Next Article in Journal
Convergence Rate of a Stable, Monotone and Consistent Scheme for the Monge-Ampère Equation
Next Article in Special Issue
Local Dynamics in an Infinite Harmonic Chain
Previous Article in Journal
Pseudospin Symmetry as a Bridge between Hadrons and Nuclei
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analytical Solutions of Temporal Evolution of Populations in Optically-Pumped Atoms with Circularly Polarized Light

Department of Physics, Chonnam National University, Gwangju 500-757, Korea
Symmetry 2016, 8(3), 17; https://doi.org/10.3390/sym8030017
Submission received: 10 December 2015 / Revised: 22 February 2016 / Accepted: 14 March 2016 / Published: 19 March 2016
(This article belongs to the Special Issue Harmonic Oscillators In Modern Physics)

Abstract

:
We present an analytical calculation of temporal evolution of populations for optically pumped atoms under the influence of weak, circularly polarized light. The differential equations for the populations of magnetic sublevels in the excited state, derived from rate equations, are expressed in the form of inhomogeneous second-order differential equations with constant coefficients. We present a general method of analytically solving these differential equations, and obtain explicit analytical forms of the populations of the ground state at the lowest order in the saturation parameter. The obtained populations can be used to calculate lineshapes in various laser spectroscopies, considering transit time relaxation.
PACS:
02.30.Hq; 32.80.Xx; 32.30.-r

1. Introduction

When an atom is illuminated by single-mode laser light, the populations of the magnetic sublevels and coherences between them exhibit complicated temporal variations. This phenomenon is called optical pumping, which is widely used in the preparation of internal atomic states of interest [1,2]. It has recently been observed that optical pumping affects the lineshapes in saturated absorption spectroscopy (SAS) [3], electromagnetically induced transparency (EIT) [4], and absorption of cold atoms with a Λ-type three-level scheme [5]. Nonlinear effects in optical pumping have also been investigated [6,7].
The temporal dynamics of the internal states of an atom are accurately described by density matrix equations [8,9]. In some special cases, however, a simpler method can be employed to solve for the dynamics of the internal states of the atom, using rate equations [10,11]. Furthermore, when the intensity of light is weak, the rate equations can be solved analytically [12,13,14,15]. These analytical solutions are practically very useful; once they are obtained, it is readily possible to obtain analytically computed quantities such as the absorption coefficient of a probe beam and lineshape functions in nonlinear laser spectroscopy. We have previously reported analytical solutions for SAS [16,17] and polarization spectroscopy (PS) [18].
Interestingly, the equations governing the temporal dynamics of populations at the weak intensity limit are homogeneous or inhomogeneous second-order linear differential equations (DEs) with constant coefficients [12,13,14,15]. Unlike the harmonic oscillator in mechanics, where under- or over-damped motions are observed [19], the equations for optical pumping show only over-damped behaviors. However, this system exhibits a variety of inhomogeneous DEs. In a recent publication, we reported the method of solving these equations analytically, in the context of a pedagogical description of the method of solving inhomogeneous DEs [15]. Although the method is straightforward in principle, it is not easy to obtain analytical solutions for complicated atomic structures, such as Cs. Extending the previous study [15], in this paper, we present a general method of analytically solving the DEs for such a complicated atom.

2. Theory

The energy level diagram under consideration is shown in Figure 1. Since alkali-metal atoms are considered, there are two ground states with F g = I + 1 / 2 and F g = I 1 / 2 (I: nuclear spin angular momentum quantum number). We consider a σ + polarized weak laser beam, whose Rabi frequency is Ω and optical frequency is ω = ω 0 + δ ( ω 0 is the resonance frequency and δ is the laser frequency detuning). We assume that the laser frequency is tuned to the transition from one of the two ground states (in Figure 1, the state F g = I + 1 / 2 ). Then, the other ground state (in Figure 1, the state F g = I 1 / 2 ) is not excited by laser light, and can be populated by spontaneous emission from the excited state when the optical transition is not cycling. The populations (and the states themselves) of the magnetic sublevels in the excited, upper ground, and lower ground states are labeled, respectively, as g i , f i , and h i with i = 1 , 2 , .
The internal dynamics of the atom can be described by the density matrix equation in the frame rotating with frequency ω:
ρ ˙ = i / H , ρ + ρ ˙ sp ,
where ρ is the density operator. In Equation (1), the Hamiltonian, H, is given by
H = j δ g j g j j Δ g h j h j Ω 2 j C j j g j f j + h . c . ,
where Δ g is the hyperfine splitting between the two ground states and h.c. denotes the harmonic conjugate. In Equation (2), the first two terms in the right-hand side represent the bare atomic Hamiltonian and the rest terms denote the atom-photon interaction Hamiltonian [20]. C i j is the normalized transition strength between the states f i and g j , and R i j C i j 2 is given below (Equation (13)). In Equation (1), ρ ˙ sp represents spontaneous emission term, whose matrix representations are given by:
g i ρ ˙ sp g j = Γ g i ρ g j , g i ρ ˙ sp f j = Γ 2 g i ρ f j , g i ρ ˙ sp h j = Γ 2 g i ρ h j , f i ρ ˙ sp f j = Γ ϵ = 2 0 C i i + ϵ C j j + ϵ g i + ϵ ρ g j + ϵ , h i ρ ˙ sp h j = Γ ϵ = 2 0 D i i + ϵ D j j + ϵ g i + ϵ ρ g j + ϵ ,
and μ ρ ˙ sp ν = ν ρ ˙ sp μ * when μ ν , where Γ is the decay rate of the excited state. D i j is the normalized transition strength between the states h i and g j , and T i j D i j 2 is also given below (Equation (13)). Inserting Equations (2) and (3) into Equation (1), we can obtain the following differential equations for the optical coherences and populations:
g i ρ ˙ f i = i δ Γ 2 g i ρ f i + i 2 C i i Ω g i f i ,
g ˙ i = Γ g i + i 2 C i i Ω g i ρ f i f i ρ g i ,
f ˙ i = Γ j = i 2 i C i j 2 g j i 2 C i i Ω g i ρ f i f i ρ g i ,
h ˙ i = Γ j = i 2 i D i j 2 g j ,
where we use simplified expressions for the populations: g i ρ g i = g i , f i ρ f i = f i , and h i ρ h i = h i . In Equations (4)–(7), we assume that g i ρ h i = 0 because Δ g is much larger than δ and Γ. We note that, because the polarization of light is σ + , and therefore the Zeeman coherences between the magnetic sublevels in the excited and ground states disappear.
In Equation (4), the characteristic decay rate of the optical coherence is Γ / 2 , which is much larger than the characteristic decay rate of the populations ( s Γ ; see Equation (12) below for definition of s). Thus, the optical coherences evolve much faster than the populations, which is called the rate equation approximation [21]. Owing to this rate equation approximation, g i ρ f i can be expressed in terms of the populations as follows by letting g i ρ ˙ f i = 0 :
g i ρ f i = C i i Ω i Γ + 2 δ f i g i .
Then, inserting Equation (8) and its complex conjugate into Equations (5)–(7), we can obtain the following rate equations for the populations:
f ˙ i = Γ 2 s R i i f i g i + j = i 2 i Γ R i j g j ,
g ˙ i = Γ 2 s R i i f i g i Γ g i ,
h ˙ i = j = i 2 i Γ T i j g j ,
for i = 1 , 2 , . In Equations (9)–(11), s is the saturation parameter, which is given by
s = Ω 2 / 2 δ 2 + Γ 2 / 4 ,
and R i j = C i j 2 and T i j = D i j 2 . We note that s is a function of both the δ and Rabi frequency. Notably, the reference of the frequency detuning differs, depending on the transition line considered. When i and j refer to the states F g , m g and F e , m e , respectively, the transition strength ( R i j ) is given by
R F g , m g F e , m e = ( 2 L e + 1 ) ( 2 J e + 1 ) ( 2 J g + 1 ) ( 2 F e + 1 ) ( 2 F g + 1 ) × L e J e S J g L g 1 J e F e I F g J g 1 F g 1 F e m g m e m g m e 2 ,
where L and S denote the orbital and electron spin angular momenta, respectively, and the curly (round) brackets represent the 6 J ( 3 J ) symbol. T i j are similarly obtained by using different F g values in Equation (13).
The explicit form of Equation (9) is given by
f ˙ i = Γ 2 s R i i g i f i + Γ R i i 2 g i 2 + R i i 1 g i 1 + R i i g i ,
and f i can be expressed in terms of g ˙ i and g i from Equation (10) at the lowest order in s as follows:
f i = 2 Γ s R i i g ˙ i + Γ g i .
Insertion of Equations (10) and (15) into Equation (14) yields the following DE for g i :
g ¨ i + Γ 1 + s 2 R i i g ˙ i + s 2 Γ 2 R i i 1 R i i g i = s 2 Γ 2 R i i 2 R i i g i 2 + s 2 Γ 2 R i i 1 R i i g i 1 .
when i = 1 , the right-hand side of Equation (16) vanishes. Therefore, Equation (16) becomes a homogeneous DE. In contrast, when i 1 , Equation (16) becomes an inhomogeneous DE because the right-hand side terms are functions of g i .
We solve Equation (16) from i = 1 consecutively. As is well-known, the solution of Equation (16) consists of two parts: a homogeneous solution and a particular solution. We first find the solutions of the homogeneous equation by inserting the equation g i e λ Γ t into Equation (16). Then, we have two values ( λ 2 i 1 , λ 2 i ) for λ as follows:
λ 2 i 1 ( 2 i ) = 1 4 2 s R i i ( + ) 4 4 s R i i + s ( 8 + s ) R i i 2 ,
which can be approximated as follows in the weak intensity limit:
λ 2 i 1 1 s 2 R i i 2 , λ 2 i s 2 R i i 1 R i i .
We consider the case of i = 1 in Equation (16). Then, the solution is given by
g 1 = C 1 , 1 e λ 1 Γ t + C 1 , 2 e λ 2 Γ t ,
where the coefficients C 1 , 1 and C 1 , 2 should be determined using the initial conditions. In the case of i = 2 , the right-hand side in Equation (16) contains the terms of e λ 1 Γ t and e λ 2 Γ t . Therefore, g 2 has four exponential terms:
g 2 = C 2 , 1 e λ 1 Γ t + C 2 , 2 e λ 2 Γ t + C 2 , 3 e λ 3 Γ t + C 2 , 4 e λ 4 Γ t ,
where the coefficients should also be determined. Therefore, we can express g j generally as follows:
g j = k = 1 2 j C j , k e λ k Γ t .
We find C j , k with k = 1 , 2 , , 2 j by means of recursion relations; i.e., C j , k are expressed in terms of C i , l with i < j and l = 1 , 2 , , 2 i . Inserting Equation (17) into Equation (16), we obtain
g i = C i , 2 i 1 e λ 2 i 1 Γ t + C i , 2 i e λ 2 i Γ t + k = 1 2 ( i 1 ) ( s / 2 ) R i i 1 R i i C i 1 , k λ k 2 + λ k + s 2 R i i 1 + λ k R i i e λ k Γ t + k = 1 2 ( i 2 ) ( s / 2 ) R i i 2 R i i C i 2 , k λ k 2 + λ k + s 2 R i i 1 + λ k R i i e λ k Γ t .
Comparing Equations (17) and (18) gives
C i , k = ( s / 2 ) R i i R i i 1 C i 1 , k + R i i 2 C i 2 , k λ k 2 + λ k + s 2 R i i 1 + λ k R i i , for k = 1 , 2 , , 2 ( i 2 ) ,
C i , k = ( s / 2 ) R i i 1 R i i C i 1 , k λ k 2 + λ k + s 2 R i i 1 + λ k R i i , for k = 2 i 3 and 2 ( i 1 ) .
The remaining two coefficients, C i , 2 i 1 and C i , 2 i , can be derived from Equation (18) using two initial conditions for g i ( 0 ) and g ˙ i ( 0 ) :
g i ( 0 ) = 0 , g ˙ i ( 0 ) = s 2 p 0 R i i ,
where p 0 is the population of each sublevel in the ground state at equilibrium, which is equal to 1 / [ 2 ( 2 I + 1 ) ] . Then, the results are given by
C i , 2 i 1 = 1 2 Q i 2 A i + 2 A i + B i + 2 B i + A i + B i 2 p 0 s R i i A i + B i 2 ,
C i , 2 i = 1 2 Q i 2 A i + 2 A i + B i + 2 B i + A i + B i 2 p 0 s R i i A i + B i 2 ,
where
Q i = 4 + s R i i 4 + 8 + s R i i , A i = k = 1 2 ( i 1 ) ( s / 2 ) R i i 1 R i i C i 1 , k λ k 2 + λ k + s 2 R i i 1 + λ k R i i , for i 2 B i = k = 1 2 ( i 2 ) ( s / 2 ) R i i 2 R i i C i 2 , k λ k 2 + λ k + s 2 R i i 1 + λ k R i i , for i 3 , A i = k = 1 2 ( i 1 ) ( s / 2 ) R i i 1 R i i λ k C i 1 , k λ k 2 + λ k + s 2 R i i 1 + λ k R i i , for i 2 B i = k = 1 2 ( i 2 ) ( s / 2 ) R i i 2 R i i λ k C i 2 , k λ k 2 + λ k + s 2 R i i 1 + λ k R i i , for i 3 ,
and
A 1 = 0 , A 1 = 0 , B 1 = B 2 = 0 , and B 1 = B 2 = 0 .
The coefficients in g i from g 1 can be obtained by successively using the recursion relations in Equations (19)–(21). Once g i are obtained, f i can be obtained using Equation (15). Up to the lowest order in s, the result is given by
f i = k = 1 i 2 C i , 2 k s R i i e λ 2 k Γ t .
Since λ k 1 for odd k, g i can be expressed as follows:
g i = k = 1 i C i , 2 k 1 e Γ t + C i , 2 k e λ 2 k Γ t .
Taking the derivative of Equation (24) with respect to time and letting t = 0 , we have
g ˙ i ( 0 ) = k = 1 i C i , 2 k 1 ,
up to the first order in s, since λ 2 k ( k = 1 , 2 , , i ) are already in the first order in s. Because one of the initial conditions is g ˙ i ( 0 ) = s p 0 R i i / 2 , and g i ( 0 ) = k = 1 i C i , 2 k 1 + C i , 2 k = 0 from the other initial condition, we obtain the following equations:
k = 1 i C i , 2 k 1 = k = 1 i C i , 2 k = s 2 p 0 R i i .
Using the relations in Equations (23) and (25), we find the simplified form of g i as follows:
g i = R i i s 2 f i p 0 e Γ t .
We obtain the populations of the sublevels in the ground state, which are not excited by laser light. The one or two magnetic sublevels with higher magnetic quantum numbers correspond to this case. We can easily obtain analytical populations by integrating the populations spontaneously transferred from the excited state, and the result is given by
f i = p 0 + k = 1 i 2 R i i 2 C i 2 , 2 k e λ 2 k Γ t 1 λ 2 k + k = 1 i 1 R i i 1 C i 1 , 2 k e λ 2 k Γ t 1 λ 2 k .
In several cases of atomic transition systems, λ k can duplicate, and the method of solving particular solutions given in Equation (18) no longer holds. We may solve for the particular solutions using the method presented in our previous paper [15]. However, it is also possible to solve by intentionally modifying λ k to satisfy the conditions that all λ k are unique. One possible method is setting R i j R i j + δ i , j j ϵ , where ϵ is a constant that is taken as zero at the final stage of the calculation. Although this method is not novel, it is very efficient.
The populations ( h i ) of the sublevels in the ground state, which are not excited by laser light, can be easily obtained analytically by integrating the populations spontaneously transferred from the excited state (Equation (11)), and the result is given by
h i = p 0 + l = 2 0 k = 1 i + l T i i + l C i + l , 2 k e λ 2 k Γ t 1 λ 2 k .

3. Calculated Results

Based on the method developed in Section 2, here we present the calculated results of the populations for the two transition schemes: (i) F g = 4 F e = 5 and (ii) F g = 3 F e = 3 for the D2 line of Cs. The energy level diagram for the Cs-D2 line is shown in Figure 2a, and the energy level diagrams for these two transitions are shown in Figure 2b,c. Owing to the large hyperfine splitting in the excited states, it is justifiable to neglect the off-resonant transitions; i.e., the F g = 4 F e = 4 and F g = 4 F e = 3 transitions can be neglected when the laser light is tuned to the F g = 4 F e = 5 transition line. Although it is in principle possible to include the off-resonant transitions in the analytical calculation of the populations [13], the complicated analytical solutions may not be practically useful.

3.1. Results for the F g = 4 F e = 5 Transition

The F g = 4 F e = 5 transition shown in Figure 2b is cycling, and is used in many experiments, such as laser cooling and trapping [22]. Because σ + polarized laser light is illuminated, the sublevels with m e = 5 and 4 are not optically excited. The normalized transition strengths, for the transitions presented in Figure 2b, are given by
R 1 1 , R 2 2 , R 3 3 , R 4 4 , R 5 5 , R 6 6 , R 7 7 , R 8 8 , R 9 9 = 1 45 , 1 15 , 2 15 , 2 9 , 1 3 , 7 15 , 28 45 , 4 5 , 1 .
For the transition for i = 1 , we obtain λ 1 1 and λ 2 22 2025 s , and
C 1 , 1 = s 1440 , C 1 , 2 = s 1440 .
Thus, using Equation (23), we obtain
f 1 = 1 16 e 22 s Γ t / 2025 .
The λ 4 for the transition for i = 2 is approximately given by 7 225 s , and the coefficients are given by
C 2 , 1 = s 240 , C 2 , 2 = s 2460 , C 2 , 3 = s 160 , C 2 , 4 = 11 6560 s .
Therefore, we have
f 2 = 1 82 e 22 s Γ t / 2025 + 33 656 e 7 s Γ t / 225 .
The remaining λ 2 k ( k = 2 , , 9 ) values are given by
λ 6 , λ 8 , λ 10 , λ 12 , λ 14 , λ 16 , λ 18 = 13 225 s , 7 81 s , s 9 , 28 225 s , 238 2025 s , 2 25 s , 0 ,
and the remaining populations are explicitly given by
f 3 = 413 31 160 e 22 τ / 2025 + 77 2624 e 7 τ / 225 + 121 6080 e 13 τ / 225 , f 4 = 2317 264 860 e 22 τ / 2025 + 693 20 992 e 7 τ / 225 + 1089 44 080 e 13 τ / 225 1001 252416 e 7 τ / 81 , f 5 = 25 577 3 072 376 e 22 τ / 2025 + 4235 125 952 e 7 τ / 225 + 5203 141 056 e 13 τ / 225 5005 504 832 e 7 τ / 81 143 22 272 e τ / 9 , f 6 = 148 693 17 666 162 e 22 τ / 2025 + 1925 47 232 e 7 τ / 225 + 2057 35 264 e 13 τ / 225 1625 63 104 e 7 τ / 81 715 16 704 e τ / 9 + 13 552 e 28 τ / 225 , f 7 = 921 751 89 260 608 e 22 τ / 2025 + 2519 41 984 e 7 τ / 225 + 891 7424 e 13 τ / 225 49 075 504 832 e 7 τ / 81 5555 7424 e τ / 9 273 736 e 28 τ / 225 + 209 192 e 238 τ / 2025 , f 8 = 39 041 249 2 119 939 440 e 22 τ / 2025 + 1561 10 496 e 7 τ / 225 + 225 071 352 640 e 13 τ / 225 + 219 275 126 208 e 7 τ / 81 + 9955 3712 e τ / 9 + 3367 3680 e 28 τ / 225 77 24 e 238 τ / 2025 459 160 e 2 τ / 25 , f 9 = 9 16 1 205 666 281 8 479 757 760 e 22 τ / 2025 74 771 188 928 e 7 τ / 225 316 701 352 640 e 13 τ / 225 404 009 252 416 e 7 τ / 81 62 953 33 408 e τ / 9 3133 5520 e 28 τ / 225 + 407 192 e 238 τ / 2025 + 459 160 e 2 τ / 25 ,
where we use a simplified notation: τ s Γ t . Since the F g = 4 F e = 5 transition is cycling, the populations in the magnetic sublevels in the F g = 3 ground state remain at their equilibrium value, 1/16. It should be also noted that the sum of the ground state populations is conserved, i.e.,
i = 1 9 f i = 9 16 .
From Equation (26), the populations of the sublevels in the excited state can be expressed in terms of the populations in the ground state as follows:
g i = R i i s 2 f i 1 16 e Γ t .
The constants in f 9 and g 9 can be accurately calculated using Equation (10). In the steady-state regime, all the populations except f 9 and g 9 vanish, and these satisfy the following equations:
Γ 2 s f 9 ( ) g 9 ( ) Γ g 9 ( ) = 0 , f 9 ( ) + g 9 ( ) = 9 16 ,
with R 9 9 = 1 . Then, we have
f 9 ( ) = 9 ( 2 + s ) 32 ( 1 + s ) , g 9 ( ) = 9 s 32 ( 1 + s ) ,
which can be used in a more accurate calculation of the SAS spectrum.

3.2. Results for the F g = 3 F e = 3 Transition

Now we present the calculated results of the populations for the F g = 3 F e = 3 transition of the D2 line of Cs. The energy level diagram for the transition is shown in Figure 2c. The sublevel of the excited state with m e = 3 is not optically excited, and thus the sublevel of the upper-ground state with m g = 4 is not filled by spontaneous emission. We also obtain the solutions for the populations in the other ground state ( F g = 4 ). To prevent the duplication of the transition strengths in this transition, we introduce ϵ so that the transition strengths are given explicitly by
R 1 1 , R 2 2 , R 3 3 , R 4 4 , R 5 5 , R 6 6 = 3 16 + ϵ , 5 16 + 2 ϵ , 3 8 + 3 ϵ , 3 8 + 4 ϵ , 5 16 + 5 ϵ , 3 16 + 6 ϵ .
We take ϵ 0 at the final stage of the calculation. The λ 2 k ( k = 1 , , 6 ) values at ϵ 0 are given by
λ 2 , λ 4 , λ 6 , λ 8 , λ 10 , λ 12 = 39 512 s , 55 512 s , 15 128 s , 15 128 s , 55 512 s , 39 512 s .
We first find various C i k values using the recursion relations in Equations (19)–(22). For the transition for i = 1 , we obtain
C 1 , 1 = 3 512 s , C 1 , 2 = 3 512 s ;
thus, using Equation (23), we obtain
f 1 = 1 16 e 39 s Γ t / 512 .
Using a similar method, we can obtain f 2 and f 3 as follows:
f 2 = 3 64 e 39 τ / 512 + 1 64 e 55 τ / 512 , f 3 = 25 448 e 39 τ / 512 + 1 64 e 55 τ / 512 1 112 e 15 τ / 128 ,
where the simplified notation, τ s Γ t , is used. In the calculation of f 4 , because λ 6 and λ 8 are equal, f 4 may contain the term τ e 15 τ / 128 . However, because the transition between g 3 and f 4 is prohibited, the particular solution for f 4 does not contain the term τ e 15 τ / 128 . In contrast, f 5 , f 6 , and f 7 contain the terms proportional to τ. The results for f 4 , f 5 , and f 6 are explicitly given by
f 4 = 15 224 e 39 τ / 512 + 3 32 e 55 τ / 512 11 112 e 15 τ / 128 , f 5 = 135 896 e 39 τ / 512 + 173 640 + 9 τ 4096 e 55 τ / 512 + 51 280 e 15 τ / 128 , f 6 = 269 12 544 + 1125 τ 114 688 e 39 τ / 512 + 19 256 45 τ 16 384 e 55 τ / 512 13 392 e 15 τ / 128 .
Since f 7 is not excited by laser light, using Equation (27) yields,
f 7 = 68 971 327 184 343 323 2 119 936 + 10 125 τ 1 490 944 e 39 τ / 512 + 1371 30 976 + 135 τ 180 224 e 55 τ / 512 3 98 e 15 τ / 128 .
The populations of the sublevels in the excited state, using Equation (26), can be expressed as follows:
g i = R i i s 2 f i 1 16 e Γ t .
The populations of the sublevels in the ground state F g = 4 can be obtained using Equation (28), and are presented in the appendix.

4. Conclusions

We have presented a general method of solving homogeneous or inhomogeneous second-order DEs corresponding to the optical pumping phenomenon with σ + polarized laser light. Unlike the harmonic oscillator in mechanics or electrical circuits, this system only exhibits over-damped behavior. Although the method of solving inhomogeneous DEs with constant coefficients is straightforward in principle, obtaining accurate analytical solutions for the equations related to optically pumped atoms, in particular, those with complicated atomic structures, such as Cs, is cumbersome. Our method of solving the DEs provides an easy way to obtain analytical solutions at the weak intensity limit. This method is general and applicable to most atoms. As stated in Section 1, the obtained analytical form of the populations can be used in the calculation of spectroscopic lineshapes such as in saturated absorption spectroscopy (SAS) [16,17] and polarization spectroscopy (PS) [18]. Calculations of SAS and PS for Cs atoms are in progress.

Acknowledgments

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT and future Planning (2014R1A2A2A01006654).

Conflicts of Interest

The authors declare no conflict of interest.

Appendix

When the laser frequency is tuned to the F g = 3 F e = 3 transition (Figure 2c), the populations of the sublevels in the ground state F g = 4 are given by
h 1 = 23 312 7 624 e 39 τ / 512 , h 2 = 93 1144 41 2496 e 39 τ / 512 5 2112 e 55 τ / 512 , h 3 = 895 10 296 1109 52 416 e 39 τ / 512 3 704 e 55 τ / 512 + 1 1008 e 15 τ / 128 , h 4 = 235 2574 685 26 208 e 39 τ / 512 19 1760 e 55 τ / 512 + 41 5040 e 15 τ / 128 , h 5 = 10 727 113 256 3475 104 832 e 39 τ / 512 2641 232 320 + 3 τ 45 056 e 55 τ / 512 + 31 2520 e 15 τ / 128 , h 6 = 143 477 1 472 328 843 497 19 079 424 + 125 τ 1 490 944 e 39 τ / 512 + 401 30 976 45 τ 180 224 e 55 τ / 512 13 3528 e 15 τ / 128 , h 7 = 293 731 2 944 656 147 347 2 725 632 + 125 τ 212 992 e 39 τ / 512 + 7889 154 880 63 τ 180 224 e 55 τ / 512 43 1260 e 15 τ / 128 , h 8 = 299 023 2 944 656 24 497 681 408 + 125 τ 53 248 e 39 τ / 512 + 959 116 160 + 21 τ 45 056 e 55 τ / 512 + 13 2520 e 15 τ / 128 .
Finally, we note that the sum of the populations is conserved, i.e.,
1 16 + i = 1 7 f i + i = 1 8 h i = 1 ,
where 1 / 16 is the population at the sublevel m g = 4 in the upper ground state.

References

  1. Happer, W. Optical pumping. Rev. Mod. Phys. 1972, 44, 169–249. [Google Scholar] [CrossRef]
  2. McClelland, J.J. Optical State Preparation of Atoms. In Atomic, Molecular, and Optical Physics: Atoms and Molecules; Dunning, F.B., Hulet, R.G., Eds.; Academic Press: San Diego, CA, USA, 1995; pp. 145–170. [Google Scholar]
  3. Smith, D.A.; Hughes, I.G. The role of hyperfine pumping in multilevel systems exhibiting saturated absorption. Am. J. Phys. 2004, 72, 631–637. [Google Scholar] [CrossRef]
  4. Magnus, F.; Boatwright, A.L.; Flodin, A.; Shiell, R.C. Optical pumping and electromagnetically induced transparency in a lithium vapour. J. Opt. B: Quantum Semiclass. Opt. 2005, 7, 109–118. [Google Scholar] [CrossRef]
  5. Han, H.S.; Jeong, J.E.; Cho, D. Line shape of a transition between two levels in a three-level Λ configuration. Phys. Rev. A 2011, 84. [Google Scholar] [CrossRef]
  6. Sydoryk, I.; Bezuglov, N.N.; Beterov, I.I.; Miculis, K.; Saks, E.; Janovs, A.; Spels, P.; Ekers, A. Broadening and intensity redistribution in the Na(3p) hyperfine excitation spectra due to optical pumping in the weak excitation limit. Phys. Rev. A 2008, 77. [Google Scholar] [CrossRef]
  7. Porfido, N.; Bezuglov, N.N.; Bruvelis, M.; Shayeganrad, G.; Birindelli, S.; Tantussi, F.; Guerri, I.; Viteau, M.; Fioretti, A.; Ciampini, D.; et al. Nonlinear effects in optical pumping of a cold and slow atomic beam. Phys. Rev. A 2015, 92. [Google Scholar] [CrossRef]
  8. McClelland, J.J.; Kelley, M.H. Detailed look at aspects of optical pumping in sodium. Phys. Rev. A 1985, 31, 3704–3710. [Google Scholar] [CrossRef] [PubMed]
  9. Farrell, P.M.; MacGillivary, W.R.; Standage, M.C. Quantum-electrodynamic calculation of hyperfine-state populations in atomic sodium. Phys. Rev. A 1988, 37, 4240–4251. [Google Scholar] [CrossRef] [PubMed]
  10. Balykin, V.I. Cyclic interaction of Na atoms with circularly polarized laser radiation. Opt. Commun. 1980, 33, 31–36. [Google Scholar] [CrossRef]
  11. Liu, S.; Zhang, Y.; Fan, D.; Wu, H.; Yuan, P. Selective optical pumping process in Doppler-broadened atoms. Appl. Opt. 2011, 50, 1620–1624. [Google Scholar] [CrossRef] [PubMed]
  12. Moon, G.; Shin, S.R.; Noh, H.R. Analytic solutions for the populations of an optically-pumped multilevel atom. J. Korean Phys. Soc. 2008, 53, 552–557. [Google Scholar] [CrossRef]
  13. Moon, G.; Heo, M.S.; Shin, S.R.; Noh, H.R.; Jhe, W. Calculation of analytic populations for a multilevel atom at low laser intensity. Phys. Rev. A 2008, 78. [Google Scholar] [CrossRef]
  14. Won, J.Y.; Jeong, T.; Noh, H.R. Analytical solutions of the time-evolution of the populations for D1 transition line of the optically-pumped alkali-metal atoms with I = 3/2. Optik 2013, 124, 451–455. [Google Scholar] [CrossRef]
  15. Noh, H.R. Analytical Study of Optical Pumping for the D1 Line of 85Rb Atoms. J. Korean Phys. Soc. 2104, 64, 1630–1635. [Google Scholar] [CrossRef]
  16. Moon, G.; Noh, H.R. Analytic solutions for the saturated absorption spectra. J. Opt. Soc. Am. B 2008, 25, 701–711. [Google Scholar] [CrossRef]
  17. Moon, G.; Noh, H.R. Analytic Solutions for the Saturated Absorption Spectrum of the 85Rb Atom with a Linearly Polarized Pump Beam. J. Korean Phys. Soc. 2009, 54, 13–22. [Google Scholar] [CrossRef]
  18. Do, H.D.; Heo, M.S.; Moon, G.; Noh, H.R.; Jhe, W. Analytic calculation of the lineshapes in polarization spectroscopy of rubidium. Opt. Commun. 2008, 281, 4042–4047. [Google Scholar] [CrossRef]
  19. Thornton, T.; Marion, J.B. Classical Dynamics of Particles and Systems, 5th ed.; Brooks/Cole: New York, NY, USA, 2004. [Google Scholar]
  20. Cohen-Tannoudji, C.; Dupont-Roc, J.; Grynberg, G. Atom–Photon Interactions Basic Processes and Applications; Wiley: New York, NY, USA, 1992. [Google Scholar]
  21. Meystre, P.; Sargent, M., III. Elements of Quantum Optics; Springer: New York, NY, USA, 2007. [Google Scholar]
  22. Metcalf, H.J.; van der Straten, P. Laser Cooling and Trapping; Springer: New York, NY, USA, 1999. [Google Scholar]
Figure 1. An energy level diagram for an optically pumped atom under the influence of circularly polarized light.
Figure 1. An energy level diagram for an optically pumped atom under the influence of circularly polarized light.
Symmetry 08 00017 g001
Figure 2. (a) Energy level diagram of the Cs-D2 line. (b) Energy level diagrams for the F g = 4 F e = 5 cycling transition line and (c) for the F g = 3 F e = 3 transition line illuminated by σ + polarized laser light.
Figure 2. (a) Energy level diagram of the Cs-D2 line. (b) Energy level diagrams for the F g = 4 F e = 5 cycling transition line and (c) for the F g = 3 F e = 3 transition line illuminated by σ + polarized laser light.
Symmetry 08 00017 g002

Share and Cite

MDPI and ACS Style

Noh, H.-R. Analytical Solutions of Temporal Evolution of Populations in Optically-Pumped Atoms with Circularly Polarized Light. Symmetry 2016, 8, 17. https://doi.org/10.3390/sym8030017

AMA Style

Noh H-R. Analytical Solutions of Temporal Evolution of Populations in Optically-Pumped Atoms with Circularly Polarized Light. Symmetry. 2016; 8(3):17. https://doi.org/10.3390/sym8030017

Chicago/Turabian Style

Noh, Heung-Ryoul. 2016. "Analytical Solutions of Temporal Evolution of Populations in Optically-Pumped Atoms with Circularly Polarized Light" Symmetry 8, no. 3: 17. https://doi.org/10.3390/sym8030017

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop