Projected Schrödinger equation second order response properties

Second-order properties

In order to calculate second-order properties, we calculate the second total derivative of the energy with respect to the corresponding perturbation: \[\begin{equation} \begin{split} \frac{ % \dd^2{E(\boldsymbol{\eta})} % }{\dd{\eta_m} \dd{\eta_n}} = % & % \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}^\star, \boldsymbol{\lambda}^\star) % }{\eta_m}{\eta_n} \\ & + \sum_i^m % \eval{ % \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}, \boldsymbol{\lambda}^\star) % }{p_i}{\eta_n} % }_{\vb{p}^\star} % \qty( % \pdv{ % p_i^\star(\boldsymbol{\eta}) % }{\eta_m} % ) \\ % &+ % \sum_a^l % \eval{ % \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}^\star, \boldsymbol{\lambda}) % }{\lambda_a}{\eta_n} % }_{\boldsymbol{\lambda}^\star} % \qty( % \pdv{ % \lambda_a^\star(\boldsymbol{\eta}) % }{\eta_m} % ) \thinspace , \end{split} \label{eq:molecular_Hessians} \end{equation}\] from which we see that we require the first-order response of the wave function parameters \[\begin{equation} \vb{x} = % \pdv{ % \vb{p}^\star(\boldsymbol{\eta}) % }{\boldsymbol{\eta}} \end{equation}\] and of the Lagrangian multipliers: \[\begin{equation} \vb{y} = % \pdv{ % \boldsymbol{\lambda}^\star(\boldsymbol{\eta}) % }{\boldsymbol{\eta}} % \thinspace , \end{equation}\] where we have introduced the short-hand symbols \(\vb{x}\) and \(\vb{y}\). In order to determine the linear wave function response \(\vb{x}\), we take the total derivative of equation \(\eqref{eq:Lagrangian_var_lambda}\), yielding: \[\begin{equation} \vb{k}_{\vb{p}} \thinspace \vb{x} % = - \vb{F}_{\vb{p}} % \thinspace , \end{equation}\] in which we will call \(\vb{k}_{\vb{p}}\) the parameter response constant: \[\begin{equation} \qty( % \vb{k}_{\vb{p}} )_{ai} % = % \eval{ % \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}, \boldsymbol{\lambda}) % }{\lambda_a}{p_i} % }_{\vb{p}^\star, \boldsymbol{\lambda}^\star} % = \eval{ % \pdv{ % f_a(\boldsymbol{\eta}, \vb{p}) % }{p_i} % }_{\vb{p}^\star} \end{equation}\] and \(\vb{F}_{\vb{p}}\) the parameter response force: \[\begin{equation} \qty( % \vb{F}_{\vb{p}} )_{am} % = % \eval{ % \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}^\star, \boldsymbol{\lambda}) % }{\lambda_a}{\eta_m} % }_{\boldsymbol{\lambda}^\star} % = % \pdv{ % f_a(\boldsymbol{\eta}, \vb{p}^\star) % }{\eta_m} % \thinspace , \end{equation}\] in analogy with Hooke’s law. We note that the parameter response constant \(\vb{k}_{\vb{p}}\) is related to the Jacobian of the PSEs. After having solved the linear equation for the first-order wave function response \(\vb{x}\), we are able to solve a linear equation for the first-order Lagrange multiplier response \(\vb{y}\): \[\begin{align} \vb{k}_{\boldsymbol{\lambda}} \thinspace \vb{y} % &= - \vb{F}_{\boldsymbol{\lambda}} \\ % &= - \vb{A}_{\boldsymbol{\lambda}} % - \vb{B}_{\boldsymbol{\lambda}} \vb{x} % \thinspace , \end{align}\] where \(\vb{k}_{\boldsymbol{\lambda}}\) is then accordingly called the multiplier response constant: \[\begin{equation} \qty( % \vb{k}_{\boldsymbol{\lambda}} )_{ia} % = \qty( % \vb{k}_{\vb{p}} )_{ai} \end{equation}\] and \(\vb{F}_{\boldsymbol{\lambda}}\) the multiplier response force: \[\begin{equation} \qty( % \vb{F}_{\vb{p}} )_{am} % = % \eval{ % \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}^\star, \boldsymbol{\lambda}) % }{\lambda_a}{\eta_m} % }_{\boldsymbol{\lambda}^\star} % = % \pdv{ % f_a(\boldsymbol{\eta}, \vb{p}^\star) % }{\eta_m} \end{equation}\] that can be calculated using the intermediaries \[\begin{align} \qty( % \vb{A}_{\boldsymbol{\lambda}} )_{im} % & = % \eval{ % \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}, \boldsymbol{\lambda}^\star) % }{p_i}{\eta_m} % }_{\vb{p}^\star} \\ & = % \eval{ % \pdv{ % E(\boldsymbol{\eta}, \vb{p}) % }{p_i}{\eta_m} % }_{\vb{p}^\star} % + \sum_a^l \lambda_a^\star % \eval{ % \pdv{ % f_a(\boldsymbol{\eta}, \vb{p}) % }{p_i}{\eta_m} % }_{\vb{p}^\star} \end{align}\] and \[\begin{align} \qty( % \vb{B}_{\boldsymbol{\lambda}} )_{ij} % &= % \eval{ % \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}, \boldsymbol{\lambda}) % }{p_i}{p_j} % }_{\vb{p}^\star, \boldsymbol{\lambda}^\star} \\ &= % \eval{ % \pdv{ % E(\boldsymbol{\eta}, \vb{p}) % }{p_i}{p_j} % }_{\vb{p}^\star} % + \sum_a^l \lambda_a^\star % \eval{ % \pdv{ % f_a(\boldsymbol{\eta}, \vb{p}) % }{p_i}{p_j} % }_{\vb{p}^\star} % \thinspace . \end{align}\] After having solved these two sets of linear response equations, we can calculate second-order properties as: \[\begin{equation} \begin{split} \frac{ % \dd^2{E(\boldsymbol{\eta})} % }{\dd{\eta_m} \dd{\eta_n}} % = % & \pdv{ % \mathscr{L}(\boldsymbol{\eta}, \vb{p}^\star, \boldsymbol{\lambda}^\star) % }{\eta_m}{\eta_n} % + \qty( % \vb{x}^{\text{T}} \vb{A}_{\boldsymbol{\lambda}} % )_{mn} \\ & + \qty( % \vb{y}^{\text{T}} \vb{F}_{\vb{p}} % )_{mn} % \thinspace . \end{split} \end{equation}\] For the PSE framework, we list here the following energy derivatives: \[\begin{equation} \eval{ % \pdv{ % E(\boldsymbol{\eta}, \vb{p}) % }{\eta_m}{\eta_n} % }_{\vb{p}^\star} % = % \frac{ % \matrixel**{0}{ % \pdv[2]{ % \hat{\mathcal{H}}(\boldsymbol{\eta}) % }{\eta_m}{\eta_n} % }{\Psi(\vb{p}^\star)} % }{ \braket{0}{\Psi(\vb{p}^\star)} } % \end{equation}\]