The Quantum Vacuum Self-Consistency Principle: Emergent Dynamics of Spacetime and the Standard Model

Tao Huang [email protected] Hechi, Guangxi, China
(November 6, 2025)
Abstract

The principle of self-consistency of the quantum vacuum postulates that the classical backgrounds we observe—spacetime geometry, gauge fields, and the Higgs condensate—are macroscopic order parameters of a single quantum state whose existence is sustained by the vacuum expectation values of all quantum fluctuations living on it. Building on this postulate, a background-field, heat-kernel based derivation is developed that yields the coupled low-energy effective field equations for the metric, gauge fields, and the Higgs field as vacuum “equations of state.” The resulting framework rigorously recovers the Einstein, Yang-Mills, and Higgs equations, augmented by the higher-derivative operators required by quantum consistency, with renormalized couplings determined by the content of quantum fields. A Renormalization Group (RG) structure is obtained in which all couplings—including Newton’s constant, the cosmological constant, the coefficients of R2R^{2} and Cμνρσ2C_{\mu\nu\rho\sigma}^{2}, gauge couplings, Yukawas, and the Higgs quartic—run coherently with a single effective action that respects background diffeomorphism and gauge invariance. A simplified solvable model is analyzed: an O(N)O(N) scalar sector with nonminimal coupling on a constant-curvature background, for which the one-loop effective potential including the leading curvature corrections is computed and the vacuum gap equations are solved explicitly. Phenomenological consequences follow. First, a robust prediction: the anomaly- and loop-induced R2R^{2} operator generically drives Starobinsky-type inflation, with ns12/Nen_{s}\simeq 1-2/N_{e} and r12/Ne2r\simeq 12/N_{e}^{2}, compatible with Planck data; the required coefficient corresponds to a scalaron mass M(1.3±0.1)×105MPlM\simeq(1.3\pm 0.1)\times 10^{-5}M_{\rm Pl}. Second, the universally calculable quantum correction to Newton’s potential and the Yukawa tails from the massive spin-0 and spin-2 modes are quantified and shown to satisfy laboratory bounds. Third, constraints from GW170817 enforce luminal gravitational wave speed for the massless graviton in this framework, while higher-derivative effects remain suppressed at LIGO/Virgo/KAGRA frequencies. The theory is predictive in its inflationary sector and in its universal low-energy corrections to gravity, while it remains honest about open issues: the smallness of the observed cosmological constant, the nonperturbative completion of the higher-derivative sector, and the determination of threshold-matched couplings beyond one loop. The structure is sufficiently complete to be testable across cosmology, astrophysics, and precision gravity, and it reduces to General Relativity and the Standard Model at accessible scales with controlled corrections.

I Introduction

The Standard Model (SM) of particle physics and General Relativity (GR) are profoundly successful descriptions of nature [Glashow1961, Weinberg1967, Salam1968, Misner1973]. Yet their unification remains elusive. Quantizing GR perturbatively leads to nonrenormalizable divergences [tHooft1974], which signal the need for a deeper organizing principle. A natural path, inspired by Sakharov’s induced gravity [Sakharov1968], is that the classical backgrounds we observe are emergent, macroscopic order parameters of an underlying quantum vacuum. In this view, geometry, gauge fields, and the Higgs condensate are not separate axioms; rather, they are coherent manifestations of one vacuum that sustains itself through self-consistency.

The central postulate is that the vacuum state |Ω|\Omega\rangle determined by classical backgrounds must, through quantum fluctuations, generate precisely the sources that maintain those same backgrounds. This self-consistency is enforced by the quantum effective action computed in the presence of background fields. The background field method and the heat-kernel expansion provide the technical backbone [DeWitt2003, Vassilevich2003, Barvinsky1985, BirrellDavies1982, ParkerToms2009, Buchbinder1992, Avramidi2000]. At one loop, the Seeley–DeWitt coefficient a2a_{2} fixes the divergences and therefore the RG running of all couplings that can appear in a local effective action.

This work develops that program into a predictive framework. The self-consistency postulate is kept intact and is expressed via vacuum gap equations that are precisely the stationarity conditions of the renormalized 1PI effective action with respect to the background fields. We compute and reorganize the one-loop effective action to derive: (i) the renormalized Einstein equations plus controlled higher-derivative corrections, (ii) the renormalized Yang–Mills equations, and (iii) the renormalized Higgs equation and potential. We also construct the coupled RG equations for gravitational and SM couplings, which makes the framework calculable.

A simplified, analytically tractable model is solved explicitly to exhibit the mechanism: an O(N)O(N) scalar with nonminimal coupling on a constant-curvature background. The one-loop effective potential including curvature-dependent terms is computed and the gap equations are solved for the vacuum expectation value and curvature. The analysis isolates which combinations of couplings are fixed by self-consistency and which must be matched to data.

Finally, the framework is confronted with data. The loop- and anomaly-induced R2R^{2} term drives Starobinsky inflation [Starobinsky1980, Whitt1984], yielding benchmark predictions for the spectral tilt and tensor-to-scalar ratio consistent with Planck [Planck2018]. Quantum corrections to Newton’s potential are universal and match the effective field theory result [Donoghue1994]. Short-range Yukawa corrections from the massive scalar (the scalaron) and the heavy spin-2 mode are shown to obey submillimeter bounds [Kapner2007]. Gravitational wave propagation is luminal in the infrared, in agreement with GW170817 [GW170817].

The approach is conservative in spirit: it embraces effective field theory, honors all symmetries via the background field method, and leans on mathematically controlled tools. At the same time, it reframes the unification problem in terms of a single organizing idea: vacuum self-consistency.

II Unified framework and the self-consistency postulate

All fields are decomposed into classical backgrounds and quantum fluctuations,

g^μν(x)\displaystyle\hat{g}_{\mu\nu}(x) =g¯μν(x)+h^μν(x),\displaystyle=\bar{g}_{\mu\nu}(x)+\hat{h}_{\mu\nu}(x), (1)
A^μa(x)\displaystyle\hat{A}_{\mu}^{a}(x) =A¯μa(x)+a^μa(x),\displaystyle=\bar{A}_{\mu}^{a}(x)+\hat{a}_{\mu}^{a}(x), (2)
Φ^(x)\displaystyle\hat{\Phi}(x) =12(vH(x)+σ^(x)),\displaystyle=\frac{1}{\sqrt{2}}\bigl(v_{H}(x)+\hat{\sigma}(x)\bigr), (3)
Ψ^(x)\displaystyle\hat{\Psi}(x) =ψ^(x).\displaystyle=\hat{\psi}(x). (4)

Fermions are purely quantum in the vacuum since they do not condense. Backgrounds {g¯,A¯,vH}\{\bar{g},\bar{A},v_{H}\} define a vacuum |Ω=|0g¯,A¯,vH|\Omega\rangle=|0_{\bar{g},\bar{A},v_{H}}\rangle.

The unified self-consistency postulate is that the backgrounds are sustained by the renormalized vacuum expectation values (VEVs) of their source operators in |Ω|\Omega\rangle. Equivalently, in the presence of external sources JJ, the renormalized 1PI effective action Γ\Gamma obeys the background-field stationarity conditions,

2g¯δΓδg¯μν\displaystyle\frac{2}{\sqrt{-\bar{g}}}\frac{\delta\Gamma}{\delta\bar{g}^{\mu\nu}} =Tμνext,\displaystyle=T_{\mu\nu}^{\rm ext}, (5)
1g¯δΓδA¯μa\displaystyle\frac{1}{\sqrt{-\bar{g}}}\frac{\delta\Gamma}{\delta\bar{A}_{\mu}^{a}} =Jextaμ,\displaystyle=J^{a\,\mu}_{\rm ext}, (6)
1g¯δΓδvH\displaystyle\frac{1}{\sqrt{-\bar{g}}}\frac{\delta\Gamma}{\delta v_{H}} =Jσext.\displaystyle=J_{\sigma}^{\rm ext}. (7)

In the absence of external sources, these become vacuum gap equations. Writing the bare (classical) actions with bare parameters and integrating out fluctuations generates the quantum terms that renormalize both parameters and the structure of the equations. The background field method ensures background diffeomorphism and gauge invariance, so that Ward identities are preserved [Abbott1981, Buchbinder1992].

To one loop, the quantum effective action is

Γ[g¯,A¯,vH]\displaystyle\Gamma[\bar{g},\bar{A},v_{H}] =Sbare[g¯,A¯,vH]\displaystyle=S_{\rm bare}[\bar{g},\bar{A},v_{H}]
+i2Trln𝒟bosiTrln𝒟ferm\displaystyle\quad+\frac{i}{2}\mathrm{Tr}\ln\mathcal{D}_{\rm bos}-i\,\mathrm{Tr}\ln\mathcal{D}_{\rm ferm}
iTrln𝒟ghost+,\displaystyle\quad-i\,\mathrm{Tr}\ln\mathcal{D}_{\rm ghost}+\cdots, (8)

with minimal second-order operators of the form 𝒟=¯2+𝒫\mathcal{D}=-\bar{\nabla}^{2}+\mathcal{P} acting on appropriate bundles and 𝒫\mathcal{P} containing masses, curvature couplings, and background field strengths. Divergences are governed by the heat-kernel coefficient a2a_{2} [DeWitt2003, Vassilevich2003, Barvinsky1985, Avramidi2000]. For each species jj, one finds a local divergent density

div,j(1)\displaystyle\mathcal{L}^{(1)}_{\rm div,j} =116π2ϵg¯tra2,j(x),\displaystyle=\frac{1}{16\pi^{2}\epsilon}\sqrt{-\bar{g}}\;\mathrm{tr}\,a_{2,j}(x), (9)

with ϵ=4d\epsilon=4-d in dimensional regularization and the trace over internal indices. Summing over all fluctuating fields with appropriate statistics yields

Γdiv(1)\displaystyle\Gamma_{\rm div}^{(1)} =116π2ϵd4xg¯[CΛ+CRR¯\displaystyle=\frac{1}{16\pi^{2}\epsilon}\int d^{4}x\sqrt{-\bar{g}}\Bigl[C_{\Lambda}+C_{R}\bar{R}
+CC2CμνρσCμνρσ+CEE4\displaystyle\quad+C_{C^{2}}\,C_{\mu\nu\rho\sigma}C^{\mu\nu\rho\sigma}+C_{E}\,E_{4}
+iCFitr(F¯μν(i)F¯(i)μν)\displaystyle\quad+\sum_{i}C_{F_{i}}\,\mathrm{tr}\bigl(\bar{F}_{\mu\nu}^{(i)}\bar{F}^{(i)\mu\nu}\bigr)
+CRHR¯vH2+CHvH4+],\displaystyle\quad+C_{RH}\,\bar{R}\,v_{H}^{2}+C_{H}\,v_{H}^{4}+\cdots\Bigr], (10)

where E4=Rμνρσ24Rμν2+R2E_{4}=R_{\mu\nu\rho\sigma}^{2}-4R_{\mu\nu}^{2}+R^{2} is the Euler density and CμνρσC_{\mu\nu\rho\sigma} is the Weyl tensor. The coefficients CXC_{X} are calculable sums of species-dependent numbers and masses, and depend on the nonminimal couplings of scalars. The ellipsis denotes higher-derivative scalar terms and total derivatives such as R\Box R that can be discarded or absorbed into local counterterms.

Finite parts, including nonlocal form factors that resum infrared and threshold effects, are encoded in the full Γ\Gamma [BarvinskyVilkovisky1985b, BarvinskyVilkovisky1990] and are essential for precision predictions, but for many purposes the local part suffices.

III Stationarity conditions and emergent field equations

The gap equations (5)–(7) are the equations of state of the vacuum. They are obtained by varying the renormalized effective action,

Γ\displaystyle\Gamma =d4xg¯[ZR16πGR¯ZΛ8πGΛ\displaystyle=\int d^{4}x\sqrt{-\bar{g}}\Biggl[\frac{Z_{R}}{16\pi G}\bar{R}-\frac{Z_{\Lambda}}{8\pi G}\Lambda
+αCCμνρσCμνρσ+αRR¯2\displaystyle\quad+\alpha_{C}C_{\mu\nu\rho\sigma}C^{\mu\nu\rho\sigma}+\alpha_{R}\bar{R}^{2}
iZFi4gi2tr(F¯μν(i)F¯(i)μν)\displaystyle\quad-\sum_{i}\frac{Z_{F_{i}}}{4g_{i}^{2}}\mathrm{tr}\bigl(\bar{F}_{\mu\nu}^{(i)}\bar{F}^{(i)\mu\nu}\bigr)
+ZH12g¯μνμvHνvHVeff(vH,R¯)\displaystyle\quad+Z_{H}\frac{1}{2}\bar{g}^{\mu\nu}\partial_{\mu}v_{H}\partial_{\nu}v_{H}-V_{\rm eff}(v_{H},\bar{R})
+]+Γnl,\displaystyle\quad+\cdots\Biggr]+\Gamma_{\rm nl}, (11)

where ZZ-factors include finite renormalizations, αC\alpha_{C} and αR\alpha_{R} are the renormalized dimensionless couplings of the curvature-squared operators, VeffV_{\rm eff} is the renormalized effective potential for the Higgs condensate including curvature dependence, and Γnl\Gamma_{\rm nl} collects nonlocal form factors such as Rln(¯/μ2)RR\ln(-\bar{\Box}/\mu^{2})R that are known from the covariant perturbation theory of Barvinsky and Vilkovisky [BarvinskyVilkovisky1990]. Dots denote higher-order and higher-derivative terms suppressed at low energies.

Varying (11) with respect to g¯μν\bar{g}^{\mu\nu} yields

ZRGμν+ZΛΛg¯μν\displaystyle Z_{R}\,G_{\mu\nu}+Z_{\Lambda}\,\Lambda\,\bar{g}_{\mu\nu}
+16πG[αCBμν+αRHμν(R2)]\displaystyle\quad+16\pi G\Bigl[\alpha_{C}\,B_{\mu\nu}+\alpha_{R}\,H^{(R^{2})}_{\mu\nu}\Bigr]
=8πG(Tμνext+Tμν(H)+Tμνnl),\displaystyle=8\pi G\Bigl(T_{\mu\nu}^{\rm ext}+T_{\mu\nu}^{(H)}+T_{\mu\nu}^{\rm nl}\Bigr), (12)

where BμνB_{\mu\nu} is the Bach tensor,

Bμν\displaystyle B_{\mu\nu} =ρσCμρνσ+12RρσCμρνσ,\displaystyle=\nabla^{\rho}\nabla^{\sigma}C_{\mu\rho\nu\sigma}+\tfrac{1}{2}R^{\rho\sigma}C_{\mu\rho\nu\sigma}, (13)

and Hμν(R2)H^{(R^{2})}_{\mu\nu} is the metric variation of R2R^{2},

Hμν(R2)\displaystyle H^{(R^{2})}_{\mu\nu} =2RRμν12g¯μνR22μνR+2g¯μνR.\displaystyle=2RR_{\mu\nu}-\tfrac{1}{2}\bar{g}_{\mu\nu}R^{2}-2\nabla_{\mu}\nabla_{\nu}R+2\bar{g}_{\mu\nu}\Box R. (14)

The tensor Tμν(H)T_{\mu\nu}^{(H)} contains the stress tensor of vHv_{H},

Tμν(H)\displaystyle T_{\mu\nu}^{(H)} =ZHμvHνvH12ZHg¯μν(vH)2+g¯μνVeff,\displaystyle=Z_{H}\partial_{\mu}v_{H}\partial_{\nu}v_{H}-\tfrac{1}{2}Z_{H}\bar{g}_{\mu\nu}(\partial v_{H})^{2}+\bar{g}_{\mu\nu}V_{\rm eff}, (15)

and TμνnlT_{\mu\nu}^{\rm nl} encodes finite nonlocal contributions. In the long-wavelength, weak-curvature regime, the higher-derivative and nonlocal terms are suppressed, and (12) reduces to the Einstein equations with renormalized constants.

Variation with respect to the gauge backgrounds yields

(ZFigi2)D¯νF¯(i)νμ\displaystyle\Bigl(\frac{Z_{F_{i}}}{g_{i}^{2}}\Bigr)\,\bar{D}_{\nu}\bar{F}^{(i)\nu\mu} =Jext(i)μ+Jnl(i)μ,\displaystyle=J^{(i)\mu}_{\rm ext}+J^{(i)\mu}_{\rm nl}, (16)

with Jnl(i)μJ^{(i)\mu}_{\rm nl} the finite nonlocal current from integrating out matter and gravity.

Variation with respect to vHv_{H} gives the renormalized Higgs equation,

ZHg¯vHVeff(vH,R¯)vH\displaystyle Z_{H}\,\Box_{\bar{g}}v_{H}-\frac{\partial V_{\rm eff}(v_{H},\bar{R})}{\partial v_{H}} =Jσext,\displaystyle=J_{\sigma}^{\rm ext}, (17)

where Veff(vH,R¯)V_{\rm eff}(v_{H},\bar{R}) includes quantum and curvature corrections. In the vacuum, Jext=0J_{\rm ext}=0, and (12)–(17) become the coupled gap equations that determine the macroscopic order parameters.

IV Heat-kernel coefficients and renormalization

The coefficients CXC_{X} in (10) are determined by the a2a_{2} coefficient of the heat-kernel expansion. For a Laplace-type operator 𝒟=¯2+𝒫\mathcal{D}=-\bar{\nabla}^{2}+\mathcal{P} acting on a vector bundle with curvature μν=[¯μ,¯ν]\mathcal{F}_{\mu\nu}=[\bar{\nabla}_{\mu},\bar{\nabla}_{\nu}], one has [Vassilevich2003, Avramidi2000]

a2\displaystyle a_{2} =1180(Rμνρσ2Rμν2+R)𝕀+12𝒫2\displaystyle=\tfrac{1}{180}\Bigl(R_{\mu\nu\rho\sigma}^{2}-R_{\mu\nu}^{2}+\Box R\Bigr)\mathbb{I}+\tfrac{1}{2}\mathcal{P}^{2}
16R𝒫+112μν2.\displaystyle\quad-\tfrac{1}{6}R\,\mathcal{P}+\tfrac{1}{12}\mathcal{F}_{\mu\nu}^{2}. (18)

For a real scalar with nonminimal coupling ξ\xi, 𝒫=m2+ξR+\mathcal{P}=m^{2}+\xi R+\cdots. For a Dirac fermion, 𝒫=m2+14R+\mathcal{P}=m^{2}+\tfrac{1}{4}R+\cdots and one must include a minus sign from statistics and an overall factor from spinor components. For gauge vectors in background gauge, the operator acts on vectors and ghosts contribute with their own operators. Combining all species, one obtains the standard trace-anomaly coefficients for the C2C^{2} and E4E_{4} structures [Duff1994, ChristensenDuff1978]. For NsN_{s} real scalars, NfN_{f} Dirac fermions, and NvN_{v} gauge vectors,

CC2\displaystyle C_{C^{2}} =1120Ns+120Nf+110Nv,\displaystyle=\frac{1}{120}N_{s}+\frac{1}{20}N_{f}+\frac{1}{10}N_{v}, (19)
CE\displaystyle C_{E} =1360Ns+11360Nf+31180Nv.\displaystyle=\frac{1}{360}N_{s}+\frac{11}{360}N_{f}+\frac{31}{180}N_{v}. (20)

These appear as the coefficients of the 1/ϵ1/\epsilon divergences in front of C2C^{2} and E4E_{4} in (10). Mass-dependent terms contribute to CΛC_{\Lambda} and CRC_{R},

CΛ\displaystyle C_{\Lambda} =j(1)Fjnjmj4,\displaystyle=\sum_{j}(-1)^{F_{j}}\,n_{j}\,m_{j}^{4}, (21)
CR\displaystyle C_{R} =s(ξs16)ms2+f(112)mf2+v(16)mv2,\displaystyle=\sum_{s}\Bigl(\xi_{s}-\tfrac{1}{6}\Bigr)m_{s}^{2}+\sum_{f}\Bigl(-\tfrac{1}{12}\Bigr)m_{f}^{2}+\sum_{v}\Bigl(\tfrac{1}{6}\Bigr)m_{v}^{2}, (22)

where njn_{j} counts internal degrees of freedom and FjF_{j} is 0 (1) for bosons (fermions). In dimensional regularization, strictly massless fields do not contribute to CRC_{R} at one loop, while CC2C_{C^{2}} and CEC_{E} are mass-independent. The gauge-field divergences yield

CFi\displaystyle C_{F_{i}} =κijij,\displaystyle=\kappa_{i}-\sum_{j}\ell_{ij}, (23)

with κi\kappa_{i} the adjoint contribution and ij\ell_{ij} matter contributions in representation jj of gauge group ii.

Renormalization proceeds by absorbing (10) into counterterms, defining the renormalized couplings at scale μ\mu in minimal subtraction. The resulting RG equations take the schematic form

μddμ(ZR16πG)\displaystyle\mu\frac{d}{d\mu}\Bigl(\frac{Z_{R}}{16\pi G}\Bigr) =CR8π2,\displaystyle=-\frac{C_{R}}{8\pi^{2}}, (24)
μddμ(ZΛΛ8πG)\displaystyle\mu\frac{d}{d\mu}\Bigl(\frac{Z_{\Lambda}\Lambda}{8\pi G}\Bigr) =CΛ16π2,\displaystyle=\frac{C_{\Lambda}}{16\pi^{2}}, (25)
μdαCdμ\displaystyle\mu\frac{d\alpha_{C}}{d\mu} =CC28π2,μdαRdμ=CR28π2,\displaystyle=\frac{C_{C^{2}}}{8\pi^{2}},\qquad\mu\frac{d\alpha_{R}}{d\mu}=\frac{C_{R^{2}}}{8\pi^{2}}, (26)
μddμ(ZFigi2)\displaystyle\mu\frac{d}{d\mu}\Bigl(\frac{Z_{F_{i}}}{g_{i}^{2}}\Bigr) =CFi8π2,\displaystyle=-\frac{C_{F_{i}}}{8\pi^{2}}, (27)

with CR2C_{R^{2}} a linear combination of CEC_{E} and species-dependent terms that accompany R2R^{2}. The precise mapping between {CE,CC2}\{C_{E},C_{C^{2}}\} and {βαR,βαC}\{\beta_{\alpha_{R}},\beta_{\alpha_{C}}\} depends on the chosen basis, see [Buchbinder1992, Duff1994]. In the matter sector, the SM one-loop β\beta functions are standard [MachacekVaughnI, MachacekVaughnII, Buttazzo2013].

Nonlocal terms in Γ\Gamma carry logarithms of the form ln(¯/μ2)\ln(-\bar{\Box}/\mu^{2}) and govern both decoupling and threshold behavior [BarvinskyVilkovisky1990, Donoghue1994]. At scales μ\mu below a particle’s mass, matching across thresholds must be performed to ensure the decoupling of heavy species in gauge and Yukawa sectors [AppelquistCarazzone1975]. In curved space this procedure is technically involved but conceptually the same, and the covariant perturbation theory provides the right tool.

V A solvable model: O(N)O(N) scalar on constant curvature

To display the mechanism in a fully analytic setting, consider NN real scalars ϕI\phi^{I} with O(N)O(N) symmetry and classical potential V(ϕ)=λ4(ϕ2v02)2V(\phi)=\tfrac{\lambda}{4}(\phi^{2}-v_{0}^{2})^{2}, nonminimally coupled with coupling ξ\xi to the curvature, on a background of constant scalar curvature R¯\bar{R}. Decompose ϕI=(v+σ,πa)\phi^{I}=(v+\sigma,\pi^{a}) with a=1,,N1a=1,\dots,N-1, so that v=ϕv=\langle\phi\rangle plays the role of vHv_{H}. The classical Euclidean action relevant for the effective potential is

Scl\displaystyle S_{\rm cl} =d4xg¯[MPl22R¯+αRR¯212ξR¯ϕ2\displaystyle=\int d^{4}x\sqrt{\bar{g}}\Bigl[\frac{M_{\rm Pl}^{2}}{2}\bar{R}+\alpha_{R}\bar{R}^{2}-\frac{1}{2}\xi\bar{R}\,\phi^{2}
+12(¯ϕ)2+λ4(ϕ2v02)2+].\displaystyle\quad+\frac{1}{2}(\bar{\nabla}\phi)^{2}+\frac{\lambda}{4}(\phi^{2}-v_{0}^{2})^{2}+\cdots\Bigr]. (28)

In the constant background, gradients vanish and the tree-level potential density reads

Utree(v,R¯)\displaystyle U_{\rm tree}(v,\bar{R}) =MPl22R¯αRR¯2+12ξR¯v2\displaystyle=-\frac{M_{\rm Pl}^{2}}{2}\bar{R}-\alpha_{R}\bar{R}^{2}+\frac{1}{2}\xi\bar{R}\,v^{2}
+λ4(v2v02)2+ρΛ,\displaystyle\quad+\frac{\lambda}{4}(v^{2}-v_{0}^{2})^{2}+\rho_{\Lambda}, (29)

where ρΛ\rho_{\Lambda} collects the cosmological constant term and constant counterterms.

The one-loop effective potential including the leading curvature dependence is obtained by summing the fluctuation determinants of σ\sigma and πa\pi^{a}. Their field-dependent masses in the constant background are

mπ2(v,R¯)\displaystyle m_{\pi}^{2}(v,\bar{R}) =λ(v2v02)+ξR¯,\displaystyle=\lambda(v^{2}-v_{0}^{2})+\xi\bar{R}, (30)
mσ2(v,R¯)\displaystyle m_{\sigma}^{2}(v,\bar{R}) =λ(3v2v02)+ξR¯.\displaystyle=\lambda(3v^{2}-v_{0}^{2})+\xi\bar{R}. (31)

At one loop in MS¯\overline{\rm MS},

U1-loop(v,R¯)\displaystyle U_{\rm 1\mbox{-}loop}(v,\bar{R}) =i=π,σni64π2mi4(v,R¯)\displaystyle=\sum_{i=\pi,\sigma}\frac{n_{i}}{64\pi^{2}}m_{i}^{4}(v,\bar{R})
×[lnmi2(v,R¯)μ2ci]\displaystyle\quad\times\Bigl[\ln\frac{m_{i}^{2}(v,\bar{R})}{\mu^{2}}-c_{i}\Bigr]
+𝒪(R¯2ln(μ)),\displaystyle\quad+\mathcal{O}\bigl(\bar{R}^{2}\ln(\mu)\bigr), (32)

with nπ=N1n_{\pi}=N-1, nσ=1n_{\sigma}=1, and ci=32c_{i}=\tfrac{3}{2} for scalars in the MS¯\overline{\rm MS} scheme. The 𝒪(R¯2)\mathcal{O}(\bar{R}^{2}) terms in U1-loopU_{\rm 1\mbox{-}loop} can be absorbed into a running of αR(μ)\alpha_{R}(\mu) plus total derivatives. The full effective potential is

Ueff(v,R¯)\displaystyle U_{\rm eff}(v,\bar{R}) =Utree(v,R¯)+U1-loop(v,R¯).\displaystyle=U_{\rm tree}(v,\bar{R})+U_{\rm 1\mbox{-}loop}(v,\bar{R}). (33)

The vacuum gap equations are the stationarity conditions with respect to vv and R¯\bar{R},

Ueffv\displaystyle\frac{\partial U_{\rm eff}}{\partial v} =0,\displaystyle=0, (34)
UeffR¯\displaystyle\frac{\partial U_{\rm eff}}{\partial\bar{R}} =0.\displaystyle=0. (35)

Explicitly,

0\displaystyle 0 =λ(v2v02)v+ξR¯v\displaystyle=\lambda(v^{2}-v_{0}^{2})v+\xi\bar{R}\,v
+ini32π2mi2(v,R¯)mi2v[lnmi2μ21],\displaystyle\quad+\sum_{i}\frac{n_{i}}{32\pi^{2}}m_{i}^{2}(v,\bar{R})\,\frac{\partial m_{i}^{2}}{\partial v}\Bigl[\ln\frac{m_{i}^{2}}{\mu^{2}}-1\Bigr], (36)
0\displaystyle 0 =MPl222αRR¯+ξ2v2\displaystyle=-\frac{M_{\rm Pl}^{2}}{2}-2\alpha_{R}\bar{R}+\frac{\xi}{2}v^{2}
+ini64π2{2mi2mi2R¯[lnmi2μ21]}.\displaystyle\quad+\sum_{i}\frac{n_{i}}{64\pi^{2}}\Biggl\{2m_{i}^{2}\,\frac{\partial m_{i}^{2}}{\partial\bar{R}}\Bigl[\ln\frac{m_{i}^{2}}{\mu^{2}}-1\Bigr]\Biggr\}. (37)

To see the origin of the numerical factor in (36), note that for a single mode the one-loop term is ΔU=164π2m4[ln(m2/μ2)32]\Delta U=\frac{1}{64\pi^{2}}m^{4}[\ln(m^{2}/\mu^{2})-\tfrac{3}{2}], hence

ΔUv\displaystyle\frac{\partial\Delta U}{\partial v} =164π2 2m2m2v[lnm2μ21],\displaystyle=\frac{1}{64\pi^{2}}\,2m^{2}\frac{\partial m^{2}}{\partial v}\Bigl[\ln\frac{m^{2}}{\mu^{2}}-1\Bigr], (38)

which produces the overall 1/(32π2)1/(32\pi^{2}) factor after summing degeneracies nin_{i}.

Using (30)–(31), the derivatives are mi2/v=2λciv\partial m_{i}^{2}/\partial v=2\lambda c_{i}v with cπ=1c_{\pi}=1, cσ=3c_{\sigma}=3, and mi2/R¯=ξ\partial m_{i}^{2}/\partial\bar{R}=\xi. Solving (36)–(37) for (v,R¯)(v,\bar{R}) at given (λ,ξ,αR,μ)(\lambda,\xi,\alpha_{R},\mu) demonstrates the self-consistent determination of the order parameters.

Two limiting cases are instructive.

First, if R¯=0\bar{R}=0 and quantum corrections are small at μv\mu\sim v, then (36) gives vv0v\simeq v_{0} as usual, and (37) fixes the renormalized cosmological constant counterterm via ρΛ\rho_{\Lambda} so that Minkowski is a solution. This is the standard tuning familiar from semiclassical gravity.

Second, if αR\alpha_{R} is large and positive, a de Sitter solution arises with R¯>0\bar{R}>0 dominated by the R2R^{2} term. At tree level, (37) yields R¯14αR(ξv2MPl2)\bar{R}\simeq\tfrac{1}{4\alpha_{R}}\bigl(\xi v^{2}-M_{\rm Pl}^{2}\bigr), and (36) reduces to (λv2+ξR¯λv02)v0(\lambda v^{2}+\xi\bar{R}-\lambda v_{0}^{2})v\simeq 0. For ξ>0\xi>0 and v0v\neq 0, these equations define a consistent de Sitter branch, which is lifted by the loop terms in the expected way. This branch underlies the R2R^{2} inflation scenario discussed below.

This solvable model shows explicitly how the self-consistency postulate supplies coupled algebraic equations that fix the background curvature and the condensate in terms of renormalized couplings. The same logic extends to the full SM plus gravity, with the technical difference that gauge bosons and fermions contribute additional terms and thresholds.

VI Phenomenology and predictions

VI.1 Inflation from the induced R2R^{2} term

The universal generation of curvature-squared operators is unavoidable at one loop. The R2R^{2} term is particularly important because it is ghost-free in the scalar sector and, when dominant, is dynamically equivalent to GR plus a scalar field (the scalaron) [Whitt1984, Stelle1977]. In the Einstein frame, the scalaron potential is

V(φ)\displaystyle V(\varphi) =34M2MPl2(1e23φMPl)2,\displaystyle=\frac{3}{4}M^{2}M_{\rm Pl}^{2}\Bigl(1-e^{-\sqrt{\frac{2}{3}}\frac{\varphi}{M_{\rm Pl}}}\Bigr)^{2}, (39)

and the mapping from the Jordan-frame action g[MPl22R+αRR2]\int\sqrt{-g}\,\bigl[\tfrac{M_{\rm Pl}^{2}}{2}R+\alpha_{R}R^{2}\bigr] to (39) gives

αR\displaystyle\alpha_{R} =MPl212M2.\displaystyle=\frac{M_{\rm Pl}^{2}}{12M^{2}}. (40)

For slow roll with NeN_{e} e-folds,

ns\displaystyle n_{s} 12Ne,r12Ne2.\displaystyle\simeq 1-\frac{2}{N_{e}},\qquad r\simeq\frac{12}{N_{e}^{2}}. (41)

Taking Ne=50N_{e}=506060 yields ns0.962n_{s}\simeq 0.9620.9670.967 and r0.003r\simeq 0.0030.0040.004, consistent with Planck [Planck2018]. The scalar amplitude fixes M(1.3±0.1)×105MPlM\simeq(1.3\pm 0.1)\times 10^{-5}M_{\rm Pl}, hence

αR\displaystyle\alpha_{R} MPl212M2(4.55.5)×108.\displaystyle\simeq\frac{M_{\rm Pl}^{2}}{12M^{2}}\simeq(4.5\text{--}5.5)\times 10^{8}. (42)

This is a concrete, quantitative prediction: the self-consistent vacuum generates an R2R^{2} term whose coefficient is large and positive in the inflationary regime. In the late universe, the R2R^{2} term is inert and safely compatible with local tests.

VI.2 Short-distance gravity and laboratory bounds

Quadratic gravity around flat space contains, besides the massless graviton, a massive scalar (the scalaron) and a massive spin-2 mode from the C2C^{2} term [Stelle1977]. The static potential between two point masses m1m_{1} and m2m_{2} at separation rr is

V(r)\displaystyle V(r) =GNm1m2r[1+13eMr43em2r\displaystyle=-\frac{G_{N}m_{1}m_{2}}{r}\Biggl[1+\frac{1}{3}e^{-Mr}-\frac{4}{3}e^{-m_{2}r}
+4110πGNr2c3+],\displaystyle\quad+\frac{41}{10\pi}\frac{G_{N}\hbar}{r^{2}c^{3}}+\cdots\Biggr], (43)

where M2=MPl2/(12αR)M^{2}=M_{\rm Pl}^{2}/(12\alpha_{R}) and m22MPl2/(2αC)m_{2}^{2}\simeq M_{\rm Pl}^{2}/(2\alpha_{C}) are the scalar and spin-2 masses, respectively, and the last term is the universal quantum correction [Donoghue1994]. Submillimeter tests constrain extra Yukawa forces with strength 𝒪(1)\mathcal{O}(1) to have range λ0.1\lambda\lesssim 0.1 mm, implying M,m22×103M,m_{2}\gtrsim 2\times 10^{-3} eV [Kapner2007]. The inflationary value (42) corresponds to M1013M\sim 10^{13} GeV, and a loop-induced αC\alpha_{C} of order 10210^{-2}10010^{0} yields m2MPl/αC1018m_{2}\sim M_{\rm Pl}/\sqrt{\alpha_{C}}\gtrsim 10^{18} GeV, both well above laboratory reach. The sign of the m2m_{2} residue is negative (a ghost) at the perturbative level [Stelle1977]. As discussed below, this is acceptable within an effective description valid below m2m_{2}, and several proposals exist for its nonperturbative treatment [AnselmiPiva2018, Percacci2017].

VI.3 Gravitational waves

In the infrared limit where higher derivatives are negligible, gravitational waves propagate at the speed of light, cT=1c_{T}=1, consistent with GW170817 and its electromagnetic counterpart [GW170817]. The massive scalar and spin-2 modes would be excited only at frequencies ωM,m2\omega\gtrsim M,m_{2}, far above LIGO/Virgo/KAGRA bands. Nonlocal form factors induce a mild running of GNG_{N} with momentum but do not modify the luminal propagation of the massless graviton in the observed regime [Donoghue1994, BarvinskyVilkovisky1990].

VI.4 Running couplings and vacuum stability

The RG of the SM couplings is standard and can be combined with the gravitational running in (24)–(27). For the SM matter content at high scales, the trace-anomaly coefficients are

cSM\displaystyle c_{\rm SM} =1120Ns+120Nf+110Nv\displaystyle=\frac{1}{120}N_{s}+\frac{1}{20}N_{f}+\frac{1}{10}N_{v}
=1120×4+120×22.5+110×12\displaystyle=\frac{1}{120}\times 4+\frac{1}{20}\times 22.5+\frac{1}{10}\times 12
2.36,\displaystyle\simeq 2.36, (44)
aSM\displaystyle a_{\rm SM} =1360Ns+11360Nf+31180Nv\displaystyle=\frac{1}{360}N_{s}+\frac{11}{360}N_{f}+\frac{31}{180}N_{v}
=4360+11360×22.5+31180×12\displaystyle=\frac{4}{360}+\frac{11}{360}\times 22.5+\frac{31}{180}\times 12
2.77,\displaystyle\simeq 2.77, (45)

with Ns=4N_{s}=4 real scalars, Nf=22.5N_{f}=22.5 Dirac fermions (equivalently 45 Weyl), and Nv=12N_{v}=12 vectors. These numbers govern the one-loop running of the gravitational higher-derivative couplings in minimal subtraction [Duff1994, Buchbinder1992].

The Higgs quartic coupling λ\lambda runs to small or even negative values at 10910^{9}101110^{11} GeV depending on inputs [Buttazzo2013]. In this framework, curvature and nonminimal coupling ξ\xi feed back into the effective potential and can stabilize the vacuum during the inflationary epoch. In the late universe, where R¯0\bar{R}\to 0, the standard SM running applies. A full two-loop study with gravitational corrections will sharpen these statements; the one-loop structure already indicates how the feedback operates.

VII Predictivity and renormalization conditions

Predictivity arises in several ways:

  • The R2R^{2}-driven inflationary sector is sharply predictive. Once αR\alpha_{R} is matched to AsA_{s}, nsn_{s} and rr follow from (41) independently of microphysics [Starobinsky1980, Whitt1984, Planck2018].

  • The universal quantum correction to Newton’s potential in (43) has a fixed coefficient [Donoghue1994].

  • The strengths and ranges of Yukawa corrections to gravity are set by αR\alpha_{R} and αC\alpha_{C}, which are constrained by cosmology and laboratory experiments [Kapner2007].

Within the full coupled system, renormalization conditions encode the choice of vacuum. A natural set is:

(i)Ueffv|v=v,R¯=R¯=0,(ii)UeffR¯|v=v,R¯=R¯=0,\displaystyle\text{(i)}\quad\frac{\partial U_{\rm eff}}{\partial v}\Big|_{v=v_{\star},\bar{R}=\bar{R}_{\star}}=0,\quad\text{(ii)}\quad\frac{\partial U_{\rm eff}}{\partial\bar{R}}\Big|_{v=v_{\star},\bar{R}=\bar{R}_{\star}}=0,
(iii)Ueff(v,R¯)=ρvacobs.\displaystyle\text{(iii)}\quad U_{\rm eff}(v_{\star},\bar{R}_{\star})=\rho_{\rm vac}^{\rm obs}. (46)

Condition (iii) fixes the finite part of the cosmological constant counterterm. In practice, ρvacobs\rho_{\rm vac}^{\rm obs} is tiny compared to any particle physics scale, which requires a tuning familiar from semiclassical gravity. The first two conditions, however, are dynamical gap equations that constrain combinations of couplings and masses.

In applications, one matches the renormalized couplings to measured low-energy observables and then integrates the RG equations upward, solving the gap equations at the desired scale for the background. Conversely, given a high-scale boundary such as inflation, one runs down and predicts low-energy corrections. This is the standard effective field theory strategy implemented with the self-consistency postulate.

VIII Limitations and outlook

The framework developed here rests on a controlled one-loop calculation with the background field method and the heat-kernel expansion. Several limitations and tasks lie ahead.

First, while the one-loop structure is universal and robust, higher-loop and nonperturbative effects could modify the running of gravitational couplings. Functional RG studies suggest nontrivial fixed points for gravity [Weinberg1979, Reuter1998, Percacci2017, Codello2008], which, if realized, would render the entire system asymptotically safe and predictive at all scales. A careful match between the one-loop effective action and functional RG fixed points is an important next step.

Second, quadratic gravity is perturbatively renormalizable but contains a massive spin-2 ghost [Stelle1977]. As an effective theory valid below the ghost mass scale m2m_{2}, there is no inconsistency with observations. For UV completion, several avenues exist: the fakeon prescription [AnselmiPiva2018], asymptotic safety [Percacci2017], and nonlocal form factors that soften the propagator without introducing additional poles [BarvinskyVilkovisky1990]. Deciding among these requires further work.

Third, the smallness of the observed cosmological constant remains a deep puzzle. In this framework, it appears as a renormalization condition selecting the vacuum. Whether a symmetry or a dynamical mechanism enforces the tiny value is an open question not addressed here.

Fourth, threshold matching in curved space, nonlocal form factors, and the interplay with finite-temperature and nonequilibrium effects in the early universe are technically rich and phenomenologically relevant. The covariant perturbation theory provides the conceptual tools, and explicit computations in the full SM plus gravity are feasible.

Despite these caveats, the self-consistency principle, embedded in the effective action and realized with standard quantum field theory techniques, yields a cohesive and predictive picture that connects inflationary cosmology, precision gravity, and particle physics within a single calculational framework.

IX Conclusion

Starting from the postulate that the classical backgrounds of gravity and the Standard Model are macroscopic order parameters of a unified quantum vacuum sustained by its own fluctuations, a coherent, calculable framework is constructed. The renormalized Einstein, Yang–Mills, and Higgs equations emerge as vacuum equations of state derived from the stationarity of the one-particle-irreducible effective action. The heat-kernel expansion fixes the renormalization of all couplings and requires the presence of higher-derivative gravitational terms. A solvable O(N)O(N) model illustrates how the vacuum gap equations determine both curvature and condensates. The theory makes concrete predictions, notably Starobinsky-type inflation and universal quantum corrections to gravity, and it is consistent with current astrophysical and laboratory constraints. A systematic program of higher-loop calculations, functional RG studies, and phenomenological applications can now test and refine the picture.

Appendix A Conventions and useful variations

We use signature (,+,+,+)(-,+,+,+) and define Rρ=σμνμΓνσρνΓμσρ+R^{\rho}{}_{\sigma\mu\nu}=\partial_{\mu}\Gamma^{\rho}_{\nu\sigma}-\partial_{\nu}\Gamma^{\rho}_{\mu\sigma}+\cdots, Rμν=RρμρνR_{\mu\nu}=R^{\rho}{}_{\mu\rho\nu}, and R=g¯μνRμνR=\bar{g}^{\mu\nu}R_{\mu\nu}. The Weyl tensor is

Cμνρσ\displaystyle C_{\mu\nu\rho\sigma} =Rμνρσ12(g¯μρRνσg¯μσRνρ\displaystyle=R_{\mu\nu\rho\sigma}-\frac{1}{2}\bigl(\bar{g}_{\mu\rho}R_{\nu\sigma}-\bar{g}_{\mu\sigma}R_{\nu\rho}
g¯νρRμσ+g¯νσRμρ)+R6(g¯μρg¯νσg¯μσg¯νρ).\displaystyle\quad-\bar{g}_{\nu\rho}R_{\mu\sigma}+\bar{g}_{\nu\sigma}R_{\mu\rho}\bigr)+\frac{R}{6}\bigl(\bar{g}_{\mu\rho}\bar{g}_{\nu\sigma}-\bar{g}_{\mu\sigma}\bar{g}_{\nu\rho}\bigr). (47)

The basic variations used in the main text are

δ(g¯R)\displaystyle\delta\bigl(\sqrt{-\bar{g}}\,R\bigr) =g¯[(Gμν+g¯μνμν)δg¯μν],\displaystyle=\sqrt{-\bar{g}}\Bigl[\bigl(G_{\mu\nu}+\bar{g}_{\mu\nu}\Box-\nabla_{\mu}\nabla_{\nu}\bigr)\delta\bar{g}^{\mu\nu}\Bigr], (48)
δ(g¯R2)\displaystyle\delta\bigl(\sqrt{-\bar{g}}\,R^{2}\bigr) =g¯Hμν(R2)δg¯μν,\displaystyle=\sqrt{-\bar{g}}\,H^{(R^{2})}_{\mu\nu}\delta\bar{g}^{\mu\nu}, (49)
δ(g¯C2)\displaystyle\delta\bigl(\sqrt{-\bar{g}}\,C^{2}\bigr) =2g¯Bμνδg¯μν.\displaystyle=2\sqrt{-\bar{g}}\,B_{\mu\nu}\delta\bar{g}^{\mu\nu}. (50)

Appendix B Heat-kernel coefficients for standard fields

For a real scalar with nonminimal coupling ξ\xi, Dirac fermion, and gauge vector (in background gauge), the a2a_{2} densities contributing to C2C^{2} and E4E_{4} are, respectively,

a2scalar\displaystyle a_{2}^{\rm scalar} 1120C21360E4+12(ξ16)2R2,\displaystyle\supset\frac{1}{120}C^{2}-\frac{1}{360}E_{4}+\frac{1}{2}\Bigl(\xi-\tfrac{1}{6}\Bigr)^{2}R^{2}, (51)
a2Dirac\displaystyle a_{2}^{\rm Dirac} 120C211360E4,\displaystyle\supset\frac{1}{20}C^{2}-\frac{11}{360}E_{4}, (52)
a2vector\displaystyle a_{2}^{\rm vector} 110C231180E4,\displaystyle\supset\frac{1}{10}C^{2}-\frac{31}{180}E_{4}, (53)

in agreement with [Duff1994, Buchbinder1992]. Gauge and ghost contributions combine to the vector result.

Appendix C Bach tensor and equations of motion

The Bach tensor defined in (13) is traceless, Bμ=μ0B^{\mu}{}_{\mu}=0, and covariantly conserved in conformally flat spacetimes. In four dimensions, the metric variation of d4xgC2\int d^{4}x\sqrt{-g}\,C^{2} yields 2d4xgBμνδgμν2\int d^{4}x\sqrt{-g}\,B_{\mu\nu}\delta g^{\mu\nu}. The variation of R2R^{2} is given by (14).

Appendix D Standard Model one-loop beta functions

At one loop in the MS¯\overline{\rm MS} scheme, the SM gauge couplings obey

μdg1dμ\displaystyle\mu\frac{dg_{1}}{d\mu} =416g1316π2,\displaystyle=\frac{41}{6}\frac{g_{1}^{3}}{16\pi^{2}}, (54)
μdg2dμ\displaystyle\mu\frac{dg_{2}}{d\mu} =196g2316π2,\displaystyle=-\frac{19}{6}\frac{g_{2}^{3}}{16\pi^{2}}, (55)
μdg3dμ\displaystyle\mu\frac{dg_{3}}{d\mu} =7g3316π2,\displaystyle=-7\frac{g_{3}^{3}}{16\pi^{2}}, (56)

with GUT normalization for g1g_{1}. The Higgs quartic λ\lambda and the top Yukawa yty_{t} run as [MachacekVaughnI, MachacekVaughnII, Buttazzo2013]

μdλdμ\displaystyle\mu\frac{d\lambda}{d\mu} =116π2[24λ26yt4+38(2g24+(g22+g12)2)\displaystyle=\frac{1}{16\pi^{2}}\Bigl[24\lambda^{2}-6y_{t}^{4}+\frac{3}{8}\bigl(2g_{2}^{4}+(g_{2}^{2}+g_{1}^{2})^{2}\bigr)
9λg223λg12+12λyt2],\displaystyle\quad-9\lambda g_{2}^{2}-3\lambda g_{1}^{2}+12\lambda y_{t}^{2}\Bigr], (57)
μdytdμ\displaystyle\mu\frac{dy_{t}}{d\mu} =yt16π2[92yt21712g1294g228g32].\displaystyle=\frac{y_{t}}{16\pi^{2}}\Bigl[\frac{9}{2}y_{t}^{2}-\frac{17}{12}g_{1}^{2}-\frac{9}{4}g_{2}^{2}-8g_{3}^{2}\Bigr]. (58)

Appendix E One-loop effective potential with curvature dependence

For a scalar field with field-dependent mass m2(ϕ,R)=m02(ϕ)+ξRm^{2}(\phi,R)=m_{0}^{2}(\phi)+\xi R, the one-loop contribution to the effective potential in the limit of slowly varying backgrounds is

ΔU\displaystyle\Delta U =164π2m4(ϕ,R)[lnm2(ϕ,R)μ232]\displaystyle=\frac{1}{64\pi^{2}}m^{4}(\phi,R)\Bigl[\ln\frac{m^{2}(\phi,R)}{\mu^{2}}-\frac{3}{2}\Bigr]
+1192π2(ξ16)m2(ϕ,R)R\displaystyle\quad+\frac{1}{192\pi^{2}}\Bigl(\xi-\frac{1}{6}\Bigr)m^{2}(\phi,R)R
×[lnm2(ϕ,R)μ21]+,\displaystyle\quad\times\Bigl[\ln\frac{m^{2}(\phi,R)}{\mu^{2}}-1\Bigr]+\cdots, (59)

where dots denote higher-order terms in curvature and total derivatives [ParkerToms2009, Buchbinder1992, BirrellDavies1982]. This formula underlies (32) and the gap equations (36)–(37).