HTML
--> --> -->The principal method used to date has primarily focused on the asymptotic states, which are somewhat similar to the usual Minkowskian particle or antiparticle states, as in the case of the adiabatic Bunch-Davies vacuum type [21] largely used in applications. The problem arising in this vacuum is that one cannot reach the rest limit for vanishing momentum, as this limit is undefined for the corresponding scalar mode functions or Dirac spinors. This problem has been considered sporadically by other authors, who have found similar results [2, 4, 7, 18].
Hence, we propose a method of separating the frequencies in rest frames, thus defining the rest frame vacuum (r.f.v.) of the massive Klein-Gordon [22], Dirac [23], and Proca [24] fields. The novelty of this approach is that the bosons can have either a tardyonic behavior or even a tachyonic one, if the mass is less than a given limit, which depends on the type of coupling (minimal or conformal). Fortunately, the tachyonic modes are eliminated in a natural manner as all their mode functions have null norms [22]. In contrast, the Dirac field of any non-vanishing mass survives in this vacuum [23]. In other respects, we must specify that the r.f.v. can be defined only for massive particles as the massless ones do not have rest frames. This is not an impediment because the massless fields of physical interest, namely the Maxwell and neutrino ones, have conformally covariant field equations for which the solutions in the co-moving de Sitter chart with conformal time can be drawn from special relativity [16, 25].
Technically speaking, to define the r.f.v., we introduced suitable phases depending on momentum to ensure the correct limits of the mode functions in rest frames [22-24]. Unfortunately, these phases are not sufficient to define the other important limit, namely the flat one, when the de Sitter Hubble constant tends to zero. In general, this limit is undefined because of singularities arising in the phases of the mode functions defined in the adiabatic vacuum. To remove these, a regularization procedure was applied by adding the convenient phase factors which, in general, depend on momentum [2, 4, 7, 18].
As this problem has not yet been considered for the recently defined r.f.v., the aim of this paper is to study the flat limits of the Klein-Gordon and Dirac fundamental solutions in this vacuum, deriving the regularized phases that guarantee that these limits are well-defined. In this manner, the rest and flat limits completely determine the form of the mode functions or spinors of the de Sitter QFT. This is important because there are quantities with expressions that are strongly dependent on the momentum-dependent phase factors, such as the one-particle energy (or Hamiltonian) operator [2, 6, 16]. We prove that the phase factors derived here determine the correct Minkowskian flat limit of this operator.
We specify that we recover previous results [2, 4, 7, 18] concerning the regularized phases, but the complete expressions of the scalar mode functions or Dirac spinors are presented here for the first time. Moreover, the approximations we obtain here are also new results that can be used in concrete calculations. However, the principal new result is that in the r.f.v., the flat limit is reached naturally, including for vanishing momenta, in contrast to the adiabatic vacuum in which the rest limit is undefined, forcing one to redefine the time modulation functions [7].
In the next section, we present the framework of our attempt to briefly review the theory of covariant fields on local-Minkowskian curved backgrounds, observing that these fields transform under isometries according to covariant representations (reps.) induced by the finite-dimensional reps. of the Lorentz group. In the second part of this section, the covariant reps. on the de Sitter space-time are considered focusing on their generators, which are the principal conserved observables of the quantum theory. The third section is devoted to the plane wave fundamental solutions of the Klein-Gordon and Dirac fields in the adiabatic and rest frame vacua, observing that the last vacuum naturally solves the rest limits. In the next section, we derive the flat limits of the aforementioned fields using a new uniform asymptotic expansion we propose here based on numerical arguments. This enables us to derive the regularized phases that ensure the flat limits of the fundamental solutions. Thus, we obtain, for the first time, complete expressions of the scalar mode functions and Dirac spinors in the r.f.v. as well as useful approximations of these quantities. Moreover, we show that in this approach, the flat limit of the one-particle energy operator is simply the corresponding operator of special relativity. Finally, we present our concluding remarks.
The presence of the Dirac field requires orthogonal unholonomic (or local) frames in which the Dirac matrices make sense. These frames are defined by the tetrad vector fields
2
A.Covariant fields
To obtain a gauge-invariant theory, we define the covariant fields $ A(\omega) = \exp\left(-\frac{{\rm i}}{2}\omega^{\hat\alpha\hat\beta} S_{\hat\alpha\hat\beta}\right) \in SL(2,{\mathbb{C}}) $ ![]() | (1) |
The Lagrangian theory of the covariant field
$ D_{\hat\alpha}^{(\rho)} = e_{\hat\alpha}^{\mu}D_{\mu}^{(\rho)} = \hat\partial_{\hat\alpha}+\frac{{\rm i}}{2}\, \rho(S^{\hat\beta\, \cdot} _{\cdot\, \hat\gamma})\,\hat\Gamma^{\hat\gamma}_{\hat\alpha \hat\beta}\,, $ ![]() | (2) |
$ \hat\Gamma^{\hat\sigma}_{\hat\mu \hat\nu} = e_{\hat\mu}^{\alpha} e_{\hat\nu}^{\beta}(\hat e_{\gamma}^{\hat\sigma} \Gamma^{\gamma}_{\alpha \beta} -\hat e^{\hat\sigma}_{\beta, \alpha})\,, $ ![]() | (3) |
When
$ x\to x' = \phi_{{\frak g}(\xi)}(x) = x+\xi^a k_a(x) +... $ ![]() | (4) |
In general, the isometries may change the relative position of the local frames, thus affecting the physical interpretation. For this reason, we propose the theory of external symmetry [26], in which we introduce the combined transformations
$ \Lambda^{\hat\alpha\,\cdot}_{\cdot\,\hat\beta}[A_{\frak g}(x)] = \hat e_{\mu}^{\hat\alpha}[\phi_{\frak g}(x)]\frac{\partial \phi^{\mu}_{\frak g}(x)} {\partial x^{\nu}}\,e^{\nu}_{\hat\beta}(x)\,, $ ![]() | (5) |
$ ({A_{\frak g}},{\phi _{\frak g}}):\quad \begin{aligned}{e(x)}& \to &{e'(x')}& = &{e[{\phi _{\frak g}}(x)]{\mkern 1mu} ,}\\{\hat e(x)}& \to &{\hat e'(x')}& = &{\hat e[{\phi _{\frak g}}(x)]{\mkern 1mu} ,}\end{aligned}$ ![]() | (6) |
$ (A_{\frak g},\phi_{\frak g}):\quad \psi_{(\rho)}(x)\to\psi_{(\rho)}'(x') = \rho[A_{\frak g}(x)]\psi_{(\rho)}(x)\,, $ ![]() | (7) |
$ (T_{\frak g}^{(\rho)}\psi_{(\rho)})[\phi_{\frak g}(x)] = \rho[A_{\frak g}(x)]\psi_{(\rho)}(x)\,, $ ![]() | (8) |
We have shown that the pairs
For each Killing vector
$ \begin{aligned}[b]X^{(\rho)}_{a}& = {\rm i}{\partial_{\xi^a} T_{{\frak g}(\xi)}^{(\rho)}}_{|\xi = 0}\\ \;\;\;\;\;\;\;& = -{\rm i}k^{\mu}_{a}D_{\mu}^{(\rho)}+\frac{1}{2}\, k_{a\, \mu;\,\nu}\,e^{\mu}_{\hat\alpha}\,e^{\nu}_{\hat\beta}\, \rho(S^{\hat\alpha\hat\beta})\,,\end{aligned}$ ![]() | (9) |
Another advantage of the Lagrangian theory is the possibility of introducing the relativistic scalar product
In this framework, for any conserved operator X and field
$ C[X]\; \; \to\; \; {\cal{X}} = :\langle\psi,X\psi\rangle: $ ![]() | (10) |
2
B.Covariant fields on the de Sitter expanding universe
The $ \eta^5_{AB}z^A(x) z^B(x) = -\frac{1}{\omega^2}\,. $ ![]() | (11) |
The group of external symmetry,
$ L_{AB}^{5} = i\left[\eta_{AC}^5 z^{C}\partial_{B}- \eta_{BC}^5 z^{C} \partial_{A}\right] = -iK_{(AB)}^C\partial_C , $ ![]() | (12) |
$ \xi_{(AB)} \; \; \to\; \; k_{(AB)\mu} = -z_A(x)\stackrel{\leftrightarrow}{\partial}_{\mu}z_B(x) , $ ![]() | (13) |
Exploiting the special coset structure of the de Sitter manifold,
The hyperboloid equation is solved simply in the conformal chart
$ \begin{aligned}[b]z^0(x) &= -\frac{1}{2\omega^2 t_c}\left[1-\omega^2({t_c}^2 - \vec{x}^2)\right]\,, \\ z^i(x) &= -\frac{1}{\omega t_c}x^i \,, \\ z^4(x) &= -\frac{1}{2\omega^2 t_c}\left[1+\omega^2({t_c}^2 - \vec{x}^2)\right]\,,\end{aligned} $ ![]() | (14) |
$ {\rm d}s^{2} = \eta^5_{AB}{\rm d}z^A(x){\rm d}z^B(x) = \frac{1}{\omega^2 {t_c}^2}\left({{\rm d}t_c}^{2}-{\rm d}\vec{ x}\cdot {\rm d}\vec{x}\right)\,. $ ![]() | (15) |
$ M_+ \; :\; \; -\infty<t_c = -\frac{1}{\omega}{\rm e}^{-\omega t}<0 \,, $ ![]() | (16) |
$ M_- \; :\; \; \; \; \; \; \; 0<t_c = \frac{1}{\omega}{\rm e}^{\omega t}<\infty\,. $ ![]() | (17) |
$ {\rm d}s^2 = {\rm d}t^2-{\rm e}^{2\omega t} {\rm d}{\vec x}\cdot {\rm d}{\vec x}\,, $ ![]() | (18) |
The isometry group
$ H = \omega X_{(04)}^{(\rho)} = -i\omega(t_c\partial_{t_c}+x^i\partial_i) = i\partial_t-i\omega x^i\partial_i \,,\; \; \; \; $ ![]() | (19) |
$ P^i = \omega(X_{(4i)}^{(\rho)}-X_{(0i)}^{(\rho)}) = -i\partial_i\,, $ ![]() | (20) |
Respecting ad litteram the principles of the quantum theory, we assume that the quantum states of
Finally, we note that another attempt using de Sitter space-times considers only the unitary and irreducible linear reps. of the
On
2
A.Klein-Gordon field
In the chart $ \left( \partial_t^2-{\rm e}^{-2\omega t}\Delta +3\omega \partial_t+m^2\right)\Phi(x) = 0\,, $ ![]() | (21) |
$ \begin{aligned}[b] \Phi(x) =& \Phi^{(+)}(x)+\Phi^{(-)}(x)\\ =& \int {\rm d}^3p \left[f_{\vec{p}}(x)a(\vec{p})+f_{\vec{p}}^*(x)b^{\dagger}(\vec{p})\right] \,, \end{aligned} $ ![]() | (22) |
$ \langle f_{\vec{p}},f_{\vec{p}'}\rangle_{KG} = -\langle f_{\vec{p}}^*,f_{\vec{ p}'}^*\rangle_{KG} = \delta^3(\vec{p}-\vec{p}')\,, $ ![]() | (23) |
$ \langle f_{\vec{p}},f_{\vec{p}'}^*\rangle_{KG} = 0\,, $ ![]() | (24) |
$ \langle f,f'\rangle_{KG} = {\rm i}\int {\rm d}^3x\, a(t)^3\, f^*(x) \stackrel{\leftrightarrow}{\partial_{t}} f'(x)\,. $ ![]() | (25) |
$f \in \left\{ \begin{aligned}&{{{\cal{H}}_ + } \subset {{\cal{K}}_ + }}&{{\rm{if}}}&{{{\langle f,f\rangle }_{KG}} > 0{\mkern 1mu} ,}\\&{{{\cal{H}}_0} \subset {{\cal{K}}_0}}&{{\rm{if}}}&{{{\langle f,f\rangle }_{KG}} = 0{\mkern 1mu} ,}\\&{{{\cal{H}}_ - } \subset {{\cal{K}}_ - }}&{{\rm{if}}}&{{{\langle f,f\rangle }_{KG}} < 0{\mkern 1mu} .}\end{aligned} \right.$ ![]() | (26) |
The fundamental mode functions can be expressed as
$ f_{\vec{p}}(t,\vec{x}) = \frac{{\rm e}^{{\rm i} \vec{x}\cdot \vec{p}}}{[2\pi a(t)]^{\frac{3}{2}}}{\cal{F}}_p(t)\,, $ ![]() | (27) |
$ \left[\frac{{\rm d}^2}{{\rm d}t^2}+\frac{p^2}{a(t)^2}+m^2-\frac{9}{4}\,\omega^2\right] {\cal{F}}_p(t) = 0\,, $ ![]() | (28) |
$ \left({\cal{F}}_p, {\cal{F}}_p\right)\equiv i\,{\cal{F}}_p^*(t)\stackrel{\leftrightarrow}{\partial}_{t}{\cal{F}}_p(t) = 1\,, $ ![]() | (29) |
The general solution of Eq. (28) can be derived easily in the chart
$ {\cal{F}}_p(t_c) = \ c_1\phi_p(t_c) + c_2\phi_p^*(t_c)\,,\quad \phi_p(t_c) = \frac{1}{\sqrt{\pi\omega}}\,K_{\nu}({\rm i}pt_c)\,, $ ![]() | (30) |
$\nu= \left\{ {\begin{aligned}{\sqrt {\frac{9}{4} - {\mu ^2}} }\quad {{\rm{for}}}\quad {\mu < \frac{3}{2}}\\{{\rm i}\kappa {\mkern 1mu} ,\kappa = \sqrt {{\mu ^2} - \frac{9}{4}} }\quad {{\rm{for}}}\quad {\mu > \frac{3}{2}}\end{aligned}} \right.{\mkern 1mu} , $ ![]() | (31) |
$ \left|c_1\right|^2-\left|c_2\right|^2 = 1\,. $ ![]() | (32) |
The most popular vacuum is the adiabatic Bunch-Davies type [21], with
$ K_{{\rm i}\kappa} ({\rm i}pt_c)\propto \frac{1}{\Gamma\left(\dfrac{1}{2}-{\rm i}\kappa\right)}\left(\frac{{\rm i}pt_c}{2}\right)^{-{\rm i}\kappa}-\frac{1}{\Gamma\left(\dfrac{1}{2}+{\rm i}\kappa\right)}\left(\frac{{\rm i}pt_c}{2}\right)^{{\rm i}\kappa}\,, $ ![]() | (33) |
To avoid these difficulties, we recently defined the r.f.v., separating the frequencies in the rest frames just as in special relativity [22]. Thus, we found that the rest energy,
$ M = \kappa\omega = \sqrt{m^2-\frac{9}{4}\omega^2}\,, \quad m>\frac{3}{2}\omega\,, $ ![]() | (34) |
$ c_1 = -{\rm i}\left(\frac{p}{2\omega}\right)^{-{\rm i}\kappa}\frac{{\rm e}^{\pi\kappa}}{\sqrt{{\rm e}^{2\pi\kappa}-1}}\,, $ ![]() | (35) |
$ c_2 = {\rm i}\left(\frac{p}{2\omega}\right)^{-{\rm i}\kappa}\frac{1}{\sqrt{{\rm e}^{2\pi\kappa}-1}}\,, $ ![]() | (36) |
$ {\cal{F}}_p(t_c) = \sqrt{\frac{\pi}{\omega}}\left(\frac{p}{2\omega}\right)^{-{\rm i}\kappa}\frac{I_{{\rm i}\kappa}({\rm i}pt_c)}{\sqrt{{\rm e}^{2\pi\kappa}-1}}\,. $ ![]() | (37) |
$ \mathop {\lim }\limits_{p \to 0} {\cal{F}}_p(t_c) = \frac{1}{\sqrt{2M}}\,{\rm e}^{-{\rm i} M t}\,, $ ![]() | (38) |
Finally, we must specify that the adiabatic and rest frame vacua are different from a physical perspective. First, they differ because in the adiabatic vacuum, the scalar particles may have any mass, whereas in the r.f.v. only the particles with
2
B.Dirac field
The Dirac field transforms under isometries according to a rep. induced by the Dirac rep. $ \rho_D(S^{\hat\alpha\hat\beta}) = \frac{{\rm i}}{4}\left[\gamma^{\hat\alpha}, \gamma^{\hat\beta}\right]\,. $ ![]() | (39) |
$ e_0 = \partial_t = \frac{1}{a(t_c)}\,\partial_{t_c}\,,\quad \; \; \; \; \omega^0 = {\rm d}t = a(t_c){\rm d}t_c\,, $ ![]() | (40) |
$ e_i = \frac{1}{a(t)}\,\partial_i = \frac{1}{a(t_c)}\,\partial_i\,, \quad \omega^i = a(t){\rm d}x^i = a(t_c){\rm d}x^i\,,\; \; \; \; \; $ ![]() | (41) |
$ D_x = {\rm i}\gamma^0\partial_{t}+ {\rm i} {\rm e}^{-\omega t}\gamma^i\partial_i +\frac{3{\rm i} \omega}{2} \gamma^{0}\,, $ ![]() | (42) |
The general solution of the Dirac equation may be written as a mode integral,
$ \begin{aligned}[b]\psi({x}\,)& = \psi^{(+)}({x}\,)+\psi^{(-)}({x}\,)\\ \; \; \; \; & = \int {\rm d}^{3}p \sum\limits_{\sigma}[U_{\vec{p},\sigma}(x){\frak a}(\vec{p},\sigma) +V_{\vec{p},\sigma}(x){\frak b}^{\dagger}(\vec{p},\sigma)]\,,\end{aligned} $ ![]() | (43) |
$ P^i U_{\vec{p},\sigma}(x) = p_iU_{\vec{p},\sigma}(x)\,, \quad P^i V_{\vec{p},\sigma}(x) = -p_iV_{\vec{p},\sigma}(x)\,, $ ![]() | (44) |
$ V_{\vec{p},\sigma}(t,\vec{x}) = U^c_{\vec{p},\sigma}(t,\vec{x}) = C\left[{U}_{\vec{p},\sigma}(t,\vec{x})\right]^* \,, \quad C = {\rm i}\gamma^2\,, $ ![]() | (45) |
$ \langle U_{\vec{p},\sigma}, U_{{\vec{p}\,}',\sigma'}\rangle_D = \langle V_{\vec{p},\sigma}, V_{{\vec{p}\,}',\sigma'}\rangle_D = \delta_{\sigma\sigma^{\prime}}\delta^{3}(\vec{p}-\vec{p}\,^{\prime})\; \; \; $ ![]() | (46) |
$ \langle U_{\vec{p},\sigma}, V_{{\vec{p}\,}',\sigma'}\rangle _D = \langle V_{\vec{p},\sigma}, U_{{\vec{p}\,}',\sigma'}\rangle_D = 0\,, $ ![]() | (47) |
$ \langle \psi, \psi'\rangle_D = \int {\rm d}^{3}x\, a(t)^{3}\bar{\psi}(x)\gamma^{0}\psi(x)\,, $ ![]() | (48) |
In the standard rep. of the Dirac matrices (with diagonal
$ U_{\vec{p},\sigma}(t,\vec{x}\,) = \frac{{\rm e}^{{\rm i}\vec{p}\cdot\vec{x}}}{[2\pi a(t)]^{\frac{3}{2}}}\left( \begin{array}{c} u^+_p(t) \, \xi_{\sigma}\\ u^-_p(t) \, \dfrac{{p}^i{\sigma}_i}{p}\,\xi_{\sigma} \end{array}\right)\,, $ ![]() | (49) |
$ V_{\vec{p},\sigma}(t,\vec{x}\,) = \frac{{\rm e}^{-{\rm i}\vec{p}\cdot\vec{x}}}{[2\pi a(t)]^{\frac{3}{2}}} \left( \begin{array}{c} v^+_p(t)\, \dfrac{{p}^i{\sigma}_i}{p}\,\eta_{\sigma}\\ v^-_p(t) \,\eta_{\sigma} \end{array}\right) \,, $ ![]() | (50) |
The time modulation functions
$ \left[{\rm i}\partial_{t_c}\mp m\, a(t_c)\right]u_p^{\pm}(t_c) = {p}\,u_p^{\mp}(t_c)\,, $ ![]() | (51) |
$ \left[{\rm i}\partial_{t_c} \mp m\, a(t_c)\right]v_p^{\pm}(t_c) = -{p}\,v_p^{\mp}(t_c)\,, $ ![]() | (52) |
$ v_p^{\pm} = \left[u_p^{\mp}\right]^*\,, $ ![]() | (53) |
$ |u_p^+|^2+|u_p^-|^2 = |v_p^+|^2+|v_p^-|^2 = 1 \\ $ ![]() | (54) |
$ u^{+}_p(t_c) = \sqrt{-\frac{p t_c}{\pi}}\left[c_1 K_{\nu_{-}}\left({\rm i} p t_c\right)+c_2 K_{\nu_{-}}\left(-{\rm i} p t_c\right)\right]\,,\; \; \; \; $ ![]() | (55) |
$ u^{-}_p(t_c) = \sqrt{-\frac{p t_c}{\pi}}\left[c_1 K_{\nu_{+}}\left({\rm i} p t_c\right)-c_2 K_{\nu_{+}}\left(-{\rm i} p t_c\right)\right]\,, $ ![]() | (56) |
$ |c_1|^2+|c_2|^2 = 1\,, $ ![]() | (57) |
Thus, we derive the general structure of the covariant Dirac field (43), for which the field operators
To completely determine our solutions we must adopt the criterion of frequency separation. The adiabatic vacuum may be defined simply by choosing
The solution is to adopt the r.f.v., imposing the conditions [23]
$ \mathop {\lim }\limits_{p \to 0} u_p^ - (t) = \mathop {\lim }\limits_{p \to 0} v_p^ + (t) = 0{\mkern 1mu} , $ ![]() | (58) |
$ c_1 = \frac{{\rm e}^{\pi\kappa}p^{-{\rm i}\kappa}}{\sqrt{1+{\rm e}^{2\pi \kappa}}}\,, \quad c_2 = \frac{{\rm i}\,p^{-{\rm i}\kappa}}{\sqrt{1+{\rm e}^{2\pi \kappa}}}\,, $ ![]() | (59) |
$ u_p^{\pm}(t_c) = \pm \frac{\sqrt{-\pi t_c}\, p^{\nu_-}}{\sqrt{1+{\rm e}^{2\pi\kappa}}}\, I_{\mp\nu_{\mp}}({\rm i}pt_c)\,. $ ![]() | (60) |
$ \mathop {\lim }\limits_{\vec p \to 0} {U_{\vec p,\sigma }}(t,\vec x) = \frac{{{{\rm e}^{ -{\rm i}mt}}}}{{{{[2\pi a(t)]}^{\frac{3}{2}}}}}\left( {\begin{array}{*{20}{c}}{{\xi _\sigma }}\\0\end{array}} \right){\mkern 1mu} , $ ![]() | (61) |
$ \mathop {\lim }\limits_{\vec p \to 0} {V_{\vec p,\sigma }}(t,\vec x) = \frac{{{{\rm e}^{{\rm i}mt}}}}{{{{[2\pi a(t)]}^{\frac{3}{2}}}}}\left( {\begin{array}{*{20}{c}}0\\{{\eta _\sigma }}\end{array}} \right){\mkern 1mu} , $ ![]() | (62) |
In contrast to the Klein-Gordon field, the r.f.v. of the Dirac field holds for any mass, separating the frequencies just as in special relativity. Nevertheless, its relation with the adiabatic vacuum is similar to that in the scalar case because the basis of the adiabatic vacuum is related to that of the r.f.v. through a non-trivial Bogolyubov transformation, the coefficients of which are proportional to the constants (59).
$ x = -p t_c = \frac{p}{\omega}{\rm e}^{-\omega t} $ ![]() | (63) |
2
A.Approximation method
We propose a generalization of the standard uniform asymptotic expansion (A8) to $ \begin{aligned}[b]J_{i\kappa+\lambda}(\kappa x) \simeq {\cal{J}}(\kappa,\lambda, x) =& \frac{{\rm e}^{\frac{\pi\kappa}{2}-\frac{{\rm i}\pi\lambda}{2}}}{\sqrt{2\kappa\pi}}\\ \; \; \; \; \;& \times\frac{{\rm e}^{{\rm i} \kappa\sqrt{1+x^2}-\frac{{\rm i}\pi}{4}}}{ (1+x^2)^{\frac{1}{4}}} \left(\frac{1}{x}+\sqrt{1+\frac{1}{x^2}}\right)^{-{\rm i}\kappa-\lambda}\,,\end{aligned} $ ![]() | (64) |
We start with the observation that, in our case, the variable (63) is positively defined without reaching the value

$ {\cal{E}}(\kappa,\lambda,x) = 1-\frac{{|{\cal{J}}|}(\kappa,\lambda,x)|}{|J_{i\kappa+\lambda}(\kappa x)|}\,, $ ![]() | (65) |

The conclusion is that our approximation is numerically satisfactory even for
$ J_{{\rm i}\kappa+\lambda}(x)\simeq\frac{{\rm e}^{\frac{\pi\kappa}{2}-\frac{{\rm i}\pi\lambda}{2}}}{\sqrt{2\pi}}\,\frac{{\rm e}^{{\rm i} \sqrt{\kappa^2+x^2}-\frac{{\rm i}\pi}{4}}}{ (\kappa^2+x^2)^{\frac{1}{4}}} \left(\frac{\kappa}{x}+\sqrt{1+\frac{\kappa^2}{x^2}}\right)^{-{\rm i}\kappa-\lambda}\,, $ ![]() | (66) |
2
B.Flat limit of the Klein-Gordon field in r.f.v.
Let us briefly analyze the flat limits, for $ {\cal{F}}_p(t) = {\rm e}^{{\rm i}\delta_{KG}(p)}\left(\frac{p}{2\omega}\right)^{-{\rm i}\kappa}\sqrt{\frac{\pi}{\omega}}\frac{{\rm e}^{\frac{1}{2}\pi\kappa}}{\sqrt{{\rm e}^{2\pi\kappa}-1}}\,J_{{\rm i}\kappa}\left(\frac{p}{\omega}{\rm e}^{-\omega t}\right)\,. $ ![]() | (67) |
$ {\cal{F}}_p(t_c)\simeq\rho(p,t){\rm e}^{{\rm i}\theta(p,t)}\,, $ ![]() | (68) |
$ \rho(p,t) = \frac{1}{\sqrt{2}(M^2+p^2{\rm e}^{-2\omega t})^{\frac{1}{4}}}\frac{{\rm e}^{\pi\kappa}}{\sqrt{{\rm e}^{2\pi\kappa}-1}}\,, $ ![]() | (69) |
$ \begin{aligned}[b]\theta(p,t) = &\delta_{KG}(p)-\frac{\pi}{4}-\frac{M}{\omega} \ln\left(\frac{1}{2\omega}\right)-Mt\\ &-\frac{M}{\omega}\ln\left(M+\sqrt{M^2 +p^2{\rm e}^{-2\omega t}}\right)\\ &+\frac{1}{\omega}\sqrt{M^2+p^2{\rm e}^{-2\omega t}}\,.\end{aligned}$ ![]() | (70) |
$ \delta_{KG}(p) = \frac{\pi}{4}+\frac{M}{\omega}\ln \frac{M+\sqrt{M^2+p^2}}{2\omega}-\frac{\sqrt{M^2+p^2}}{\omega}\,. $ ![]() | (71) |
$ {\rm e}^{{\rm i}\delta_{KG}(p)} = {\rm e}^{\frac{{\rm i} \pi}{4}}\left(\frac{M+\sqrt{M^2+p^2}}{2\omega} \right)^{\frac{iM}{\omega}}{\rm e}^{-{\rm i}\frac{\sqrt{M^2+p^2}}{\omega}}\,, $ ![]() | (72) |
$ {\cal{F}}_p(t)\simeq\frac{{\rm e}^{\pi\kappa}}{\sqrt{{\rm e}^{2\pi\kappa}-1}}\frac{{\rm e}^{{\rm i}\theta_{KG}(p,t)}}{\sqrt{2}(M^2+p^2{\rm e}^{-2\omega t})^{\frac{1}{4}}}\,. $ ![]() | (73) |
$ \begin{aligned}[b] \theta_{KG}(p,t) =& -Mt -\frac{M}{\omega}\ln\left(\frac{M+\sqrt{M^2+p^2{\rm e}^{-2\omega t}}}{M+\sqrt{M^2+p^2}}\right)\\ \;\;\;\;\;\;\;\;\;\;\;\;\;&+\frac{1}{\omega}\left(\sqrt{M^2+p^2{\rm e}^{-2\omega t}}-\sqrt{M^2+p^2}\right)\,, \end{aligned}$ ![]() | (74) |
$ \begin{aligned}[b] \theta_{KG}(p,t) =& -\sqrt{M^2+p^2}\,t+\frac{\omega p^2 t^2}{2\sqrt{M^2+p^2}}\\ &-\frac{\omega^2 p^2(2M^2+p^2)t^3}{6(M^2+p^2)^\frac{3}{2}}+{\cal{O}}(\omega^3)\,, \end{aligned}$ ![]() | (75) |
$ \mathop {\lim }\limits_{\omega \to 0} {\cal{F}}_p(t) = \frac{{\rm e}^{-{\rm i}E(p)t}}{\sqrt{2E(p)}}\,, \quad E(p) = \sqrt{m^2+p^2}\,, $ ![]() | (76) |
2
C.Flat limit of the Dirac field in r.f.v.
To analyze how the flat limit can be reached in the case of the Dirac field, it is convenient to rewrite the time modulation functions (60) in the chart $\begin{aligned}[b] u_p^{\pm}(t) =& \pm {\rm e}^{{\rm i}\delta_D(p)\pm \frac{{\rm i}\pi}{4}} p^{-{\rm i}\kappa}\\ \;\;\;\;\;\;\;\;\; &\times \frac{{\rm e}^{\frac{\pi\kappa}{2}}}{\sqrt{{\rm e}^{2\pi \kappa}+1}}\sqrt{\frac{\pi}{\omega}\, p {\rm e}^{-\omega t}}\, J_{{\rm i}\kappa\mp\frac{1}{2}}\left(\frac{p}{\omega}{\rm e}^{-\omega t}\right)\,, \end{aligned}$ ![]() | (77) |
$ u_p^{\pm}(t)\simeq\rho^{\pm}(p,t) {\rm e}^{{\rm i}\theta(p,t)}\,, $ ![]() | (78) |
$ \rho^+(p,t) = \frac{{\rm e}^{\pi\kappa}}{\sqrt{{\rm e}^{2\pi \kappa}+1}}\,\frac{\sqrt{\sqrt{m^2+p^2{\rm e}^{-2\omega t}}+m}}{\sqrt{2}(m^2+p^2 {\rm e}^{-2\omega t})^{\frac{1}{4}}}\,, $ ![]() | (79) |
$\begin{aligned}[b]\rho^-(p,t) =& \frac{{\rm e}^{\pi\kappa}}{\sqrt{{\rm e}^{2\pi \kappa}+1}}\\ \; \;& \times\frac{p {\rm e}^{-\omega t}}{\sqrt{2}(m^2+p^2 {\rm e}^{-2\omega t})^{\frac{1}{4}}\sqrt{\sqrt{m^2+p^2{\rm e}^{-2\omega t}}+m}}\,,\; \; \; \;\end{aligned} $ ![]() | (80) |
$ \begin{aligned}[b] \theta(p,t) =& \delta_D(p)+\frac{\pi}{4}-mt\\ &-\frac{m}{\omega}\ln\left(m+\sqrt{m^2+p^2{\rm e}^{-2\omega t}}\right)\\ &+\frac{1}{\omega}\sqrt{m^2+p^2{\rm e}^{-2\omega t}}\,. \end{aligned} $ ![]() | (81) |
$ p{\rm e}^{-\omega t} = \sqrt{\sqrt{m^2+p^2{\rm e}^{-2\omega t}}+m}\,\sqrt{\sqrt{m^2+p^2{\rm e}^{-2\omega t}}-m} $ ![]() | (82) |
$ \delta_D(p) = -\frac{\pi}{4}+\frac{m}{\omega}\ln(\sqrt{m^2+p^2}+m)-\frac{1}{\omega}\sqrt{m^2+p^2}\,, $ ![]() | (83) |
$ u_p^{\pm}(t)\simeq \frac{{\rm e}^{\frac{\pi m}{\omega}}}{\sqrt{{\rm e}^{2 \frac{\pi m}{\omega}}+1}}\frac{\sqrt{\sqrt{m^2+p^2{\rm e}^{-2\omega t}}\pm m}}{\sqrt{2}(m^2+p^2 {\rm e}^{-2\omega t})^{\frac{1}{4}}}\, {\rm e}^{{\rm i}\theta_D(p,t)}\,. $ ![]() | (84) |
$ \begin{aligned}[b] \theta(p,t) =& -mt-\frac{m}{\omega}\ln\left(\frac{m+\sqrt m^2+p^2e^{-2\omega t}}{m+\sqrt{m^2+p^2}}\right)\\ &+\frac{1}{\omega}(\sqrt{m^2+p^2e^{-2\omega t}}-\sqrt{m^2+p^2}),\end{aligned} $ ![]() | (85) |
$ \begin{aligned}[b] \theta_{D}(p,t) =& -\sqrt{m^2+p^2}\,t+\frac{\omega p^2 t^2}{2\sqrt{m^2+p^2}}\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;&-\frac{\omega^2 p^2(2m^2+p^2)t^3}{6(m^2+p^2)^\frac{3}{2}}+{\cal{O}}(\omega^3)\,, \end{aligned} $ ![]() | (86) |
Finally, we verify that, in the flat limit, we obtain the usual Minkowskian time modulation functions
$ \mathop {\lim }\limits_{\omega \to 0} u_p^{\pm}(t) = \sqrt{\frac{E(p)\pm m}{2 E(p)}}\, {\rm e}^{-{\rm i}E(p)t}\,. $ ![]() | (87) |
2
D.Physical consequences
Solving the problem of the flat limit, we derive suitable phases that complete the phases we introduced previously to define the r.f.v. We thus obtain the definitive form of the scalar time modulation functions, $ {\cal{F}}_p(t) = {\rm e}^{{\rm i}\alpha_{KG}(p)}\sqrt{\frac{\pi}{\omega}}\frac{{\rm e}^{\frac{1}{2}\pi\kappa}}{\sqrt{{\rm e}^{2\pi\kappa}-1}}\,J_{i\kappa}\left(\frac{p}{\omega}{\rm e}^{-\omega t}\right)\,, $ ![]() | (88) |
$ \begin{aligned}[b] \alpha_{KG}(p) =& \delta_{KG}(p)-\kappa \ln\left(\frac{p}{2\omega}\right) = \frac{\pi}{4}\\ \;\;\;\;\;\;\;\;\;\;\;&+\frac{M}{\omega}\ln \frac{M+\sqrt{M^2+p^2}}{p}-\frac{\sqrt{M^2+p^2}}{\omega}\,,\end{aligned} $ ![]() | (89) |
For the Dirac time modulation functions, we may write a similar result,
$ u_p^{\pm}(t) = \pm {\rm e}^{{\rm i}\alpha_D(p)\pm \frac{i\pi}{4}} \frac{{\rm e}^{\frac{\pi\kappa}{2}}}{\sqrt{{\rm e}^{2\pi \kappa}+1}}\sqrt{\frac{\pi}{\omega}\, p {\rm e}^{-\omega t}}\, J_{{\rm i}\kappa\mp\frac{1}{2}}\left(\frac{p}{\omega}{\rm e}^{-\omega t}\right)\,, $ ![]() | (90) |
$ \begin{aligned}[b]\alpha_D(p) = &\delta_{D}(p)-\kappa \ln p \\ = & -\frac{\pi}{4}+\frac{m}{\omega}\ln\left(\frac{\sqrt{m^2+p^2}+m}{p}\right)-\frac{\sqrt{m^2+p^2}}{\omega}\,, \end{aligned} $ ![]() | (91) |
These phases are so important in the de Sitter space-time because they do not affect the scalar products or contribute to the expressions of the transition probabilities. A specific feature of the de Sitter QFT is that the forms of some one-particle operators, including the energy one, are strongly dependent on the phases, which are functions of p. We remind the reader that, after canonical quantization, the one-particle energy operator is (19). In the case of the Klein-Gordon field, we consider the normalized mode functions (27) with the time modulation functions (88). Then, it is not difficult to verify the identity
$ (Hf_{\vec{p}}) = \left[-i\omega \left(p^i\partial_{p_i}+{\frac{3}{2}}\right)-\omega p^i\partial_{p^i}\alpha_{KG}(p)\right]f_{\vec{p}} , $ ![]() | (92) |
$ \begin{aligned}[b] {\cal{H}}_{KG} =& :\langle \Phi, H\Phi\rangle_{KG}: \\ \;\;\;\;\;\;\;\;= &\int {\rm d}^3 p\, \sqrt{M^2+p^2} \left[a^{\dagger}(\vec{p}) a(\vec{p}) +{b}^{\dagger}(\vec{p}){b}(\vec{p})\right]\\ \;\;\;\;\;\;\;\;&+\frac{i\omega}{2}\int {\rm d}^3p\, p^i \left\{ \left[\, a^{\dagger}(\vec{p})\stackrel{\leftrightarrow}{\partial}_{p_i} a(\vec{p})\right]\right.\\ & \left.+ \left[\, b^{\dagger}(\vec{p}) \stackrel{\leftrightarrow}{\partial}_{p_i} b(\vec{p})\right]\right\}\,, \end{aligned}$ ![]() | (93) |
$ \omega p^i\partial_{p^i}\alpha_{KG}(p) = -\sqrt{M^2+p^2}\,. $ ![]() | (94) |
$ \mathop {\lim }\limits_{\omega \to 0} {\cal{H}}_{KG} = \int {\rm d}^3 p\, \sqrt{m^2+p^2} \left[a^{\dagger}(\vec{p}) a(\vec{p}) +{b}^{\dagger}(\vec{p}){b}(\vec{p})\right]\,, $ ![]() | (95) |
Similar results concerning the phase factors or mode expansions as the second term of Eq. (93) were obtained for the scalar and spinor fields previously in Ref. [2], in which the adiabatic vacuum was considered. Subsequent studies refined these results [4, 7, 18] such that we can now conclude that the regularized phases derived so far are very similar to those obtained here in the r.f.v. This is because in the adiabatic vacuum, where the rest limits are undefined, before performing the flat limit, one must force the rest limit to change ad hoc the form of the time modulation functions to introduce phase factors proportional to
Apart from the regularized phases recovered here, we report important new results such as the final forms of the time modulation functions (88) and (90) in the r.f.v. and the approximations (73) and (84) that can be used in applications for deriving transition amplitudes between states, defined in this vacuum instead of the adiabatic one that has been considered previously [41-50].
However, the principal new result is that, in the r.f.v., the flat limit occurs naturally without the forced artifices used in the case of the adiabatic vacuum we presented above. In our opinion, this result is a crucial argument in favor of the hypothesis that the r.f.v. could be the principal candidate for applying Feynman rules in the de Sitter expanding universe.
$\begin{aligned}[b] K_{\nu}(z) = K_{-\nu}(z) = \frac{\pi}{2}\frac{I_{-\nu}(z)-I_{\nu}(z)}{\sin\pi \nu}\,, \end{aligned} \tag{A1} $ ![]() | (A1) |
$ \begin{aligned}[b] I_{\pm\nu}(z) &= {\rm e}^{\mp {\rm i}\pi\nu}I_{\pm\nu}(-z) \\ \;\;\;\;\;\;\;\;\;\;&= \frac{{\rm i}}{\pi}\left[K_{\nu}(-z)-{\rm e}^{\mp {\rm i}\pi\nu}K_{\nu}(z)\right]\,. \end{aligned} \tag{A2} $ ![]() | (A2) |
$ \begin{aligned}[b] {\rm i} I_{{\rm i}\kappa}({\rm i} s) \stackrel{\leftrightarrow}{\partial_{s}}I_{-{\rm i}\kappa}({\rm i}s) = \frac{2\, {\rm{sinh}}\,\pi\kappa}{\pi s}\,, \end{aligned}\tag{A3} $ ![]() | (A3) |
$ \begin{aligned}[b] {\rm i} K_{\nu}(-{\rm i} s) \stackrel{\leftrightarrow}{\partial_{s}}K_{\nu}({\rm i}s) = \frac{\pi}{|s|}\,, \end{aligned}\tag{A4} $ ![]() | (A4) |
For
$ \begin{aligned}[b] I_{\nu}(z) \to \sqrt{\frac{\pi}{2z}}{\rm e}^{z}\,, \quad K_{\nu}(z) \to K_{\frac{1}{2}}(z) = \sqrt{\frac{\pi}{2z}}{\rm e}^{-z}\,. \end{aligned}\tag{A5} $ ![]() | (A5) |
$ \begin{aligned}[b] I_{\nu}(z)\sim \frac{1}{\Gamma(\nu+1)} \left(\frac{z}{2}\right)^{\nu}\,, \end{aligned}\tag{A6} $ ![]() | (A6) |
The modified Bessel functions I are related to the usual ones as
$ \begin{aligned}[b] I_{\nu}(-ix) = {\rm e}^{-\frac{i\pi\nu}{2}}J_{\nu}\left(x\right)\,,\quad \forall\, x\in{\mathbb{R}},\, \nu\in{{\mathbb{C}}}\,, \end{aligned}\tag{A7}$ ![]() | (A7) |
$ \begin{aligned}[b] J_{i\kappa}(\kappa x) =& \frac{{\rm e}^{\frac{\pi\kappa}{2}}}{\sqrt{2\kappa}\,\pi}\,\frac{{\rm e}^{{\rm i} \kappa\sqrt{1+x^2}-\frac{{\rm i}\pi}{4}}}{ (1+x^2)^{\frac{1}{4}}} \left(\frac{1}{x}+\sqrt{1+\frac{1}{x^2}}\right)^{-{\rm i}\kappa}\\\; \; \; \; \; \;& \times\left[\sum\limits_{n = 0}^{n = N} \frac{(2 i)^n\Gamma(n+\frac{1}{2})}{\kappa^n (1+x^2)^{\frac{n}{2}}}\,a_n (x)+ {\cal{O}}(\kappa^{-N-1})\right]\,,\end{aligned} \tag{A8} $ ![]() | (A8) |
$\begin{aligned}[b] a_0(x) =& 1\,, \\ a_1(x)=& -\frac{1}{8}+\frac{5}{24}(1+x^2)^{-1}\,, \\a_2(x) =& \frac{3}{128}-\frac{77}{576}(1+x^2)^{-1}\\&+\frac{385}{3456}(1+x^2)^{-2},... \end{aligned}$ ![]() |