Skip to main content
Log in

Time-dependent wave packet propagation using quantum hydrodynamics

  • Regular Article
  • Published:
Theoretical Chemistry Accounts Aims and scope Submit manuscript

An Erratum to this article was published on 03 June 2014

Abstract

A new approach for propagating time-dependent quantum wave packets is presented based on the direct numerical solution of the quantum hydrodynamic equations of motion associated with the de Broglie–Bohm formulation of quantum mechanics. A generalized iterative finite difference method (IFDM) is used to solve the resulting set of non-linear coupled equations. The IFDM is 2nd-order accurate in both space and time and exhibits exponential convergence with respect to the iteration count. The stability and computational efficiency of the IFDM is significantly improved by using a “smart” Eulerian grid which has the same computational advantages as a Lagrangian or Arbitrary Lagrangian Eulerian (ALE) grid. The IFDM is generalized to treat higher-dimensional problems and anharmonic potentials. The method is applied to a one-dimensional Gaussian wave packet scattering from an Eckart barrier, a one-dimensional Morse oscillator, and a two-dimensional (2D) model collinear reaction using an anharmonic potential energy surface. The 2D scattering results represent the first successful application of an accurate direct numerical solution of the quantum hydrodynamic equations to an anharmonic potential energy surface.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Notes

  1. A one-dimensional IFDM Fortran computer code (Online Resource 1) is available for download from the Theoretical Chemistry Accounts supplementary material website: www.springer.com.

  2. Two MPEG movie files (Online Resources 2 and 3) are available for download from the Theoretical Chemistry Accounts supplementary material website: www.springer.com

References

  1. Madelung E (1926) Z Phys 40:322

    Google Scholar 

  2. de Broglie L (1926) CR Acad Sci Paris 183:447

    Google Scholar 

  3. de Broglie L (1927) CR Acad Sci Paris 184:273

    Google Scholar 

  4. Bohm D (1952) Phys Rev 85:166

    Article  CAS  Google Scholar 

  5. Bohm D (1952) Phys Rev 85:180

    Article  CAS  Google Scholar 

  6. Holland PR (1993) The quantum theory of motion. Cambridge University Press, New York

    Book  Google Scholar 

  7. Lopreore C, Wyatt RE (1999) Phys Rev Lett 82:5190

    Article  CAS  Google Scholar 

  8. Hughes KH, Wyatt RE (2002) Chem Phys Lett 366:336

    Article  CAS  Google Scholar 

  9. Trahan CJ, Wyatt RE (2003) J Chem Phys 118:4784

    Article  CAS  Google Scholar 

  10. Kendrick BK (2003) J Chem Phys 119:5805

    Article  CAS  Google Scholar 

  11. Pauler DK, Kendrick BK (2004) J Chem Phys 120:603

    Article  CAS  Google Scholar 

  12. Kendrick BK (2004) J Chem Phys 121:2471

    Article  CAS  Google Scholar 

  13. Derrickson SW, Bittner ER, Kendrick BK (2005) J Chem Phys 123:54107-1

    Article  Google Scholar 

  14. Kendrick BK (2010) J Mol Struct Theochem 943:158

    Article  CAS  Google Scholar 

  15. Kendrick BK (2010) The direct numerical solution of the quantum hydrodynamic equations of motion. In: Chattaraj PK (eds) Quantum trajectories. CRC Press/Taylor & Francis Group, USA, p 325

    Google Scholar 

  16. Kendrick BK (2011) An iterative finite difference method for solving the quantum hydrodynamic equations of motion. In: Hughes KH, Parlant G (eds) Quantum trajectories. CCP6: Dynamics of Open Quantum Systems, Warrington, p 13

  17. Wyatt RE, Bittner ER (2000) J Chem Phys 113:8898

    Article  CAS  Google Scholar 

  18. Rassolov VA, Garashchuk S (2004) J Chem Phys 120:6815

    Article  CAS  Google Scholar 

  19. Garashchuk S (2009) J Phys Chem A 113:4451

    Article  CAS  Google Scholar 

  20. Poirier B (2004) J Chem Phys 121:4501

    Article  CAS  Google Scholar 

  21. Babyuk D, Wyatt RE (2004) J Chem Phys 121:9230

    Article  CAS  Google Scholar 

  22. Burghardt I, Cederbaum LS (2001) J Chem Phys 115:10303

    Article  CAS  Google Scholar 

  23. Burghardt I, Moller KB, Hughes K (2007) In: Micha DA (eds) Springer series in chemical physics, p 391

  24. Goldfarb Y, Degani I, Tannor DJ (2006) J Chem Phys 125:231103

    Article  Google Scholar 

  25. Rowland BA, Wyatt RE (2008) Chem Phys Lett 461:155

    Article  CAS  Google Scholar 

  26. Chou CC, Sanz AS, Miret-Artés S, Wyatt RE (2009) Phys Rev Lett 102:250401-1

    Google Scholar 

  27. Garashchuk S (2010) J Chem Phys 132:014112

    Article  Google Scholar 

  28. Garashchuk S (2010) Chem Phys Lett 491:96

    Article  CAS  Google Scholar 

  29. Wyatt RE (2005) Quantum dynamics with trajectories: introduction to quantum hydrodynamics. Springer, New York

    Google Scholar 

  30. Garashchuk S, Rassolov V, Prezhdo O (2011) Review in computational chemistry, vol 27. Wiley, London, pp 111–210

    Google Scholar 

  31. Scannapeico E, Harlow FH (1995) Los Alamos national laboratory report LA-12984

  32. Press WH, Flannery BP, Teukolsky SA, Vetterling WT (1986) Numerical recipes; the art of scientific computing. Cambridge University Press, New York

    Google Scholar 

  33. Patankar SV (1980) Numerical heat transfer and fluid flow. Hemisphere Publishing Co., New York

    Google Scholar 

  34. Morse PM (1929) Phys Rev 34:57

    Article  CAS  Google Scholar 

  35. Kais S, Levine RD (1990) Phys Rev A 41:2301

    Article  CAS  Google Scholar 

  36. VonNeumann J, Richtmyer RD (1950) J Appl Phys 21:232

    Article  Google Scholar 

  37. Harlow FH (1960) Los Alamos scientific laboratory report LA-2412

  38. Harlow FH, Welch JE (1965) Phys Fluids 8:2182

    Article  Google Scholar 

  39. Leonard BP (1979) Comput Meth Appl Mech Eng 19:59

    Article  Google Scholar 

  40. HSL (2011) A collection of Fortran codes for large scale scientific computation. http://www.hsl.rl.ac.u

  41. Tannor DJ, Weeks DE (1993) J Chem Phys 98:3884

    CAS  Google Scholar 

  42. Wolfram Research, Inc (2010) Mathematica, version 8.0. Champaign, IL

Download references

Acknowledgments

This work was done under the auspices of the US Department of Energy at Los Alamos National Laboratory. Los Alamos National Laboratory is operated by Los Alamos National Security, LLC, for the National Nuclear Security Administration of the US Department of Energy under contract DE-AC52-06NA25396.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Brian K. Kendrick.

Additional information

Published as part of the special collection of articles celebrating the 50th anniversary of Theoretical Chemistry Accounts/Theoretica Chimica Acta.

An erratum to this article is available at http://dx.doi.org/10.1007/s00214-014-1500-y.

Electronic supplementary material

Below is the link to the electronic supplementary material.

(tar 110 kb)

(mpg 9.31 mb)

(mpg 8.07 mb)

Appendices

Appendix 1: Analytic derivatives of F 1 and \(\tilde F_{2}\)

The functions F 1(C,v) and \(\tilde F_2(C,v)\) are defined by Eqs. 9 and 26, respectively. The C and v are short-hand notation for C = C n+1 i and v = v n+1 i . Thus, the analytic derivatives of F 1(C, v) are obtained by taking the derivative of Eq. 9 with respect to C n+1 i and v n+1 i

$$ {\partial F_1\over\partial C} = 1, $$
(52)
$$ {\partial F_1\over\partial v} = {\Updelta t\over \Updelta x}\left [1/16 + (C^{n+1}_{i+1} - C^{n+1}_{i-1})/8\right]. $$
(53)

Similarly, the analytic derivatives of \(\tilde F_2(C,v)\) are obtained by taking the derivative of Eq. 26 with respect to C n+1 i and v n+1 i

$$ {\partial{\tilde F}_2\over\partial C} = -{\Updelta t\over 2 m} {\partial f^{n+1}_{q i'}\over\partial C}, $$
(54)
$$ {\partial{\tilde F}_2\over\partial v} = 1 + {\Updelta t\over \Updelta x}\left[(v^{n+1}_{i'+1} - v^{n+1}_{i'-1})/4 + \Updelta t^2 \tilde\gamma^{n+1}_{i'}/(m \Updelta x)\right]. $$
(55)

The derivative of f n+1 q i with respect to C n+1 i is given by

$$ {\partial f^{n+1}_{q i'}\over\partial C} ={\hbar^2\over 2m}\left[ 1/(2\Updelta x^3) - \langle \partial^2_x C \rangle_{i'}/(2\Updelta x) - \langle \partial_x C \rangle_{i'}/\Updelta x^2\right]. $$
(56)

Substituting the 2nd-order finite difference expressions for the derivatives of C into Eq. 56 gives

$$ {\partial f^{n+1}_{q i'}\over\partial C} ={\hbar^2\over 2m}\left [ 1 + C^{n+1}_{i} - C^{n+1}_{i+2}\right]/(2\Updelta x^3). $$
(57)

In the above derivations, the derivatives of the artificial viscosity coefficient \(\tilde\gamma\) with respect to C and v are ignored.

Appendix 2: Derivation of γ(Cv)

2.1 One-dimensional case

The second-order truncation error with respect to \(\Updelta t\) in the finite-difference representation of the velocity equation (Eq. 6) is given by \(\varepsilon\) defined in Eq. 24. If the classical potential does not depend explicitly on time, then the time derivatives of the total force become equal to the time derivatives of just the quantum force, and Eq. 24 becomes

$$ \varepsilon =\frac{1}{12}\Updelta t^2 \left [ \partial^2_t f^{n'}_{q i'} - m \partial^2_t(v^{n'}_{i'}\partial_x v^{n'}_{i'})\right]. $$
(58)

By expanding the time derivatives of the two terms in brackets in Eq. 58 and using the definition of the quantum force (Eq. 8), Eq. 58 becomes a complicated expression involving many terms of mixed time and spacial derivatives of C and v. By taking the appropriate time and spacial derivatives of the equations of motion for both C (Eq. 4) and v (Eq. 3), all of the various time derivatives of C and v can be eventually expressed in terms of derivatives of x only. The coefficient of the second derivative of v can then be identified as the effective numerical diffusion coefficient (i.e., \(\Upgamma(C,v) \partial^2_x v\)). The algebra is straightforward but tedious, and for one dimension, it is feasible to pursue. However, it is much faster and straightforward to implement the analysis on a symbolic algebra solver such as Mathematica [42]. The resulting expression for \(\Upgamma(C,v) = \Updelta t^2 \gamma\) is given by

$$ \begin{aligned} \Upgamma(C,v) =& \frac{1}{12}\Updelta t^2 [ 3 f v - 9 m v^2 \partial_x v +18 {\hbar^2\over 2m} (\partial_x C)^2 \partial_x v \\ +& 30 {\hbar^2\over 2m} v \partial_x C \partial^2_x C +36 {\hbar^2\over 2m} \partial_x v \partial^2_x C \\ +& 21 {\hbar^2\over 2m} v \partial^3_x C + 8 {\hbar^2\over 2m} \partial^3_x v ], \end{aligned} $$
(59)

where for clarity, the superscripts (n′) and subscripts (i′) of the C and v have been dropped. The force f in the first term in brackets in Eq. 59 is the total force (f = f q  + f c ). Each term in brackets will be identified as γ i where i = 1–7 so that γ = ∑ 7 i=1 γ i /12. Equation 59 shows that γ contains both classical and quantum contributions. The classical contributions correspond to 3 f c v + γ2. The remaining terms are all proportional to \(\hbar^2\) and are therefore quantum in nature. For a Gaussian wave packet, C is quadratic and v is linear with respect to x − x 0 where x 0 is the center of the packet. Thus, the first four γ1–4 are quadratic (i.e., they are proportional to (x − x 0)2). The fifth term γ5 is a constant term proportional to the product of the width of the wave packet and slope of the velocity field. At the edges of the grid, this term is much smaller than the quadratic terms. The last two terms γ6 and γ7 involve third derivatives of C and v and are zero for a Gaussian wave packet. For the above reasons, the original IFDM included only the γ1–3 [14]. Note that for a Gaussian wave packet, the fourth term γ4 has the same functional form as the quantum force term in γ1. Due to the accumulation of numerical noise, the γ6 and especially γ7 can become significant at the edges of the grid even for a Gaussian wave packet. As shown in Sect. 3.2, the long-term stability of the method is improved by including these terms (especially γ7). Furthermore, for anharmonic potentials (such as the Morse potential), the third derivative of C in γ6 can be non-zero even without numerical noise.

In practice, \(\tilde\gamma\) in Eq. 26 is computed from evaluating the following expression at each grid point

$$ \tilde\gamma = \frac{1}{12} \{\gamma_0 {\rm MAX}[\vert\gamma_1\vert, \vert\gamma_2\vert,\vert\gamma_3\vert,\vert\gamma_4\vert,\vert\gamma_5\vert] + \gamma'_0 {\rm MAX}[\vert\gamma_6\vert,\vert\gamma_7\vert ]\}, $$
(60)

where the MAX function returns the maximum value of its arguments. The first five γ i are grouped separately from the γ6 and γ7 in order to allow for separate scaling via the two independent constants γ0 and γ′0. The artificial viscosity constants γ0 and γ′0 are determined from convergence and stability tests. The goal is to choose them as small as possible but still large enough to maintain the stability of the calculation.

2.2 Two-dimensional case

The second-order truncation error with respect to \(\Updelta t\) in the finite-difference representation of the velocity equation (Eq. 30) is given by \(\varepsilon_{\rm 2D}\) defined in Eq. 43. If the classical potential does not depend explicitly on time, then the time derivatives of the total force become equal to the time derivatives of just the quantum force, and Eq. 43 becomes

$$ \varepsilon_{\rm 2D} =\frac{1}{12}\Updelta t^2 \left [ \partial^2_t f^{n'}_{q i',j} - m \partial^2_t(v^{n'}_{i',j}\partial_x v^{n'}_{i',j}) - m \partial^2_t(u^{n'}_{i',j}\partial_y v^{n'}_{i',j})\right]. $$
(61)

By expanding the time derivatives of the two terms in brackets in Eq. 61 and using the definition of the quantum force (Eq. 32), Eq. 61 becomes a complicated expression involving many terms of mixed time and spacial derivatives of C,  v, and u. By taking the appropriate time and spacial derivatives of the equations of motion for both C (Eq. 29), v (Eq. 27), and u (Eq. 28), all of the various time derivatives of C,  v, and u can be eventually expressed in terms of derivatives of x and y only. The coefficient of the second derivatives of v with respect to x and y can then be identified as the effective numerical diffusion coefficients: \(^1\Upgamma(C,v,u)\) and \(^2\Upgamma(C,v,u)\) (and similarly for u). The algebra is straightforward but tedious, and for two dimensions, a symbolic algebra solver such as Mathematica [42] is much preferred. The resulting expression for \(^1\Upgamma(C,v,u) = {^1}\gamma \Updelta t^2\) is given by

$$ ^1\Upgamma(C,v,u) = \frac{1}{12}\Updelta t^2 \sum_i {^1\gamma_i }$$
(62)

where 1γ = ∑ i 1γ i /12 and the individual terms are

$$ \begin{aligned} ^1\gamma_1 =& +3 f v , {^1}\gamma_2 = -9 m v^2 \partial_x v , {^1}\gamma_3 = +18 {h} (\partial_x C)^2 \partial_x v ,\\ {^1}\gamma_4 =& +30 {h} v \partial_x C \partial^2_x C , {^1}\gamma_5 = +36 {h} \partial_x v \partial^2_x C , {^1}\gamma_6 = +21 {h} v \partial^3_x C , \\ {^1}\gamma_7 =& +8 {h} \partial^3_x v , {^1}\gamma_8 = -6 m v u \partial_y v , {^1}\gamma_9 = +4 {h} \partial_x C \partial_y C\partial_y v ,\\ {^1}\gamma_{10} =&+8 {h} \partial_x C \partial_y C\partial_x u , {^1}\gamma_{11} = +6 {h} v \partial_y C \partial^2_{xy} C , {^1}\gamma_{12} = +8 {h} \partial^2_{xy} C \partial_y v ,\\ {^1}\gamma_{13} =&+12 {h} u \partial_x C \partial^2_{xy} C , {^1}\gamma_{14} =+16 {h} \partial^2_{xy} C \partial_x u , {^1}\gamma_{15} = +4 {h} \partial^2_{xy} v \partial_y C ,\\ {^1}\gamma_{16} =&+4 {h} \partial^2_{xy} u \partial_x C , {^1}\gamma_{17} = +8 {h} \partial^2_x u \partial_y C , {^1}\gamma_{18} =+12 {h} u \partial^3_{x^2 y} C ,\\ {^1}\gamma_{19} =&+4 {h} \partial^3_{x^2 y} u , {^1}\gamma_{20} =+2 {h} \partial^2_{y} v \partial_x C , {^1}\gamma_{21} =+3 {h} v \partial^3_{x y^2} C ,\\ {^1}\gamma_{22} =&+2 {h} \partial^3_{x y^2} v, \end{aligned} $$
(63)

where for clarity, the superscripts and subscripts on the C,  v, and u have been dropped and \(h={\hbar^2/2m}\). The first seven terms are identical to those derived for the one-dimensional case (see Eq. 59). The remaining fifteen new terms involve mixed derivatives, or x derivatives of u and y derivatives of v. For the 2D application considered in this work, most of these new terms for i > 7 are small relative to those for i ≤ 7.

The resulting expression for \(^2\Upgamma(C,v,u) = {^2}\gamma \Updelta t^2\) is given by

$$ ^2\Upgamma(C,v,u) = \frac{1}{12}\Updelta t^2 \sum_i {^2\gamma_i} , $$
(64)

where 2γ = ∑ i 2γ i /12 and the individual terms are

$$ \begin{aligned} {^2}\gamma_1 =& 3 g u , {^2}\gamma_2 = -3 m u^2 \partial_x v , {^2}\gamma_3 = -6 m u^2 \partial_y u ,\\ {^2}\gamma_4 = &+6 {h} u \partial_y C \partial^2_{x} C , {^2}\gamma_5 = +5 {h} \partial_x v \partial^2_{x} C , {^2}\gamma_6 = +3 {h} v \partial^3_{x} C ,\\ {^2}\gamma_7 = &+{h} \partial^3_{x} v , {^2}\gamma_8 = -6 m u v \partial_x u , {^2}\gamma_{9} = +2 {h} \partial_x C \partial_y C \partial_y v ,\\ {^2}\gamma_{10} = &+4 {h} \partial_x C \partial_y C \partial_x u , {^2}\gamma_{11} = +4 {h} \partial_y u \partial^2_{x} C , {^2}\gamma_{12} = +3 {h} \partial^2_{xy} C \partial_y v ,\\ {^2}\gamma_{13} =& +6 {h} u \partial_x C \partial^2_{xy} C , {^2}\gamma_{14} = +10 {h} \partial^2_{xy} C \partial_x u , {^2}\gamma_{15} = +2 {h} \partial^2_{x} v \partial_x C ,\\ {^2}\gamma_{16} = &+4 {h} \partial^2_{xy} u \partial_x C , {^2}\gamma_{17} = +4 {h} \partial^2_{x} u \partial_y C , {^2}\gamma_{18} = +9 {h} u \partial^3_{x^2 y} C ,\\ {^2}\gamma_{19} = &+3 {h} \partial^3_{x^2 y} u. \end{aligned} $$
(65)

The various terms in Eqs. 63 and 65 have been ordered so that they match each other in form except for the terms with i = 3,  11, and 15. For the 2D application considered in this work, most of the terms in Eq. 65 for i > 7 are small relative to those for i ≤ 7.

In practice \(^1\tilde\gamma\) in Eq. 45 is computed from evaluating the following expression at each grid point

$$ ^1\tilde\gamma = \frac{1}{12} \left\{ ^1\gamma_0 {\rm MAX}\left[\sum_{i\ne 6,7}\vert ^1\gamma_i\vert\right] + {^1}\gamma'_0 {\rm MAX}\left[\vert ^1\gamma_6\vert,\vert ^1\gamma_7\vert\right]\right\}, $$
(66)

and similarly for \(^2\tilde\gamma\)

$$ ^2\tilde\gamma = \frac{1}{12} \left\{ ^2\gamma_0 {\rm MAX}\left [\sum_{i\ne 6,7}\vert ^2\gamma_i\vert\right] + {^2}\gamma'_0 {\rm MAX}\left[\vert ^2\gamma_{6}\vert,\vert ^2\gamma_{7}\vert\right]\right\}, $$
(67)

where the MAX function returns the maximum value of its arguments. The artificial viscosity constants 1γ0,  1γ′0,  2γ0, and 2γ′0 are determined from convergence and stability tests. The goal is to choose them as small as possible but still large enough to maintain the stability of the calculation. For the 2D application considered in this work, stability was obtained using only the first seven terms in Eqs. 63 and 65 (i.e., i ≤ 7) for the diffusion coefficients \(^1\Upgamma\) and \(^2\Upgamma\) multiplying \(\partial^2_x v\) and \(\partial^2_y v,\) respectively. Similar expressions are used for the diffusion coefficients multiplying ∂ 2 x u and ∂ 2 y u. These expressions can be obtained from those in Eqs. 63 and 65 by interchanging \(x\leftrightarrow y,\,v\leftrightarrow u,\) and \(f\leftrightarrow g\). In evaluating the various 1γ i and 2γ i , it is important to use the appropriate four-point averages so that they are properly evaluated on the staggered v or u grids (see Fig. 1).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Kendrick, B.K. Time-dependent wave packet propagation using quantum hydrodynamics. Theor Chem Acc 131, 1075 (2012). https://doi.org/10.1007/s00214-011-1075-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00214-011-1075-9

Keywords

Navigation