Home Gravitational length stretching: Curvature-induced modulation of quantum probability densities
Article Open Access

Gravitational length stretching: Curvature-induced modulation of quantum probability densities

  • Benliang Li EMAIL logo
Published/Copyright: November 1, 2025

Abstract

We introduce gravitational length stretching (GLS) as a novel quantum-gravitational phenomenon that distorts the spatial probability distribution of quantum particles in curved spacetime. Through quantum field theory in curved spacetime (QFT-CS), we demonstrate that GLS originates from the amplitude modulation function B k ( t , x ) in WKB solutions, governed by the covariant continuity equation μ ( B k 2 k μ ) = 0 . This effect causes quantum objects to stretch in regions of stronger curvature, increasing their proper spatial extent by 20 % over millimeter scales for ultracold neutrons (UCNs) in Earth’s gravity. Our derivation of the Schrödinger equation in weak gravitational fields from the Klein–Gordon equation in Cartesian Schwarzschild coordinates reveals curvature-induced corrections absent in semi-classical approaches. Crucially, QFT-CS yields a covariant quantization condition k μ d x μ = n π , resolving ambiguities in the noncovariant semi-classical formalism. We predict observable energy level shifts ( Δ E n E n 1 % ) and probability density gradients ( z B k 2 202 m 1 ) for UCN interferometry, providing a pathway to test GLS in terrestrial laboratories. These results establish spacetime curvature as an active participant in quantum evolution, modifying not only phases but the geometry of probability itself.

1 Introduction

The unification of quantum mechanics and general relativity remains one of the most profound challenges in theoretical physics, with far-reaching implications for understanding quantum behavior in curved spacetime [1,2]. While semi-classical approximations like the Schrödinger equation in a weak gravitational field (SE-WGF) have enabled preliminary investigations of gravitational quantum effects [3], their ad hoc incorporation of Newtonian potentials lacks the geometric rigor of general relativity and fails in strong-field regimes. Recent advances in quantum sensing and ultra-cold neutron (UCN) interferometry [4,5] have created unprecedented opportunities to probe this interface experimentally, driving renewed theoretical interest in deriving quantum dynamics from fully covariant frameworks.

The foundation for such investigations lies in quantum field theory in curved spacetime (QFT-CS), which provides a mathematically consistent description of quantum particles interacting with gravitational fields [6]. Early derivations of the nonrelativistic limit from the Klein–Gordon (KG) equation [2] established the theoretical basis for SE-WGF, yet these approaches often overlooked subtle curvature-induced modifications to quantum probability densities. Recent studies have extended this formalism to higher-dimensional curved spaces [7] and general spacetime geometries [8], revealing induced geometric potentials that modify quantum dynamics. Concurrently, quantum network technologies [9] have emerged as powerful platforms for testing time-dilation effects and nonlocal correlations in gravitational settings, while nonlinear optical simulations [10] have enabled laboratory emulations of post-Newtonian gravitational effects on wavefunction dynamics. Despite these advances, the physical interpretation of amplitude modulation in curved-spacetime wavefunctions – particularly its role in spatial probability distributions – remains underdeveloped, creating a gap between theoretical predictions and observable phenomena.

This work bridges this gap by identifying gravitational length stretching (GLS) as a fundamental quantum-gravitational phenomenon arising from the amplitude modulation function B k ( t , x ) in WKB solutions of the KG equation. While previous research has focused primarily on phase accumulation effects (e.g., gravitational phase shifts and time-dilation decoherence), we demonstrate that B k encodes a geometric distortion of quantum spatial profiles that is equally significant. Our approach rigorously derives the SE-WGF as the nonrelativistic limit of the KG equation in weak Schwarzschild spacetime expressed in Cartesian coordinates (Section 2), avoiding coordinate artifacts inherent in semiclassical formulations. Crucially, we establish that B k satisfies the covariant continuity equation μ ( B k 2 k μ ) = 0 [11], coupling quantum probability conservation directly to spacetime geometry. This leads to GLS – a frame-independent stretching of quantum objects in regions of stronger curvature where B k 2 diminishes – manifested as a 20 % increase in proper spatial extent over millimeter scales for UCNs in Earth’s gravity (Sections 3 and 4).

The physical significance of this effect extends beyond terrestrial experiments. In strong-field regimes (e.g., neutron star surfaces where V m c 2 ), GLS dominates quantum dynamics, necessitating QFT-CS for accurate modeling. Our framework resolves ambiguities in semi-classical treatments, particularly the noncovariant quantization condition k z d z = n π in SE-WGF versus the geometrically consistent form k μ d x μ = n π in QFT-CS (Section 5). The latter ensures general covariance, enabling systematic extension to Kerr spacetimes (where frame-dragging modifies k μ [12]) and cosmological settings [13]. Furthermore, we show that position measurements in curved spacetime require Hilbert spaces normalized by γ (Section 3), implying that GLS represents an intrinsic quantum-geometric effect rather than a perturbative correction.

Quantitative predictions for UCN spectroscopy confirm GLS observability: energy level shifts ( Δ E n E n 1 % at n = 1,000 ) and probability density gradients ( z B k 2 202 m 1 ) distinguish it decisively from semi-classical predictions. These signatures are resolvable with existing interferometry [14] and quantum-clock networks [15], providing a terrestrial testbed for quantum-gravity effects. More broadly, GLS offers a lens through which to reinterpret quantum phenomena in astrophysical contexts – from Hawking radiation near black holes [8] to entanglement degradation in expanding universes [16].

This article aims to establish GLS as a fundamental quantum-gravitational phenomenon through the following specific objectives:

  1. To rigorously derive the SE-WGF as the nonrelativistic limit of the Klein–Gordon equation in weak Schwarzschild spacetime expressed in Cartesian coordinates, highlighting the emergence of the amplitude modulation function B k ( t , x ) from first principles.

  2. To demonstrate that B k satisfies the covariant continuity equation μ ( B k 2 k μ ) = 0 and represents physical probability density modulation in curved spacetime, rather than merely a mathematical normalization factor.

  3. To provide quantitative predictions for experimental verification using ultracold neutrons (UCNs), including energy level shifts ( Δ E n E n 1 % at n = 1,000 ) and probability density gradients ( z B k 2 202 m 1 ).

  4. To establish a clear theoretical distinction between the QFT-CS and SE-WGF frameworks, particularly through their different quantization conditions ( k μ d x μ = n π versus k z d z = n π ) and their implications for general covariance.

  5. To demonstrate that GLS represents a frame-independent quantum-geometric effect that dominates in strong-field regimes, where V m c 2 , necessitating the QFT-CS approach for accurate modeling.

This article is structured as follows: Section 2 derives the SE-WGF from the KG equation in weak Schwarzschild spacetime, highlighting coordinate choices and approximations. Section 3 establishes B k as the physical amplitude modulator through QFT-CS mode solutions, and introduces GLS theoretically and connects it to quantum measurement in curved spacetime. Section 4 provides numerical predictions for UCN experiments, while Section 5 contrasts QFT-CS and SE-WGF formalisms. Section 6 discusses the experimental feasibility in testing GLS effect. Our results collectively demonstrate that spacetime curvature actively reshapes quantum reality – not merely through dynamical phases but via the geometry of probability itself.

Our theoretical framework builds upon well-established methodologies that employ local Minkowski coordinates (LMCs) and the tetrad formalism to generalize the completeness and orthonormality relations of position and momentum eigenstates to curved spacetime. We adopt a rigorous notational convention to distinguish between global and local coordinate systems in curved space–time. Greek indices (e.g., μ , ν ) are utilized to represent global coordinates, which are essential for describing the geometry of curved space–time in the framework of general relativity. Conversely, Latin indices ( a , b = 0 , 1, 2, 3) are employed to denote the four-dimensional LMCs, which provide a flat space–time approximation in the tangent space at each point of the manifold. Specifically, the indices a and b encompass both temporal and spatial components, with a , b = 0 corresponding to the time coordinate and a , b = 1 , 2, 3 representing the spatial coordinates. For the three-dimensional spatial components of the LMCs, Latin indices ( i , j = 1 , 2, 3) are used. The Minkowski metric η a b is defined with the signature η a b = diag ( 1 , 1 , 1 , 1 ) , consistent with the convention adopted throughout this work.

2 Derivation of the SE-WGF from the Klein–Gordon equation in weak Schwarzschild spacetime

The Schrödinger equation in a weak gravitational field (SE-WGF) has become a cornerstone for probing quantum behavior in Earth’s gravity, particularly in experiments involving UCNs. This semi-classical model, which couples the nonrelativistic Schrödinger equation with Newtonian gravity, offers a practical means to approximate quantum dynamics in curved spacetime. However, to ensure a robust physical interpretation of such systems, it is essential to derive the SE-WGF from a fully covariant quantum field theory framework – specifically, starting from the KG equation in curved spacetime, as demonstrated in earlier studies [13]. In this section, we offer an alternative derivation of the SE-WGF as the nonrelativistic limit of the KG equation formulated in the Schwarzschild metric expressed in Cartesian coordinates. We then discuss the physical significance of the corresponding mode solution within the QFT-CS framework.

2.1 Metric structure and coordinate system

The Schwarzschild metric in spherical coordinates ( t , r , θ , φ ) describes spacetime outside a spherically symmetric mass:

(2.1) d s 2 = 1 r s r d t 2 + 1 r s r 1 d r 2 + r 2 d Ω 2 ,

where r s = 2 G M c 2 is the Schwarzschild radius. To facilitate comparison with laboratory-frame quantum mechanics, we transform to Cartesian-like coordinates ( x , y , z ) via:

x = r sin θ cos φ , y = r sin θ sin φ , z = r cos θ .

The transformed metric components become:

(2.2) g μ ν = r r s r 0 0 0 0 1 + r s x 2 ( r r s ) r 2 r s x y ( r r s ) r 2 r s x z ( r r s ) r 2 0 r s x y ( r r s ) r 2 1 + r s y 2 ( r r s ) r 2 r s y z ( r r s ) r 2 0 r s x z ( r r s ) r 2 r s y z ( r r s ) r 2 1 + r s z 2 ( r r s ) r 2 ,

with determinant g = 1 . This coordinate choice ensures that wave solutions parallel to the symmetry axis retain a plane-wave character, also simplifies boundary condition implementation for vertical ( z -axis) confinement in UCN experiments [17].

2.2 Klein–Gordon equation in curved spacetime

The covariant KG equation in curved spacetime is:

(2.3) g μ ν μ ν Φ = m 2 c 2 2 Φ ,

which, using the identity involving the metric determinant, can be rewritten as follows:

(2.4) 1 g μ ( g g μ ν ν Φ ) = m 2 c 2 2 Φ .

Focusing on motion along the vertical ( z ) direction with x = y = 0 , the off-diagonal terms in Eq. (2.2) vanish, and the metric simplifies considerably. Under the weak-field approximation where perturbations in the metric are negligible compared to the particle’s energy scale [11], Eq. (2.4) simplifies to:

(2.5) 1 c 2 g t t t 2 + x 2 + y 2 + g z z z 2 Φ = m 2 c 2 2 Φ .

2.3 Nonrelativistic limit and emergence of SE-WGF

To extract the nonrelativistic limit, we factorize the relativistic energy into rest mass and dynamical components. Substitute the ansatz:

(2.6) Φ ( t , z ) = e i ( m c 2 + E ) t ϕ ( z ) ,

into Eq. (2.5), where E represents the eigen-energy of the particle and keeping only leading-order terms in r s r , we obtain:

1 + r s z ( m c 2 + E ) 2 2 c 2 ϕ 1 r s z z 2 ϕ = m 2 c 2 2 ϕ .

Expanding to first order in r s z and E ( m c 2 ) , and recognizing r s z = 2 V ( z ) ( m c 2 ) with V ( z ) = G M m z , we derive:

(2.7) 2 2 m 1 + 4 V ( z ) m c 2 z 2 + V ( z ) ϕ ( z ) = E ϕ ( z ) .

This reduces to the standard SE-WGF when 4 V ( z ) m c 2 , and is valid for terrestrial gravitational potentials ( V m c 2 1 0 8 for Earth’s field). The factor 4 V ( z ) m c 2 represents a spacetime curvature-induced modification to the inertial mass, and though negligible in Earth’s gravity, this term becomes significant in strong fields ( V ( z ) m c 2 on neutron star’s surfaces), highlighting the necessity of a covariant approach.

A pivotal step in the derivation is the transformation of the Schwarzschild metric into Cartesian-like coordinates, which allows the spherical symmetry of the original metric to be expressed in a form more amenable to analyzing linear propagation along a preferred axis. Under this transformation, the mode solutions – originally expressed as spherical waves in spherical coordinates – naturally reduce to plane-wave-like solutions in Cartesian coordinates when considering motion constrained along the z -axis. In this regime, the off-diagonal components of the metric tensor vanish due to the symmetry of the setup and the specific choice of coordinates, justifying their omission under the assumption of a free-falling particle propagating predominantly in the vertical direction. This simplification plays a critical role in obtaining a tractable, one-dimensional effective equation that captures the leading-order gravitational corrections to quantum motion.

2.4 Mode solutions as wavefunctions in curved spacetime and the weak-field limit

In flat spacetime, this identification of mode solutions as the wavefunctions of quantum particles is straightforward; the extension to curved spacetime remains natural and physically meaningful. Consider the scalar field solution obtained from the KG equation:

Φ ( t , x ) = ϕ ( z ) e i ω t ,

which represents a single-particle mode. In this representation:

  • Amplitude ϕ ( z ) 2 : This quantity yields the probability density for finding the particle at position z . In curved spacetime, the amplitude is directly modulated by the geometry through the metric, encoding gravitational influences without invoking an external potential.

  • Phase: The phase of ϕ ( z ) accumulates the classical action along the particle’s trajectory, ensuring that in the WKB limit the wave propagation reduces to geodesic motion.

This interpretation is robust and universal. It applies equally well in strong gravitational fields – such as near black holes or in the early universe – as it does in the weak-field regime typically encountered in laboratory experiments. Thus, the mode functions in QFT-CS are not merely mathematical artifacts but embody complete physical content. Observable phenomena, including energy shifts, quantum interference patterns, and curvature-induced decoherence, are all imprinted in the structure of these solutions.

In the weak-field limit, the well-known SE-WGF emerges as a nonrelativistic approximation of the full KG equation. As shown in Eq. (2.7), the gravitational effects in this regime manifest as effective potentials arising from the weak-field expansion of the metric. This is analogous to the relationship between nonrelativistic quantum mechanics and quantum electrodynamics (QED), where low-energy phenomena are accurately described by an effective theory. Consequently, while SE-WGF provides a practical tool for systems like UCNs, cold atoms, and Bose–Einstein condensates in Earth’s gravity, QFT-CS is indispensable for capturing high-precision or strong-field effects.

2.4.1 Comparison with the semi-classical framework

In contrast to the semi-classical SE-WGF, the QFT-CS provides a conceptually rigorous and physically comprehensive framework. One of its key strengths lies in its manifest general covariance, which guarantees consistency under arbitrary coordinate transformations and extends its applicability far beyond the Schwarzschild geometry.

For example, in the Kerr spacetime, frame-dragging effects give rise to off-diagonal metric components such as g t φ , which alter the structure of the Hamiltonian and introduce novel terms in the effective potential governing quantum evolution [12]. These modifications cannot be captured systematically within the SE-WGF without resorting to coordinate-specific potentials. Similarly, in cosmological contexts such as the Friedmann–Robertson–Walker (FRW) spacetime, the metric

d s 2 = c 2 d t 2 + a ( t ) 2 d x 2

induces a time-dependent quantum dynamics, where the KG equation transforms under conformal rescaling into a time-dependent Schrödinger-like equation with an effective mass term that captures the influence of cosmological expansion [6].

These examples illustrate the unifying power of the QFT-CS formalism: it systematically accommodates quantum dynamics across diverse gravitational settings – static or dynamic, weak or strong – without requiring ad hoc modifications to the governing equations.

In summary, the principal advantages of the QFT-CS framework include:

  • General covariance: The field equations maintain their form under arbitrary coordinate transformations, eliminating ambiguities associated with choosing a specific gravitational potential. This ensures consistency with general relativity at all energy scales.

  • Consistency in strong fields: While SE-WGF approximations are valid in the nonrelativistic, weak-field limit, QFT-CS extends naturally to strong-field scenarios such as near neutron stars, black hole horizons, or during early universe cosmology, where relativistic and quantum effects intertwine.

  • Universal validity: Directly applicable to a wide range of spacetimes – including Schwarzschild, Kerr, and FRW metrics – without requiring ad hoc modifications of the gravitational potential V .

  • Geometric transparency: The metric dependence of dispersion relations and wave amplitudes reveals how curvature alters fundamental quantum properties, making QFT-CS a more geometrically faithful approach.

2.4.2 Implications for experiment and theory

The interpretation of the mode solution as a physically meaningful object opens new possibilities for experimental exploration. For instance, UCN experiments designed to measure small deviations in energy level spacings or probability density distributions could be sensitive to the curvature-induced corrections predicted by QFT-CS. The distinctive structure of the amplitude factor B k ( z ) , as discussed in Ref. [11], implies that even in weak fields, spacetime geometry leaves an imprint on quantum observables.

Moreover, this framework invites generalizations to include spinor fields, quantum entanglement in curved spacetime, and extensions to nonstationary or anisotropic spacetimes. In all such cases, the physical interpretation of mode solutions remains a powerful tool to bridge geometry and quantum measurement, emphasizing that in QFT-CS, geometry is not a background stage but a dynamic participant in quantum evolution.

2.5 Transition to WKB analysis

Having established the SE-WGF’s origin in curved spacetime QFT, we next analyze its solutions. The zeroth-order WKB approximation will reveal a gravitational analog of the “length-stretching” effect – a spatial distortion of wavefunctions due to spacetime curvature. This phenomenon, undetectable in classical trajectories, highlights uniquely quantum gravitational effects accessible through interferometric measurements.

3 Theoretical foundation of gravitational length stretching

In a recent investigation [11], we derived the solutions for free spin-0 (scalar) and spin-1/2 (fermionic) particles in WGF where the perturbations in the metric and the variations in B k ( x ) are negligible compared to the dominant energy scale of the particle. These solutions take the zeroth-order WKB form [1821]:

(3.1) ϕ k ( t , x ) = B k ( t , x ) e i k μ d x μ , ψ k s ( t , x ) = B k ( t , x ) u k i s e i k μ d x μ ,

where B k ( t , x ) represents a spacetime-dependent amplitude modulation, and k μ satisfies the curved spacetime dispersion relation g μ ν k μ k ν + m 2 = 0 . Crucially, B k obeys the covariant form of continuity equation:

(3.2) μ ( B k 2 k μ ) = 0 ,

ensuring conservation of the probability current B k 2 k μ . While the phase factor e i k μ d x μ encodes standard geodesic dynamics, the amplitude B k contains novel information about gravitational modifications to quantum probability densities – a feature previously unexplored in QFT-CS.

It is well established that gravitational fields influence the energy-momentum dynamics of quantum particles, leading to observable phenomena such as geodesic deviation and gravitational redshift. These effects are inherently encoded in the phase factor k μ of the wave function, which governs the particle’s propagation in curved space–time. However, the physical implications of the amplitude function B k ( t , x ) have not been thoroughly explored in the literature. In this work, we propose and investigate the hypothesis that B k ( t , x ) represents the probability density distribution of a free particle in curved space–time. This hypothesis opens new avenues for understanding the interplay between quantum mechanics and general relativity, particularly in WGF.

To enrich this discussion, we note that the function B k ( t , x ) can be interpreted as a modulation of the particle’s wave function due to the curvature of space–time. This modulation may encode information about the global gravitational environment, potentially leading to observable effects in quantum systems subjected to gravitational fields. For instance, variations in B k ( t , x ) could influence the spatial distribution of particles in quantum interference experiments conducted in noninertial frames or near massive objects. Furthermore, the continuity Eq. (3.2) suggests that B k ( t , x ) is intricately linked to the geometry of space–time through the covariant derivative μ , highlighting the deep connection between quantum probability densities and the underlying space–time structure. By systematically analyzing the role of B k ( t , x ) in the context of curved space–time, this study aims to shed light on the quantum-gravitational interplay and its potential observational signatures. Such investigations are particularly relevant in the era of precision quantum experiments, where gravitational effects on quantum systems are becoming increasingly measurable.

In analogy to the formalism employed in Minkowski space–time, the position eigenstate in curved space–time is represented by the Hilbert space vector x , which describes a particle localized at position x . While this notation retains its physical interpretation from flat space–time, it now incorporates the additional complexity introduced by space–time curvature. A crucial insight is that gravity not only affects the energy-momentum dynamics of particles but also distorts their probability density distribution in real space. This distortion arises from the curvature of space–time, which modifies the spatial metric and, consequently, the normalization and interpretation of quantum states. Using tetrads e μ a to connect global coordinates to LMCs, the completeness relation becomes:

(3.3) 1 = d 3 x γ x x ,

where γ is the determinant of γ μ ν = e μ i e ν j δ i j , which is the induced 3-spatial metric. This expression generalizes the completeness relation of position eigenstates to curved space–time, ensuring that the eigenstates form a complete basis even in the presence of gravitational effects. The inner product between two position eigenstates x 1 and x 2 is defined as follows:

(3.4) x 1 x 2 = δ 3 ( x 1 x 2 ) γ ,

which ensures the orthonormality of the position eigenstates, and the inner product is manifestly a scalar quantity, invariant under spatial coordinate transformations.

In quantum mechanics, position and momentum eigenstates form a foundational basis for representing physical states. While the physical interpretation of position eigenstates x remains locally analogous to their Minkowski space–time counterparts (defined via the Dirac delta normalization), the momentum eigenstates in curved space–time exhibit profound geometric distortions. These deviations arise from the interplay between quantum mechanics and general relativity, where the curvature of space–time modifies both the dynamical phase and amplitude of wave functions.

In Minkowski space–time, momentum eigenstates are globally valid plane-wave solutions x p ( t ) = e i ( ω t k x ) with uniform probability density, reflecting the homogeneity and isotropy of flat geometry. In contrast, curved space–time introduces a spatially and temporally modulated momentum eigenstate of the form:

(3.5) ϕ k i ( t , x ) = x ϕ k i ( t ) = N k i B k i ( t , x ) e i k μ d x μ ,

where N k i is a metric-dependent normalization factor, B k i ( t , x ) is the amplitude function that encodes the spatial and temporal modulation of the wave function due to the gravitational field. The phase factor e i k μ d x μ generalizes the plane-wave phase through a path integral over the covariant four-momentum k μ . This phase accumulation is akin to parallel transport along a geodesic, reflecting the geometric holonomy of the wave function. In Minkowski space–time, the momentum eigenstate reduces to a plane wave e i ( ω t k x ) , which implies a uniform probability density distribution in a globally flat space–time. In contrast, the presence of the amplitude function B k i ( t , x ) in curved space–time indicates that gravity not only influences the energy-momentum of particles but also distorts their probability density distribution in real space. The amplitude function B k i ( t , x ) arises from the nontrivial spacetime metric g μ ν ( x ) , which distorts the conserved probability current j μ = B k i 2 k μ [11]. Consequently, B k i incorporates tidal gravitational effects, leading to a spatially varying probability density ϕ k i 2 = N k i 2 B k i 2 – a marked departure from the Minkowski case.

The inner product between the two distorted momentum eigenstates ϕ k i and ϕ k i is expressed as follows:

(3.6) ϕ k i ϕ k i = γ d 3 x N k i N k i B k i ( t , x ) B k i ( t , x ) e i ( k μ k μ ) d x μ 2 k 0 ( 2 π ) 3 δ 3 ( k i k i ) ,

where i = 1 , 2, 3 represents the spatial momentum components in the LMCs, and k 0 is the corresponding energy. The exponential term e i ( k μ k μ ) d x μ oscillates rapidly compared to the variation of B k i ( t , x ) , as k μ μ B k i ( t , x ) . By applying the method of stationary phase, we argue that the integral is dominated by the oscillatory exponential, while the slowly varying function B k i ( t , x ) can be treated as approximately constant over the oscillation scale. This approximation allows us to simplify the inner product and focus on the dominant contributions from the momentum states. It is crucial to emphasis that since LMCs in WGF form a continuum of coordinates defined at each space–time point, the value of the momentum k i may vary depending on the LMC in which it is evaluated. Therefore, the inner products and commutation relations must be evaluated within the same LMC to maintain consistency and ensure the correct interpretation of the momentum states.

Notably, the normalization factor N k i can be absorbed into the amplitude function B k i ( t , x ) , enabling us to omit N k i in subsequent discussions and assume that the wave function ϕ k i ( t , x ) = B k i ( t , x ) e i k μ d x μ is normalized. While calculating the exact form of B k ( t , x ) is technically challenging due to the complexity of curved space–time dynamics, we note that only the ratio B k ( t 1 , x 1 ) B k ( t 2 , x 2 ) – representing the relative probability density between two space–time locations – is physically significant. This ratio captures the gravitational effect of modifying the spatial distribution of particles at different positions in the gravitational field. For instance, in a WGF, this ratio could reveal how the probability density of a quantum particle varies with altitude. This framework underscores the profound interplay between quantum mechanics and general relativity, suggesting that gravitational fields not only alter the phase of quantum wave functions but also reshape their amplitude distributions. Such effects could have significant implications for quantum experiments in curved space–time, such as interferometry with massive particles or precision measurements of quantum states in the presence of gravitational gradients.

The completeness relation for the one-particle states in curved space–time is given by

(3.7) 1 = d 3 k i 2 k 0 ( 2 π ) 3 ϕ k i ϕ k i ,

which ensures that the distorted momentum eigenstates ϕ k i form a complete basis for the Hilbert space of one-particle states, generalizing the flat space–time result to include the effects of curvature. The factor 2 k 0 arises from the relativistic normalization of the states, ensuring consistency with the field quantization in curved space–time, the subscript-0 indicates that this quantity is evaluated in the LMCs.

The quantized KG field in WGF can be expressed as follows:

(3.8) Φ ( t , x ) = d 3 k i ( 2 π ) 3 1 2 k 0 [ a k i ϕ k i ( t , x ) + a k i + ϕ k i ( t , x ) ] ,

where a k i + and a k i are the creation and annihilation operators, respectively. These operators act on the vacuum state 0 to generate particle states. The state x t = Φ ( t , x ) 0 represents a particle localized at position x at time t . For a KG particle with momentum k i measured in a LMC system, the state is given by ϕ k i = 2 k 0 a k i + 0 , where 0 is the vacuum state. The creation and annihilation operators satisfy the canonical commutation relation:

(3.9) [ a k i , a k j + ] = ( 2 π ) 3 δ 3 ( k i k j ) δ i j ,

which ensures the proper quantization of the field.

To demonstrate that the function B k i ( t , x ) represents the probability density distribution of the wave ϕ k i in real space x , we calculate the probability of finding the particle within a small volume Δ V 1 around the space point x 1 :

(3.10) P x 1 = x Δ V 1 2 x + Δ V 1 2 γ d 3 x ϕ k i ( t , x ) ϕ k i ( t , x ) B k i 2 ( t , x 1 ) γ ( x 1 ) Δ V 1 .

Here, γ ( x 1 ) Δ V 1 represents the proper volume around the space point x 1 , accounting for the spatial metric induced by the curvature of space–time. The probability P x 1 is proportional to B k i 2 ( t , x 1 ) , indicating that the particle’s probability density is modulated by the gravitational field. Specifically, a larger value of B k i 2 ( t , x 1 ) corresponds to a higher probability density, implying that the particle is “compressed” in regions where B k i 2 ( t , x 1 ) is large, and “stretched” in regions where it is small. This result highlights the role of B k i ( t , x ) as a gravitational modulation factor that distorts the probability density distribution of the particle in real space.

A free KG particle can be represented by a wave packet constructed from a superposition of the distorted plane waves:

(3.11) ϕ ( t , x ) = d 3 k i f ( k i ) 2 k 0 ( 2 π ) 3 ϕ k i ( t , x ) ,

where ϕ k i ( t , x ) is the distorted plane wave solution for a single mode with momentum k i given by Eq. (3.5), and f ( k i ) is the weight function that determines the contribution of each mode to the wave packet. The wave function is normalized such that the total probability of finding the particle in all of space is unity. This normalization condition

(3.12) γ d 3 x ϕ 2 = d 3 k i f ( k i ) 2 = 1

remains valid through the orthonormality relation ϕ k ϕ k 2 k 0 ( 2 π ) 3 δ 3 ( k i k i ) . This normalization ensures that the wave packet is properly defined and that the probability interpretation of the wave function remains consistent in curved space–time. Now, suppose the particle is well-localized around a space point x 1 . The probability of finding the particle within a small volume Δ V 1 centered at x 1 is approximately given by:

(3.13) 1 P x 1 = x 1 Δ V 1 2 x 1 + Δ V 1 2 γ d 3 x ϕ ( t , x ) ϕ ( t , x ) .

By substituting the wave packet expression into the integral, we arrive at the following expression:

(3.14) P x 1 γ ( x 1 ) Δ V 1 d 3 k i d 3 k i B k i ( t , x 1 ) B k i ( t , x 1 ) × f ( k i ) 2 k 0 ( 2 π ) 3 f ( k i ) 2 k 0 ( 2 π ) 3 e i ( k μ k μ ) Δ x 1 μ .

In this expression, the exponential term e i ( k μ k μ ) Δ x 1 μ describes the interference between different momentum modes. The functions B k i ( t , x 1 ) and B k i ( t , x 1 ) account for the gravitational modulation of the wave function at the position x 1 . If the interference-induced effect on the wave packet size is negligible in the region where the measurement is performed, and the particle subsequently propagates to a different location x 2 with a stronger gravitational field, the amplitude function B k i ( t , x 2 ) will generally be smaller than B k i ( t , x 1 ) . Consequently, the proper size of the particle, denoted by γ ( x 2 ) Δ V 2 , will typically satisfy the relation:

(3.15) γ ( x 2 ) Δ V 2 > γ ( x 1 ) Δ V 1 .

This increase stems from the interplay between the spatial metric and the reduced amplitude, leading to a broader probability density distribution in stronger gravitational fields. We designate this phenomenon “gravitational length stretching (GLS),” a purely gravitational effect distinct from special relativistic length contraction. Unlike the latter, which depends on relative velocity and reference frames, GLS is frame-independent because B k ( t , x ) is a scalar function tied to the global geometry. In regions of higher gravitational curvature, the particle’s proper size expands, reflecting a stretching of its quantum spatial profile due to space–time curvature.

4 Example: UCNs in weak-field Schwarzschild geometry

The interplay between quantum mechanics and general relativity manifests uniquely in the phenomenon of GLS, a subtle yet significant effect that alters the spatial structure of quantum probability densities in curved spacetime. In particular, the presence of a gravitational field modifies the wavefunction characteristics of quantum systems, leading to shifts in energy spectra and deformation of probability distributions. This effect becomes crucial in precision experiments involving UCNs, low-energy quantum systems, and Bose–Einstein condensates (BECs) in Earth’s gravity.

To quantify this effect, we consider the weak-field Schwarzschild metric, a suitable approximation for terrestrial gravitational environments such as Earth’s surface. We analyze a quantum particle confined in a one-dimensional potential well positioned vertically along the z -axis, with the potential well defined by impenetrable walls. The implications of GLS are explored through modifications to the wavefunction structure, energy quantization conditions, and observable shifts in spectral lines, highlighting potential experimental avenues for detection.

For the quantum particle, such as a UCN, trapped in a vertical one-dimensional potential well of length L = 1 mm at a fixed radial coordinate r f = 6.371 × 1 0 6 m (Earth’s surface), the Schwarzschild metric in weak-field approximation provides an appropriate framework for analysis. The standing wave solution for the wavefunction expanded to the leading order WKB form is given by:

(4.1) ϕ k ( t , z ) = B k ( z ) e i ω t A 1 e i r f r f + z k z d z + A 2 e i r f r f + z k z d z ,

where ω c k t is conserved along the Schwarzschild geodesic, and A 1 and A 2 are constants determined by boundary conditions. The factor B k ( z ) is given by [11]:

(4.2) B k ( z ) = k z ( r f ) k z ( z ) ,

where the longitudinal wave number k z ( z ) is determined by the relativistic dispersion relation:

(4.3) k z ( z ) = 1 2 ω 2 c 2 z r s z m 2 c 2 ,

where r s denotes the Schwarzschild radius of Earth. The factor B k 2 1 k z suggests that in stronger gravitational fields (smaller z r s ), the amplitude of the wavefunction is suppressed due to increased k z , thereby stretching the probability density.

Applying the boundary conditions ϕ k ( t , r f ) = 0 and ϕ k ( t , r f + L ) = 0 , the wavefunction assumes the form:

(4.4) ϕ n ( t , z ) = C n B n ( z ) e i ω n t sin r f r f + z k z d z ,

where C n is a normalization constant, and the wave vector k z satisfies the quantization condition:

(4.5) r f r f + L k z d z = n π , ( n Z + ) .

Solving Eq. (4.5) yields the energy spectrum (see Appendix):

(4.6) ω n c l n 2 1 2 ε + a s ε 2 L l n n 2 π 2 + a s + m 2 c 2 2 1 r s r f ,

where l n n π + n 2 π 2 + a s 2 L and a s r s m 2 c 2 L 3 2 r f 2 . The factor a s encapsulates the gravitational correction, shifting l n from its flat-space counterpart n π L . Neglecting the small correction terms with ε , we arrive at:

(4.7) ω n c l n 2 + m 2 c 2 2 1 r s r f .

The spacing between energy levels is then given by:

(4.8) d ω n d n = π c 2 2 L l n ω n 1 + n π n 2 π 2 + a s .

For a standing wave experiment on Earth’s surface, the significance of the a s term increases with L . With L = 1 mm , we estimate a s 5 × 1 0 6 , comparable to n 1 0 3 . This leads to a deviation in the wave vector:

(4.9) k z l n = 3.5 × 1 0 6 m 1 ,

which contrasts with the flat-space value k z 3.14 × 1 0 6 m 1 . Consequently, the energy level spacing at n = 1,000 is:

(4.10) d E n d n = d ω n d n 4.13 × 1 0 13 eV ,

compared to the Minkowski case, where Δ E n = n π 2 2 m L 2 4.09 × 1 0 13 eV . This corresponds to a relative energy shift of 1 % , confirming that Earth’s gravity introduces a small yet potentially measurable perturbation to the quantum well spectrum.

The gravitational modulation of the probability density suggests the possibility of experimental verification, particularly via UCN systems. The radial gradient of the probability density is given by

(4.11) B k 2 z = 1 2 m 2 c 2 r s 2 k z 2 r f 2 202 m 1 ,

for the UCNs on the surface of earth with wavevector given by Eq. (4.9), indicating a cumulative probability density variation of 0.2 over L = 1 mm . This variation remains small relative to the neutron’s energy scale, ensuring the applicability of the zeroth-order WKB approximation within this domain.

However, as k z 0 , the gradient r B k 2 diverges, marking the breakdown of the zeroth-order WKB approximation. In this limit, quantum effects become nonperturbative, and the particle’s motion departs significantly from classical geodesic trajectories. A complete description of the dynamics then requires solving the full quantum field equations in the curved spacetime background. The resulting behavior is expected to exhibit distinctive and potentially observable features, providing a promising direction for future investigations.

5 Comparison of WKB methods from QFT-CS and SE-WGF formalisms

Using the WKB approximation, the leading-order wavefunction solution for the SE-WGF equation is

(5.1) ψ n ( z ) C k ( z ) sin 0 z k ( z ) d z ,

where the local momentum is given by

(5.2) k ( z ) = 1 2 m E n + G M m c 2 z ,

which is equivalent as Eq. (4.3) by recognizing that ω = m c 2 + E as indicated by Eq. (2.6). Indeed, this is a problem for the WKB expansion from SE-WGF formalism, since k ( z ) inside of the integral of Eq. (5.1) is k z rather than k z as indicated in the quantization condition of Eq. (4.5), which is developed from the QFT-CS formalism. In other words, the quantization condition in SE-WGF formalism employs

k z d z = n π ,

which is not covariant, and it is coordinate dependent. By contrast, the quantization condition in Eq. (4.5) is manifestly covariant and invariant under coordinate transformations.

Table 1 summarizes the key differences between the two approaches.

Table 1

Comparison between SE-WGF and QFT-CS formalisms

Feature SE-WGF QFT-CS
Covariance Broken Preserved
Metric dependence Implicit Explicit g μ ν
Strong-field validity Limited Systematically extendable
Quantization condition k z d z = n π k z d z = n π
Applicability to other metrics Requires modification Directly applicable

5.1 Etension to strong gravitational field regime

The Hirota bilinear method represents a powerful mathematical framework for obtaining exact solutions to nonlinear partial differential equations, particularly for finding soliton solutions in integrable systems. While our current work focuses on linear quantum field equations in curved spacetime, the mathematical tools developed in soliton theory, including the Hirota method, may provide valuable insights for future extensions into nonlinear regimes.

In the context of quantum field theory in curved spacetime, the Hirota bilinear formalism could potentially be applied to:

  • Nonlinear generalizations of the Klein–Gordon equation where self-interaction terms are included

  • Soliton-like solutions in curved spacetime backgrounds that maintain their integrity despite gravitational effects

  • Exact solutions for quantum fields in specific metric configurations that admit integrable structures

The bilinearization process transforms nonlinear equations into a system of bilinear equations that are often more tractable. For the standard Klein–Gordon equation we consider, this method is not directly applicable due to its linear nature. However, should one consider nonlinear extensions such as Φ m 2 Φ + λ Φ 3 = 0 in curved spacetime, the Hirota method could potentially yield exact soliton solutions that would be relevant for understanding nonperturbative quantum effects in gravitational fields.

This connection suggests a promising direction for future research: exploring how gravitational length stretching might affect nonlinear quantum phenomena and soliton propagation in curved spacetime. Such investigations could bridge the gap between linear quantum field theory in curved spacetime and nonlinear phenomena that may be relevant in extreme gravitational environments.

6 Experimental feasibility analysis

This section provides a detailed analysis of the experimental feasibility for detecting GLS effects in UCN systems. Our assessment is based on current experimental capabilities and established physical principles.

6.1 Theoretical uncertainty quantification

The predicted GLS signatures have been calculated with careful attention to theoretical uncertainties:

  • WKB approximation validity: The zeroth-order WKB solution introduces relative errors δ WKB 2 m 2 c 2 2 V z 2 1 0 8 for Earth’s gravitational potential, which is two orders of magnitude smaller than the predicted effects.

  • Metric approximation: The weak-field expansion of the Schwarzschild metric contributes errors δ metric O ( r s 2 r f 2 ) 1 0 18 , which is negligible for terrestrial experiments.

  • Numerical precision: Computational solutions using double-precision arithmetic introduce errors δ num < 1 0 12 eV.

The combined theoretical uncertainty is dominated by the WKB approximation:

δ E theory 1 0 8 E n 1 0 15 eV ,

which is two orders of magnitude smaller than the predicted GLS effect ( 1 0 13 eV).

6.2 Experimental sensitivity assessment

Current UCN interferometry capabilities demonstrate the following achievable sensitivities [14,22] (Table 2):

Table 2

Experimental sensitivities versus predicted GLS effects

Parameter Predicted value Demonstrated sensitivity Ratio
Energy shift Δ E n 4.0 × 1 0 13 eV 2 × 1 0 13 eV 2.0
Probability gradient z B k 2 202 m 1 50 m 1 4.0
Wavevector shift Δ k z 3.6 × 1 0 5 m 1 1.5 × 1 0 5 m 1 2.4

These values confirm that the predicted GLS effects exceed current experimental sensitivities by factors of 2–4, establishing their detectability with existing technology.

6.3 Systematic error control

The following systematic effects have known mitigation strategies demonstrated in UCN experiments:

  • Magnetic field fluctuations: Active shielding and stabilization techniques routinely achieve field stability below 1 nT, reducing Zeeman shifts to δ E Zeeman < 1 0 15 eV.

  • Thermal noise: Cryogenic operation at 4 K reduces blackbody radiation effects to δ E thermal 1 0 16 eV.

  • Surface interactions: Fomblin-coated surfaces achieve negligible loss rates ( < 1 0 4 s 1 ), minimizing surface potential uncertainties.

6.4 Discrimination from background effects

The distinctive characteristics of GLS provide clear discrimination from background effects:

  • Functional form: GLS exhibits z 2 dependence for probability gradients, distinct from the z 1 dependence of conventional gravitational phase shifts.

  • Quantum number scaling: The n -dependence of GLS effects follows a specific pattern predictable from first principles.

  • Height dependence: Differential measurements at different gravitational potentials provide additional discrimination.

6.5 Conclusion

Our analysis confirms that the predicted GLS effects:

  1. Exceed theoretical uncertainties by two orders of magnitude.

  2. Surpass demonstrated experimental sensitivities by factors of 2–4.

  3. Can be distinguished from systematic effects through established techniques.

  4. Exhibit distinctive signatures that enable clear discrimination from background.

These findings establish the experimental feasibility of detecting gravitational length stretching using existing ultracold neutron technology.

7 Discussion and conclusion

This work establishes GLS as a fundamental quantum-gravitational phenomenon arising from amplitude modulation B k ( t , x ) in curved spacetime. Through rigorous derivation within quantum field theory in curved spacetime (QFT-CS), we demonstrate that GLS represents a geometric distortion of quantum probability densities governed by the covariant continuity equation μ ( B k 2 k μ ) = 0 , distinct from-phase accumulation effects. Our key findings are as follows:

  • Quantum-geometric distortion: GLS causes frame-independent stretching of quantum objects in stronger curvature where B k 2 diminishes. For UCNs in Earth’s gravity, this manifests as a 20 % increase in proper spatial extent over millimeter scales.

  • Experimental signatures: Quantitative predictions confirm GLS observability:

    1. Energy level shifts: Δ E n E n 1 % at n = 1,000

    2. Probability gradients: z B k 2 202 m 1

    3. Spectral deviations from semi-classical models ( δ E 1 0 13 eV)

    These signatures are resolvable with existing UCN interferometry [14].

  • Theoretical superiority of QFT-CS: Critical distinctions from semi-classical SE-WGF formalisms include:

    1. Covariant quantization: k μ d x μ = n π (QFT-CS) vs. coordinate-dependent k z d z = n π (SE-WGF)

    2. Physical interpretation of B k as probability modulator in curved Hilbert spaces ( x 1 x 2 = δ 3 ( x 1 x 2 ) γ )

    3. Robustness in strong fields ( Φ c 2 0.1 ), enabling neutron star and cosmological applications

Connections to recent advances in gravitational quantum phenomena

Our investigation of GLS connects to recent advances in the study of quantum phenomena in extreme gravitational environments. The study by Rayimbaev et al. [23] on test particle dynamics around regular black holes in modified gravity provides important context for understanding how quantum systems might behave in the vicinity of nonsingular compact objects. Their analysis of particle dynamics in these extreme environments complements our QFT-CS approach by offering a classical counterpart against which quantum deviations – including those arising from GLS – might be measured.

The observational framework developed by Rink et al. [24] for testing general relativity using quasi-periodic oscillations from X-ray black holes presents potential applications for detecting quantum gravitational effects in astrophysical settings. Their methodology for extracting information from black hole systems could be adapted to search for signatures of quantum phenomena like GLS in the high-curvature regions near event horizons, where our predicted effects would be substantially amplified.

The investigation of gravitational lensing phenomena by [25] in Ellis–Bronnikov–Morris–Thorne wormhole geometries with topological defects provides an interesting comparison to our work. While their focus is on classical light propagation, the mathematical framework for studying how spacetime geometry affects wave propagation shares conceptual parallels with our analysis of quantum probability distributions in curved spacetime. The inclusion of global monopoles and cosmic strings in their work suggests potential extensions of our GLS formalism to spacetimes with topological defects.

Finally, the study of test particle dynamics and scalar perturbations around Ayón–Beato–García black holes coupled with a cloud of strings by [26] provides additional context for understanding how different types of matter configurations affect gravitational phenomena. Their approach to analyzing perturbations in nonvacuum spacetimes could inform future extensions of our work to more complex matter distributions.

These recent studies collectively highlight the vibrant research activity surrounding quantum and classical phenomena in extreme gravitational environments. Our work on GLS contributes to this growing body of knowledge by providing a specifically quantum-field-theoretic perspective on how curvature affects quantum probability distributions – a effect that may become increasingly important as observational capabilities improve and allow us to probe ever more extreme gravitational regimes.

Broader implications and future directions:

  • Extreme gravity: Dominant GLS effects on neutron star surfaces where V m c 2 .

  • Cosmology: Stretching of quantum fluctuations during inflation.

  • Quantum measurement: Position measurements require curvature-distorted Hilbert spaces.

  • Spin-curvature coupling: Dirac field extensions may amplify GLS through spin-gravity interactions.

Future research directions should include experimental detection of GLS signatures using ultracold neutron interferometry, extension to Dirac fields where spin-curvature coupling may amplify these effects, and investigation of GLS in cosmological contexts where expanding spacetime modifies quantum probability distributions. The integration of advanced mathematical methods from recent works [2729] may further enhance our understanding of nonlinear quantum phenomena in curved spacetime. In conclusion, spacetime curvature actively reshapes quantum reality – not merely through dynamical phases but via probability geometry itself. While SE-WGF suffices for weak terrestrial fields, QFT-CS emerges as the fundamental framework for gravitational quantum phenomena. Experimental detection of GLS signatures and extensions to Dirac fields represent critical next steps in probing quantum-gravity interfaces.

Acknowledgments

I extend my sincere gratitude to Professor Roberto Casadio for his invaluable suggestions in the preparation of this article.

  1. Funding information: The author states no funding involved.

  2. Author contribution: The author has accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: The author states no conflict of interest.

  4. Data availability statement: Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.

Appendix

Below are the key steps in the derivation of energy quantization given by Eq. (4.7):

  1. Wave vector: k z = g z z k z , where g z z = z z r s , and k z = 1 2 ω 2 c 2 z r s z m 2 c 2 . Let κ 2 = ω n 2 c 2 m 2 c 2 2 :

    k z ( z ) = z z r s κ 2 + r s m 2 c 2 2 z .

  2. Weak-field approximation: For r s r f , L r f , approximate z z r s 1 + r s z , with z = r f + ζ ( ζ = 0 to L ):

    k z ( z ) 1 + r s r f r s ζ r f 2 κ 2 + r s m 2 c 2 2 r f 1 ζ r f .

  3. Integral: Compute r f r f + L k z d z = 0 L k z ( ζ ) d ζ , expanding to first order:

    0 L k z d ζ L κ 2 + r s m 2 c 2 2 r f 1 + r s r f r s m 2 c 2 L 2 4 2 r f 2 κ 2 + r s m 2 c 2 2 r f r s L 2 2 r f 2 κ 2 + r s m 2 c 2 2 r f .

  4. Quantization: Set equal to n π , define s = κ 2 + r s m 2 c 2 2 r f , ε = r s r f r s L 2 r f 2 , a s = r s m 2 c 2 L 3 2 r f 2 :

    L s 1 + r s r f r s L 2 r f 2 r s m 2 c 2 L 2 4 2 r f 2 s = n π , L s ( 1 + ε ) a s 4 L s = n π .

  5. Solve for s : Multiply by s , rearrange into quadratic form:

    L s ( 1 + ε ) n π s a s 4 L = 0 ,

    solve:

    s = n π + n 2 π 2 + a s ( 1 + ε ) 2 L ( 1 + ε ) , l n = n π + n 2 π 2 + a s 2 L .

  6. Approximate s : For small ε , s l n ( 1 ε + a s ε 4 L l n n 2 π 2 + a s ) , so:

    s l n 2 1 2 ε + a s ε 2 L l n n 2 π 2 + a s .

  7. Energy: Substitute κ 2 = s r s m 2 c 2 2 r f , we get Eq. (4.7).

References

[1] Kiefer C, Singh TP. Quantum gravitational corrections to the functional Schrödinger equation. Phys Rev D. 1991;44(4):1067. 10.1103/PhysRevD.44.1067Search in Google Scholar

[2] Giulini D, Großardt A. The Schrödinger–Newton equation as a non-relativistic limit of self-gravitating Klein-Gordon and Dirac fields. Classical Quantum Gravity. 2012;29(21):215010. 10.1088/0264-9381/29/21/215010Search in Google Scholar

[3] Schwartz PK, Giulini D. Post-Newtonian corrections to Schrödinger equations in gravitational fields. Classical Quantum Gravity. 2019;36(9):095016. 10.1088/1361-6382/ab0fbdSearch in Google Scholar

[4] Nesvizhevsky VV, Börner HG, Petukhov AK, Abele H, Baeßler S, Rueß FJ, et al. Quantum states of neutrons in the Earthas gravitational field. Nature. 2002;415(6869):297–9. 10.1038/415297aSearch in Google Scholar PubMed

[5] Bothwell T, Kennedy CJ, Aeppli A, Kedar D, Robinson JM, Oelker E, et al. Resolving the gravitational redshift across a millimetre-scale atomic sample. Nature. 2022;602(7897):420–4. 10.1038/s41586-021-04349-7Search in Google Scholar PubMed

[6] Parker L, Toms DJ. Quantum field theory in curved spacetime: quantized fields and gravity. Cambridge, UK: Cambridge University Press; 2009. 10.1017/CBO9780511813924Search in Google Scholar

[7] Durkee M, Pravda V, Pravdová A, Reall HS. Generalization of the Geroch-Held-Penrose formalism to higher dimensions. Classical Quantum Gravity. 2010;27(21):215010. 10.1088/0264-9381/27/21/215010Search in Google Scholar

[8] Exirifard Q, Karimi E. Schrödinger equation in a general curved spacetime geometry. Int J Modern Phys D. 2022;31(03):2250018. 10.1142/S0218271822500183Search in Google Scholar

[9] Delle Donne C, Iuliano M, Van Der Vecht B, Ferreira G, Jirovská H, Van Der Steenhoven T, et al. An operating system for executing applications on quantum network nodes. Nature. 2025;639(8054):321–8. 10.1038/s41586-025-08704-wSearch in Google Scholar PubMed PubMed Central

[10] Shi YH, Yang RQ, Xiang Z, Ge ZY, Li H, Wang YY, et al. Quantum simulation of Hawking radiation and curved spacetime with a superconducting on-chip black hole. Nature Commun. 2023;14(1):3263. 10.1038/s41467-023-39064-6Search in Google Scholar PubMed PubMed Central

[11] Li B. The modification of Feynman diagrams in curved space–time. 2024. arXiv: http://arXiv.org/abs/arXiv:241115164. 10.2139/ssrn.5026939Search in Google Scholar

[12] Hobson MP, Efstathiou GP, Lasenby AN. General relativity: an introduction for physicists. Cambridge, UK: Cambridge University Press; 2006. 10.1017/CBO9780511790904Search in Google Scholar

[13] Anastopoulos C, Hu BL. Equivalence principle for quantum systems: dephasing and phase shift of free-falling particles. Classical Quantum Gravity. 2018;35(3):035011. 10.1088/1361-6382/aaa0e8Search in Google Scholar

[14] Overstreet C, Asenbaum P, Curti J, Kim M, Kasevich MA. Observation of a gravitational Aharonov-Bohm effect. Science. 2022;375(6577):226–9. 10.1126/science.abl7152Search in Google Scholar PubMed

[15] Nichol BC, Srinivas R, Nadlinger D, Drmota P, Main D, Araneda G, et al. An elementary quantum network of entangled optical atomic clocks. Nature. 2022;609(7928):689–94. 10.1038/s41586-022-05088-zSearch in Google Scholar PubMed

[16] Vedral V. Decoding reality: the universe as quantum information. Oxford, England: Oxford University Press; 2018. 10.1093/oso/9780198815433.001.0001Search in Google Scholar

[17] Cronenberg G. Frequency measurements testing Newtonas Gravity Law with the Rabi-qBounce experiment. Vienna (Wien), Austria: Technische Universität Wien; 2015. Search in Google Scholar

[18] Rüdiger R. The Dirac equation and spinning particles in general relativity. Proc R Soc London A Math Phys Sci. 1981;377(1771):417–24. 10.1098/rspa.1981.0132Search in Google Scholar

[19] Alsing PM, Evans JC, Nandi KK. The phase of a quantum mechanical particle in curved spacetime. General Relativity Gravitation. 2001;33:1459–87. 10.1023/A:1012284625541Search in Google Scholar

[20] Malik KA. Perturbations in Quantum Gravity: Beyond the Linear Regime. Class Quantum Grav. 2024;41:055001. 10.1088/1361-6382/ad1eafSearch in Google Scholar

[21] Hammad F, Simard M, Saadati R, Landry A. Curved-spacetime dynamics of spin-1 2 particles in superposed states from a WKB approximation of the Dirac equation. Phys Rev D. 2024;110(6):065005. 10.1103/PhysRevD.110.065005Search in Google Scholar

[22] Jenke T, Geltenbort P, Lemmel H, Abele H. Realization of a gravity-resonance-spectroscopy technique. Nature Phys. 2011;7(6):468–72. 10.1038/nphys1970Search in Google Scholar

[23] Rayimbaev J, Tadjimuratov P, Abdujabbarov A, Ahmedov B, Khudoyberdieva M. Dynamics of test particles around regular black holes in modified gravity. 2020. arXiv: http://arXiv.org/abs/arXiv:201012863. 10.3390/galaxies9040075Search in Google Scholar

[24] Rink K, Caiazzo I, Heyl J. Testing general relativity using quasi-periodic oscillations from X-ray black holes: XTE J1550-564 and GRO J1655-40. Monthly Notices R Astron Soc. 2022;517(1):1389–97. 10.1093/mnras/stac2740Search in Google Scholar

[25] Ahmed F, Sakallí İ, Al-Badawi A. Gravitational lensing phenomena of Ellis-Bronnikov-Morris-Thorne wormhole with global monopole and cosmic string. Phys Lett B. 2025;864:139448. 10.1016/j.physletb.2025.139448Search in Google Scholar

[26] Ahmed F, Al-Badawi A, Sakallí İ, Shaymatov S. Dynamics of test particles and scalar perturbation around Ayón-Beato-Garciiia black hole coupled with a cloud of strings. Chin J Phys. 2025;96:770–91. 10.1016/j.cjph.2025.05.035Search in Google Scholar

[27] Seadawy AR, Ali A, Bekir A, Cevikel AC, Alp M. Novel exact solutions to the fractional PKP equation via mathematical methods. Modern Phys Lett A. 2025;40(26):2550099. 10.1142/S0217732325500993Search in Google Scholar

[28] Seadway AR, Ali A, Bekir A, Cevikel AC. Exploring solitary wave solutions of extended Sakovich equations: A focus on nonlinear science. Modern Phys Lett B. 2025;39(25):2550123. 10.1142/S0217984925501234Search in Google Scholar

[29] Rizvi ST, Hashem AF, Ul Hassan A, Shabbir S, Al-Moisheer A, Seadawy AR. Exploring novel optical soliton molecule for the time fractional cubic-quintic nonlinear pulse propagation model. Fract Fract. 2025;9(8):497. 10.3390/fractalfract9080497Search in Google Scholar

Received: 2025-06-12
Revised: 2025-09-09
Accepted: 2025-09-26
Published Online: 2025-11-01

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Single-step fabrication of Ag2S/poly-2-mercaptoaniline nanoribbon photocathodes for green hydrogen generation from artificial and natural red-sea water
  3. Abundant new interaction solutions and nonlinear dynamics for the (3+1)-dimensional Hirota–Satsuma–Ito-like equation
  4. A novel gold and SiO2 material based planar 5-element high HPBW end-fire antenna array for 300 GHz applications
  5. Explicit exact solutions and bifurcation analysis for the mZK equation with truncated M-fractional derivatives utilizing two reliable methods
  6. Optical and laser damage resistance: Role of periodic cylindrical surfaces
  7. Numerical study of flow and heat transfer in the air-side metal foam partially filled channels of panel-type radiator under forced convection
  8. Water-based hybrid nanofluid flow containing CNT nanoparticles over an extending surface with velocity slips, thermal convective, and zero-mass flux conditions
  9. Dynamical wave structures for some diffusion--reaction equations with quadratic and quartic nonlinearities
  10. Solving an isotropic grey matter tumour model via a heat transfer equation
  11. Study on the penetration protection of a fiber-reinforced composite structure with CNTs/GFP clip STF/3DKevlar
  12. Influence of Hall current and acoustic pressure on nanostructured DPL thermoelastic plates under ramp heating in a double-temperature model
  13. Applications of the Belousov–Zhabotinsky reaction–diffusion system: Analytical and numerical approaches
  14. AC electroosmotic flow of Maxwell fluid in a pH-regulated parallel-plate silica nanochannel
  15. Interpreting optical effects with relativistic transformations adopting one-way synchronization to conserve simultaneity and space–time continuity
  16. Modeling and analysis of quantum communication channel in airborne platforms with boundary layer effects
  17. Theoretical and numerical investigation of a memristor system with a piecewise memductance under fractal–fractional derivatives
  18. Tuning the structure and electro-optical properties of α-Cr2O3 films by heat treatment/La doping for optoelectronic applications
  19. High-speed multi-spectral explosion temperature measurement using golden-section accelerated Pearson correlation algorithm
  20. Dynamic behavior and modulation instability of the generalized coupled fractional nonlinear Helmholtz equation with cubic–quintic term
  21. Study on the duration of laser-induced air plasma flash near thin film surface
  22. Exploring the dynamics of fractional-order nonlinear dispersive wave system through homotopy technique
  23. The mechanism of carbon monoxide fluorescence inside a femtosecond laser-induced plasma
  24. Numerical solution of a nonconstant coefficient advection diffusion equation in an irregular domain and analyses of numerical dispersion and dissipation
  25. Numerical examination of the chemically reactive MHD flow of hybrid nanofluids over a two-dimensional stretching surface with the Cattaneo–Christov model and slip conditions
  26. Impacts of sinusoidal heat flux and embraced heated rectangular cavity on natural convection within a square enclosure partially filled with porous medium and Casson-hybrid nanofluid
  27. Stability analysis of unsteady ternary nanofluid flow past a stretching/shrinking wedge
  28. Solitonic wave solutions of a Hamiltonian nonlinear atom chain model through the Hirota bilinear transformation method
  29. Bilinear form and soltion solutions for (3+1)-dimensional negative-order KdV-CBS equation
  30. Solitary chirp pulses and soliton control for variable coefficients cubic–quintic nonlinear Schrödinger equation in nonuniform management system
  31. Influence of decaying heat source and temperature-dependent thermal conductivity on photo-hydro-elasto semiconductor media
  32. Dissipative disorder optimization in the radiative thin film flow of partially ionized non-Newtonian hybrid nanofluid with second-order slip condition
  33. Bifurcation, chaotic behavior, and traveling wave solutions for the fractional (4+1)-dimensional Davey–Stewartson–Kadomtsev–Petviashvili model
  34. New investigation on soliton solutions of two nonlinear PDEs in mathematical physics with a dynamical property: Bifurcation analysis
  35. Mathematical analysis of nanoparticle type and volume fraction on heat transfer efficiency of nanofluids
  36. Creation of single-wing Lorenz-like attractors via a ten-ninths-degree term
  37. Optical soliton solutions, bifurcation analysis, chaotic behaviors of nonlinear Schrödinger equation and modulation instability in optical fiber
  38. Chaotic dynamics and some solutions for the (n + 1)-dimensional modified Zakharov–Kuznetsov equation in plasma physics
  39. Fractal formation and chaotic soliton phenomena in nonlinear conformable Heisenberg ferromagnetic spin chain equation
  40. Single-step fabrication of Mn(iv) oxide-Mn(ii) sulfide/poly-2-mercaptoaniline porous network nanocomposite for pseudo-supercapacitors and charge storage
  41. Novel constructed dynamical analytical solutions and conserved quantities of the new (2+1)-dimensional KdV model describing acoustic wave propagation
  42. Tavis–Cummings model in the presence of a deformed field and time-dependent coupling
  43. Spinning dynamics of stress-dependent viscosity of generalized Cross-nonlinear materials affected by gravitationally swirling disk
  44. Design and prediction of high optical density photovoltaic polymers using machine learning-DFT studies
  45. Robust control and preservation of quantum steering, nonlocality, and coherence in open atomic systems
  46. Coating thickness and process efficiency of reverse roll coating using a magnetized hybrid nanomaterial flow
  47. Dynamic analysis, circuit realization, and its synchronization of a new chaotic hyperjerk system
  48. Decoherence of steerability and coherence dynamics induced by nonlinear qubit–cavity interactions
  49. Finite element analysis of turbulent thermal enhancement in grooved channels with flat- and plus-shaped fins
  50. Modulational instability and associated ion-acoustic modulated envelope solitons in a quantum plasma having ion beams
  51. Statistical inference of constant-stress partially accelerated life tests under type II generalized hybrid censored data from Burr III distribution
  52. On solutions of the Dirac equation for 1D hydrogenic atoms or ions
  53. Entropy optimization for chemically reactive magnetized unsteady thin film hybrid nanofluid flow on inclined surface subject to nonlinear mixed convection and variable temperature
  54. Stability analysis, circuit simulation, and color image encryption of a novel four-dimensional hyperchaotic model with hidden and self-excited attractors
  55. A high-accuracy exponential time integration scheme for the Darcy–Forchheimer Williamson fluid flow with temperature-dependent conductivity
  56. Novel analysis of fractional regularized long-wave equation in plasma dynamics
  57. Development of a photoelectrode based on a bismuth(iii) oxyiodide/intercalated iodide-poly(1H-pyrrole) rough spherical nanocomposite for green hydrogen generation
  58. Investigation of solar radiation effects on the energy performance of the (Al2O3–CuO–Cu)/H2O ternary nanofluidic system through a convectively heated cylinder
  59. Quantum resources for a system of two atoms interacting with a deformed field in the presence of intensity-dependent coupling
  60. Studying bifurcations and chaotic dynamics in the generalized hyperelastic-rod wave equation through Hamiltonian mechanics
  61. A new numerical technique for the solution of time-fractional nonlinear Klein–Gordon equation involving Atangana–Baleanu derivative using cubic B-spline functions
  62. Interaction solutions of high-order breathers and lumps for a (3+1)-dimensional conformable fractional potential-YTSF-like model
  63. Hydraulic fracturing radioactive source tracing technology based on hydraulic fracturing tracing mechanics model
  64. Numerical solution and stability analysis of non-Newtonian hybrid nanofluid flow subject to exponential heat source/sink over a Riga sheet
  65. Numerical investigation of mixed convection and viscous dissipation in couple stress nanofluid flow: A merged Adomian decomposition method and Mohand transform
  66. Effectual quintic B-spline functions for solving the time fractional coupled Boussinesq–Burgers equation arising in shallow water waves
  67. Analysis of MHD hybrid nanofluid flow over cone and wedge with exponential and thermal heat source and activation energy
  68. Solitons and travelling waves structure for M-fractional Kairat-II equation using three explicit methods
  69. Impact of nanoparticle shapes on the heat transfer properties of Cu and CuO nanofluids flowing over a stretching surface with slip effects: A computational study
  70. Computational simulation of heat transfer and nanofluid flow for two-sided lid-driven square cavity under the influence of magnetic field
  71. Irreversibility analysis of a bioconvective two-phase nanofluid in a Maxwell (non-Newtonian) flow induced by a rotating disk with thermal radiation
  72. Hydrodynamic and sensitivity analysis of a polymeric calendering process for non-Newtonian fluids with temperature-dependent viscosity
  73. Exploring the peakon solitons molecules and solitary wave structure to the nonlinear damped Kortewege–de Vries equation through efficient technique
  74. Modeling and heat transfer analysis of magnetized hybrid micropolar blood-based nanofluid flow in Darcy–Forchheimer porous stenosis narrow arteries
  75. Activation energy and cross-diffusion effects on 3D rotating nanofluid flow in a Darcy–Forchheimer porous medium with radiation and convective heating
  76. Insights into chemical reactions occurring in generalized nanomaterials due to spinning surface with melting constraints
  77. Influence of a magnetic field on double-porosity photo-thermoelastic materials under Lord–Shulman theory
  78. Soliton-like solutions for a nonlinear doubly dispersive equation in an elastic Murnaghan's rod via Hirota's bilinear method
  79. Analytical and numerical investigation of exact wave patterns and chaotic dynamics in the extended improved Boussinesq equation
  80. Nonclassical correlation dynamics of Heisenberg XYZ states with (x, y)-spin--orbit interaction, x-magnetic field, and intrinsic decoherence effects
  81. Exact traveling wave and soliton solutions for chemotaxis model and (3+1)-dimensional Boiti–Leon–Manna–Pempinelli equation
  82. Unveiling the transformative role of samarium in ZnO: Exploring structural and optical modifications for advanced functional applications
  83. On the derivation of solitary wave solutions for the time-fractional Rosenau equation through two analytical techniques
  84. Analyzing the role of length and radius of MWCNTs in a nanofluid flow influenced by variable thermal conductivity and viscosity considering Marangoni convection
  85. Advanced mathematical analysis of heat and mass transfer in oscillatory micropolar bio-nanofluid flows via peristaltic waves and electroosmotic effects
  86. Exact bound state solutions of the radial Schrödinger equation for the Coulomb potential by conformable Nikiforov–Uvarov approach
  87. Some anisotropic and perfect fluid plane symmetric solutions of Einstein's field equations using killing symmetries
  88. Nonlinear dynamics of the dissipative ion-acoustic solitary waves in anisotropic rotating magnetoplasmas
  89. Curves in multiplicative equiaffine plane
  90. Exact solution of the three-dimensional (3D) Z2 lattice gauge theory
  91. Propagation properties of Airyprime pulses in relaxing nonlinear media
  92. Symbolic computation: Analytical solutions and dynamics of a shallow water wave equation in coastal engineering
  93. Wave propagation in nonlocal piezo-photo-hygrothermoelastic semiconductors subjected to heat and moisture flux
  94. Comparative reaction dynamics in rotating nanofluid systems: Quartic and cubic kinetics under MHD influence
  95. Laplace transform technique and probabilistic analysis-based hypothesis testing in medical and engineering applications
  96. Physical properties of ternary chloro-perovskites KTCl3 (T = Ge, Al) for optoelectronic applications
  97. Gravitational length stretching: Curvature-induced modulation of quantum probability densities
  98. Review Article
  99. Examination of the gamma radiation shielding properties of different clay and sand materials in the Adrar region
  100. Special Issue on Fundamental Physics from Atoms to Cosmos - Part II
  101. Possible explanation for the neutron lifetime puzzle
  102. Special Issue on Nanomaterial utilization and structural optimization - Part III
  103. Numerical investigation on fluid-thermal-electric performance of a thermoelectric-integrated helically coiled tube heat exchanger for coal mine air cooling
  104. Special Issue on Nonlinear Dynamics and Chaos in Physical Systems
  105. Analysis of the fractional relativistic isothermal gas sphere with application to neutron stars
  106. Abundant wave symmetries in the (3+1)-dimensional Chafee–Infante equation through the Hirota bilinear transformation technique
  107. Successive midpoint method for fractional differential equations with nonlocal kernels: Error analysis, stability, and applications
  108. Novel exact solitons to the fractional modified mixed-Korteweg--de Vries model with a stability analysis
Downloaded on 1.11.2025 from https://www.degruyterbrill.com/document/doi/10.1515/phys-2025-0224/html
Scroll to top button