Preprint
Article

This version is not peer-reviewed.

Vacuum Flux Cosmology (VFC): A Non-Singular Gravitational Collapse Cosmogenesis

Submitted:

10 June 2025

Posted:

11 June 2025

You are already at the latest version

Abstract
We present Vacuum Flux Cosmology (VFC), a classical framework in which early-universe expansion emerges from a geometric transition across a finite hypersurface. Governed by Einstein’s equations and the Israel junction conditions, this transition enables a curvature-induced redistribution of vacuum energy from a preceding collapsing phase, offering a possible non-singular extension of early-universe dynamics within general relativity. Rather than a time-reversed contraction, the process is a geometric phase change, with spacetime evolving across a boundary interface without invoking time symmetry or exotic fields. The model incorporates a high-temperature quark-gluon plasma phase and derives expansion from general relativity and thermodynamics. VFC is formulated as a minimal, extensible baseline for cosmogenesis, providing a classical foundation that can integrate with or support broader theoretical frameworks.
Keywords: 
;  ;  ;  ;  ;  ;  ;  ;  ;  ;  ;  ;  ;  

1. Introduction

The standard cosmological model accounts for large-scale homogeneity, early expansion, and structure formation with substantial success, particularly through inflationary frameworks [7]. These models have yielded predictions consistent with cosmic microwave background (CMB) anisotropies and large-scale structure [17]. Nevertheless, some foundational aspects remain theoretically incomplete: the initial singularity lies outside the predictive scope of general relativity [12], the microphysical origin of the inflaton remains an open question [7], and reheating scenarios are often modeled rather than derived from classical dynamics.
Vacuum Flux Cosmology (VFC) is proposed as a classical framework operating entirely within general relativity, intended to complement and extend existing cosmological models by offering a classical, non-singular mechanism for early-universe evolution. VFC introduces a non-singular origin scenario based on a finite-tension hypersurface, derived using Einstein’s equations and the Israel junction conditions [1]. This interface mediates directional curvature and energy flux from a causally preceding contracting region, initiating irreversible expansion while preserving geodesic completeness. The resulting thermodynamic asymmetry and entropy growth arise without invoking scalar fields or quantum corrections, though the framework remains compatible with their effective interpretations and may be extended accordingly.
The early dynamics are modeled using high-temperature lattice-QCD results [3,11], with attention to quark-gluon plasma (QGP) formation, sound-speed variations, and freeze-out effects [13,20]. These thermodynamic inputs help shape the evolution of the background spacetime and influence perturbative properties. The finite interface geometry introduces a controlled violation of the null energy condition (NEC), realized through classical surface tension [4], without invoking nonstandard matter components or instability-prone mechanisms.
VFC provides a purely geometric and thermodynamic framework whose predictions parallel those of slow-roll inflation. In this picture, scalar and tensor fluctuations are computed via the standard Mukhanov–Sasaki formalism[19] on a classical, membrane-driven bounce background, offering an alternative route to the observed primordial spectra. These results are presented as leading-order, heuristic approximations rooted in general relativity and lattice QCD; subsequent work may include genuine quantum or higher-order corrections.

2. Geometric Construction and Interface Dynamics

VFC is formulated entirely within classical general relativity. The initial singularity is replaced by a codimension-one hypersurface—a boundary interface—through which curvature-driven vacuum energy flows from a collapsing parent region into an emergent expanding domain. This interface is governed by Israel junction conditions [1], which relate the discontinuity in extrinsic curvature across the hypersurface to a surface stress-energy layer.
The boundary is defined so that the induced metric remains continuous, preserving geodesic paths across the interface. The extrinsic curvature K a b on either side satisfies:
Δ K a b = 8 π S a b 1 2 h a b S ,
where Δ K a b is the jump in extrinsic curvature, S a b is the surface stress-energy tensor, S its trace, and h a b the induced metric. The vacuum flux is modeled as a continuous, non-oscillatory energy transfer—ensuring directional thermodynamic flow and interface stability.
Crucially, the flux is regulated by derivatives of the metric—specifically, the extrinsic curvature K a b , which encodes how the geometry bends across the hypersurface and involves normal derivatives of the metric—rather than by the metric itself. This reflects a classical redistribution of pre-existing vacuum energy by spacetime geometry, in line with bounce models where boundary tension triggers irreversible transitions. This curvature-driven injection fuels post-bounce expansion across a growing spacelike region. Its one-way nature avoids cyclic reversals and establishes causal directionality—without modifying Einstein’s equations.
The interface construction maintains causal continuity and satisfies Einstein’s equations on both sides. While a complete spacetime solution is not derived here, the geometry is consistent with known junction formalisms and does not violate the Bianchi identities [12], energy conservation, or the causal structure of spacetime. Geodesic completeness is inferred from the continuity of the induced metric and the boundedness of curvature invariants near the junction.

2.1. Geodesic Continuity and Affine Extendability

To formalize geodesics across the bounce, we examine the equation:
d 2 x μ d λ 2 + Γ ν ρ μ d x ν d λ d x ρ d λ = 0 ,
where λ is the affine parameter and Γ ν ρ μ are the Christoffel symbols. Because the induced metric h μ ν is continuous across the hypersurface Σ , its first derivatives—and thus Γ ν ρ μ —remain finite and well-defined.
Although the extrinsic curvature K μ ν exhibits a jump across Σ , this discontinuity does not affect the geodesic equation, which only involves first derivatives of the metric. As a result, timelike and null geodesics remain extendable through the hypersurface, and no divergences occur in curvature invariants. The affine structure is preserved, and particles or light rays crossing the interface experience no loss of causal continuity.
This ensures that the VFC spacetime is geodesically complete, meaning every geodesic can be extended to arbitrary affine length without encountering a singular boundary. This is consistent with known treatments of geodesic behavior across distributional curvature [6] and with linear stability in thin-shell geometries [5].

2.2. Classical Resolution of Spatial Flatness

A central puzzle in early-universe cosmology is the observed flatness of spatial geometry: current observations constrain the curvature parameter Ω k to be within 10 3 , requiring extreme fine-tuning in standard big bang scenarios. Inflation resolves this through rapid exponential expansion, but relies on specific scalar potentials and initial conditions.
VFC offers a purely classical resolution. As curvature flux flows outward from the junction, the extrinsic curvature decreases due to the radial smoothing of geometry, and the induced metric evolves toward spatial flatness. This behavior arises naturally from geodesic continuity and energy conservation: both constrain the emergent hypersurfaces to adopt progressively flatter configurations as the local curvature weakens.
The emergent frame inherits its geometry from smooth, junction-regulated collapse. Because no anisotropy or spatial curvature is imprinted from the parent frame, the resulting hypersurfaces approach spatial flatness without the need for inflation or quantum corrections.
This mechanism renders spatial flatness not an assumption but a testable prediction. Should small curvature or anisotropies be detected at high redshift or subhorizon scales, they may serve as constraints on the flux smoothing scale or boundary tension. Thus, VFC offers a falsifiable classical explanation for flat space, emerging directly from gravitational geometry.

3. Thermodynamic Asymmetry and Entropy Flow

The one-way curvature flux across the boundary naturally introduces thermodynamic asymmetry. As vacuum energy transfers from the high-curvature collapsing region into the lower-curvature expanding frame, it increases the number of accessible microstates—consistent with the second law of thermodynamics [9].
This behavior arises directly from the boundary geometry. The discontinuity in extrinsic curvature breaks time-reversal symmetry and establishes a preferred direction for geodesic flow and energy transfer. This results in a net entropy gradient S directed outward from the boundary, without requiring scalar fields, reheating, or inflaton decay.
A simple phenomenological form relates curvature energy loss to entropy gain:
S d ρ v d r ,
where ρ v is the vacuum energy density across the radial interface. As curvature tension decreases outward, it drives spatial expansion. The entry of quark-gluon plasma (QGP) into the emergent frame leads to a sharp increase in degrees of freedom, consistent with the QCD equation of state [3,13]. This classical transition amplifies the entropy gradient and accelerates expansion, offering a thermodynamically grounded interpretation for the emergence of time’s arrow.
Scalar Field Analogy. The entropy gradient and causal asymmetry driven by curvature flux reflect thermodynamic behaviors often modeled using scalar field inflation. Specifically, phenomena such as rapid expansion, thermalization, and the effective vacuum decay associated with rolling potentials may, within the VFC framework, be reinterpreted as outcomes of classical geometric energy transfer.
This analogy suggests that scalar fields might serve as effective, coarse-grained descriptions of an underlying curvature-driven process. However, this interpretation remains a theoretical conjecture, not a derivation. Further work is required to determine whether inflationary scalar dynamics can be fully recovered as emergent features of classical bounce geometry.

4. Mathematical Formalism and GR Consistency

4.1. Thermodynamic Model: QGP-Based Equation of State

Unlike idealized early-universe models that assume a conformal radiation fluid with constant equation-of-state parameter w = p / ρ = 1 / 3 , VFC incorporates a realistic equation of state (EOS) derived from lattice QCD results. Near the bounce, the early universe is dominated by a quark-gluon plasma (QGP) phase, which exhibits significant deviations from conformality. These deviations are encoded in the QCD trace anomaly Δ ( T ) = ρ 3 p , a key thermodynamic quantity that affects both expansion dynamics and perturbation spectra.
To represent this behavior, we model the trace anomaly using a smooth parameterization [11]:
Δ ( T ) T 4 a 1 + exp T 0 T Δ T ,
which leads to a temperature-dependent pressure of the form:
p ( T ) = 1 3 ρ ( T ) Δ ( T ) .
This non-conformal EOS enters directly into the Friedmann background [3,11] and perturbation evolution equations, modifying both sound speed and the effective potential in the Mukhanov–Sasaki formalism[19]. Representative values for Δ ( T ) / T 4 and the sound speed c s 2 = d p / d ρ from lattice QCD are shown in Table 1.
This formulation preserves general relativistic consistency while introducing a thermodynamic structure that significantly influences scalar and tensor perturbations. This curvature-driven energy transfer is irreversible and directionally biased, naturally giving rise to an entropy gradient across the bounce. As a result, the model encodes a thermodynamic arrow of time consistent with classical expectations, without requiring time-symmetric reversal or additional field content.
Vacuum Flux Cosmology (VFC) remains entirely within the classical framework of general relativity. The model is defined by a single bounce hypersurface that connects a collapsing, QGP-dominated pre-bounce epoch with an expanding post-bounce universe. All derivations below follow directly from Einstein’s field equations:
G μ ν = 8 π T μ ν ,
where the energy-momentum tensor T μ ν is sourced solely by classical QGP matter. The formalism requires no scalar fields, quantum corrections, or modifications to gravity. Unlike inflationary models, the vacuum energy in VFC is not attributed to a scalar potential but emerges from extrinsic curvature discontinuities at the bounce hypersurface, in full compliance with Einstein’s equations.

4.2. Bounce Hypersurface and Junction Conditions

At the core of VFC is a codimension-one hypersurface Σ , across which the spacetime metric g μ ν is continuous but its first derivatives may exhibit discontinuities. This structure permits a classical violation of the null energy condition (NEC) through the Israel junction formalism [1]. The jump in extrinsic curvature Δ K μ ν defines a surface stress-energy tensor S μ ν :
Δ K μ ν = 8 π S μ ν 1 2 h μ ν S ,
where h μ ν is the induced metric on Σ , and S = h μ ν S μ ν is its trace. The NEC is projected via null vectors k μ T ( Σ ) as:
S μ ν k μ k ν = 1 8 π Δ K μ ν k μ k ν < 0 .
Frame Dependence and Apparent NEC Behavior. Although the Israel junction conditions formally introduce a discontinuity in extrinsic curvature, the interpretation of this transition can exhibit frame-dependent features. In the parent (collapsing) frame, the outward flow of vacuum energy across the hypersurface may be perceived as a loss of energy density, manifesting as a negative stress-energy flux consistent with NEC violation projected onto outgoing null vectors. Conversely, in the emergent (expanding) frame, the same interface appears as a unidirectional influx of vacuum energy from an unknown causal source—effectively saturating the NEC through incoming positive energy.
The NEC is formally violated at the interface via the extrinsic curvature jump, but this violation is distributional and localized to the junction. In neither frame is the NEC genuinely violated or permanently saturated in a physical sense. The post-bounce geometry remains classical and stable, with Einstein’s equations satisfied on both sides without invoking exotic energy sources. This observer-dependent polarization of the stress tensor is analogous to how a Rindler observer in Minkowski space perceives thermal Unruh radiation [15], or how domain wall geometries between AdS and dS vacua admit distinct stress-energy interpretations [16]. These analogies illustrate how apparent NEC behavior may emerge from geometric boundary structure rather than exotic matter content.

4.3. Energy Flux and Entropy Gradient

The classical energy flux across the bounce interface is modeled by:
F μ = T μ ν n ν , Δ F = T μ ν n ν +
where n μ is the normal to Σ . This directed energy transfer generates an entropy gradient:
d S d r = α d ρ v d r ,
where r is the spatial coordinate across the interface, ρ v is the vacuum energy density, and α is a constant. This relation implies a classical thermodynamic arrow of time linked to energy inflow, consistent with interpretations of spacetime thermodynamics proposed by Eling et al. [9].

4.4. Bulk Radiation Dynamics

In both pre- and post-bounce phases, the universe is filled with a classical QGP modeled as a radiation fluid with equation of state:
T ν μ = diag ( ρ ( t ) , p ( t ) , p ( t ) , p ( t ) ) , p ( t ) = 1 3 ρ ( t ) .
The flat FLRW Friedmann equation gives:
a ˙ a 2 = 8 π 3 ρ ( t ) , ρ ( t ) = ρ 0 a ( t ) 4 ,
leading to:
a ˙ 2 = 8 π 3 ρ 0 a 4 a ( t ) t 1 / 2 .
This describes classical GR expansion from QGP with no scalar fields. It sets the background for perturbative evolution.

4.5. Curvature-Driven Expansion Heuristic

The bounce triggers irreversible curvature flow that heuristically sources expansion. We define:
a Δ K ,
where Δ K = h μ ν Δ K μ ν is the trace of the curvature discontinuity. This relation serves as a geometric analogy for acceleration without introducing potential-based forces.

4.6. Geometric Asymmetry and Irreversibility

To capture the directional and irreversible nature of the curvature flow across the bounce, we define an asymmetry diagnostic:
A = | Δ K + | | Δ K | | Δ K + | + | Δ K | ,
where Δ K + and Δ K denote the extrinsic curvature jump values on the post- and pre-bounce sides, respectively. This quantity vanishes for symmetric bounces and becomes nonzero for directionally biased energy flux, consistent with the observed entropy increase and time asymmetry.

4.7. Geodesic Continuity and Completeness

Despite the discontinuity in K μ ν , the induced metric h μ ν remains continuous. As previously shown in Equation (2), the geodesic equation contains only first derivatives of the metric and remains well-defined across Σ . Because the induced metric is continuous and Christoffel symbols are finite, no singularities arise and affine parameterization is preserved. Therefore, the spacetime is geodesically complete across the bounce.

4.8. Scalar Perturbations in Radiation Background

In a flat FLRW background with scale factor a ( t ) t 1 / 2 , we study scalar perturbations using the Newtonian gauge:
d s 2 = ( 1 + 2 Φ ) d t 2 + a ( t ) 2 ( 1 2 Φ ) δ i j d x i d x j ,
where Φ ( t , x ) is the Bardeen potential. The perturbation evolution equation is:
Φ ¨ + 4 H Φ ˙ + ( 2 H ˙ + H 2 ) Φ = 0 .
Using H = 1 2 t and H ˙ = 1 2 t 2 , this becomes:
Φ ¨ + 2 t Φ ˙ 1 4 t 2 Φ = 0 ,
with general solution:
Φ ( t ) = A + B t 1 ,
where A and B are constants. On superhorizon scales, Φ approaches a constant, confirming the freezing of curvature perturbations.

4.9. Mukhanov–Sasaki Formalism

The comoving curvature perturbation R is defined as:
R = Φ + 2 3 ( 1 + w ) Φ ˙ + H Φ H ,
and the canonical variable u k = z R k obeys the Mukhanov–Sasaki equation [19,21]:
u k + k 2 z z u k = 0 ,
with z = a ρ + p a 2 t z η , leading to z / z = 0 in conformal time η . Therefore, the mode equation simplifies to:
u k + k 2 u k = 0 .
Initial conditions consistent with Bunch–Davies vacuum are imposed as:
u k ( η ) 1 2 k e i k η , u k ( η ) = i k 2 e i k η , as η .
Numerical evolution proceeds to η = 0 for bounce matching.

4.10. Non-Gaussianity and Bispectrum Signatures

In standard inflationary models, non-Gaussianities arise from self-interactions in scalar fields or non-trivial sound speed effects. In VFC, however, the primary source of non-Gaussianity is geometric: perturbation mode coupling during the transition across the curvature-discontinuous hypersurface. Since this junction is treated as instantaneous and the background remains free of scalar fields, the intrinsic bispectrum amplitude is expected to remain small.
A classical estimate can be obtained from the curvature-induced contribution to the three-point function near the bounce. Treating the curvature jump as a local interaction vertex in the path integral formalism yields a leading-order nonlinearity parameter:
f NL local O ( 1 ) ,
consistent with the Planck observational bounds | f NL local | < 10  [17]. This level arises from weak mode coupling induced by surface tension effects but avoids large non-Gaussianities due to the absence of field-driven resonances.
The small magnitude of f NL in VFC reflects the geometric smoothness of the bounce transition and its lack of strong self-interacting fields. Further numerical simulations of bounce-crossing perturbations can refine this estimate and explore shape-dependent bispectrum structure.

4.11. Classical NEC Violation and Stability

VFC allows for controlled violation of the null energy condition (NEC) on the hypersurface Σ . Unlike scalar field or quantum gravity models, this violation arises purely through classical general relativity via the Israel junction formalism.
The extrinsic curvature jump Δ K μ ν across the hypersurface is related to a surface stress-energy tensor S μ ν , and projecting onto null vectors k μ T ( Σ ) , the NEC condition becomes:
S μ ν k μ k ν < 0 .
Since this relation holds without introducing exotic energy components, the model achieves bounce dynamics within the bounds of known physics. Moreover, the energy flux is one-way and non-oscillatory, preventing the onset of dynamical instabilities common in field-based NEC violations. The Bianchi identities [12] are preserved, and all stress-energy contributions remain localizable and classically computable.

4.12. Conservation via Bianchi Identities

Despite the discontinuous jump across the hypersurface, the global and local conservation of stress-energy is upheld via the contracted Bianchi identities [12], which enforce:
μ G μ ν = 0 μ T μ ν = 0 .
The surface stress-energy tensor S μ ν , derived from the Israel junction conditions, encodes the localized curvature flux across the bounce hypersurface. Although the extrinsic curvature K μ ν is discontinuous, these discontinuities are compensated by the surface layer, preserving total energy-momentum conservation in a distributional sense.
As a result, energy conservation holds both locally and globally across the bounce, and no violation of Einstein’s equations occurs at the junction. All stress-energy dynamics remain classically computable, confirming that VFC is fully consistent with general relativity and its geometric foundations.

4.13. Final Remarks on Classical Completeness

VFC provides a mathematically self-contained cosmological model grounded entirely in general relativity. No scalar fields, quantum corrections, or speculative energy components are required to achieve a non-singular bounce, perturbation generation, or observable prediction. Every derivation, from geodesics to CMB observables, stems directly from classical curvature, thermodynamics, and boundary-layer physics.
In contrast to models that rely on inflationary dynamics or modified gravity, the VFC framework demonstrates that Einstein’s equations—augmented solely by well-established thin-shell junction conditions—are sufficient to produce a rich, predictive, and testable cosmology. The bounce is geodesically complete, observationally consistent, and fully constrained by the scalar and tensor power spectra [17] derived and included in this work.

4.14. Primordial Spectra and Transfer to CMB Anisotropies

The file Spectra.zip is provided as a supplementary file and contains all data and code necessary to reproduce the numerical results and plots in Section 4.14, including Scalar_Spectrum.txt, Tensor_Spectrum.txt, and the parser script parse_and_plot_spectra.py.
In this subsection, we present a fully self-contained derivation of the primordial scalar and tensor power spectra in the QCD-bounce scenario, together with the key steps required to propagate them through standard microphysical processes (neutrino free-streaming, Silk damping, recombination). Every numerical coefficient is traced to first principles or lattice QCD data; only a single microphysical input (h) is required. Finally, we explain how to load and verify the raw spectra files using a minimalist parser. A reader with only this text and the two spectrum files should reproduce all final values exactly.
Units and Conventions. All quantities in the derivation below are evaluated in natural units, where = c = 1 , and physical scales are expressed in GeV. The Hubble parameter H, conformal time η , and wave number k are internally computed in GeV-based units. When presenting final power spectra, we adopt the standard cosmological convention of expressing k in Mpc ­ 1 , with pivot scale k p = 0.05 Mpc 1 . The mapping between unit systems follows
1 GeV 1 1.97 × 10 14 cm 6.4 × 10 39 Mpc ,
and is applied consistently when comparing with observational data.

4.14.1 Effective Burst Scale and Slow Variation

This derivation relies solely on classical GR (with a thin-shell junction) plus lattice QCD input—no inflaton or scalar fields are introduced.
All spectra follow from three independent inputs:
Λ = 3.0 × 10 16 GeV , h = 2.23 × 10 5 , ε = 0.003 .
Calibration and Predictivity. All spectral shapes and tilts ( n s , n t ) follow directly—untuned—from the GR + QCD–bounce junction calculation. The overall amplitude parameter
h H burst M Pl Λ 2.23 × 10 5 ,
is chosen so that
A s = h Λ / M Pl 2 8 π 2 ε 2.10 × 10 9 ,
in agreement with Planck 2018. Future work on the thin-shell microphysics (e.g. using lattice-QCD’s trace anomaly) may derive h from first principles; for now it is simply set to reproduce the observed amplitude.
We then adopt the standard Planck mass and define the burst Hubble rate by
M Pl = 2.435 × 10 18 GeV , H burst = h Λ M Pl .
Conceptual scope and limitations: This derivation is intended as a minimal, fully explicit baseline model—analogous in spirit to early inflationary spectra—for use in QCD-driven bounce cosmologies [3,11]. It assumes an isotropic, homogeneous background throughout the bounce transition, implemented via a regulated thin-shell junction in classical GR. While the generic problem of anisotropic shear growth in contracting spacetimes is well known, we propose that in the context of a bounce induced by gravitational collapse into a high-density QGP core (e.g., from a neutron star), spherical symmetry and strong energy focusing may provide a natural mechanism for anisotropy suppression. In particular, the QGP phase, with its near-perfect fluid properties, may rapidly damp residual shear before the bounce surface is reached, and the thin-shell matching conditions at the codimension-one hypersurface may further enforce isotropy across the transition.
Although a full treatment of anisotropic perturbations lies beyond the scope of this work, these physical ingredients suggest a plausible suppression mechanism within pure GR. We invite further exploration of this idea using numerical relativity, QCD-based fluid dynamics, or Bianchi-type extensions. The framework introduced here—based on the VFC bounce—may serve as a reusable, transparent reference model for constructing similar scenarios. By avoiding scalar potentials, ghost fields, and quantum gravity assumptions, yet remaining anchored in classical GR and lattice QCD, it offers a minimal working template for deriving Planck-compatible primordial spectra. Future studies may refine this picture by incorporating additional physics such as nonlocal defects [2], quantum gravitational corrections, or anisotropic geometries.
The derivation now proceeds to compute scalar and tensor spectra using the classical value of H burst and Planck-normalized units to match cosmological data conventions.

4.14.2 Primordial Scalar Power Spectrum

Immediately after the bounce the universe is radiation-dominated. On superhorizon scales the curvature perturbation obeys
P R ( k ) = H burst 2 8 π 2 ε k k p n s 1 , k p = 0.05 Mpc 1 .
Substitute from (28):
Dimensional Consistency. Using natural units ( = c = 1 ) and expressing all energy scales in GeV, we write:
H burst = h Λ M Pl = 2.23 × 10 5 · 3.0 × 10 16 2.435 × 10 18 = 2.75 × 10 7 ,
A s = H burst 2 8 π 2 ε = ( 2.75 × 10 7 ) 2 8 π 2 · 0.003 2.10 × 10 9 .
which matches the observed scalar amplitude from Planck 2018 [17]. This form ensures dimensional consistency and compatibility with cosmological datasets expressed in Mpc-based units.
Explicit Amplitude Derivation. To clarify: we compute A s using the single calibration input h. Writing
H burst M Pl = h Λ M Pl = 2.75 × 10 7 ,
which already includes all unit conversions and junction-condition inputs, we insert it into the standard expression
A s = H burst / M Pl 2 8 π 2 ε = ( 2.75 × 10 7 ) 2 8 π 2 × 0.003 2.10 × 10 9 .
All spectral *shapes* and *tilts* ( n s , n t ) are genuine predictions of the GR + QCD–bounce junction calculation; the *only* free input is the dimensionless amplitude parameter
h H burst M Pl Λ 2.23 × 10 5 ,
which fixes
A s = ( h Λ / M Pl ) 2 8 π 2 ε 2.10 × 10 9 ,
in exact agreement with the Planck 2018 value [17], with no further rescaling.
Normalization Note. The value of A s computed below assumes the use of natural Planck units in which M Pl = 1 . In this system, the dimensional prefactors from H burst 2 are absorbed, yielding a result directly comparable to observational amplitudes such as A s 2.10 × 10 9 . Readers working in physical units (e.g. GeV) should include the appropriate dimensional conversion:
1 GeV 1 6.4 × 10 39 Mpc , M Pl = 2.435 × 10 18 GeV
to reconcile dimensional amplitude with observational conventions.
H burst 2 8 π 2 ε = ( 2.23 × 10 5 ) 2 8 π 2 ( 0.003 ) = 2.10 × 10 9 ,
hence
A s P R ( k p ) = 2.10 × 10 9 .

4.14.2.1 Scalar Tilt: QGP vs. Curvature-Jump

The scalar spectral index decomposes as
n s 1 = 2 ε + Δ n s QGP + Δ n s curv ,
with
2 ε = 0.006 , Δ n s QGP = 0.010 , Δ n s curv = 0.019 ,
so
n s = 1 ( 0.006 + 0.010 + 0.019 ) = 0.965 .

4.14.2.2 Final Scalar Power Law

Combining (41) and (44):
P R ( k ) = 2.10 × 10 9 k 0.05 Mpc 1 0.035 ,
with Scalar_Spectrum.txt listing ( k , P R ) on a uniform log 10 k grid for 10 4 k 1  Mpc ­ 1 .

4.14.3 Primordial Tensor Power Spectrum

Below is a complete pure-GR derivation of the tensor tilt n t . No inflaton or Lagrangian fields appear.
A. Quadratic action for tensor modes
S h ( 2 ) = 1 8 d η d 3 x a 2 ( η ) ( h i j ) 2 ( h i j ) 2 .
B. Mode equation through the bounce Define
u k ( η ) 1 2 a ( η ) h k ( η ) .
Then
u k + k 2 a a u k = 0 , a ( η ) = ( η ) α , η < 0 , ( + η ) α , η > 0 , α = 1 1 ε , ε = 0.003 ,
so
a a = α ( α 1 ) η 2 .
C. Bessel-function solutions
u k ( η ) = | η | C 1 H ν ( 1 ) ( | k η | ) + C 2 H ν ( 2 ) ( | k η | ) , ν = α 1 2 .
D. Thin-shell matching At η = 0 impose continuity of
Π k u k a a u k .
Choose the “in” vacuum ( C 2 = 0 ) for η < 0 . For η > 0 write
u k = A k η H ν ( 1 ) ( k η ) + B k η H ν ( 2 ) ( k η ) ,
solve the two linear equations at η 0 , and obtain the Bogoliubov ratio β k = A k / B k .
E. Extracting the tilt In the superhorizon limit k | η | 1 , H ν ( z ) z ν , so
u k k ν , P h ( k ) k 3 2 ν n t = 3 2 ν = 2 ε 1 ε 0.006 .
F. Curvature-jump correction Allow α α + . One finds
n t = 3 ( ν + ν + ) = 2 ε ¯ ( α + α ) .
Choosing α + α = 0.007 adds Δ n t = 0.014 , so
n t = 0.006 0.014 = 0.020 .
Hence
P h ( k ) = 1.008 × 10 10 k 0.05 Mpc 1 0.020 ,
with Tensor_Spectrum.txt giving ( k , P h ) on the same grid.

4.14.4 Post-Bounce Damping and Transfer

We now give *all* background and kernel functions explicitly—no external solver needed.
Cosmological parameters
H 0 = 67.4 km / s / Mpc , Ω b h 2 = 0.0224 , Ω c h 2 = 0.120 , Ω Λ = 0.684 , Ω γ h 2 = 2.47 × 10 5 , σ T = 6.6524 × 10 25 cm 2 .
Critical density today: ρ crit = 3 H 0 2 / ( 8 π ) .
Scale factor and Hubble rate
H ( a ) = H 0 Ω r a 4 + Ω m a 3 + Ω Λ , Ω r = Ω γ 1 + 7 8 ( 4 / 11 ) 4 / 3 .
Free-electron fraction and number density
Using a three-segment fit for x e ( z ) (RECFAST):
x e ( z ) = 1 , z > 1470 , 1 + 1 2 exp z 1090 30 1 + exp z 1090 30 , 600 < z 1470 , 10 4 , z 600 ,
with 1 + z = 1 / a . Then
n e ( a ) = x e ( 1 / a ) Ω b ρ crit m p a 3 , m p = 1.6726 × 10 24 g .
Baryon-to-photon ratio
R ( a ) = 3 ρ b ( a ) 4 ρ γ ( a ) = 3 Ω b 4 Ω γ a .
Damping factor
D tot ( k ) = D γ ( k ) D ν ( k ) , D γ ( k ) = exp ( k / k D , γ ) 2 , D ν ( k ) = exp α ν k 2 , α ν = 0.0025 Mpc 2 , k D , γ 2 ( a ) = 0 a d a a 3 n e ( a ) σ T R ( a ) 2 + 16 15 [ 1 + R ( a ) ] 6 H ( a ) [ 1 + R ( a ) ] 2 .
Projection kernels For scalars X = R and polarization Y = T (temperature),
T T ( k ) = 0 η 0 d η g ( η ) Θ 0 ( k , η ) + Ψ ( k , η ) Sachs Wolfe j k [ η 0 η ] + 0 η 0 d η e τ ( η ) Ψ + Φ j k [ η 0 η ] ,
where
g ( η ) = τ ˙ e τ , τ ( η ) = η η 0 d η a n e σ T ,
Θ 0 is the monopole, Ψ , Φ are metric potentials, and primes denote d / d η . For tensor Y = E , B one uses the standard spherical-Bessel projection (e.g. see [18]).
Finally,
C X Y = 4 π d k k P X ( k ) D tot ( k ) T Y ( k ) .

4.14.5 Gaussianity and Non-Gaussianity

Thermal QGP fluctuations give
f NL O ( g * 1 ) 0.025 , f NL 1 ,
consistent with Planck’s f NL local = 0.9 ± 5.1  [17].

4.14.6 Entropy Production and Arrow of Time

Hadronization at T 200 MeV yields Δ S O ( 1 ) , so
S post S pre .

4.14.7 Verification: Parser

Run parse_spectra.py on Scalar_Spectrum.txt/Tensor_Spectrum.txt:
  • Checks the k-grid is uniform in log 10 k .
  • Fits log 10 P vs. log 10 ( k / k p ) .
  • Recovers ( A s , n s , A t , n t , r ) = ( 2.10 × 10 9 , 0.965 , 1.008 × 10 10 , 0.020 , 0.048 ) .
  • Plots raw vs. analytic spectra to < 10 12 .
No hidden parameters or external references remain; this section alone suffices for exact reproduction and direct comparison to CMB data,1.

5. Interpretive Context and Theoretical Extensions

Vacuum Flux Cosmology (VFC) offers a classically complete, general relativistic model of a non-singular bounce, constructed without scalar fields, quantum corrections, or modified gravity. Within the wider landscape of bounce cosmologies, this positions VFC as a minimalist alternative that remains mathematically self-contained while preserving observational compatibility.
  Comparison with Other Bounce Models. Most bounce frameworks invoke quantum gravity mechanisms—such as Loop Quantum Cosmology (LQC) [14]—or rely on scalar field dynamics to realize a bounce or generate primordial perturbations [7,22]. VFC, by contrast, achieves these outcomes via classical junction conditions and thermodynamic inputs derived from lattice QCD. Its energy flux, entropy gradient, and perturbation generation are all traceable to geometric features of the bounce hypersurface.
  • Scalar Field Equivalence. While VFC does not invoke fundamental scalar fields, it can reproduce behaviors traditionally modeled using them—such as red-tilted spectra, gravitational waves, and thermalization [17]. In this view, scalar field dynamics may serve as effective descriptions of curvature-driven transitions, but VFC refrains from presupposing their physical necessity. This perspective does not diminish their utility but offers a complementary geometric interpretation.
  • Classical Realization of Bounce. The bounce in VFC is realized via the Israel junction conditions without invoking quantized spacetime, higher-derivative curvature terms, or violations of Einstein’s equations [4]. While the model describes energy flow from higher to lower curvature regions, all curvature dynamics arise from standard GR geometry and surface stress-energy without modifying the Einstein-Hilbert action.
  • No Exotic Matter Requirements. The formalism relies solely on classical QGP matter and gravitational collapse, avoiding phantom fields, ghost instabilities, or nonstandard energy components.
  • Causal Continuity. Geodesics remain continuous and well-defined across the hypersurface, ensuring causal completeness and avoiding singular disconnection or spacetime truncation.

6. Limitations and Future Directions

The following limitations are acknowledged to clarify the model’s current scope and highlight directions for future research:
  • Boundary Microphysics: The codimension-one interface is modeled classically via Israel junction conditions and curvature-based stress-energy flow. A complete microphysical account—potentially involving quantum gravity or effective material analogs—remains beyond the present scope. However, the classical structure supports well-posed evolution, spectral predictions, and causal continuity, rendering the bounce mathematically tractable and observationally testable.
  • QGP as a Physical Anchor: The use of quark–gluon plasma (QGP) as the dominant matter component near the bounce is based on well-established expectations from high-energy collapse. Its thermodynamic behavior follows the QCD equation of state, approximated as p 1 3 ρ Δ ( T ) , with the trace anomaly Δ ( T ) obtained from lattice QCD [10]. While direct observation in cosmological settings is unavailable, the QGP’s role provides a physically grounded, equation-of-state-driven curvature source.
  • Geometric Role of QGP Transfer: In the emergent frame, the geometry of the incoming QGP determines the flux conduit’s area, influencing energy transfer and entropy deposition. The resulting spacetime leaves no residual mass in the parent frame, only a localized curvature imprint. Reheating is unnecessary: the post-bounce QGP supplies thermal energy and entropy via classical propagation under GR.
  • Observational Fingerprints: VFC reproduces scalar and tensor spectra consistent with Planck and BICEP/Keck observations, including n s = 0.965 , r = 0.048 , and A s 2.1 × 10 9 . However, these observables overlap with predictions from inflationary models. While no distinct signatures are currently predicted, potential discriminants may include deviations in high-k tensor modes, bispectrum anomalies, or entropy-correlated relics—each subject to future observational and numerical investigation.
Future work may explore whether scalar-field models—such as slow-roll inflation—can be reinterpreted as effective coarse-grained descriptions of underlying curvature flux. Such interpretations could clarify the geometric origin of scalar field behavior, spontaneous symmetry breaking, and large-scale expansion dynamics [7]. Numerical simulations of bounce-crossing perturbations, interface stability under rotation, and relic entropy gradients may further refine the predictions and test the model’s robustness [8].

7. Conclusions

We have introduced Vacuum Flux Cosmology (VFC): a classically consistent cosmological model formulated entirely within general relativity. The model replaces the initial singularity with a codimension-one hypersurface that irreversibly transfers curvature and vacuum energy into an expanding spacetime domain. Governed by Israel junction conditions, the bounce features a discontinuity in extrinsic curvature while maintaining metric continuity and preserving geodesic completeness.
VFC satisfies energy conservation via the Bianchi identities [12], admits localized and stable NEC violation through surface stress-energy [1,4], and generates thermodynamic asymmetry through irreversible vacuum energy flux driven by curvature discontinuity and the classical thermodynamics of QGP governed by the QCD equation of state [3,11].
By reframing gravitational collapse as a geometric transition rather than a singular endpoint, the model offers a classical foundation for early-universe dynamics. It reproduces inflation-era features—such as expansion, entropy generation, and thermalization—through purely geometric and thermodynamic mechanisms grounded in Einstein’s equations.
One of the model’s most robust predictions is the classical emergence of spatial flatness. As curvature flux propagates outward, the extrinsic curvature naturally smooths, flattening the induced metric without requiring fine-tuning or exponential inflation. This arises directly from geometric continuity and conservation principles and provides a falsifiable explanation for the observed near-flatness of the universe.
All numerical values and spectra presented in this work are fully reproducible via the included supplemental fileSpectra.zip.
Companion Framework.
Although the present work remains focused on early-universe dynamics, its geometric structure serves as a foundation for broader extensions. These are developed in the companion framework—the Dark Tension Field Framework (DTFF): A Geometric Lens for Dark Sector Dynamics [23]—which generalizes the classical principles of VFC to explore large-scale structure, early galaxy formation, and unresolved phenomena in modern cosmology. A prototype R200 estimator tool derived from vacuum flux geometry is available via the Zenodo open-access archive: doi.org/10.5281/zenodo.14890207. DTFF remains grounded in reversible curvature dynamics and extends the testable structure of VFC into new observational domains.

References

  1. Israel, W. (1966). Singular Hypersurfaces and Thin Shells in General Relativity. Il Nuovo Cimento B, 44(1), 1–14. [CrossRef]
  2. Sabine Hossenfelder, Phenomenology of space-time imperfection. I. Nonlocal defects, Phys. Rev. D 88, 124030 (2013). [CrossRef]
  3. Sabine Hossenfelder, Phenomenology of space-time imperfection. I. Nonlocal defects, Phys. Rev. D 88, 124030 (2013). [CrossRef]
  4. Visser, M. (1995). Lorentzian Wormholes: From Einstein to Hawking. AIP Press.
  5. Poisson, E., & Visser, M. (1995). Thin-shell wormholes: Linearization stability. Physical Review D, 52(12), 7318–7321. [CrossRef]
  6. Clarke, C. J. S., & Dray, T. (1987). Junction conditions for null hypersurfaces. Classical and Quantum Gravity, 4(1), 265–275.
  7. Baumann, D. (2009). TASI Lectures on Inflation. arXiv preprint arXiv:0907.5424. https://arxiv.org/abs/0907.5424.
  8. Brandenberger, R., & Peter, P. (2017). Bouncing Cosmologies: Progress and Problems. Foundations of Physics, 47, 797–850. [CrossRef]
  9. Eling, C., Guedens, R., & Jacobson, T. (2006). Non-equilibrium thermodynamics of spacetime. Physical Review Letters, 96(12), 121301. [CrossRef]
  10. Borsányi, S., Fodor, Z., Katz, S. D., Krieg, S., Ratti, C., & Szabó, K. K. (2014). Full result for the QCD equation of state with 2+1 flavors. Physics Letters B, 730, 99–104. [CrossRef]
  11. Borsányi, S., Fodor, Z., Hoelbling, C., Katz, S. D., Krieg, S., Ratti, C., & Szabó, K. K. (2010). The QCD equation of state with dynamical quarks. Journal of High Energy Physics, 2010(11), 077. [CrossRef]
  12. Hawking, S. W., & Ellis, G. F. R. (1973). The Large Scale Structure of Space-Time. Cambridge University Press. [CrossRef]
  13. CERN. (2024). LHC Experiments Overview. https://home.cern/science/experiments.
  14. Ashtekar, A. (2009). Loop quantum cosmology: an overview. General Relativity and Gravitation, 41(4), 707–741. [CrossRef]
  15. Unruh, W. G. (1976). Notes on black-hole evaporation. Physical Review D, 14(4), 870–892. [CrossRef]
  16. Martín-Moruno, P., & Visser, M. (2013). Classical and semi-classical energy conditions. arXiv preprint. arXiv:1306.2076.
  17. Planck Collaboration, Aghanim, N., Akrami, Y., et al. (2020). Planck 2018 results. VI. Cosmological parameters. Astronomy & Astrophysics, 641, A6. https://www.aanda.org/10.1051/0004-6361/201833910e.
  18. U. Seljak and M. Zaldarriaga, “A Line of Sight Approach to Cosmic Microwave Background Anisotropies,” Astrophys. J. 469, 437 (1996).
  19. Mukhanov, V. (2005). Physical Foundations of Cosmology. Cambridge University Press. ISBN 978-0-521-61939-0.
  20. ALICE Collaboration. (2010). First proton–proton collisions at the LHC as observed with the ALICE detector. European Physical Journal C, 65, 111–125. [CrossRef]
  21. Ma, C.-P., & Bertschinger, E. (1995). Cosmological perturbation theory in the synchronous and conformal Newtonian gauges. Astrophysical Journal, 455, 7. [CrossRef]
  22. Khoury, J., Ovrut, B. A., Steinhardt, P. J., & Turok, N. (2001). The Ekpyrotic Universe: Colliding branes and the origin of the hot big bang. Physical Review D, 64(12), 123522.
  23. Koritar, B. (2025). Dark Tension Field Framework (DTFF): A Geometric Lens for Dark Sector Dynamics. Zenodo. [CrossRef]
1
Spectral shapes and tilts are direct predictions; the overall amplitude is fixed by the single parameter h.
Table 1. Representative values from lattice QCD for the trace anomaly and sound speed during the QGP phase [3,11].
Table 1. Representative values from lattice QCD for the trace anomaly and sound speed during the QGP phase [3,11].
Temperature T [MeV] Δ ( T ) / T 4 c s 2 = d p / d ρ
150 1.2 0.15
200 0.5 0.20
250 0.2 0.25
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

Disclaimer

Terms of Use

Privacy Policy

Privacy Settings

© 2025 MDPI (Basel, Switzerland) unless otherwise stated