Sign up with your email address to be the first to know about new products, VIP offers, blog features & more.

Probing Multi-Field Inflation with Supersymmetric Quantum Mechanics

Multi-scalar field cosmology is essential for solving the Wheeler-DeWitt equation in the context of quantum gravity. Here, I will test MFI with supersymmetric quantum mechanics based on Witten’s axiomatic approach. One can axiomatize multi-scalar field theory by the following conditions: 1) a Lagrangian containing up to second order derivatives of the fields, and 2) field equations that contain up to second order derivatives of the fields obeying:

with:

    \[{{{\bar X}_{ij}} = \frac{1}{2}{\partial _a}{\pi _i}{\partial ^a}{\pi _j}}\]

and \frac{{\partial {A^{{i_1}...{i_m}}}}}{{\partial {X_{kl}}}} symmetric in all of its indices {i_1}...{i_m},k,l

With the multi-field action in D dimensions having the form:

    \[S = \int {{d^D}} x\hat L\left( {{\pi _i},{\partial _a}{\pi _j},{\partial _b}{\partial _c}{\pi _k}} \right)\]

whose Euler-Lagrange equations are given by:

    \[\frac{{\partial \hat L}}{{\partial {\pi _i}}} - {\partial _a}\left( {\frac{{\partial \hat L}}{{\partial {\pi _{ia}}}}} \right) + {\partial _a}{\partial _b}\left( {\frac{{\partial \hat L}}{{\partial {\pi _{iab}}}}} \right) = 0\]

with a fourth derivatives constraint:

    \[\frac{{\partial \hat L}}{{\partial {\pi _{icd}}\partial {\pi _{iab}}}}{\pi _{i,abcd}}\]

Thus, the universal multi-field action is:

for the multi-fields \left( {\sigma ,\phi } \right), hence the corresponding field equations:

    \[\begin{array}{l}{G_{\alpha \beta }} + {g_{\alpha \beta }}\Lambda = + \frac{1}{2}\left( {{\nabla _\alpha }\phi {\nabla _\beta }\phi - \frac{1}{2}{g_{\alpha \beta }}{g^{\mu \nu }}{\nabla _\mu }\phi {\nabla _\nu }\phi } \right)\\ + \frac{1}{2}\left( {{\nabla _\alpha }\sigma {\nabla _\beta }\sigma - \frac{1}{2}{g_{\alpha \beta }}{g^{\mu \nu }}{\nabla _\mu }\sigma {\nabla _\nu }\sigma } \right)\\ - \frac{1}{2}{g_{\alpha \beta }}V\left( {\phi ,\sigma } \right) - 8\pi G{{\rm T}_{\alpha \beta }}\end{array}\]

with:

    \[\begin{array}{c}{g^{\mu \nu }}{\phi _{,\mu \nu }} - {g^{\alpha \beta }}\Gamma _{\alpha \beta }^\nu {\nabla _\nu }\phi - \frac{{\partial V}}{{\partial \phi }} = 0\\ \Leftrightarrow \\{{\hat \bigcirc }_L}\phi - \frac{{\partial V}}{{\partial \phi }} = 0\end{array}\]

    \[\begin{array}{c}{g^{\mu \nu }}{\sigma _{,\mu \nu }} - {g^{\alpha \beta }}\Gamma _{\alpha \beta }^\nu {\nabla _\nu }\sigma - \frac{{\partial V}}{{\partial \sigma }} = 0\\ \Leftrightarrow \\{{\hat \bigcirc }_L}\sigma - \frac{{\partial V}}{{\partial \sigma }} = 0\end{array}\]

with:

    \[{\rm T}_{;\mu }^{\mu \nu } = 0\,\;;\,\;{\rm T} = {\rm{P}}{g_{\mu \nu }} + \left( {{\rm{P}} + \rho } \right){\tilde w_\mu }{\tilde w_\nu }\]

\rho the energy density, {\rm{P}} the pressure, and {\tilde w_\mu } the velocity, satisfying {\tilde w_\mu }{\tilde w^\mu } = - 1.

The multi-scalar field cosmological paradigm requires the two canonical fields \left( {\sigma ,\phi } \right), the action of a universe based on such fields, the cosmological term contribution, and matter as a perfect fluid content, and is given by:

Our metric has the form:

    \[{\rm{d}}{{\rm{s}}^2} = - {\rm{N}}\left( t \right){\rm{d}}{{\rm{t}}^2} + {e^{2\Omega \left( {\rm{t}} \right)}}{\left( {{e^{2\beta \left( {\rm{t}} \right)}}} \right)_{{\rm{ij}}}}{\omega ^{\rm{i}}}{\omega ^{\rm{j}}}\]

with {\beta _{{\rm{ij}}}}\left( {\rm{t}} \right) a 3 x 3 diagonal matrix,

    \[{\beta _{{\rm{ij}}}} = {\rm{diag}}\left( {{\beta _ + } + \sqrt 3 {\beta _ - },{\beta _ + } - \sqrt 3 {\beta _ - }, - 2{\beta _ + }} \right)\]

and \Omega \left( t \right) is a scalar and {\omega ^{\rm{i}}} are one-forms that characterize each cosmological Bianchi type model, and obey the form:

    \[{\rm{d}}{\omega ^{\rm{i}}} = \frac{1}{2}C_{{\rm{jk}}}^{\rm{i}}{\omega ^{\rm{j}}} \wedge {\omega ^{\rm{k}}}\]

and C_{{\rm{jk}}}^{\rm{i}} are structure constants of the corresponding model. Hence, in Misner’s parametrization, we get:

    \[\begin{array}{l}{\rm{ds}}_{\rm{I}}^2 = - {{\rm{N}}^2}{\rm{d}}{{\rm{t}}^2} + {e^{2\Omega + 2{\beta _ + } + 2\sqrt 3 \beta }} - {\rm{d}}{{\rm{x}}^2}\\ + \,{e^{2\Omega + 2{\beta _ + } - 2\sqrt 3 \beta }} - {\rm{d}}{{\rm{y}}^2} + {e^{2\Omega - 4{\beta _ + }}}{\rm{d}}{{\rm{z}}^2}\end{array}\]

with the anisotropic conditions:

    \[\left\{ {\begin{array}{*{20}{c}}{{{\rm{R}}_{\rm{1}}} = {e^{\Omega + {\beta _ + } + \sqrt 3 {\beta _ - }}}}\\{{{\rm{R}}_{\rm{2}}} = {e^{\Omega + {\beta _ + } - \sqrt 3 {\beta _ - }}}}\\{{{\rm{R}}_{\rm{3}}} = {e^{\Omega - 2{\beta _ + }}}}\end{array}} \right.\]

So the lagrangian density above can be written as:

    \[\begin{array}{l}{{\hat L}_{\rm{I}}} = {e^{3\Omega }}\left[ {6\frac{{{{\dot \Omega }^2}}}{{\rm{N}}} - 6\frac{{\dot \beta _ - ^2}}{{\rm{N}}} - 6\frac{{\dot \beta _ - ^2}}{{\rm{N}}} - 6\frac{{{{\dot \varphi }^2}}}{{\rm{N}}} - 6\frac{{{{\dot \varsigma }^2}}}{{\rm{N}}}} \right.\\ + {\rm{N}}\left. {\left( {V\left( {\varphi ,\varsigma } \right) + 2\Lambda + 16\pi G\rho } \right)} \right]\end{array}\]

with overdot denotes time derivative, with the re-scaling:

    \[\left\{ {\begin{array}{*{20}{c}}{\phi = \sqrt {12} \varphi }\\{\sigma \sqrt {12} \varsigma }\end{array}} \right.\]

And the momenta are defined as:

    \[\left\{ {\begin{array}{*{20}{c}}{\prod\nolimits_{{q^i}} = \frac{{\partial \hat L}}{{\partial {q^i}}}}\\{{q^i} = \left( {{\beta _ \pm },\Omega ,\varphi ,\varsigma } \right)}\end{array}} \right.\]

and:

    \[\left\{ {\begin{array}{*{20}{c}}{{\Pi _\Omega } = \frac{{\partial \hat L}}{{\partial \dot \Omega }} = \frac{{12{e^{3\Omega }}\dot \Omega }}{{\rm{N}}}}\\{ \to \dot \Omega = \frac{{{\rm{N}}{\Pi _\Omega }}}{{12}}{e^{ - 3\Omega }}}\end{array}} \right.\]

    \[\left\{ {\begin{array}{*{20}{c}}{{\Pi _ \pm } = \frac{{\partial \hat L}}{{\partial {{\dot \beta }_ \pm }}} = \frac{{12{e^{3\Omega }}{{\dot \beta }_ \pm }}}{{\rm{N}}}}\\{ \to {{\dot \beta }_ \pm } = - \frac{{{\rm{N}}{\Pi _ \pm }}}{{12}}{e^{ - 3\Omega }}}\end{array}} \right.\]

    \[\left\{ {\begin{array}{*{20}{c}}{{\Pi _\varphi } = \frac{{\partial \hat L}}{{\partial \dot \varphi }} = \frac{{12{e^{3\Omega }}\dot \varphi }}{{\rm{N}}}}\\{ \to \dot \varphi = - \frac{{{\rm{N}}{\Pi _\varphi }}}{{12}}{e^{ - 3\Omega }}}\end{array}} \right.\]

    \[\left\{ {\begin{array}{*{20}{c}}{{\Pi _\varsigma } = \frac{{\partial \hat L}}{{\partial \dot \varsigma }} = \frac{{12{e^{3\Omega }}\dot \varsigma }}{{\rm{N}}}}\\{ \to \dot \varsigma = - \frac{{{\rm{N}}{\Pi _\varsigma }}}{{12}}{e^{ - 3\Omega }}}\end{array}} \right.\]

Thus, our Hamiltonian density is given by:

and by use of the above covariant derivative, we have:

    \[3\dot \Omega \rho + 3\dot \Omega p + p = 0\]

with:

    \[{\hat L_{{\rm{canonical}}}} = \prod\nolimits_{\rm{q}} {{\rm{\dot q}}} - {\rm{N}}{{\rm H}_{\rm{I}}}\]

    \[\frac{{\delta {{\hat L}_{{\rm{canonical}}}}}}{{\delta {\rm{N}}}} = 0\]

and:

    \[{{\rm H}_{\rm{I}}} = 0\]

and the density solution:

    \[\rho = {{\rm{M}}_\gamma }{e^{ - 3\left( {1 + \gamma } \right)\,\Omega }}\]

The Wheeler-DeWitt equation

Hence, our first approximation of the Wheeler-DeWitt equation is:

with:

    \[{\hat \bigcirc _L} = - \frac{{{\partial ^2}}}{{\partial {\Omega ^2}}} + \frac{{{\partial ^2}}}{{\partial {\varsigma ^2}}} + \frac{{{\partial ^2}}}{{\partial {\varphi ^2}}} + \frac{{{\partial ^2}}}{{\partial \beta _ - ^2}} + \frac{{{\partial ^2}}}{{\partial \beta _ + ^2}}\]

the d’Alambertian in the coordinates:

    \[{q^\mu } = \left( {\Omega ,{\beta _ \pm },\varsigma ,\varphi } \right)\]

with the {\rm{U}} potential that couples to the wave-function \psi and gives the whole quantum dynamics by the following equation:

    \[{\rm H}\psi = \left( {{g^{\mu \nu }}{\nabla _\mu }{\nabla _\nu } - V\left( {{q^\mu }} \right)} \right)\psi = 0\]

with:

    \[\psi = {\rm{R}}\left( {{q^\mu }} \right){e^{\frac{i}{\hbar }S\left( {{q^\mu }} \right)}}\]

Using the following ansatz for the wavefunction:

Hence, our Wheeler–DeWitt equation equation is:

with:

    \[\left\{ {\begin{array}{*{20}{c}}{{c^2} = \left( {a_1^2 + a_2^2} \right)}\\{{{\hat \bigcirc }_L} = {\ell ^\mu }\left( {\Omega ,\varsigma ,\varphi } \right)}\end{array}} \right.\]

and:

    \[\Xi \left( {{\ell ^\mu }} \right) = {\rm{W}}\left( {{\ell ^\mu }} \right){e^{ - \frac{{{S_\hbar }}}{\hbar }\left( {{\ell ^\mu }} \right)}}\]

where {S_\hbar }\left( {{\ell ^2}} \right) is the superpotential function, and {\rm{W}} is the probability amplitude.

Let us now utilize the mathematics of supersymmetric quantum mechanics to probe the Wheeler–DeWitt equation and the superpotential via Witten’s formalism of finding the supersymmetric supercharges operators Q and \bar Q that produce a super-Hamiltonian {{\rm{H}}_{{\rm{ss}}}}, where the Wheeler–DeWitt equation can be derived as the bosonic sector of this super-Hamiltonian in the superspace. The right method to supersymmetrize a bosonic Lagrangian is to consider the true supersymmetry transformation in the superfield scheme into the bosonic Lagrangian, then the fermionic terms will emerge in a natural way.

In this Witten-method, our supercharges for the 3-D case are:

    \[Q = {\psi ^\mu }\left[ { - \hbar {\partial _{{{\rm{q}}^\mu }}} + \frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}}} \right]\]

    \[\bar Q = {\bar \psi ^\nu }\left[ { - \hbar {\partial _{{{\rm{q}}^\nu }}} + \frac{{\partial S}}{{\partial {{\rm{q}}^\nu }}}} \right]\]

where S is defined implicitly by the following equation:

    \[S = \frac{{{e^{3\Omega }}}}{\mu }g\left( \varphi \right){\rm{h}}\left( \varsigma \right)\]

and the super-algebra for the variables \left( {{\psi ^\mu },{{\bar \psi }^\nu }} \right) is given by:

    \[\left\{ {\begin{array}{*{20}{c}}{\left\{ {{\psi ^\mu },{{\bar \psi }^\nu }} \right\} = {\eta ^{\mu \nu }}}\\{\left\{ {{\psi ^\mu },{\psi ^\nu }} \right\} = 0}\\{\left\{ {{{\bar \psi }^\mu },{{\bar \psi }^\nu }} \right\} = 0}\end{array}} \right.\]

Under the representation:

    \[{\psi ^\nu } = {\theta ^\nu }{\rm{y}}{\bar \psi ^\mu } = {\eta ^{\mu \nu }}\frac{\partial }{{\partial {\theta ^\nu }}}\]

the superspace Hamiltonian takes the form:

    \[{{\rm{H}}_{{\rm{ss}}}} = \left\{ {Q,\bar Q} \right\} = {H_0} + \hbar \frac{{{\partial ^2}S}}{{\partial {{\rm{q}}^\mu }\partial {{\rm{q}}^\nu }}}\left[ {{\psi ^\mu },{{\bar \psi }^\nu }} \right]\]

with:

    \[{H_0} = {\hat \bigcirc _L} - {\rm{U}}\left( {{{\rm{q}}^\mu }} \right)\]

being the standard Wheeler–DeWitt equation, {\hat \bigcirc _L} the 3-D d’Alambertian in the {{\rm{q}}^\mu } coordinates with {\eta _{\mu \nu }} = {\rm{diag}}\left( { - 1,1,1} \right), and \left\{ {\,,\,} \right\} and \left[ {\,,\,} \right] represent the anticommutator and the commutator respectively. The supercharges and the super-Hamiltonian satisfy the following algebra:

    \[\left\{ {\begin{array}{*{20}{c}}{\left\{ {Q,\bar Q} \right\} = {{\rm{H}}_{{\rm{ss}}}}}\\{\left[ {{{\rm{H}}_{{\rm{ss}}}},Q} \right] = \left[ {{{\rm{H}}_{{\rm{ss}}}},\bar Q} \right] = 0}\end{array}} \right.\]

Hence, our supersymmetric physical states are selected by the constraints:

    \[\left\{ {\begin{array}{*{20}{c}}{Q\Psi = 0}\\{\bar Q\Psi = 0}\end{array}} \right.\]

which reduces the problem of finding supersymmetric ground states because the energy is known a priori and the factorization of:

    \[{{\rm{H}}_{{\rm{ss}}}}\left| \Psi \right\rangle = 0\]

into

    \[\left\{ {\begin{array}{*{20}{c}}{Q\Psi = 0}\\{\bar Q\Psi = 0}\end{array}} \right.\]

yields a first-order equation for the ground state wave-function due to the sovability of the bosonic Hamiltonians and normalization just means that supersymmetry is quantum mechanically unbroken.

In the 3-D Grassmannian variable-representation, the wave-function has the following decomposition:

    \[\begin{array}{c}\Psi = {{\tilde A}_ + } + {{\tilde B}_\nu }{\theta ^\mu } + \frac{1}{2}{\varepsilon _{\mu \nu \lambda }}{C^\lambda }{\theta ^\mu }{\theta ^\nu }\\ + {{\tilde A}_ - }{\theta ^0}{\theta ^1}{\theta ^2}\end{array}\]

and with the ansatz:

    \[{\tilde B_\nu } = \frac{{\partial {{\rm{f}}_ + }\left( {{{\rm{q}}^\nu }} \right)}}{{\partial {{\rm{q}}^\nu }}}{e^{\frac{{{S_{\left( {\rm{q}} \right)}}}}{\hbar }}}\]

introduced into

    \[\left\{ {\begin{array}{*{20}{c}}{Q\Psi = 0}\\{\bar Q\Psi = 0}\end{array}} \right.\]

and

    \[\begin{array}{c}\Psi = {{\tilde A}_ + } + {{\tilde B}_\nu }{\theta ^\mu } + \frac{1}{2}{\varepsilon _{\mu \nu \lambda }}{C^\lambda }{\theta ^\mu }{\theta ^\nu }\\ + {{\tilde A}_ - }{\theta ^0}{\theta ^1}{\theta ^2}\end{array}\]

where S is the superpotential function obtained as a solution for the Einstein-Hamilton-Jacobi equation, the following identity:

    \[{\left( {\nabla {S_\hbar }} \right)^2} - \tilde U = 0\]

yields the master equation for the auxiliary function {{\rm{f}}_ + }:

    \[\hbar {\hat \bigcirc _L}{{\rm{f}}_ + } + 2{\eta ^{\mu \nu }}\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}}\frac{{\partial {{\rm{f}}_ + }}}{{\partial {{\rm{q}}^\nu }}} = 0\]

with:

    \[\frac{1}{2}{\varepsilon _{\mu \nu \lambda }}{C^\lambda }{\theta ^\alpha }{\theta ^\mu }{\theta ^\nu } = {C^\alpha }{\theta ^0}{\theta ^1}{\theta ^2}\]

and with the following ansatz:

    \[{C^\mu } = {\eta ^{\mu \nu }}\frac{{\partial {{\rm{f}}_ - }}}{{\partial {{\rm{q}}^\nu }}}{e^{ - \frac{{{S_{\left( {\rm{q}} \right)}}}}{\hbar }}}\]

we get the second master equation in the form:

    \[\hbar {\hat \bigcirc _L}{{\rm{f}}_ - } - 2{\eta ^{\mu \nu }}\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}}\frac{{\partial {{\rm{f}}_ - }}}{{\partial {{\rm{q}}^\nu }}} = 0\]

allowing us to get the reduction to:

    \[\hbar {\hat \bigcirc _L}{{\rm{f}}_ \pm } \pm 2{\eta ^{\mu \nu }}\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}}\frac{{\partial {{\rm{f}}_ \pm }}}{{\partial {{\rm{q}}^\nu }}} = 0\]

and the equations for the functions {\tilde A_ \pm } are:

    \[\left[ {\hbar \frac{\partial }{{\partial {{\rm{q}}^\mu }}} \pm \frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}}} \right]{\tilde A_ \pm } = 0\]

with solutions:

    \[{\tilde A_ \pm } = {a_{0 \pm }}{e^{ \pm \frac{1}{\hbar }S}}\]

with {a_{0 \pm }} the integration constants.

To solve our equation:

    \[\hbar {\hat \bigcirc _L}{{\rm{f}}_ \pm } \pm 2{\eta ^{\mu \nu }}\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}}\frac{{\partial {{\rm{f}}_ \pm }}}{{\partial {{\rm{q}}^\nu }}} = 0\]

we need to write it as a homogeneous linear equation of second degree:

    \[{\hat \bigcirc _L}{{\rm{W}}_ \pm } = {{\rm{W}}_ \pm }g\left( {{{\rm{q}}^\mu }} \right)\]

and we do this by introducing into it the ansatz:

    \[{{\rm{f}}_ \pm } = {{\rm{W}}_ \pm }\left( {{{\rm{q}}^\mu }} \right){e^{ \pm \phi \left( {{{\rm{q}}^\mu }} \right)/\hbar }}\]

This way, we obtain a wave-like equation:

    \[{\hat \bigcirc _L}{{\rm{W}}_ \pm } \pm {{\rm{W}}_ \pm }{\hat \bigcirc _L}S - {{\rm{W}}_ \pm }{\left( {\nabla S} \right)^2} = 0\]

with:

    \[g\left( {{{\rm{q}}^\mu }} \right) = \left( {\nabla S} \right) \mp {\hat \bigcirc _L}S\]

and:

    \[{\hat \bigcirc _L}{{\rm{W}}_ \pm } = g\left( {{{\rm{q}}^\mu }} \right){{\rm{W}}_ \pm }\]

The following wave-like ansatz:

    \[{{\rm{W}}_ \pm } = {\beta _ \pm }{e^{ \mp * }}\]

suffices to solve, yielding a condition on the {\rm{s}} function:

    \[\left[ {{{\left( {\nabla {\rm{s}}} \right)}^2} \mp \,{{\hat \bigcirc }_L}{\rm{s}}} \right] = \left[ {{{\left( {\nabla {\rm{S}}} \right)}^2} \mp \,{{\hat \bigcirc }_L}{\rm{S}}} \right]\]

where the following conditions hold:

    \[\left\{ {\begin{array}{*{20}{c}}{{\rm{s}} = S \mp {\rm{h}}\left( {{{\rm{q}}^\mu }} \right)}\\{{\rm{h}}\left( {{{\rm{q}}^\mu }} \right) \equiv {{\rm{m}}_\mu }{{\rm{q}}^\mu }}\end{array}} \right.\]

Allowing us to construct the following term:

    \[\left[ {{{\left( {\nabla {\rm{s}}} \right)}^2} \mp \,{{\hat \bigcirc }_L}{\rm{s}}} \right]\]

satisfying:

    \[\begin{array}{l}{\left( {\nabla {\rm{s}}} \right)^2} \mp {{\hat \bigcirc }_L}{\rm{s}}{\left( {\nabla {\rm{S}}} \right)^2} \mp \,{{\hat \bigcirc }_L}S \pm \\2{\eta ^{\mu \alpha }}{{\rm{m}}_\alpha }\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}} + {{\rm{m}}^\mu }{{\rm{m}}_\mu }\end{array}\]

Now, for:

    \[2{{\rm{m}}^\mu }\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}} \mp {{\rm{m}}^\mu }{{\rm{m}}_\mu } = 0\]

we must consider the two cases, one with taking the constant c into account and one without.

For c \ne 0 and with our superpotential. In this situation,

    \[2{{\rm{m}}^\mu }\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}} \mp {{\rm{m}}^\mu }{{\rm{m}}_\mu } = 0\]

gives the following equation:

    \[\begin{array}{l}\frac{{{e^{3\Omega }}g{\rm{h}}}}{\mu }\left[ { - 6{{\rm{m}}_0} + 6{{\rm{m}}_1}\frac{{{\eta _1}}}{{{{\rm{b}}_3}}}} \right] - \\2c\left( { - {{\rm{m}}_0}{{\rm{b}}_1} + {{\rm{m}}_1}{{\rm{b}}_2} + {{\rm{m}}_2}{{\rm{b}}_3}} \right) + {\rm{m}}_0^2 - {\rm{m}}_1^2 - {\rm{m}}{2^2} = 0\end{array}\]

with vector solutions:

    \[\left\{ {\begin{array}{*{20}{c}}{{{\rm{m}}_\mu } = \left( {2c{{\rm{b}}_1},2c{{\rm{b}}_2},2c{{\rm{b}}_3}} \right)}\\{{{\rm{b}}_1} \equiv {\eta _1} + {\eta _2}}\end{array}} \right.\]

For c = 0 and with our superpotential. In this situation, we need to separate the following two independent equations:

    \[\left\{ {\begin{array}{*{20}{c}}{{{\rm{m}}^\mu }{{\rm{m}}_\mu } = 0}\\{{\eta ^{\mu \alpha }}{{\rm{m}}_\alpha }\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}} = 0}\end{array}} \right.\]

where:

    \[{{\rm{m}}^\mu }{{\rm{m}}_\mu } = 0\]

entails that {{\rm{m}}_\mu } is a vector of null measure, and:

    \[{\eta ^{\mu \alpha }}{{\rm{m}}_\alpha }\frac{{\partial S}}{{\partial {{\rm{q}}^\mu }}} = 0\]

entails:

    \[\frac{{\partial S}}{{\partial \Omega }}{{\rm{m}}_0} = \frac{{\partial S}}{{\partial \phi }}{{\rm{m}}_1} + \frac{{\partial S}}{{\partial \sigma }}{{\rm{m}}_2}\]

Now, when we use the superpotential function:

    \[S = \frac{{{e^{3\Omega }}}}{\mu }g\left( \varphi \right){\rm{h}}\left( \varsigma \right)\]

we get the following structural relations:

    \[\left\{ {\begin{array}{*{20}{c}}{{\rm{g}}\left( \phi \right) = {g_0}{e^{{\varepsilon _1}\Delta \phi }}}\\{{\rm{h}}\left( \sigma \right) = {{\rm{h}}_0}{g_0}{e^{{\varepsilon _2}\Delta \phi }}}\end{array}} \right.\]

for \left( {g\left( \varphi \right){\rm{;h}}\left( \varsigma \right)} \right), where the constants are:

    \[\left\{ {\begin{array}{*{20}{c}}{{\varepsilon _1} = \frac{{3{{\rm{m}}_0}{{\rm{n}}_1}}}{{{{\rm{m}}_2}}}}\\{{\varepsilon _2} = \frac{{3{{\rm{m}}_0}{{\rm{n}}_2}}}{{{{\rm{m}}_1}}}}\end{array}} \right.\]

Hence, supersymmetric quantum mechanics puts the exact constraints on the family of potential fields corresponding to the inflaton exponential Hubble-Fredholm integral.

In the scenario where both equations have no null solution, the solution for the function {{\rm{f}}_ \pm } has the following structure:

    \[{{\rm{f}}_ \pm } = {{\rm{b}}_ \pm }{e^{{{\rm{m}}_\alpha }{{\rm{q}}^\alpha }}}\]

thus {\tilde B_\mu } and {C^\mu } reduce to:

    \[{\tilde B_\mu } = {{\rm{b}}_ + }{{\rm{m}}_\mu }{e^{{{\rm{m}}_\mu }{{\rm{q}}^\mu }}}{e^{\frac{S}{\hbar }}}\]

    \[{C^\mu } = {\eta ^{\mu \nu }}{{\rm{b}}_ - }{{\rm{m}}_\nu }{e^{{{\rm{m}}_\alpha }{{\rm{q}}^\alpha }}}{e^{ - \frac{S}{\hbar }}}\]

Our method above was used to obtain supersymmetric quantum solutions for all cosmological bianchi class A models in Sáez-Ballester theory.

From our superpotential function above, it follows that the only form for S in which our equations are fulfilled, is when the functions {\rm{g}} and {\rm{h}} have exponential behavior. So in a supersymmetric way, the calculation by means of the Grassmannian variables of {\left| \Psi \right|^2} given by:

    \[\begin{array}{c}\Psi = {{\tilde A}_ + } + {{\tilde B}_\nu }{\theta ^\mu } + \frac{1}{2}{\varepsilon _{\mu \nu \lambda }}{C^\lambda }{\theta ^\mu }{\theta ^\nu }\\ + {{\tilde A}_ - }{\theta ^0}{\theta ^1}{\theta ^2}\end{array}\]

is:

where * is implicitly defined by:

    \[{\left( {C{\theta _1}...{\theta _n}} \right)^ * } = \theta _n^ * ...\theta _1^ * {C^ * }\]

with the standard algebra for the Grassmannian numbers {\theta _i}{\theta _j} = - {\theta _j}{\theta _i}. The integration rules over these numbers are given by:

    \[\int {{\theta _1}\theta _1^ * ...{\theta _n}\theta _n^ * d\theta _n^ * d{\theta _n}...d\theta _1^ * d{\theta _1} = 1} \]

with:

    \[\int {d\theta _i^ * } = \int {d{\theta _i}} = 0\]

and we get:

    \[{\Psi _1} = {\Psi _2} = \Psi \]

Thus, Grassmannian integration yields:

    \[\begin{array}{l}{\left| \Psi \right|^2} = {{\tilde A}_ + }{A_ + } + {{\tilde A}_ - }{A_ - } + {{\tilde B}_0}{B_0} + \\{{\tilde B}_1}{B_1} + {{\tilde B}_2}{B_2} + {{\tilde C}^0}{C^0} + {{\tilde C}^1}{C^1} + {{\tilde C}^2}{C^2}\end{array}\]

thus supersymmetric quantum mechanics yields the required probability density:

giving us a supersymmetric quantum canonical quantization of the multi-scalar field cosmology of the anisotropic Bianchi type I model and the exact supersymmetric quantum solutions to the Wheeler-DeWitt equation are derived under the ansatz to the wave function:

    \[\Psi \left( {{\ell ^\mu }} \right) = {e^{\frac{{{{\rm{a}}_1}}}{\hbar }{\beta _ + } + {\rm{i}}\frac{{{{\rm{a}}_1}}}{\hbar }{\beta _ - }}}{\rm{W}}\left( {{\ell ^\mu }} \right){e^{ - \frac{{S\left( {{\ell ^\mu }} \right)}}{\hbar }}}\]

which is central for solving the Einstein-Hamilton-Jacobi equation.