Skip to main content
Log in

Phase-space treatment of the driven quantum harmonic oscillator

  • Published:
Pramana Aims and scope Submit manuscript

Abstract

A recent phase-space formulation of quantum mechanics in terms of the Glauber coherent states is applied to study the interaction of a one-dimensional harmonic oscillator with an arbitrary time-dependent force. Wave functions of the simultaneous values of position q and momentum p are deduced, which in turn give the standard position and momentum wave functions, together with expressions for the ηth derivatives with respect to q and p, respectively. Afterwards, general formulae for momentum, position and energy expectation values are obtained, and the Ehrenfest theorem is verified. Subsequently, general expressions for the cross-Wigner functions are deduced. Finally, a specific example is considered to numerically and graphically illustrate some results.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8

Similar content being viewed by others

References

  1. D Campos, Pramana – J. Phys. 87, 2, 27 (2016) 10.1007/s12043-016-1230-x

    Article  ADS  Google Scholar 

  2. R J Glauber, Phys. Rev. 131, 6, 2766 (1963) 10.1103/PhysRev.131.2766

    Article  ADS  MathSciNet  Google Scholar 

  3. K Husimi, Prog. Theor. Phys. 9, 4, 381 (1953) 10.1143/ptp/9.4.381

    Article  ADS  MathSciNet  Google Scholar 

  4. E H Kerner, Can. J. Phys. 36, 3, 371 (1958) 10.1139/p58-038

    Article  ADS  Google Scholar 

  5. C E Treanor, J. Chem. Phys. 43, 2, 532 (1965) http://dx.doi.org/10.1063/1.1696777

    Article  ADS  Google Scholar 

  6. P Carruthers and M M Nieto, Am. J. Phys. 33, 7, 537 (1965) http://goo.gl/puaQL7

    Article  ADS  Google Scholar 

  7. D M Gilbey and F O Goodman, Am. J. Phys. 34, 143 (1966) 10.1119/1.1972813

    Article  ADS  Google Scholar 

  8. D Rapp Quantum mechanics (Holt, Reinehart and Winston, New York, 1971) Chap. 24; The forced harmonic oscillator, SBN: 03-081294-1, pp. 445–455

  9. G W Parker, Am. J. Phys. 40, 120 (1972) 10.1119/1.1986457

    Article  ADS  Google Scholar 

  10. S Blanes, F Casas, J A Oteo, and J Ros, Phys. Rep. 470, 151 (2009) 10.1016/j.physrep.2008.11.001

    Article  ADS  MathSciNet  Google Scholar 

  11. K E Cahill and R J Glauber, Phys. Rev. 177, 5, 1857 (1969) 10.1103/PhysRev.177.1857

    Article  ADS  Google Scholar 

  12. S Blanes, F Casas, J A Oteo, and J Ros, Eur. J. Phys. 31, 4, 907 (2010) 10.1088/0143-0807/31/4/020

    Article  Google Scholar 

  13. W H Louisell Quantum statistical properties of radiation, Pure & Applied Optics (John Wiley & Sons, New York, 1973), ISBN-10:0471547859

  14. O I Marichev and M Trott The Wolfram functions site, http://functions.wolfram.com/ (a) http://functions.wolfram.com/05.01.23.0002.01 (b) http://functions.wolfram.com/05.01.16.0006.01 (c) http://functions.wolfram.com/05.01.16.0004.01

  15. E W Weisstein Hermite polynomials, http://goo.gl/Ae3I20

  16. P Ehrenfest, Z. Phys. 45, 7–8, 455 (1927) 10.1007/BF01329203

    Article  ADS  Google Scholar 

  17. R G Littlejohn, Phys. Rep. 138, 4 & 5, 193 (1986) http://dx.doi.org/10.1016/0370-1573(86)90103-1

    Article  ADS  MathSciNet  Google Scholar 

  18. Y I Salamin, J. Phys. A 28, 4, 1129 (1995) http://stacks.iop.org/0305-4470/28/i=4/a=032

    Article  ADS  Google Scholar 

  19. W Magnus, F Oberhettinger, and R P Soni Formulas and theorems for the special functions of mathematical physics, Die Grundlehren der matematischen Wissenschaft (Springer Verlag, 1966), Vol. 52, third enlarged edition, http://goo.gl/u0eUk4

  20. I S Gradshteyn and I M Ryzhik Table of integrals, series and products, 7th edn (Academic Press, 2007), http://www.mathtable.com/gr/

Download references

Acknowledgements

The author is thankful to the anonymous referees for the careful reading of the manuscript and valuable recommendations. The author would like to acknowledge the Universidad Nacional de Colombia for his designation as Emeritus Professor.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to DIÓGENES CAMPOS.

Appendices

Appendix A. Some formulae used in this work

The Hermite polynomials satisfy the following relations:

  1. 1.

    [19, Section 5.6.],

    $$\begin{array}{@{}rcl@{}} &&{} H_{\eta}(y + \sigma) = {\sum}_{k=0}^{\eta} \left( \!\begin{array}{l}\eta\\ k\end{array}\!\right) H_{k}(y) (2 \sigma)^{\eta - k}\\ &&\qquad\quad =\! {\sum}_{k=0}^{\eta} \!\left( \!\begin{array}{c}\eta\\ \eta-k\end{array}\!\right) \!H_{\eta- k}(\sigma)(2 y)^{k}\!. \end{array} $$
    (A.1)
  2. 2.

    [14(b)],

    $$\begin{array}{@{}rcl@{}} {\kern-.9pc}H_{m}(y) H_{k}(y) &=& {\sum}_{r = 0}^{\min(m, k)} 2^{r} r! \left( \!\begin{array}{l}m\\ r\end{array}\!\right) \left( \!\begin{array}{l}k\\ r\end{array}\!\right) \\&& \times H_{m+k-2r}(y). \end{array} $$
    (A.2)
  3. 3.

    [20, 8.958.2],

    $$\begin{array}{@{}rcl@{}} &&{} {\sum}_{m = 0}^{n} \left( \!\begin{array}{l}n\\ m\end{array}\!\right) H_{n - m}(x ) H_{m}(y) = 2^{n/2} H_{n}\!\left( \frac{x + y}{\sqrt{2}}\right)\!.\\ \end{array} $$
    (A.3)
  4. 4.

    The Fourier transformation of the Hermite polynomials is given by

    $$\begin{array}{@{}rcl@{}} &&{} {\int}_{-\infty}^{\infty} \exp\!\left( i x y\right) \exp\!\left( - \frac{1}{2} y^{2}\right) H_{m}(y)\mathrm{d}y \\ &&\quad = (+i)^{m} \sqrt{2 \pi} \exp\!\left( - \frac{1}{2} x^{2}\right) H_{m}(x). \end{array} $$
    (A.4)
  5. 5.

    Then, given a Hermite polynomial H n (y + c), one can write the identity

    $$\begin{array}{@{}rcl@{}} &&{} ({\kern-.6pt}y{\kern-.6pt} +{\kern-.6pt} b{\kern-.6pt})^{m} {\kern-.6pt}H_{n}{\kern-.6pt}({\kern-.6pt}y\! +\! c{\kern-.6pt})\! =\! {\sum}_{k = 0}^{[n/2]} \frac{(-1)^{k} \, n! \, 2^{n - 2k}}{k! \, (n-2k)!} \\ &&\qquad\qquad\qquad\; \times(y + b)^{m} (y + c)^{n - 2k} \\ &&{}\qquad\qquad\qquad\quad =\! {\sum}_{k = 0}^{[n/2]} {\sum}_{\ell = 0}^{m + n - 2k} \frac{(-1)^{k} \, n! \, 2^{n - 2k}}{k! \, (n-2k)!} \\ &&\qquad\qquad\qquad\; \!\times\, A_{\ell}{\kern-.5pt}({\kern-.5pt}m{\kern-.5pt},\! n\! -\! 2{\kern-.5pt} k{\kern-.5pt};\! b{\kern-.5pt},\! c{\kern-.5pt}){\kern-.5pt} y^{\ell}{\kern-.5pt}, \end{array} $$
    (A.5)

where the coefficients A (m,n−2k;b,c) are given by expression (D.2).

Appendix B. On the polynomials π n,k (x)

From the definition of π n,k (x) given by (37), and formula (A.2), it follows that

$$\begin{array}{@{}rcl@{}} {\Pi}_{n, k}(x)\! & := &\! {\sum}_{\lambda = 0}^{n} {\sum}_{r = 0}^{\min(\lambda, k)}\! \left( \!\begin{array}{l}n\\ \lambda\end{array}\!\right)\! (-2)^{r} r! \! \left( \!\begin{array}{l}\lambda\\ r\end{array}\!\right)\!\! \left( \!\begin{array}{l}k\\ r\end{array}\!\right)\! H_{n - \lambda}(x)\\ &&\times H_{\lambda+k - 2r}(x)\\ & = &\! {\sum}_{\lambda = 0}^{n} {\sum}_{r = 0}^{\min(\lambda, k)} \, {\sum}_{u = 0}^{\min(n - \lambda, \lambda + k - 2r)} {\Theta}(n, k; \lambda; r, u)\!\!\\ &&\times H_{n + k - 2r -2u}(x), \end{array} $$
(B.1)

with coefficients

$$\begin{array}{@{}rcl@{}} {\Theta}(n, k; \lambda; r, u) & := & (-1)^{r} 2^{r+u} r! u! \! \left( \!\begin{array}{l}n\\ \lambda\end{array}\!\right)\!\! \left( \!\begin{array}{l}\lambda\\ r\end{array}\!\right)\! \! \left( \!\begin{array}{l}k\\ r\end{array}\!\right)\! \\ &&\times \! \left( \!\begin{array}{c}n-\lambda\\ u\end{array}\!\right)\!\! \left( \!\begin{array}{c} \lambda\! +\! k\! -\! 2r\\u\end{array}\!\right)\!. \end{array} $$
(B.2)

Equation (B.1) can be written in a more elegant way, namely

$$\begin{array}{@{}rcl@{}} &&{}{\Pi}_{n, k}(x) = {\sum}_{N = 0 }^{n+k} \theta_{N}(n, k) H_{N}(x), \qquad\\ &&{}{\Pi}_{n, k}(-x) = (-1)^{n + k}\, {\Pi}_{n, k}(x), \end{array} $$
(B.3)

where the values 𝜃 N (n,k) are determined by equating the coefficients of x N in both sides of the equality

$$\begin{array}{@{}rcl@{}} {\sum}_{N = 0}^{n+k}\! \theta_{N}{\kern-.5pt}({\kern-.5pt}n{\kern-.5pt},{\kern-.5pt} k{\kern-.5pt}) x^{N}\! & = &\! {\sum}_{\lambda = 0}^{n} {\sum}_{r = 0}^{\min(\lambda, k)} \, {\sum}_{u = 0}^{\min(n - \lambda, \lambda + k - 2r)}\\ && \times {\Theta}{\kern-.5pt}({\kern-.5pt}n{\kern-.5pt},{\kern-.5pt} k{\kern-.5pt};{\kern-.5pt} \lambda{\kern-.5pt};{\kern-.5pt} r,{\kern-.5pt} u{\kern-.5pt}) x^{n + k - 2r -2u}. \end{array} $$
(B.4)

After exploring various examples, one finds that 𝜃 N (n,k)=0, if (n + k) is even and N is odd, or if (n + k) is odd and N is even. Using this fact and the reflection formula H N (−x) = (−1)N H N (x) for the Hermite polynomials, one confirms that the polynomial π n,k (x) is even or odd depending on the value of n + k.

At this point, in (B.3) and (B.4) consider the particular situation in which k=0 and, from (38), recall that \({\Pi }_{n, 0}(x) = 2^{n/2} H_{n}(x \sqrt {2})\). Then, a formula linking the Hermite polynomials \(H_{n}(x \sqrt {2})\) with H N (x) [14(c)] allows one to write

$$\begin{array}{@{}rcl@{}} {\Pi}_{n, 0}(x) & = & 2^{n/2} H_{n}(x \sqrt{2}) = {\sum}_{j=0}^{[n/2]} \frac{n!}{j! (n - 2j)! } 2^{n- j}\\ &&\times H_{n - 2j}(x). \end{array} $$
(B.5)

Hence, by comparing this relation with (B.3), one finds the coefficients

$$\begin{array}{@{}rcl@{}} \theta_{N}(n, 0) = \theta_{j}(n) := \frac{n!}{j! (n - 2j)! }\, 2^{n- j}, \end{array} $$
(B.6)

if N = n−2j and j=0,…,[n/2], and 𝜃 N (n,0)=0, otherwise.

Appendix C. Evaluation of the integral \(F_{M N}(\mathcal {X}, a, b, c)\)

For a>0, consider the integral

$$\begin{array}{@{}rcl@{}} &&{\kern-.6pc} F_{M N}(\mathcal{X}, a, b, c)\\ &&\quad := {\int}_{-\infty}^{\infty}{\kern-.6pt} \mathrm{d}y \exp(- i \mathcal{X} y {\kern-.6pt})\exp(- a y^{2})\\ &&\qquad \times H_{M}{\kern-.6pt}(y{\kern-.6pt} +{\kern-.6pt} b{\kern-.6pt}){\kern-.6pt} H_{N}{\kern-.6pt}(y{\kern-.6pt} +{\kern-.6pt} c). \end{array} $$
(C.1)

By using (A.1) and (A.2), and the integral

$$\begin{array}{@{}rcl@{}} &&{\kern-.6pc}{\int}_{-\infty}^{\infty} \exp\!\left( - i x y\right) y^{k} \exp(- a y^{2}) \mathrm{d}y \\ &&\quad= (-i)^{k} \sqrt{\frac{\pi}{a}} \left( \frac{1}{2 \sqrt{a}}\right)^{k} \exp\!\left( - \frac{x^{2}}{4 a}\right)\!\! H_{k}\!\!\left( \frac{x}{2 \sqrt{a}}\right),\\ \end{array} $$
(C.2)

one finds that

$$\begin{array}{@{}rcl@{}} &&{} F_{M N}(\mathcal{X}, a, b, c) = \sqrt{\frac{\pi}{a}}\, \exp\!\left( - \frac{\mathcal{X}^{2}}{4 a}\right) \\ &&{}\qquad\quad \times {\sum}_{k=0}^{M} {\sum}_{\ell =0}^{N}\! \left( \!\begin{array}{c}M\\ M\! -\! k\end{array}\!\right)\! \! \left( \!\begin{array}{c}N\\ N-\ell\end{array}\!\right)\! \left( \! - \frac{i}{ \sqrt{a}}\right)^{k + \ell}\\ &&{}\qquad\quad \times H_{k + \ell}\left( \frac{ \mathcal{X} }{2 \sqrt{a}}\right)\! H_{M - k}(b) H_{N - \ell}(c). \end{array} $$
(C.3)

Note that \(F_{M N}(-\mathcal {X}, a, b, c) {=} F_{M N}^{\star }(\mathcal {X}, a, b, c)\), where denotes the complex conjugate.

Appendix D. Evaluation of the integral \(\mathcal {F}(m, M, n, N | \mathcal {X}, a, b, c, {\Lambda }, \lambda )\)

$$\begin{array}{@{}rcl@{}} &&{\kern-.6pc}\mathcal{F}(\mu, M, \eta, N | \mathcal{X}, a, b, c, {\Lambda}, \lambda)\\ &&\quad := {\int}_{-\infty}^{\infty} \exp\!\left( - i \mathcal{X} y \right) \exp(- a y^{2}) (y + {\Lambda})^{\mu}\\ && {\kern.4pc}\quad\; \times H_{M}(y + b) (y + \lambda)^{\eta} H_{N}(y + c) \mathrm{d}y. \end{array} $$
(D.1)

On the one hand, for μ = η=0, one has \(\mathcal {F}(0, M, 0,\) \( N | \mathcal {X}, a, b, c, {\Lambda }, \lambda ) \,=\, F_{M N}(\mathcal {X}, a, b, c)\). On the other hand, using the binomial theorem followed by a change of indices : = μ + ν, for μ and η integers, one gets the relation

$$\begin{array}{@{}rcl@{}} (y + {\Lambda})^{\mu} (y+\lambda)^{\eta} = {\sum}_{r = 0}^{\mu + \eta} A_{r}(\mu, \eta; {\Lambda}, \lambda) y^{r}, \end{array} $$
(D.2)

where, for Λ ≠ 0 and λ ≠ 0, the coefficients A r (μ,η; Λ,λ) are given by

$$\begin{array}{@{}rcl@{}} A_{r}{\kern-.5pt}({\kern-.5pt}\mu{\kern-.5pt}, \eta; {\Lambda}, \lambda)\! =\! {\sum}_{j = 0}^{r}\!\left( \!\begin{array}{l}\mu\\ j\end{array}\!\right)\! \left( \!\begin{array}{c}\eta\\ r\! -\! j\end{array}\!\right)\! {\Lambda}^{\mu - j} \lambda^{\eta -r + j}. \end{array} $$
(D.3)

In addition, the following special cases should be considered: (i) for Λ=0 and λ ≠ 0,

$$\begin{array}{@{}rcl@{}} A_{r}{\kern-.5pt}({\kern-.4pt}\mu,{\kern-.5pt} \eta;{\kern-.5pt} 0,{\kern-.5pt} \lambda{\kern-.4pt}){\kern-.5pt} =\! \left( \!\begin{array}{c}\eta\\ r-\mu\end{array}\!\right)\! \lambda^{\eta -r+ \mu}, \quad \text{if}\, \mu{\kern-.5pt} \le{\kern-.5pt} r{\kern-.5pt} \le{\kern-.5pt} \mu{\kern-.5pt} +{\kern-.5pt} \eta{\kern-.4pt},\!\!\!\!\!\!\\ \end{array} $$
(D.4)

(ii) for Λ≠0 and λ=0,

$$\begin{array}{@{}rcl@{}} A_{r}{\kern-.5pt}(\mu{\kern-.5pt},{\kern-.5pt} \eta{\kern-.5pt};{\kern-.5pt} {\Lambda}{\kern-.5pt},{\kern-.5pt} 0{\kern-.5pt})\! =\!\left( \!\begin{array}{c}\mu\\ r-\eta\end{array}\!\right){\Lambda}^{\eta -r+ \mu}{\kern-.5pt}, \quad \text{if}\, \eta{\kern-.5pt} \le{\kern-.5pt} r{\kern-.5pt} \le{\kern-.5pt} \mu{\kern-.5pt}+{\kern-.5pt}\eta,\!\!\!\!\!\!\\ \end{array} $$
(D.5)

(iii) for Λ=0 and λ=0,

$$\begin{array}{@{}rcl@{}} A_{r}(\mu, \eta; 0, 0) = 1 \quad \text{if}\, r = \mu+\eta, \end{array} $$
(D.6)

where A r (μ,η;Λ,λ)=0, otherwise.

Then, the identity \((i \partial /\partial \mathcal {X})^{r} \exp (- i \mathcal {X} y)\, {=}\, y^{r}\) \(\times \exp (- i\mathcal {X} y)\) implies that

$$\begin{array}{@{}rcl@{}} &&{\kern-.6pc}\mathcal{F}(\mu, M, \eta, N | \mathcal{X}, a, b, c, {\Lambda}, \lambda)\\ &&{\kern.6pc}= {\sum}_{r = 0}^{\mu + \eta} A_{r}(\mu, \eta; {\Lambda}, \lambda) \bigg(i \frac{\partial}{\partial \mathcal{X}} \bigg)^{r} F_{M N}(\mathcal{X}, a, b, c).\\ \end{array} $$
(D.7)

Thus, because \(F_{M N}(\mathcal {X}, a, b, c)\) is given by expression (C.3), one can define an auxiliary function \(\mathbb {G}_{k + \ell , r}(\chi , a)\) by the relation

$$\begin{array}{@{}rcl@{}} \mathbb{G}_{k + \ell, r}{\kern-.5pt}({\kern-.5pt}\chi{\kern-.5pt},{\kern-.5pt} a{\kern-.5pt})\! & :=& \! \exp\!\left( \! + \frac{\chi^{2}}{4a}\! \right)\!\! \left( \! i \,\frac{\partial}{\partial \mathcal{X}}\! \right)^{r} F(\mathcal{X}) G_{k + \ell}(\mathcal{X}) \\ &=&\!\! \left( \frac{i}{2 \sqrt{a}} \right)^{r} {\sum}_{s=0}^{r} \left( \begin{array}{l}r\\ s\end{array}\right)\theta(k + \ell- s)\\ &&\times\, \frac{2^{s} (k+\ell)!}{(k+ \ell -s)!} (-1)^{r-s}\, H_{r -s}\bigg(\frac{\chi}{2\sqrt{a}} \bigg)\\ && \times\, H_{k+\ell-s}\bigg(\frac{\chi}{2\sqrt{a}} \bigg), \end{array} $$
(D.8)

where \(F(\mathcal {X}) = \exp (- {\chi ^{2}}/{(4a)})\) and \(G_{k + \ell }(\mathcal {X}) = H_{k +\ell }(\chi /(2 \sqrt {a}\,))\). For getting (D.8), one uses Leibnitz’s rule for the rth derivative of a product of functions, Rodrigues’s formula for the Hermite polynomials, and the expression

$$\begin{array}{@{}rcl@{}} \left( \!\frac{\partial}{\partial x}\!\right)^{s}\! H_{k + \ell}(x)\! =\! \theta(k{\kern-.5pt} +{\kern-.5pt} \ell{\kern-.5pt}-{\kern-.5pt} s{\kern-.5pt}) \frac{2^{s} (k{\kern-.5pt}+{\kern-.5pt}\ell)!}{(k{\kern-.5pt}+{\kern-.5pt} \ell{\kern-.5pt} -{\kern-.5pt}s{\kern-.5pt})!} H_{k+\ell-s}(x)\!\!\!\!\!\\ \end{array} $$
(D.9)

for the sth derivative of the Hermite polynomial H k + (x). Here, 𝜃(y) is the unit step function: it is equal to 0 for y<0 and 1 for y≥0.

To summarize, the result of the calculation can be cast into the form

$$\begin{array}{@{}rcl@{}} &&{\kern-.6pc}\mathcal{F}(\mu, M, \eta, N | \mathcal{X}, a, b, c, {\Lambda}, \lambda)= \sqrt{\frac{\pi}{a}}\,\exp\!\left( - \frac{\chi^{2}}{4a}\right)\\ &&\quad \times {\sum}_{k=0}^{M} {\sum}_{\ell =0}^{N}\left( \!\begin{array}{c}M\\ M-k\end{array}\!\right)\left( \!\begin{array}{c}N\\ N-\ell\end{array}\!\right) \left( - \frac{i}{ \sqrt{a}}\right)^{k + \ell}\\ &&\quad \times \bigg[ {\sum}_{r = 0}^{\mu + \eta} A_{r}(\mu, \eta; {\Lambda}, \lambda) \mathbb{G}_{k + \ell, r}(\chi, a) \bigg]\\ &&\quad \times H_{M-k}(b)H_{N-\ell}(c). \end{array} $$
(D.10)

From (C.3) and (D.10) it is immediately verificable that \(\mathcal {F}(0, M, 0, N | \mathcal {X}, a, b, c, {\Lambda }, \lambda ) = F_{M N}(\mathcal {X}, a, b, c)\), because (D.8) implies that \(\mathbb {G}_{k + \ell ,\, 0}(\chi , a) = H_{k + \ell } \times ({ \mathcal {X} }/{(2 \sqrt {a}\,)})\).

Appendix E. A system with Hamiltonian \(\hat {H}(t) = \hat {K} + {V}(t)\)

Instead of the driven harmonic oscillator, consider now a system with a Hamiltonian \(\hat {H}\) that splits into a time-independent unperturbed part \(\hat {K}\) and a perturbation \(\hat {V}(t)\), that is, \(\hat {H} = \hat {K} + \hat {V}(t)\). Instead of eq. (9), one write the Schrödinger evolution operator in the form, with τ = tt 0,

$$\begin{array}{@{}rcl@{}} \hat{U}\! (t{\kern-.5pt},{\kern-.5pt} t_{0}{\kern-.5pt})\! =\! \exp\!\left( \!-\frac{i}{\hbar} \tau \hat{K}\!\right) \hat{U}_{\mathrm{I}}(t, t_{0}), \quad \mathcal{U}_{\mathrm{I}}({\kern-.5pt}\hat{q}{\kern-.5pt}, \hat{p}; t_{0}{\kern-.5pt}, t_{0}{\kern-.5pt})\! =\! \hat{1},\!\!\!\!\!\!\\ \end{array} $$
(E.1)

where \(\hat {U}_{\mathrm {I}}(t, t_{0})\) is the evolution operator in the interaction picture and \(\hat {U}_{0}(t, t_{0}) := \exp (\!- i \tau \hat {K}/ \hbar )\) is the time-evolution operator associated with the Hamiltonian \(\hat {K}\). By inserting the identity \(\hat {1} \,=\, \exp (+\frac {i}{\hbar } \tau \hat {K})\times \exp \!\big (\!\!-\frac {i}{\hbar } \tau \hat {K}\big )\) to the right-hand side of \(\hat {U}_{\mathrm {I}}(t, t_{0})\), the generalization of (14) is given by

$$\begin{array}{@{}rcl@{}} && \hat{U}\!({\kern-.5pt}t{\kern-.5pt}, t_{0}{\kern-.5pt})\! =[ \hat{U}_{0}(t, t_{0})\hat{U}_{\mathrm{I}}(t, t_{0}) \hat{U}_{0}^{+}(t, t_{0})]\\ && \qquad\quad\;\;\;\times \exp\!\left( -\frac{i}{\hbar} \tau \hat{K}\right)\!. \end{array} $$
(E.2)

Formally, the solution of the equation of motion governing the time evolution of \(\hat {U}_{\mathrm {I}}(t, t_{0})\) can be expressed as \(\hat {U}_{\mathrm {I}}(t, t_{0}) := \mathcal {U}(\hat {q}, \hat {p}; t, t_{0})\), where \(\mathcal {U}(\hat {q}, \hat {p}; t, t_{0})\) is a function of the position and momentum operators \(\hat {q}\) and \(\hat {p}\), beside the time dependence due to \(\hat {V}(t)\). Now, according to the Weyl prescription, one can write

$$\begin{array}{@{}rcl@{}} \mathcal{U}(\hat{q}, \hat{p}; t, t_{0}) & = & \frac{1}{(2\pi \hbar)^{f}} \int \mathcal{U}_{\mathrm{W}}(q^{\prime}, p^{\prime}; t, t_{0})\\ &&\times \hat{D}(q^{\prime}, p^{\prime}) \mathrm{d}q^{\prime} \mathrm{d}p^{\prime}, \end{array} $$
(E.3)

where the scalar function \(\mathcal {U}_{\mathrm {W}}(q^{\prime }, p^{\prime }; t, t_{0}) \! =\! \text {Tr}(\mathcal {U}(\hat {q},\) \(\hat {p}; t, t_{0}) \hat {D}(-q^{\prime }, -p^{\prime }))\) is the Weyl symbol associated with \(\mathcal {U}(\hat {q}, \hat {p}; t, t_{0})\) and \( \hat {D}(q, p)\) is the Weyl operator defined in (1).

Consequently, the evolution operator in the Schrödinger picture can be written as

$$\begin{array}{@{}rcl@{}} \hat{U}(t, t_{0}) \!\!&=&\!\! \left[ \frac{1}{(2\pi \hbar)^{f}} \int \mathcal{U}_{\mathrm{W}}(q^{\prime}, p^{\prime}; t, t_{0})\right.\\ && \left. \!\!\!\! \times \hat{D}_{\bullet}(q^{\prime}, p^{\prime}, \tau) \mathrm{d}q^{\prime} \mathrm{d}p^{\prime} \vphantom{\frac{1}{(2\pi \hbar)^{f}}}\!\right]\! \exp\!\left( \!-\frac{i}{\hbar} \tau \hat{K}\!\right)\!, \end{array} $$
(E.4)

where \(\hat {D}_{\bullet }(q^{\prime }, p^{\prime }, \tau )\) is the time-dependent Weyl operator defined by the relation

$$\begin{array}{@{}rcl@{}} \hat{D}_{\bullet}(q^{\prime}, p^{\prime}, \tau) := \exp\!\left( \frac{i}{\hbar} [p^{\prime}{} \hat{Q}(\tau) - q^{\prime}{}\hat{P}(\tau)]\right), \end{array} $$
(E.5)

and

$$\begin{array}{@{}rcl@{}} &&{}\hat{Q}(\tau) := \hat{U}_{0}(t, t_{0})\, \hat{q}\, \hat{U}_{0}^{+}(t, t_{0}) = Q(\hat{q}, \hat{p}; \tau),\\ &&{}\hat{P}(\tau) := \hat{U}_{0}(t, t_{0})\, \hat{p}\, \hat{U}_{0}^{+}(t, t_{0}) = P(\hat{q}, \hat{p}; \tau) \end{array} $$
(E.6)

are time-dependent operators, which satisfy the commutation relation \([\hat {Q}(\tau ) , \hat {P}(\tau )]\, {=}\, i \hbar \). Thus, for a system with Hamiltonian \(\hat {H} {=} \hat {K} {+} \hat {V}(t)\), the state of the system at time t takes the form

$$\begin{array}{@{}rcl@{}} \left| {\Psi (t)} \right\rangle & = & \left[ \frac{1}{(2\pi \hbar)^{f}} \int \mathcal{U}_{\mathrm{W}}(q^{\prime}, p^{\prime}; t, t_{0})\hat{D}_{\bullet}(q^{\prime}, p^{\prime}, \tau)\right. \\ &&\left. \vphantom{\frac{1}{(2\pi \hbar)^{f}}} \times\! \mathrm{d}q^{\prime} \mathrm{d}p^{\prime} \right] \!\exp\!\left( -\frac{i}{\hbar} \tau \hat{K}\right)\!| {\Psi (t_{0})}\rangle\!. \end{array} $$
(E.7)

This generalization of eq. (16) is of course formal as long as explicit expressions of the operators \(\hat {Q}(\tau )\) and \(\hat {P}(\tau )\), the function \(\mathcal {U}(\hat {q}, \hat {p}; t, t_{0})\) and the corresponding Weyl symbol \(\mathcal {U}_{\mathrm {W}}(q^{\prime }, p^{\prime }; t, t_{0})\) are in general not known, and they can be difficult to obtain for a given specific system beyond the system discussed in this paper.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

CAMPOS, D. Phase-space treatment of the driven quantum harmonic oscillator. Pramana - J Phys 88, 54 (2017). https://doi.org/10.1007/s12043-016-1355-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12043-016-1355-y

Keywords

PACS Nos

Navigation