}}\cdot {\bf r}_{ba} | E}{\hbar } \end{align}</math> | |
− | + | </equation> | |
− | |||
In which case | In which case | ||
− | + | <equation id=" broad7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{ab} = \frac{\pi }{2} \Omega _ R^2 f (\omega _0 ) = \Omega _ R^2 \frac{1}{\Gamma _0} \end{align}</math> | |
− | + | </equation> | |
− | |||
If the width of the final state is due soley to spontaneous emission, <math>\Gamma _0 = A = ( 4 e^2 \omega ^3 / 3 \hbar c^3 ) | r_{ba} |^2</math>. Since <math>W_{ab}</math> is proportional to <math> | r_{ba} |^2 /A_0</math>, it is independent of <math> | r_{ba} |^2</math>. It is left as a problem to find the exact relationship, but it can readily be seen that it is of the form | If the width of the final state is due soley to spontaneous emission, <math>\Gamma _0 = A = ( 4 e^2 \omega ^3 / 3 \hbar c^3 ) | r_{ba} |^2</math>. Since <math>W_{ab}</math> is proportional to <math> | r_{ba} |^2 /A_0</math>, it is independent of <math> | r_{ba} |^2</math>. It is left as a problem to find the exact relationship, but it can readily be seen that it is of the form | ||
− | + | <equation id=" broad8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_{ab} = X\lambda ^2 I_0 /\hbar \omega \end{align}</math> | |
− | + | </equation> | |
− | |||
where X is a numerical factor. <math>I/ \hbar \omega </math> is the photon flux—i.e. the number of photons per second per unit area in the beam. Since <math>W_{ab}</math> is an excitation rate, we interpret <math>X\lambda ^2</math> as the resonance absorption cross section for the atom, <math>\sigma _0</math>. | where X is a numerical factor. <math>I/ \hbar \omega </math> is the photon flux—i.e. the number of photons per second per unit area in the beam. Since <math>W_{ab}</math> is an excitation rate, we interpret <math>X\lambda ^2</math> as the resonance absorption cross section for the atom, <math>\sigma _0</math>. | ||
Line 832: | Line 729: | ||
There is, however, an important feature of absorption that does depend on the oscillator strength. <math>\sigma _0</math> is the cross section assuming that the radiation is monochromatic compared to the natural line width. As the spontaneous decay rate becomes smaller and smaller, eventually the natural linewidth becomes narrower than the spectral width of the laser, or whatever source is used. In that case, the excitation becomes broad band. | There is, however, an important feature of absorption that does depend on the oscillator strength. <math>\sigma _0</math> is the cross section assuming that the radiation is monochromatic compared to the natural line width. As the spontaneous decay rate becomes smaller and smaller, eventually the natural linewidth becomes narrower than the spectral width of the laser, or whatever source is used. In that case, the excitation becomes broad band. | ||
− | We now discuss broad band excitation. Using the result of the last section, finding the excitation rate or the absorption cross section for broad band excitation is trivial. From | + | === Broad Band Excitation === |
+ | |||
+ | We now discuss broad band excitation. Using the result of the last section, finding the excitation rate or the absorption cross section for broad band excitation is trivial. From <xr id="broad2"/>, the absorption rate is proportional to <math>f(\omega _0 )</math>. For monochromatic excitation, <math>f (\omega _0 ) = (2/ \pi ) A^{-1} </math> and <math>W_{\rm mono}= X\lambda ^2 I_0/\hbar \omega </math>. For a spectral source having linewidth <math>\Delta \omega _ s</math>, defined so that the normalized line shape function is <math>f (\omega _0 ) = (2/ \pi ) {\Delta \omega _ s}^{-1} </math>, then the broad band excitation rate is obtained by replacing <math>\Gamma _0</math> with <math>\Delta \omega _ s</math> in <xr id="broad8"/>. Thus | ||
− | + | <equation id=" band1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ W_ B = {\left( X\lambda ^2 \frac{\Gamma _0}{\Delta \omega _ s}\right)} \frac{I_0}{\hbar \omega } \end{align}</math> | |
− | + | </equation> | |
− | |||
Similarly, the effective absorption cross section is | Similarly, the effective absorption cross section is | ||
− | + | <equation id=" band2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \sigma _{\rm eff} = \sigma _0 \frac{\Gamma _0}{\Delta \omega _ s} \end{align}</math> | |
− | + | </equation> | |
− | |||
This relation is valid provided <math>\Delta \omega _ s \gg \Gamma _0</math>. If the two widths are comparable, the problem needs to be worked out in detail, though the general behavior would be for <math>\Delta \omega _ s \rightarrow ( \Delta \omega _ s^2 + \Gamma _0^2 )^{1/2}</math>. Note that <math>\Delta \omega _ s</math> represents the actual resonance width. Thus, if Doppler broadening is the major broadening mechanism then | This relation is valid provided <math>\Delta \omega _ s \gg \Gamma _0</math>. If the two widths are comparable, the problem needs to be worked out in detail, though the general behavior would be for <math>\Delta \omega _ s \rightarrow ( \Delta \omega _ s^2 + \Gamma _0^2 )^{1/2}</math>. Note that <math>\Delta \omega _ s</math> represents the actual resonance width. Thus, if Doppler broadening is the major broadening mechanism then | ||
− | + | <equation id=" band3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \sigma _{\rm eff} = \sigma _0 \Gamma _0 /\Delta \omega _{\rm Doppler} . \end{align}</math> | |
− | + | </equation> | |
− | |||
Except in the case of high resolution laser spectroscopy, it is generally true that <math>\Delta \omega _ s \gg \Gamma _0</math>, so that <math>\sigma _{\rm eff}\ll \sigma _0</math>. | Except in the case of high resolution laser spectroscopy, it is generally true that <math>\Delta \omega _ s \gg \Gamma _0</math>, so that <math>\sigma _{\rm eff}\ll \sigma _0</math>. | ||
− | < | + | === Saturation and Saturated Absorption Rates === |
− | == | + | |
+ | When the external light intensity is strong, the population in the excited state is no longer negligible, and the transition is saturated. We define the saturated absorption rate <math>R^s </math> as the net transfer from initial state <math>a </math> to final state <math>b </math>, and <math>R^u </math> is the unsaturated rate for the stimulated absorption and emission, | ||
+ | |||
+ | <equation id=" sat1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \ R^s (n_a+n_b) = R^u (n_a-n_b). \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | When the system reaches steady state, | ||
+ | |||
+ | <equation id=" sat2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \ \dot{n_b}&=&-n_b(R^u+\Gamma)+n_aR^u =0\\ \dot{n_a}&=&n_b(R^u+\Gamma)-n_aR^u =0 \\ \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | which gives | ||
+ | |||
+ | <equation id=" sat3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \ \frac{n_b}{n_a}=\frac{R^u}{R^u+\Gamma} \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | From <xr id="sat1"/>, we have | ||
+ | |||
+ | <equation id=" sat4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \ R^s=\frac{\Gamma}{2}\frac{S}{1+S}=\frac{R^u}{1+S} \end{align}</math> | ||
+ | </equation> | ||
− | < | + | where <math>S </math> is the saturation parameter and is defined as <math>S=2R^u/\Gamma </math>. The transition rate is reduced by a factor of <math>1+S</math> due to saturation. |
− | + | For low intensity light, <math>S\ll 1</math>, and <math>R^s=R^u</math>; for very high intensity light, <math>S\gg 1</math>, <math>R^s=\Gamma/2</math>. | |
− | + | For the case of monochromatic radiation, as discussed above, the unsaturated transition rate | |
− | |||
− | |||
− | |||
− | + | <equation id=" sat5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ R^u=W_{ab}=\frac{\pi }{2}\omega_R^2 f(\omega )= \frac{\omega_R^2 }{\Gamma} \frac{1}{1+(2\delta/\Gamma)^2} \end{align}</math> | ||
+ | </equation> | ||
+ | where <math>\delta </math> the detuning with respect to the center frequency <math>\omega_0 </math>. | ||
− | + | Thus in general the saturated transition rate | |
− | |||
− | |||
− | |||
− | + | <equation id=" sat6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ R^s= \frac{\omega_R^2 }{\Gamma} \frac{1}{1+(2\delta/\Gamma)^2+2\omega_R^2/\Gamma^2} \end{align}</math> | ||
+ | </equation> | ||
+ | and the saturation parameter | ||
+ | <equation id=" sat7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \ S= \frac{S_{res}}{1+(2\delta/\Gamma)^2} \end{align}</math> | ||
+ | </equation> | ||
− | + | with the resonant saturation parameter <math>S_{res}=2\omega_R^2/\Gamma^2</math>. | |
− | |||
− | |||
− | |||
− | + | The saturated rate <math>R^s </math> has a Lorentzian line with FWHM | |
− | + | <equation id=" sat8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \delta_{FWHM }=\frac{\Gamma}{2}\sqrt{1+S_{res}} \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | === Power Broadening === | |
− | + | This resultant increase in the spectrum width is called saturation (or power) broadening. | |
− | |||
− | |||
− | |||
− | + | The saturation intensity <math>I_{sat} </math> is the light field intensity corresponding to the saturation parameter <math>S_{res}=1 </math> for a resonant light, and that is when <math>R^u=\omega_R^2/\Gamma=\Gamma/2</math>. Since the Rabi frequency <math>\omega_R^2\propto I</math>, we have the linear relation | |
− | + | <equation id=" sat9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ \omega_R^2=\frac{\Gamma^2}{2}\frac{I}{I_{sat}} \end{align}</math> | ||
+ | </equation> | ||
− | + | and that gives | |
− | + | <equation id=" sat10" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ I_{sat}=\frac{\Gamma^2}{2}\frac{I}{\omega_R^2}=\frac{\hbar \omega^3}{12\pi c^2}\Gamma \end{align}</math> | |
− | + | </equation> | |
− | } | + | for example, <math>I_{sat}=6\; mW/cm^2</math> for Na D line. |
− | + | === Saturation Intensity === | |
− | + | A quick derivation for the saturation intensity is to express the light intensity <math>I </math> and the Rabi frequency <math>\omega_R </math> in terms of the number of photons <math>n </math>, | |
− | + | <equation id=" sat11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ I=\frac{Energy}{Area\times Time}=\frac{\hbar\omega n}{V/c}=\frac{\hbar\omega nc}{V} \end{align}</math> | |
− | + | </equation> | |
− | + | <equation id=" sat12" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ \omega_R^2=(n+1)\omega_1^2\simeq n (\vec{d}\cdot \hat{e})^2 \left(\frac{2}{\hbar}\right)^2 \left(\frac{\hbar\omega}{2\epsilon_0 V}\right)=n\Gamma\frac{6\pi c^3}{\omega^2 V} \end{align}</math> | ||
+ | </equation> | ||
− | + | thus | |
+ | <equation id=" sat13" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \ \frac{I}{\omega^2_R}=\frac{\hbar \omega^3}{6\pi c^2\Gamma } \end{align}</math> | ||
+ | </equation> | ||
+ | and pluging this into <xr id=" sat9"/> gives the saturation intensity. | ||
− | + | For the case of broadband radiation, we define the average intensity per frequency interval as <math>\bar{I} </math>, and when the saturation parameter <math>S=1 </math>, <math>\bar{I}=\bar{I}_{sat} </math> | |
− | |||
− | |||
− | |||
− | + | <equation id=" sat14" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ W_{ge}=B_{ge}\frac{\bar{I}}{c}=\frac{\Gamma}{2} \end{align}</math> | ||
+ | </equation> | ||
+ | thus | ||
− | + | <equation id=" sat15" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \bar{I}_{sat}=\frac{c}{2}\frac{A}{B_{ge}}=\frac{\hbar\omega_{eg}^3}{6\pi^2 c^2} \end{align}</math> | |
− | + | </equation> | |
− | }} | + | which is independent of matrix element! For visible light, <math>\bar{I}_{sat}\approx \frac{12 \;W}{cm^2}\frac{1}{cm^{-1}}</math>, where <math>1 \;cm^{-1}\simeq 30 \;GHz </math>. |
− | + | === Absorption Cross Section === | |
− | + | Cross section is the effective area that represents the probability of some scattering or absorption event. In the case of atom-photon interaction, the absorption rate is the collision rate of an atom with the incoming photons, <math>R=n_{phot}\sigma c </math>. | |
− | + | For monochromatic radiation, | |
− | |||
− | |||
− | |||
− | + | <equation id=" sat16" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ W_{ge}=n_{phot}\sigma c=\frac{I\sigma}{\hbar\omega} \end{align}</math> | ||
+ | </equation> | ||
+ | in the low intensity limit <math>W_{ge}=R^u </math>. If we extrapolate it to saturation parameter <math>S=1 </math>, then <math>I=I_{sat} </math>, and <math>W_{ge}=R^u=\Gamma/2 </math> | ||
− | + | <equation id=" sat17" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \frac{\Gamma}{2}=\frac{I_{sat}\sigma}{\hbar \omega} \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | and from <xr id=" sat10"/>, we have | |
− | + | <equation id=" sat18" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ \sigma=6\pi\frac{c^2}{\omega^2}=6\pi (\lambda/2\pi)^2 \end{align}</math> | ||
+ | </equation> | ||
+ | This is the resonant cross section for weak radiation, and it is usually much larger than the size of the atom, and independent of matrix element. If we plot the cross section as a function of detuning, it is a Lorentzian line. Strong transitions have a larger widths, but the cross section on resonance is always the same. | ||
− | + | When the transition is saturated at high intensity, the resonant cross section goes as <math>\sigma=\sigma_0/(1+S) </math>. The transition bleaches out <math>\sigma\rightarrow 0 </math> when <math>S\gg 1 </math>. | |
− | |||
− | |||
− | |||
− | + | For broadband radiation, | |
− | + | <equation id=" sat19" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ W_{ge}&=&\int \sigma(\omega)\frac{\bar{I}(\omega)}{\hbar\omega}d\omega \\ &=& \frac{6\pi (\lambda/2\pi)^2}{\hbar\omega_{eg}}\bar{I}(\omega_{eg})\frac{\pi\Gamma}{2} \end{align}</math> | ||
+ | </equation> | ||
− | + | at saturation <math>S=1 </math>, | |
− | |||
− | |||
− | |||
− | + | <equation id=" sat20" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \frac{\Gamma}{2}=\frac{6\pi (\lambda/2\pi)^2}{\hbar\omega_{eg}}\bar{I}_{sat}\frac{\pi\Gamma}{2} \end{align}</math> | |
− | </ | + | </equation> |
+ | thus | ||
+ | <equation id=" sat21" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \ \bar{I}_{sat}=\frac{\hbar\omega_{eg}}{6\pi^2 (\lambda/2\pi)^2}=\frac{\hbar \omega_{eg}^3}{6\pi^2 c^2} \end{align}</math> | ||
+ | </equation> | ||
+ | which is the same as we have derived in <xr id=" sat15"/>. | ||
<br style="clear: both" /> | <br style="clear: both" /> | ||
− | == | + | == Higher-order radiation processes == |
+ | |||
+ | Beyond the dipole approximation: | ||
+ | Recall that the interaction Hamiltonian for an atom in an electromagnetic field is given by | ||
+ | |||
+ | <equation id="Hint" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} H_\mathrm{int} = -\frac{e}{mc} p\cdot A+\frac{e^2}{2mc^2}|A|^2+g_s\mu_B S\cdot({\bf\nabla}\times A), \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | where the last term we have so far considered only for static magnetic fields. Neglecting, as before, the <math>|A|^2</math> term, which is appreciable only for very intense fields, we now consider more fully the dominant term in the atom-field interaction, | ||
+ | |||
+ | <equation id="hor1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} H_{ba} = \frac{e}{\rm mc} \langle | ||
+ | b | p \cdot A (r) | a\rangle. \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | For concreteness, we shall take A(r) to be a plane wave of the form | ||
+ | |||
+ | <equation id="hor2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} A (r) = A\hat{z} e^{ikx}. \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | Expanding the exponential, we have | ||
+ | |||
+ | <equation id="hor3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} H_{ba} = \frac{eA}{\rm mc} \langle b | p_z (1+ikx + (ikx)^2/2 + | ||
+ | \dots ) | a\rangle. \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | Thus far in the course, we have considered only the first term, the dipole term. If dipole radiation is forbidden, for instance if <math>| a \rangle</math> and <math>| b \rangle</math> have the same parity, then the second term in the parentheses becomes important. Usually, it is <math>\alpha</math> times smaller. In particular, since | ||
+ | |||
+ | <equation id="hor4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} k r \approx \frac{\hbar\omega}{\hbar c}a_0\approx\frac{e^2/a_0}{\hbar c}a_0\approx\frac{e^2}{\hbar c}=\alpha, \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | the expansion in <xr id = "hor3"/> is effectively an expansion in <math>\alpha</math>. | ||
+ | We can rewrite the second term as follows: | ||
+ | |||
+ | <equation id="hor5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} p_z x = (p_z x - zp_x )/2 + (p_z x + zp_x )/2 . \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | The first term of <xr id="hor4"/> is <math>- \hbar L_y/2</math>, and the matrix element becomes | ||
+ | |||
+ | <equation id="hor6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} -\frac{ieAk}{2 m} \langle b | \hbar L_y | | ||
+ | a \rangle = - iAk \langle b | \mu_B | ||
+ | L_y | a \rangle, \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | where <math>\mu_B = e\hbar /2 m</math> is the Bohr magneton. | ||
+ | The magnetic field is <math>B = - i k A \hat{y}</math>. | ||
+ | Consequently, <xr id="hor5"/> can be written in the more | ||
+ | familiar form <math>-\vec{\mu} \cdot B</math>. (The orbital magnetic moment is <math>\vec{\mu} | ||
+ | = -\mu_B L</math>: the minus sign arises from our convention that <math>e</math> is | ||
+ | positive.) | ||
+ | We can readily generalize the matrix element to | ||
+ | |||
+ | <equation id="hor7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} H_{\rm int}(M1) = B \cdot \mu_B\langle b |L + g_sS| a\rangle, \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | where we have added the spin dependent term from <xr id="Hint"/>. <math>M1</math> indicates that the matrix element is for a magnetic dipole transition. The strength of the <math>M1</math> transition is set by | ||
+ | |||
+ | <equation id="hor8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \mu_B/c = \frac{1}{2}\frac{e\hbar}{mc}=\frac{1}{2}\frac{e^2}{\hbar c}\frac{\hbar^2}{e m} = \frac{1}{2}\alpha e a_0, \end{align}</math> | ||
+ | </equation> | ||
+ | |||
+ | so it is indeed a factor of <math>\alpha</math> weaker than a dipole transition, as we argued above. | ||
+ | |||
+ | The second term in <xr id ="hor4"/> involves <math>( p_z x + z p_x | ||
+ | )/2</math>. | ||
+ | Making use of the commutator relation <math>[ r, H_0 ] = i\hbar | ||
+ | p / m </math>, we | ||
+ | have | ||
− | <span | + | <equation id="hor9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> |
+ | <math>\begin{align} \frac{1}{2} (p_z x + z p_x) = \frac{m}{2i\hbar} ([z, H_0 ] x+ z[x, | ||
+ | H_0 ]) = | ||
+ | \frac{m}{2i\hbar} (- H_0 zx +zx H_0 ). \end{align}</math> | ||
+ | </equation> | ||
− | + | So, the contribution of this term to <math>H_{ba}</math> is | |
− | + | <equation id="hor10" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \frac{ieA}{mc} \frac{km}{2\hbar i} \langle b | - H_0 zx + zx H_0 | | |
− | + | a \rangle = - \frac{eAk}{2c} \frac{E_b - E_a}{\hbar} \langle b | | |
− | + | zx | a \rangle | |
+ | = \frac{ieE\omega}{2c} \langle b | zx | a \rangle, \end{align}</math> | ||
+ | </equation> | ||
− | + | where we have taken <math>E = i k A</math>. This is an electric | |
+ | quadrupole interaction, and we shall denote the matrix element by | ||
− | + | <equation id="hor11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} H_{\rm int} (E2)^\prime = \frac{ie\omega}{2c} \langle b | zx | a | ||
+ | \rangle E. \end{align}</math> | ||
+ | </equation> | ||
− | The | + | The prime indicates that we are considering only one component of a |
+ | more general expression involving the matrix element <math>\langle b |r:r|a\rangle</math> of a tensor product. It is straightforward to verify that the electric quadrupole interaction is also of order <math>\alpha</math>. | ||
− | + | The total matrix element of the second term in the expansion of <xr id="hor3"/> can be written | |
− | |||
− | |||
− | |||
− | + | <equation id="hor12" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} H_{\rm int}^{(2)} = H_{\rm int} (M1) + H_{\rm int} (E2). \end{align}</math> | ||
+ | </equation> | ||
− | + | Note that <math>H_{\rm int} (M1)</math> is | |
− | + | real, whereas <math>H_{\rm int} (E2)</math> is imaginary. Consequently, | |
− | |||
− | |||
− | + | <equation id="hor13" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} | H_{\rm int}^{(2)} |^2 = | H_{\rm int} (M1)|^2 + | | ||
+ | H_{\rm int}(E2) |^2. \end{align}</math> | ||
+ | </equation> | ||
− | + | The magnetic dipole and electric quadrupole terms do not | |
− | + | interfere. | |
− | |||
− | |||
− | + | Because transition rates depend on <math>|H_{ba} |^2</math>, the magnetic dipole and electric quadrupole rates | |
+ | are both smaller than the dipole rate by <math>\alpha^2 \sim 5 \times | ||
+ | 10^{-5}</math>. For this reason they are generally referred to as ''forbidden'' | ||
+ | processes. However, the term is used somewhat loosely, for there | ||
+ | are transitions which are much more strongly suppressed due to other | ||
+ | selection rules, as for instance triplet to singlet transitions in | ||
+ | helium. | ||
+ | {| class="wikitable" | ||
+ | |- | ||
+ | |Transition | ||
+ | | | ||
+ | |Operator | ||
+ | |Parity | ||
+ | |- | ||
+ | |Electric Dipole | ||
+ | |<math>E1</math> | ||
+ | |<math>-er</math> | ||
+ | | - | ||
+ | |- | ||
+ | |Magnetic Dipole | ||
+ | |<math>M1</math> | ||
+ | |<math>-\mu_B(L+g_sS)</math> | ||
+ | | + | ||
+ | |- | ||
+ | |Electric Quadrupole | ||
+ | |<math>E2</math> | ||
+ | |<math>-er:r</math> | ||
+ | | + | ||
+ | |- | ||
+ | |+Summary of dipole and higher-order radiation processes. | ||
+ | |} | ||
− | {{ | + | == Selection rules == |
− | + | A forbidden transition, then, is one that is weaker than an electric dipole-allowed transition by <math>\alpha^n</math> and only appears in some higher-order approximation. Examples of such higher-order effects are the magnetic dipole and electric quadrupole terms described above, multiphoton processes, the relativistic effects which allow singlet to triplet transitions in helium, and hyperfine interactions within the nucleus. | |
− | + | To derive selection rules for the transitions we have discussed above, it is useful to express the matrix elements in terms of spherical tensor operators: | |
− | }} | + | :<math> |
+ | H_{int}(T_{l,m}) = \langle n J M | T_{l,m} | n' J' M'\rangle, | ||
+ | </math> | ||
+ | where <math>T_{l,m}</math> is a spherical tensor operator of rank <math>l</math>. The operators <math>T_{l,m}</math> transform under rotations like the spherical harmonics <math>Y_{l,m}</math>, and any operator can be written as a linear combination of these spherical tensors. By the Wigner-Eckart Theorem, we can express the matrix element | ||
+ | :<math> | ||
+ | \langle n J M | T_{l,m} | n' J' M'\rangle = \frac{\langle n J \| T_l \| n' J' \rangle}{\sqrt{2J+1}}\langle J' l, M', m| J M\rangle | ||
+ | </math> | ||
+ | in terms of a reduced matrix element <math>\langle n J \| T_l \| n' J' \rangle</math> and a Clebsch-Gordan coefficient <math>\langle J' l, M', m| J M\rangle</math>. In order for the latter to be nonzero, the triangle rule requires that <math>|J'-J| \leq l \leq |J'+J|</math>, while conservation of angular momentum requires <math>M = M' + m</math>. | ||
+ | Since the operators <math>er</math> and <math>\mu_B B</math> responsible for <math>E1</math> and <math>M1</math> transitions are both vectors, i.e. tensors of rank <math>l=1</math>, these transitions are both governed by the dipole selection rules | ||
+ | :<math>\begin{align} | ||
+ | |\Delta J| &= 0, 1;\\ | ||
+ | |\Delta m| &= 0, 1. | ||
+ | \end{align}</math> | ||
+ | Since <math>\bf r</math> is a polar vector and <math>\bf L</math> is an axial vector, <math>E1</math> transitions are allowed only between states of opposite parity and <math>M</math> transitions are allowed only between states of the same parity. | ||
+ | The operator responsible for <math>E2</math> transitions is a spherical tensor of rank 2. For example, | ||
+ | :<math> | ||
+ | xz = (T_{2,-1}-T_{2,1})/4. | ||
+ | </math> | ||
− | + | In general, then, we expect that the quadrupole moment can be expressed in terms of <math>T_{2, M} ({\bf r})</math>. Thus, electric quadrupole transitions are allowed only between states connected by tensors <math>T_{2,m}(r)</math>, requiring: | |
+ | :<math>\begin{align} | ||
+ | |\Delta J| & = 0, 1, 2; \\ | ||
+ | |\Delta m| &= 0, 1, 2. | ||
+ | \end{align}</math> | ||
− | + | and parity unchanged. | |
− | |||
− | |||
− | |||
− | + | In addition, <math>J=0\rightarrow J'=0</math> transitions are forbidden in all of the cases | |
− | + | considered above, since <math>J=J'=0</math> requires <math>\Delta L=0</math> (for any interaction that | |
− | + | does not couple to spin) whereas absorption or emission of a photon implies | |
− | + | <math>|\Delta L|=1</math>. | |
− | The | + | We now illustrate the use of the spherical tensor for the case of a vector. |
+ | The dipole matrix element for a particular polarization of the field, <math>\hat{\bf {e}}</math>, is | ||
− | + | <equation id=" select1" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ {\bf \hat{e}} \cdot {\bf r}_{ba} = {\bf \hat{e}} \cdot \langle b, J_ b, m_ b | {\bf r} | a, J_ a , m_ a \rangle . \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | It is straightforward to calculate <math>x_{ba}, y_{ba}, z_{ba},</math> but a more general approach is to write '''r''' in terms of a spherical tensor. This yields the selection rules directly, and allows the matrix element to be calculated for various geometries using the Wigner-Eckart theorem as discussed above. | |
− | + | The orbital angular momentum operator of a system with total angular momentum <math>L</math> can be written in terms of a spherical harmonic <math>Y_{L,M}</math>. Consequently, the spherical harmonics constitute spherical tensor operators. A vector can be written in terms of spherical harmonics of rank 1. This permits the vector operator '''r''' to be expressed in terms of the spherical tensor <math>T_{1,M}({\bf r}).</math> | |
− | |||
− | |||
− | |||
− | + | The spherical harmonics of rank 1 are | |
− | + | <equation id=" select2" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ Y_{1,0} = \sqrt {\frac{3}{4\pi }} \cos \theta ; \qquad Y_{1, +1} = - \sqrt {\frac{3}{8\pi }} \sin \theta e^{+i\phi }\qquad Y_{1,-1} = \sqrt {\frac{3}{8\pi }} \sin \theta e^{-i\phi } \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | These are normalized so that | |
− | + | <equation id=" select3" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ \int Y_{1,m^\prime }^* Y_{1,m} \sin \theta d\theta d\phi = \delta _{m^\prime , m} \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | We can write the vector '''r''' in terms of components <math>r_ m ,\ m = +1, 0, -1</math>, | |
− | + | <equation id=" select4" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
+ | <math>\begin{align} \ r_0 = r\sqrt {\frac{4\pi }{3}} Y_{1,0} ,\qquad r_{\pm } = r\sqrt {\frac{4\pi }{3}} Y_{1,\pm 1} , \end{align}</math> | ||
+ | </equation> | ||
− | + | or, more generally | |
− | + | <equation id=" select5" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | |
− | + | <math>\begin{align} \ r_ M = rT_{1,M} (\theta , \phi ) \end{align}</math> | |
− | + | </equation> | |
− | |||
− | + | Consequently, | |
− | == | + | <equation id=" select6" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> |
+ | <math>\begin{align} \ \langle b, J_ b, m_ b | r_ M | a, J_ a, m_ a \rangle = \langle b, J_ b, m_ b | rT_{1,M} | a, J_ a, m_ a \rangle \end{align}</math> | ||
+ | </equation> | ||
− | < | + | <equation id=" select7" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> |
− | < | + | <math>\begin{align} \ = \langle b, J_ b | r | a, J_ a \rangle \langle J_ b, m_ b | T_{1,M} | J_ a, m_ a \rangle \end{align}</math> |
− | < | + | </equation> |
− | </ | ||
− | < | ||
− | |||
− | |||
− | |||
− | </ | ||
− | |||
+ | The first factor is independent of <math>m</math>. It is | ||
− | < | + | <equation id=" select8" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> |
− | < | + | <math>\begin{align} \ r_{ba} = \int _0^{\infty } R_{b,J_ b}^* (r) r R_{a,J_ a} (r) r^2 dr \end{align}</math> |
− | + | </equation> | |
− | |||
− | < | ||
− | |||
− | |||
− | </ | ||
+ | where <math>r_{ba}</math> contains the radial part of the matrix element. It vanishes unless <math>| b \rangle </math> and <math>| a \rangle </math> have opposite parity. The second factor in <xr id="select7"/> yields the selection rule | ||
− | </ | + | <equation id=" select9" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> |
− | </ | + | <math>\begin{align} \ | J_ b - J_ a | = 0, 1; ~ ~ ~ m_ b = m_ a \pm M = m_ a, m_ a \pm 1 \end{align}</math> |
+ | </equation> | ||
− | < | + | Similarly, for magnetic dipole transition, <xr id="hor6"/>, we have |
− | </ | + | <equation id=" select10" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> |
+ | <math>\begin{align} \ H_{ba} (M1) = \mu _ B B \langle b, J_ b, m_ b , | T_{LM} (L) | a, J_ a , m_ a \rangle \end{align}</math> | ||
+ | </equation> | ||
+ | It immediately follows that parity is unchanged, and that | ||
+ | <equation id=" select11" noautocaption><span style="float:right; display:block;"><caption>(%i)</caption></span> | ||
+ | <math>\begin{align} \ | \Delta J | = 0,1 ~ ~ ~ (J=0\rightarrow J= 0~ \mbox{forbidden}); ~ ~ | \Delta m | = 0,1 \end{align}</math> | ||
+ | </equation> | ||
+ | This discussion of matrix elements, selection rules, and radiative processes barely skims the subject. For an authoritative treatment, the books by Shore and Manzel, and Sobelman are recommended. | ||
− | + | == References == | |
− | + | '''JAC63''' E.T. Jaynes and F.W. Cummings, Proc. IEEE, '''51''', 89 (1963). | |
− | IEEE, 51, 89 (1963). | ||
− | + | '''EIN17''' A. Einstein, Z. Phys. '''18''', 121 (1917), reprinted in English by D. ter Haar, ''The Old Quantum Theory'', Pergammon, Oxford. | |
− | in English by D. | ||
− | + | '''EIN17a''' A. Einstein, Z. Phys. '''18''', 121 (1917), translated in ''Sources of Quantum Mechanics'', B. L. Van der Waerden, Cover Publication, Inc., New York, 1967. This book is a gold mine for anyone interested in the development of quantum mechanics. | |
[[Category:8.421]] | [[Category:8.421]] |
Latest revision as of 18:04, 10 March 2016
This section introduces the interaction of atoms with radiative modes of the electromagnetic field.
Contents
- 1 Introduction: Spontaneous and Stimulated Emission
- 2 Quantum Theory of Absorption and Emission
- 3 Quantization of the radiation field
- 4 Interaction of a two-level system and a single mode of the radiation field
- 5 Absorption and emission
- 6 Spontaneous emission rate
- 7 Line Strength
- 8 Excitation by narrow and broad band light sources
- 9 Higher-order radiation processes
- 10 Selection rules
- 11 References
Introduction: Spontaneous and Stimulated Emission
Einstein's 1917 paper on the theory of radiation [1] provided seminal concepts for the quantum theory of radiation. It also anticipated devices such as the laser, and pointed the way to the field of laser-cooling of atoms. In it, he set out to answer two questions:
1) How do the internal states of an atom that radiates and absorbs energy come into equilibrium with a thermal radiation field? (In answering this question Einstein invented the concept of spontaneous emission)
2) How do the translational states of an atom in thermal equilibrium (i.e. states obeying the Maxwell-Boltzmann Law for the distribution of velocities) come into thermal equilibrium with a radiation field? (In answering this question, Einstein introduced the concept of photon recoil. He also demonstrated that the field itself must obey the Planck radiation law.)
The first part of Einstein's paper, which addresses question 1), is well known, but the second part, which addresses question 2), is every bit as germane to contemporary atom/optical physics. Because the paper preceded the creation of quantum mechanics there was no way for him to calculate transition rates. However, his arguments are based on general statistical principles and provide the foundation for interpreting the quantum mechanical results.
Einstein considered a system of atoms in thermal equilibrium with a radiation field. The system has two levels (an energy level consists of all of the states that have a given energy; the number of quantum states in a given level is its multiplicity.) with energies and , with , and . The numbers of atoms in the two levels are related by . Einstein assumed the Planck radiation law for the spectral energy density temperature. For radiation in thermal equilibrium at temperature , the energy per unit volume in wavelength range is:
<equation id="erad1" noautocaption>(%i) </equation>
The mean occupation number of a harmonic oscillator at temperature , which can be interpreted as the mean number of photons in one mode of the radiation field, is
<equation id="erad2" noautocaption>(%i) </equation>
According to the Boltzmann Law of statistical mechanics, in thermal equilibrium the populations of the two levels are related by
<equation id="erad3" noautocaption>(%i) </equation>
Here and are the multiplicities of the two levels. The last step assumes the Bohr frequency condition, . However, Einstein's paper actually derives this relation independently.
According to classical theory, an oscillator can exchange energy with the radiation field at a rate that is proportional to the spectral density of radiation. The rates for absorption and emission are equal. The population transfer rate equation is thus predicted to be
<equation id="erad4" noautocaption>(%i) </equation>
This equation is compatible with <xr id="erad1"/>, <xr id="erad2"/>, <xr id="erad3"/> it follows that
<equation id=" erl5" noautocaption>(%i) </equation>
Consequently, the rate of transition is
<equation id=" erl6" noautocaption>(%i) </equation>
while the rate of absorption is
<equation id=" erl7" noautocaption>(%i) </equation>
If we consider emission and absorption between single states by taking , then the ratio of rate of emission to rate of absorption is .
This argument reveals the fundamental role of spontaneous emission. Without it, atomic systems could not achieve thermal equilibrium with a radiation field. Thermal equilibrium requires some form of dissipation, and dissipation is equivalent to having an irreversible process. Spontaneous emission is the fundamental irreversible process in nature. The reason that it is irreversible is that once a photon is radiated into the vacuum, the probability that it will ever be reabsorbed is zero: there are an infinity of vacuum modes available for emission but only one mode for absorption. If the vacuum modes are limited, for instance by cavity effects, the number of modes becomes finite and equilibrium is never truly achieved. In the limit of only a single mode, the motion becomes reversible.
The identification of the Einstein coefficient with the rate of spontaneous emission is so well established that we shall henceforth use the symbol to denote the spontaneous decay rate from state to . The radiative lifetime for such a transition is .
Here, Einstein came to a halt. Lacking quantum theory, there was no way to calculate .
Quantum Theory of Absorption and Emission
We shall start by describing the behavior of an atom in a classical electromagnetic field. Although treating the field classically while treating the atom quantum mechanically is fundamentally inconsistent, it provides a natural and intuitive approach to the problem. Furthermore, it is completely justified in cases where the radiation fields are large, in the sense that there are many photons in each mode, as for instance, in the case of microwave or laser spectroscopy. There is, however, one important process that this approach cannot deal with satisfactorily. This is spontaneous emission, which we shall treat later using a quantized field. Nevertheless, phenomenological properties such as selection rules, radiation rates and cross sections, can be developed naturally with this approach.
The classical E-M field
Our starting point is Maxwell's equations (S.I. units):
<equation id="Maxwell" noautocaption>(%i) </equation>
The charge density and current density J obey the continuity equation
<equation id=" wd2" noautocaption>(%i) </equation>
Introducing the vector potential A and the scalar potential , we have
<equation id=" wd3" noautocaption>(%i) </equation>
We are free to change the potentials by a gauge transformation:
<equation id=" wd4" noautocaption>(%i) </equation>
where is a scalar function. This transformation leaves the fields invariant, but changes the form of the dynamical equation. We shall work in the Coulomb gauge (often called the radiation gauge), defined by
<equation id=" wd5" noautocaption>(%i) </equation>
In free space, A obeys the wave equation
<equation id=" wd6" noautocaption>(%i) </equation>
Because , A is transverse. We take a propagating plane wave solution of the form
<equation id="A-field" noautocaption>(%i) </equation>
where and . For a linearly polarized field, the polarization vector is real. For an elliptically polarized field it is complex, and for a circularly polarized field it is given by , where the + and signs correspond to positive and negative helicity, respectively. (Alternatively, they correspond to left and right hand circular polarization, respectively, the sign convention being a tradition from optics.) The electric and magnetic fields are then given by
<equation id="E-field" noautocaption>(%i) </equation>
<equation id="B-field" noautocaption>(%i) </equation>
The time average Poynting vector is
<equation id=" wd9" noautocaption>(%i) </equation>
The average energy density in the wave is given by
<equation id="energy-density" noautocaption>(%i) </equation>
Interaction of an electromagnetic wave and an atom
The behavior of charged particles in an electromagnetic field is correctly described by Hamilton's equations provided that the canonical momentum is redefined:
<equation id=" int1" noautocaption>(%i) </equation>
The kinetic energy is . Taking , the Hamiltonian for an atom in an electromagnetic field in free space is
<equation id=" int2" noautocaption>(%i) </equation>
where describes the potential energy due to internal interactions. We are neglecting spin interactions.
Expanding and rearranging, we have
<equation id=" int3" noautocaption>(%i) </equation>
Here, . Consequently, describes the unperturbed atom. describes the atom's interaction with the field. , which is second order in A, plays a role only at very high intensities. (In a static magnetic field, however, gives rise to diamagnetism.)
Because we are working in the Coulomb gauge, so that A and p commute. We have
<equation id=" int4" noautocaption>(%i) </equation>
It is convenient to write the matrix element between states and in the form
<equation id=" int5" noautocaption>(%i) </equation>
where
<equation id=" int6" noautocaption>(%i) </equation>
Atomic dimensions are small compared to the wavelength of radiation involved in optical transitions. The scale of the ratio is set by . Consequently, when the matrix element in <xr id="int6"/> is evaluated, the wave function vanishes except in the region where . It is therefore appropriate to expand the exponential:
<equation id=" int7" noautocaption>(%i) </equation>
Unless vanishes, for instance due to parity considerations, the leading term dominates and we can neglect the others. For reasons that will become clear, this is called the dipole approximation. This is by far the most important situation, and we shall defer consideration of the higher order terms. In the dipole approximation we have
<equation id=" int8" noautocaption>(%i) </equation>
where we have used, from <xr id="E-field"/>, . It can be shown (i.e. left as exercise) that the matrix element of p can be transfomred into a matrix element for :
<equation id=" int9" noautocaption>(%i) </equation>
This results in
<equation id=" int10" noautocaption>(%i) </equation>
We will be interested in resonance phenomena in which . Consequently,
<equation id=" int11" noautocaption>(%i) </equation>
where d is the dipole operator, . Displaying the time dependence explictlty, we have
<equation id=" int12" noautocaption>(%i) </equation>
However, it is important to bear in mind that this is only the first term in a series, and that if it vanishes the higher order terms will contribute a perturbation at the driving frequency.
appears as a matrix element of the momentum operator p in <xr id="int8"/>, and of the dipole operator r in <xr id="int11"/>. These matrix elements look different and depend on different parts of the wave function. The momentum operator emphasizes the curvature of the wave function, which is largest at small distances, whereas the dipole operator evaluates the moment of the charge distribution, i.e. the long range behavior. In practice, the accuracy of a calculation can depend significantly on which operator is used.
Quantization of the radiation field
We shall consider a single mode of the radiation field. This means a single value of the wave vector k, and one of the two orthogonal transverse polarization vectors . The radiation field is described by a plane wave vector potential of the form <xr id="A-field"/>. We assume that k obeys a periodic boundary or condition, , etc. (For any k, we can choose boundaries to satisfy this.) The time averaged energy density is given by <xr id="energy-density"/>, and the total energy in the volume V defined by these boundaries is
<equation id="energy-total" noautocaption>(%i) </equation>
where is the mean squared value of averaged over the spatial mode. We now make a formal connection between the radiation field and a harmonic oscillator. We define variables Q and P by
<equation id=" qrd5" noautocaption>(%i) </equation>
Then, from <xr id="energy-total"/>, we find
<equation id=" qrd6" noautocaption>(%i) </equation>
This describes the energy of a harmonic oscillator having unit mass. We quantize the oscillator in the usual fashion by treating Q and P as operators, with
<equation id=" qrd7" noautocaption>(%i) </equation>
We introduce the operators and defined by
<equation id=" qrd8" noautocaption>(%i) </equation>
<equation id=" qrd9" noautocaption>(%i) </equation>
The fundamental commutation rule is
<equation id=" qrd10" noautocaption>(%i) </equation>
from which the following can be deduced:
<equation id=" qrd11" noautocaption>(%i) </equation>
where the number operator obeys
<equation id=" qrd12" noautocaption>(%i) </equation>
We also have
<equation id=" qrd13" noautocaption>(%i) </equation>
The operators and are called the annihilation and creation operators, respectively. We can express the vector potential and electric field in terms of and as follows
<equation id=" part1" noautocaption>(%i) </equation>
<equation id=" part2" noautocaption>(%i) </equation>
<equation id=" part3" noautocaption>(%i) </equation>
In the dipole limit we can take . Then
<equation id=" part3" noautocaption>(%i) </equation>
The interaction Hamiltonian is,
<equation id=" qrd16" noautocaption>(%i) </equation>
where we have written the dipole operator as .
Interaction of a two-level system and a single mode of the radiation field
We consider a two-state atomic system , and a radiation field described by The states of the total system can be taken to be
<equation id=" vac1" noautocaption>(%i) </equation>
We shall take . Then
<equation id=" vac2" noautocaption>(%i) </equation>
The first term in the bracket obeys the selection rule . This corresponds to loss of one photon from the field and absorption of one photon by the atom. The second term obeys . This corresponds to emission of a photon by the atom. Using <xr id="qrd13"/>, we have
<equation id=" vac3" noautocaption>(%i) </equation>
Transitions occur when the total time dependence is zero, or near zero. Thus absorption occurs when , or . As we expect, energy is conserved. Similarly, emission occurs when , or .
A particularly interesting case occurs when , i.e. the field is initially in the vacuum state, and . Then
<equation id=" vac4" noautocaption>(%i) </equation>
The situation describes a constant perturbation coupling the two states and . The states are degenerate because . Consequently, is the upper of the two atomic energy levels.
The system is composed of two degenerate eigenstates, but due to the coupling of the field, the degeneracy is split. The eigenstates are symmetric and antisymmetric combinations of the initial states, and we can label them as
<equation id=" vac5" noautocaption>(%i) </equation>
The energies of these states are
<equation id=" vac6" noautocaption>(%i) </equation>
If at , the atom is in state which means that the radiation field is in state then the system is in a superposition state:
<equation id=" vac7" noautocaption>(%i) </equation>
The time evolution of this superposition is given by
<equation id=" vac8" noautocaption>(%i) </equation>
where . The probability that the atom is in state at a later time is
<equation id=" vac9" noautocaption>(%i) </equation>
The frequency is called the vacuum Rabi frequency.
The dynamics of a 2-level atom interacting with a single mode of the vacuum were first analyzed in [2] and the oscillations are sometimes called Jaynes-Cummings oscillations.
The atom-vacuum interaction , <xr id="vac4"/>, has a simple physical interpretation. The electric field amplitude associated with the zero point energy in the cavity is given by
<equation id=" vac10" noautocaption>(%i) </equation>
Consequently, . The interaction frequency is sometimes referred to as the vacuum Rabi frequency, although, as we have seen, the actual oscillation frequency is .
Absorption and emission are closely related. Because the rates are proportional to , it is evident from <xr id="vac3"/> that
<equation id=" vac11" noautocaption>(%i) </equation>
This result, which applies to radiative transitions between any two states of a system, is general. In the absence of spontaneous emission, the absorption and emission rates are identical.
The oscillatory behavior described by <xr id="vac8"/> is exactly the opposite of free space behavior in which an excited atom irreversibly decays to the lowest available state by spontaneous emission. The distinction is that in free space there are an infinite number of final states available to the photon, since it can go off in any direction, but in the cavity there is only one state. The natural way to regard the atom-cavity system is not in terms of the atom and cavity separately, as in <xr id="vac1"/>, but in terms of the coupled states and (<xr id="vac5"/>). Such states, called dressed atom states, are the true eigenstates of the atom-cavity system.
Absorption and emission
In Chapter 6, first-order perturbation theory was applied to find the response of a system initially in state to a perturbation of the form . The result is that the amplitude for state is given by
<equation id=" abem1" noautocaption>(%i) </equation>
There will be a similar expression involving the time-dependence . The term gives rise to resonance at ; the term gives rise to resonance at . One term is responsible for absorption, the other is responsible for emission.
The probability that the system has made a transition to state at time is
<equation id=" abem2" noautocaption>(%i) </equation>
In the limit , we have
<equation id=" abem3" noautocaption>(%i) </equation>
So, for short time, increases quadratically. This is reminiscent of a Rabi resonance in a 2-level system in the limit of short time.
However, <xr id="abem2"/> is only valid provided , or for time . For such a short time, the incident radiation will have a spectral width . In this case, we must integrate <xr id="abem2"/> over the spectrum. In doing this, we shall make use of the relation
<equation id=" abem4" noautocaption>(%i) </equation>
<xr id="abem2"/> becomes
<equation id=" abem5" noautocaption>(%i) </equation>
The -function requires that eventually be integrated over a spectral distribution function. Absorbing an into the delta function, can be written
<equation id=" abem6" noautocaption>(%i) </equation>
Because the transition probability is proportional to the time, we can define the transition rate
<equation id=" abem7a" noautocaption>(%i) </equation>
<equation id=" abem7b" noautocaption>(%i) </equation>
The -function arises because of the assumption in first order perturbation theory that the amplitude of the initial state is not affected significantly. This will not be the case, for instance, if a monochromatic radiation field couples the two states, in which case the amplitudes oscillate between 0 and 1. However, the assumption of perfectly monochromatic radiation is in itself unrealistic.
Radiation always has some spectral width. is proportional to the intensity of the radiation field at resonance. The intensity can be written in terms of a spectral density function
where is the incident Poynting vector, and f() is a normalized line shape function centered at the frequency which obeys . We can define a characteristic spectral width of by
<equation id=" abem8" noautocaption>(%i) </equation>
Integrating <xr id="abem7b"/> over the spectrum of the radiation gives
<equation id=" abem9" noautocaption>(%i) </equation>
If we define the effective Rabi frequency by
<equation id=" abem10" noautocaption>(%i) </equation>
then
<equation id=" abem11" noautocaption>(%i) </equation>
Another situation that often occurs is when the radiation is monochromatic, but the final state is actually composed of many states spaced close to each other in energy so as to form a continuum. If such is the case, the density of final states can be described by
<equation id=" abem12" noautocaption>(%i) </equation>
where is the number of states in range . Taking in <xr id="abem7b"/>, and integrating gives
<equation id=" abem13" noautocaption>(%i) </equation>
This result remains valid in the limit , where . In this static situation, the result is known as Fermi's Golden Rule .
Note that <xr id="abem9"/> and <xr id="abem13"/> both describe a uniform rate process in which the population of the initial state decreases exponentially in time. If the population of the initial state is , then
<equation id=" abem14" noautocaption>(%i) </equation>
Applying this to the dipole transition described in <xr id="int11"/>, we have
<equation id=" abem15" noautocaption>(%i) </equation>
The arguments here do not distinguish whether or (though the sign of obviously does). In the former case the process is absorption, in the latter case it is emission.
Spontaneous emission rate
The rate of absorption, in CGS units, for the transition , where , is, from <xr id="qrd16"/> and <xr id="abem7b"/>,
<equation id=" sem1" noautocaption>(%i) </equation>
where . To evaluate this we need to let , where is the number of photons in the frequency interval , and integrate over the spectrum. The result is
<equation id=" sem2" noautocaption>(%i) </equation>
To calculate , we first calculate the mode density in space by applying the usual periodic boundary condition
<equation id=" sem3" noautocaption>(%i) </equation>
The number of modes in the range is
<equation id=" sem4" noautocaption>(%i) </equation>
Letting be the average number of photons per mode, then
<equation id=" sem5" noautocaption>(%i) </equation>
Introducing this into <xr id="sem2"/> gives
<equation id=" sem6" noautocaption>(%i) </equation>
We wish to apply this to the case of isotropic radiation in free space, as, for instance, in a thermal radiation field. We can take to lie along the axis and describe k in spherical coordinates about this axis. Since the wave is transverse, for one polarization, and zero for the other one. Consequently,
<equation id=" sem7" noautocaption>(%i) </equation>
Introducing this into <xr id="sem6"/> yields the absorption rates
<equation id=" sem8" noautocaption>(%i) </equation>
It follows that the emission rate for the transition is
<equation id=" sem9" noautocaption>(%i) </equation>
If there are no photons present, the emission rate—called the rate of spontaneous emission—is
<equation id=" sem10" noautocaption>(%i) </equation>
In atomic units, in which , we have
<equation id=" sem11" noautocaption>(%i) </equation>
Taking, typically, , and , we have . The “ of a radiative transition is . The dependence of indicates that radiation is fundamentally a weak process: hence the high and the relatively long radiative lifetime of a state, . For example, for the transition in hydrogen (the transition), we have , and taking , we find atomic units, or 0.8 ns. The actual lifetime is 1.6 ns.
The lifetime for a strong transition in the optical region is typically 10–100 ns. Because of the dependence of , the radiative lifetime for a transition in the microwave region—for instance an electric dipole rotational transition in a molecule—is longer by the factor , yielding lifetimes on the order of months. Furthermore, if the transition moment is magnetic dipole rather than electric dipole, the lifetime is further increased by a factor of , giving a time of thousands of years.
Line Strength
Because the absorption and stimulated emission rates are proportional to the spontaneous emission rate, we shall focus our attention on the Einstein A coefficient:
<equation id=" lines1" noautocaption>(%i) </equation>
where
<equation id=" lines2" noautocaption>(%i) </equation>
For an isolated atom, the initial and final states will be eigenstates of total angular momentum. (If there is an accidental degeneracy, as in hydrogen, it is still possible to select angular momentum eigenstates.) If the final angular momentum is , then the atom can decay into each of the final states, characterized by the azimuthal quantum number . Consequently,
<equation id=" lines3" noautocaption>(%i) </equation>
The upper level, however, is also degenerate, with a ()–fold degeneracy. The lifetime cannot depend on which state the atom happens to be in. This follows from the isotropy of space: depends on the orientation of with respect to some direction in space, but the decay rate for an isolated atom can't depend on how the atom happens to be oriented. Consequently, it is convenient to define the line strength , given by
<equation id=" lines4" noautocaption>(%i) </equation>
Then,
<equation id=" lines5" noautocaption>(%i) </equation>
The line strength is closely related to the average oscillator strength . is obtained by averaging over the initial state , and summing over the values of in the final state, . For absorption, , and
<equation id=" line11" noautocaption>(%i) </equation>
It follows that
<equation id=" line12" noautocaption>(%i) </equation>
In terms of the oscillator strength, we have
<equation id=" line13" noautocaption>(%i) </equation>
<equation id=" line14" noautocaption>(%i) </equation>
Excitation by narrow and broad band light sources
We have calculated the rate of absorption and emission of an atom in a thermal field, but a more common situation involves interaction with a light beam, either monochromatic or broad band. Here broad band means having a spectral width that is broad compared to the natural line width of the system—the spontaneous decay rate.
For an electric dipole transition, the radiation interaction is
<equation id=" broad1" noautocaption>(%i) </equation>
where is the amplitude of the field. The transition rate, from <xr id="sem7"/>, is
<equation id=" broad2" noautocaption>(%i) </equation>
where and is the normalized line shape function, or alternatively, the normalized density of states, expressed in frequency units. The transition rate is proportional to the intensity of a monochromatic radiation source. is given by the Poynting vector, and can be expressed by the electric field as . Consequently,
<equation id=" broad3" noautocaption>(%i) </equation>
In the case of a Lorentzian line having a FWHM of centered on frequency ,
<equation id=" broad4" noautocaption>(%i) </equation>
In this case,
<equation id=" broad5" noautocaption>(%i) </equation>
Note that is the rate of transition between two particular quantum states, not the total rate between energy levels. Naturally, we also have .\
An alternative way to express <xr id="broad2"/> is to introduce the Rabi frequency,
<equation id=" broad6" noautocaption>(%i) </equation>
In which case
<equation id=" broad7" noautocaption>(%i) </equation>
If the width of the final state is due soley to spontaneous emission, . Since is proportional to , it is independent of . It is left as a problem to find the exact relationship, but it can readily be seen that it is of the form
<equation id=" broad8" noautocaption>(%i) </equation>
where X is a numerical factor. is the photon flux—i.e. the number of photons per second per unit area in the beam. Since is an excitation rate, we interpret as the resonance absorption cross section for the atom, .
At first glance it is puzzling that does not depend on the structure of the atom; one might expect that a transition with a large oscillator strength—i.e. a large value of —should have a large absorption cross section. However, the absorption rate is inversely proportional to the linewidth, and since that also increases with , the two factors cancel out. This behavior is not limited to electric dipole transitions, but is quite general.
There is, however, an important feature of absorption that does depend on the oscillator strength. is the cross section assuming that the radiation is monochromatic compared to the natural line width. As the spontaneous decay rate becomes smaller and smaller, eventually the natural linewidth becomes narrower than the spectral width of the laser, or whatever source is used. In that case, the excitation becomes broad band.
Broad Band Excitation
We now discuss broad band excitation. Using the result of the last section, finding the excitation rate or the absorption cross section for broad band excitation is trivial. From <xr id="broad2"/>, the absorption rate is proportional to . For monochromatic excitation, and . For a spectral source having linewidth , defined so that the normalized line shape function is , then the broad band excitation rate is obtained by replacing with in <xr id="broad8"/>. Thus
<equation id=" band1" noautocaption>(%i) </equation>
Similarly, the effective absorption cross section is
<equation id=" band2" noautocaption>(%i) </equation>
This relation is valid provided . If the two widths are comparable, the problem needs to be worked out in detail, though the general behavior would be for . Note that represents the actual resonance width. Thus, if Doppler broadening is the major broadening mechanism then
<equation id=" band3" noautocaption>(%i) </equation>
Except in the case of high resolution laser spectroscopy, it is generally true that , so that .
Saturation and Saturated Absorption Rates
When the external light intensity is strong, the population in the excited state is no longer negligible, and the transition is saturated. We define the saturated absorption rate as the net transfer from initial state to final state , and is the unsaturated rate for the stimulated absorption and emission,
<equation id=" sat1" noautocaption>(%i) </equation>
When the system reaches steady state,
<equation id=" sat2" noautocaption>(%i) </equation>
which gives
<equation id=" sat3" noautocaption>(%i) </equation>
From <xr id="sat1"/>, we have
<equation id=" sat4" noautocaption>(%i) </equation>
where is the saturation parameter and is defined as . The transition rate is reduced by a factor of due to saturation.
For low intensity light, , and ; for very high intensity light, , .
For the case of monochromatic radiation, as discussed above, the unsaturated transition rate
<equation id=" sat5" noautocaption>(%i) </equation> where the detuning with respect to the center frequency .
Thus in general the saturated transition rate
<equation id=" sat6" noautocaption>(%i) </equation> and the saturation parameter <equation id=" sat7" noautocaption>(%i) </equation>
with the resonant saturation parameter .
The saturated rate has a Lorentzian line with FWHM
<equation id=" sat8" noautocaption>(%i) </equation>
Power Broadening
This resultant increase in the spectrum width is called saturation (or power) broadening.
The saturation intensity is the light field intensity corresponding to the saturation parameter for a resonant light, and that is when . Since the Rabi frequency , we have the linear relation
<equation id=" sat9" noautocaption>(%i) </equation>
and that gives
<equation id=" sat10" noautocaption>(%i) </equation> for example, for Na D line.
Saturation Intensity
A quick derivation for the saturation intensity is to express the light intensity and the Rabi frequency in terms of the number of photons ,
<equation id=" sat11" noautocaption>(%i) </equation> <equation id=" sat12" noautocaption>(%i) </equation>
thus <equation id=" sat13" noautocaption>(%i) </equation> and pluging this into <xr id=" sat9"/> gives the saturation intensity.
For the case of broadband radiation, we define the average intensity per frequency interval as , and when the saturation parameter ,
<equation id=" sat14" noautocaption>(%i) </equation> thus
<equation id=" sat15" noautocaption>(%i) </equation> which is independent of matrix element! For visible light, , where .
Absorption Cross Section
Cross section is the effective area that represents the probability of some scattering or absorption event. In the case of atom-photon interaction, the absorption rate is the collision rate of an atom with the incoming photons, .
For monochromatic radiation,
<equation id=" sat16" noautocaption>(%i) </equation> in the low intensity limit . If we extrapolate it to saturation parameter , then , and
<equation id=" sat17" noautocaption>(%i) </equation>
and from <xr id=" sat10"/>, we have
<equation id=" sat18" noautocaption>(%i) </equation> This is the resonant cross section for weak radiation, and it is usually much larger than the size of the atom, and independent of matrix element. If we plot the cross section as a function of detuning, it is a Lorentzian line. Strong transitions have a larger widths, but the cross section on resonance is always the same.
When the transition is saturated at high intensity, the resonant cross section goes as . The transition bleaches out when .
For broadband radiation,
<equation id=" sat19" noautocaption>(%i) </equation>
at saturation ,
<equation id=" sat20" noautocaption>(%i) </equation> thus <equation id=" sat21" noautocaption>(%i) </equation> which is the same as we have derived in <xr id=" sat15"/>.
Higher-order radiation processes
Beyond the dipole approximation: Recall that the interaction Hamiltonian for an atom in an electromagnetic field is given by
<equation id="Hint" noautocaption>(%i) </equation>
where the last term we have so far considered only for static magnetic fields. Neglecting, as before, the term, which is appreciable only for very intense fields, we now consider more fully the dominant term in the atom-field interaction,
<equation id="hor1" noautocaption>(%i) </equation>
For concreteness, we shall take A(r) to be a plane wave of the form
<equation id="hor2" noautocaption>(%i) </equation>
Expanding the exponential, we have
<equation id="hor3" noautocaption>(%i) </equation>
Thus far in the course, we have considered only the first term, the dipole term. If dipole radiation is forbidden, for instance if and have the same parity, then the second term in the parentheses becomes important. Usually, it is times smaller. In particular, since
<equation id="hor4" noautocaption>(%i) </equation>
the expansion in <xr id = "hor3"/> is effectively an expansion in . We can rewrite the second term as follows:
<equation id="hor5" noautocaption>(%i) </equation>
The first term of <xr id="hor4"/> is , and the matrix element becomes
<equation id="hor6" noautocaption>(%i) </equation>
where is the Bohr magneton. The magnetic field is . Consequently, <xr id="hor5"/> can be written in the more familiar form . (The orbital magnetic moment is : the minus sign arises from our convention that is positive.) We can readily generalize the matrix element to
<equation id="hor7" noautocaption>(%i) </equation>
where we have added the spin dependent term from <xr id="Hint"/>. indicates that the matrix element is for a magnetic dipole transition. The strength of the transition is set by
<equation id="hor8" noautocaption>(%i) </equation>
so it is indeed a factor of weaker than a dipole transition, as we argued above.
The second term in <xr id ="hor4"/> involves . Making use of the commutator relation , we have
<equation id="hor9" noautocaption>(%i) </equation>
So, the contribution of this term to is
<equation id="hor10" noautocaption>(%i) </equation>
where we have taken . This is an electric quadrupole interaction, and we shall denote the matrix element by
<equation id="hor11" noautocaption>(%i) </equation>
The prime indicates that we are considering only one component of a more general expression involving the matrix element of a tensor product. It is straightforward to verify that the electric quadrupole interaction is also of order .
The total matrix element of the second term in the expansion of <xr id="hor3"/> can be written
<equation id="hor12" noautocaption>(%i) </equation>
Note that is real, whereas is imaginary. Consequently,
<equation id="hor13" noautocaption>(%i) </equation>
The magnetic dipole and electric quadrupole terms do not interfere.
Because transition rates depend on , the magnetic dipole and electric quadrupole rates are both smaller than the dipole rate by . For this reason they are generally referred to as forbidden processes. However, the term is used somewhat loosely, for there are transitions which are much more strongly suppressed due to other selection rules, as for instance triplet to singlet transitions in helium.
Transition | Operator | Parity | |
Electric Dipole | - | ||
Magnetic Dipole | + | ||
Electric Quadrupole | + |
Selection rules
A forbidden transition, then, is one that is weaker than an electric dipole-allowed transition by and only appears in some higher-order approximation. Examples of such higher-order effects are the magnetic dipole and electric quadrupole terms described above, multiphoton processes, the relativistic effects which allow singlet to triplet transitions in helium, and hyperfine interactions within the nucleus. To derive selection rules for the transitions we have discussed above, it is useful to express the matrix elements in terms of spherical tensor operators:
where is a spherical tensor operator of rank . The operators transform under rotations like the spherical harmonics , and any operator can be written as a linear combination of these spherical tensors. By the Wigner-Eckart Theorem, we can express the matrix element
in terms of a reduced matrix element and a Clebsch-Gordan coefficient . In order for the latter to be nonzero, the triangle rule requires that , while conservation of angular momentum requires . Since the operators and responsible for and transitions are both vectors, i.e. tensors of rank , these transitions are both governed by the dipole selection rules
Since is a polar vector and is an axial vector, transitions are allowed only between states of opposite parity and transitions are allowed only between states of the same parity. The operator responsible for transitions is a spherical tensor of rank 2. For example,
In general, then, we expect that the quadrupole moment can be expressed in terms of . Thus, electric quadrupole transitions are allowed only between states connected by tensors , requiring:
and parity unchanged.
In addition, transitions are forbidden in all of the cases considered above, since requires (for any interaction that does not couple to spin) whereas absorption or emission of a photon implies .
We now illustrate the use of the spherical tensor for the case of a vector. The dipole matrix element for a particular polarization of the field, , is
<equation id=" select1" noautocaption>(%i) </equation>
It is straightforward to calculate but a more general approach is to write r in terms of a spherical tensor. This yields the selection rules directly, and allows the matrix element to be calculated for various geometries using the Wigner-Eckart theorem as discussed above.
The orbital angular momentum operator of a system with total angular momentum can be written in terms of a spherical harmonic . Consequently, the spherical harmonics constitute spherical tensor operators. A vector can be written in terms of spherical harmonics of rank 1. This permits the vector operator r to be expressed in terms of the spherical tensor
The spherical harmonics of rank 1 are
<equation id=" select2" noautocaption>(%i) </equation>
These are normalized so that
<equation id=" select3" noautocaption>(%i) </equation>
We can write the vector r in terms of components ,
<equation id=" select4" noautocaption>(%i) </equation>
or, more generally
<equation id=" select5" noautocaption>(%i) </equation>
Consequently,
<equation id=" select6" noautocaption>(%i) </equation>
<equation id=" select7" noautocaption>(%i) </equation>
The first factor is independent of . It is
<equation id=" select8" noautocaption>(%i) </equation>
where contains the radial part of the matrix element. It vanishes unless and have opposite parity. The second factor in <xr id="select7"/> yields the selection rule
<equation id=" select9" noautocaption>(%i) </equation>
Similarly, for magnetic dipole transition, <xr id="hor6"/>, we have
<equation id=" select10" noautocaption>(%i) </equation>
It immediately follows that parity is unchanged, and that
<equation id=" select11" noautocaption>(%i) </equation>
This discussion of matrix elements, selection rules, and radiative processes barely skims the subject. For an authoritative treatment, the books by Shore and Manzel, and Sobelman are recommended.
References
JAC63 E.T. Jaynes and F.W. Cummings, Proc. IEEE, 51, 89 (1963).
EIN17 A. Einstein, Z. Phys. 18, 121 (1917), reprinted in English by D. ter Haar, The Old Quantum Theory, Pergammon, Oxford.
EIN17a A. Einstein, Z. Phys. 18, 121 (1917), translated in Sources of Quantum Mechanics, B. L. Van der Waerden, Cover Publication, Inc., New York, 1967. This book is a gold mine for anyone interested in the development of quantum mechanics.