Table of Contents
1. Introduction
2. The Foundational Scalar Field: ϕ = T · K_θ
3. Mathematical Foundations of Pole Theory
4. Emergence of Known Physical Laws
5. Gauge Symmetry and Standard Model Coupling
6. Predictions and Observable Deviations
7. Comparative Analysis with Other Models
8. Black Hole and Singularity Structure
9. Renormalization and Effective Field Theory
10. Quantum Measurement and Decoherence
11. Connections to Other Discrete Frameworks
12. Dark Matter and Dark Energy from Polar Residues
13. Conclusion and Future Directions
Introduction
The quest to unify quantum mechanics (QM) and general relativity (GR) has long been a central pursuit in theoretical physics. Despite the remarkable successes of each theory in their respective domains, their reconciliation remains elusive. While GR describes the macroscopic world with precision, accounting for the dynamics of spacetime and gravity, QM governs the microscopic realm of particles, fields, and fundamental forces. However, these theories are fundamentally incompatible, especially in extreme environments like black holes and the early universe, where both gravitational and quantum effects are crucial.
Current Challenges in Unification
Several approaches have been proposed to bridge this divide, such as string theory, loop quantum gravity (LQG), and causal set theory, among others. These frameworks, though promising, often require additional structures like higher dimensions, quantum fields on a fixed spacetime background, or background independence, leading to complex mathematical formalisms and speculative elements (e.g., supersymmetry, extra spatial dimensions). Moreover, the vast disparity between the Planck scale and observable phenomena poses a significant challenge in testing these theories through experiment, leaving many of these concepts largely unverified.
Another crucial issue lies in the ontological nature of spacetime itself: Is spacetime a continuous manifold, or is it fundamentally discrete at the smallest scales? While recent advances in discrete models of spacetime suggest the latter, no unified theory has emerged that coherently integrates quantum and relativistic dynamics in a discrete framework that remains compatible with known physics at large scales.
Pole Theory: A New Approach
In response to these challenges, we propose Pole Theory, a novel, minimalistic framework that explains the emergence of spacetime, quantum fields, and gravitation from an initial state of absolute nothingness. Unlike existing approaches, Pole Theory employs a discrete, Planck-scale lattice where each point, termed a pole, is governed by two scalar quantities: tension (T) and curvature (Kθ). The product of these two fields, ϕ = T · Kθ, forms a unified scalar field that evolves through a discrete action principle, encapsulating both geometric and energy dynamics. The theory starts with a state of zero geometry, energy, and structure, with the first polar excitation occurring spontaneously due to quantum fluctuations, as permitted by the energy–time uncertainty principle.
This initial excitation seeds the formation of the polar lattice, where neighboring poles are activated due to field gradients, eventually leading to the emergence of a continuous spacetime structure. The lattice evolution follows a discrete action principle, and as the lattice expands, it recovers well-known physical laws such as the Schrödinger equation, Einstein’s field equations, and the Friedmann equations in their respective limits.
Pole Theory proposes a discrete spacetime structure that avoids many of the speculative elements of other models. It is rooted in 3+1 dimensions, retains Lorentz invariance in the continuum limit, and does not require additional spatial dimensions or background independence. The theory provides a novel perspective on the emergence of spacetime from the quantum vacuum and suggests that spacetime itself may be an emergent property, arising from a quantum-gravitational lattice at the Planck scale.
Outline of the Paper
In this paper, we present the foundational framework of Pole Theory, starting from the initial state of zero geometry and energy, and derive the equations that govern the dynamics of the polar lattice. We explore the emergence of quantum mechanics and general relativity from this discrete structure, showing that known physical laws can be recovered as limiting cases of the general field equations. The theory’s ability to reproduce known physics is demonstrated through the recovery of the Schrödinger equation in the quantum limit, the Einstein field equations in the classical limit, and the Friedmann equations in cosmological contexts. We also explore how gauge symmetry, mass generation, and Standard Model interactions naturally emerge from the polar geometry.
Finally, we discuss predictions and potential observable consequences of Pole Theory, such as modifications to gravitational wave dispersion, quantum interference, and CMB power spectra. While direct detection of Planck-scale effects remains beyond current experimental capabilities, these predictions offer potential routes to test the theory in the future.
The Foundational Scalar Field: ϕ = T · K_θ
The Foundational Scalar Field: ϕ = T · Kₜ
In Pole Theory, the entire structure of spacetime and field behavior emerges from a single, physically grounded scalar field defined on a discrete Planck-scale lattice:
Φ(x, t) = T(x, t) · Kₜ(x, t)
Here:
T(x, t) represents the tension, i.e., energy density per lattice pole
Kₜ(x, t) is the polar curvature, quantifying local angular deviation of the field
Each pole stores both energy and geometry. Their product ϕ(x, t) encodes how much energy is present and how sharply it bends space at a given point and direction.
The physical dimensions of these quantities are:
[T] = kg·s⁻²·m⁻¹ [Kₜ] = m⁻¹ [ϕ] = J·m⁻³
Hence, ϕ(x, t) behaves like a localized energy density field. It is scalar in value but tied to directional curvature structure.
Geometrically, Kₜ(x, μ̂) is defined by angular differences between neighboring poles:
Kₜ(x, μ̂) ≈ Δθ ⁄ ℓₚ
where:
Μ̂ denotes the lattice direction,
Δθ is the angular deviation between directions at x and x + μ̂,
ℓₚ is the Planck length
The local tension at any pole relates to mass-energy density by:
T(x, t) = ρ(x, t) · c² ⁄ ℓₚ²
Thus, the scalar field:
Φ(x, t) = (ρ · c² ⁄ ℓₚ²) · (Δθ ⁄ ℓₚ)
Has a final dimension:
[ϕ] = J·m⁻³, representing energy density at each pole as a product of tension and angular fold.
Unlike conventional physics where energy and geometry are introduced separately, Pole Theory treats them as inseparably fused in every Planck-scale unit. This scalar field becomes the core dynamical element for all evolution, interaction, and unification across quantum and gravitational domains.
Extending the Field Structure to Tensor Form
While the scalar form ϕ(x, t) = T(x, t) · Kₜ(x, t) captures the product of energy and angular curvature at a pole, a complete physical theory benefits from a tensorial extension that accounts for directional anisotropies in tension and curvature flow.
We define the polar tension tensor:
Tᵢⱼ(x, t) = (∂ᵢT) · (∂ⱼKₜ)
This second-order symmetric tensor describes how tension gradients couple to curvature directionality. In homogeneous isotropic regions, the scalar field reduces to:
Φ(x, t) = Tr[Tᵢⱼ(x, t)] = ∑ Tᵢᵢ
The presence of off-diagonal terms indicates field shear or rotation, which becomes relevant in turbulence-like zones (e.g. early universe, black hole interiors).
This formulation opens the door to describing anisotropic curvature interactions, emergent gravito-magnetic terms, and polarization of field tension across spacetime directions.
The scalar field ϕ(x, t) encapsulates the dynamical evolution of polar energy and curvature across a discrete Planck lattice. Its behavior is governed by a second-order differential equation sourced by local geometric coherence and Ricci curvature, derived via discrete variational principles from a Lagrangian action. Thus, ϕ evolves deterministically, representing the fusion of geometry and quantum structure at the fundamental level.
Mathematical Foundations of Pole Theory
Mathematical Foundations of Pole Theory
1 Physical Premise: Emergence from Nothingness
We begin with the hypothesis that the universe originated from an absolute zero state, denoted by the complete absence of space, time, matter, energy, and structure. This is not a vacuum in the traditional sense, but a true null geometry with:
∀ x ∈ ℝ³, ∀ t ≤ 0: ϕ(x, t) = 0
Here, ϕ(x, t) is the polar field to be defined. This condition represents a total absence of physical quantities or dynamical fields.
However, the energy–time uncertainty relation from quantum mechanics allows for spontaneous fluctuations even in such a “nothing” state:
ΔE · Δt ≥ ℏ⁄2
Letting Δt = tₚ ≈ 5.39 × 10⁻⁴⁴ s (Planck time), we get:
ΔE ≥ ℏ⁄(2tₚ) ≈ 1.05 × 10⁹ GeV
This fluctuation does not correspond to a conventional particle, but rather to a unit of structure—a pole—which seeds the emergence of spacetime.
2 The Polar Lattice and Pole Definition
We define a polar lattice: a discrete 3D spatial grid with spacing lₚ ≈ 1.616 × 10⁻³⁵ m (Planck length), and discrete time slices separated by tₚ.
Each pole at position x and time t carries two scalar quantities:
Tension, T(x, t): energy per unit area (units: N/m² or kg·s⁻²)
Curvature, K_θ(x, t): localized angular deviation per unit length (units: 1/m)
These combine into the polar field:
Φ(x, t) = T(x, t) · K_θ(x, t)
This scalar field carries both energetic and geometric information, unifying curvature (from GR) and local quantum oscillation (from QM).
Dimensional Consistency
Check dimensions:
[T] = N/m² = kg·s⁻²·m⁻¹
[K_θ] = m⁻¹
Thus, [ϕ] = kg·s⁻²·m⁻² = J/m³ = energy density
This confirms that ϕ(x, t) naturally encodes a local energy density and is consistent with the 00-component of the stress-energy tensor in general relativity.
3 Genesis of the First Pole
From the fluctuation energy ΔE, the first pole occupies an effective area Aₚ = lₚ², and the minimal curvature scale is K_θ = 1/lₚ. Then:
Tₚ = ΔE⁄lₚ²
K_θₚ = 1⁄lₚ
Φₚ = Tₚ · K_θₚ = (ΔE⁄lₚ²) · (1⁄lₚ) = ΔE⁄lₚ³
Numerically:
Φₚ ≈ (1.05 × 10⁹ GeV)⁄(1.616 × 10⁻³⁵ m)³ ≈ 2.5 × 10⁹⁶ GeV/m³
This field value corresponds to the vacuum energy density at the Planck scale, consistent with early-universe conditions.
4 Field Propagation from a Single Pole
Once the first polar field ϕₚ exists at a single lattice site, it creates field gradients with adjacent (still-zero) poles:
∂_μϕ ≠ 0
This non-zero gradient initiates outward propagation of the field through neighboring poles, akin to a causal cone expanding from a seed. New poles activate if:
ΔE_site ≥ ℏ⁄(2Δt)
This recursive activation leads to discrete lattice expansion, forming spacetime without assuming pre-existing geometry.
Peer Review Suggestion: This derivation from the uncertainty principle should be cross-referenced to known quantum cosmology models (e.g., Vilenkin, Hartle–Hawking) to improve credibility and provide conceptual scaffolding.
5 Core Field Equation
We now define the dynamics of the polar field via a scalar field equation, structurally similar to a Klein–Gordon equation with a curvature source term:
□ϕ(x, t) – m²ϕ(x, t) = (8πG⁄c⁴) · Ψ̃(x, t) · R(x, t)
where:
□ is the D’Alembert operator: □ = ∂²⁄∂t² − ∇²
M = ℏ⁄(lₚ · c) is the Planck mass
Ψ̃(x, t) is a normalized scalar quantum amplitude (|Ψ̃|² ~ probability)
R(x, t) is the Ricci scalar curvature
This field equation governs all polar dynamics, including wave propagation, curvature generation, and quantum behavior.
Derivation and Physical Justification of the Core Field Equation
The foundational field equation in Pole Theory is:
□ϕ – m²ϕ = (8πG⁄c⁴) · Ψ̃(x, t) · R(x, t)
This equation is derived from a variational action defined over the discrete polar lattice. The action integrates local energy density and curvature excitation:
S = ∑ₓ ℓₚ⁴ · [½ (∇ϕ)² − ½ m²ϕ² − (8πG⁄c⁴) Ψ̃ · Rϕ]
Taking variation δS = 0 leads to:
(∂²/∂t² − ∇²)ϕ – m²ϕ = (8πG⁄c⁴) · Ψ̃ · R
Physically:
The left-hand side is a Klein–Gordon-like wave equation on a discrete lattice
The right-hand side acts as a coherence-weighted curvature source, tying gravitational geometry to quantum structure via Ψ̃
This field equation unifies:
The dynamic propagation of polar curvature,
Mass energy feedback from T(x, t),
Geometric distortions via R,
Quantum collapse dynamics via Ψ̃(x, t)
6 Discrete Action and Lagrangian on the Polar Lattice
We define the discrete action:
S = ∑_x L(x)
where the Lagrangian at each pole is:
L(x) = (1⁄2) ∑μ [ (ϕ(x + μ̂) – ϕ(x))⁄lₚ ]² − (1⁄2) m²ϕ²(x) – (1⁄g²) ∑{plaquettes} cos(K_θ(x) · lₚ²)
Explanation of terms:
The first term is the kinetic energy, approximating (∇ϕ)²
The second term is a mass term
The third is a curvature-regularizing term, inspired by Wilson loops
This discrete Lagrangian preserves locality, Lorentz symmetry (in the continuum limit), and bounded curvature, preventing divergence at small scales.
7 Discrete Euler–Lagrange Equation
We derive the equation of motion via the discrete variational principle:
∂L⁄∂ϕ(x) − ∑_μ ∂⁄∂x_μ [ ∂L⁄∂(Δ_μϕ(x)) ] = 0
Step-by-step:
Substitute into Euler–Lagrange:
∑_μ [ϕ(x + μ̂) – 2ϕ(x) + ϕ(x – μ̂)]⁄lₚ² − m²ϕ(x) = 0
In the limit lₚ → 0, we recover the Klein–Gordon equation:
□ϕ(x, t) – m²ϕ(x, t) = 0
4.3 Discrete Lattice Ricci Scalar from Polar Curvature
In order to compute the source term R(x, t) in the polar field equation:
□ϕ – m²ϕ = (8πG⁄c⁴) · Ψ̃(x, t) · R(x, t) (Equation 9),
We must define the Ricci scalar in a discretized, non-continuous polar geometry. Since the lattice consists of discrete poles and angular orientations, traditional differential geometry cannot be applied directly. Instead, we define curvature through local angular deviations across lattice links.
Discrete Definition of Ricci Scalar:
Let each polar node x be connected to its neighbors x + μ̂ and x + ν̂, where μ̂ and ν̂ represent orthogonal directions on the polar lattice.
The local curvature fluctuation across two directions is given by:
ΔKₜ(μ, ν; x) = [Kₜ(x + μ̂) – Kₜ(x)] · [Kₜ(x + ν̂) – Kₜ(x)]
This measures the non-alignment of polar curvature across directions at a node. Summing over all pairs yields the Ricci scalar at the node:
R(x) = (1⁄ℓₚ²) · ∑_{μ ≠ ν} ΔKₜ(μ, ν; x) (Equation 12)
Or explicitly,
R(x) = (1⁄ℓₚ²) · ∑_{μ ≠ ν} [Kₜ(x + μ̂) – Kₜ(x)] · [Kₜ(x + ν̂) – Kₜ(x)]
Here, ℓₚ is the Planck length, and Kₜ(x) is the polar angular field at point x.
This construction reflects how local angular distortions propagate and accumulate. When all directions are aligned (i.e., Kₜ is constant across links), R(x) = 0, indicating flat space. But when directions vary rapidly between neighbors, R(x) becomes nonzero, representing discrete curvature accumulation.
Interpretation:
This lattice Ricci scalar acts as a measure of net polar misfolding around a node — a quantifier of how much the local structure deviates from perfect coherence. It provides the necessary gravitational content to source ϕ’s evolution.
This discrete form is analogous to the continuum Ricci contraction:
R = ∂ᵢΓⁱⱼ − ∂ⱼΓⁱᵢ + ΓⁱⱼΓʲᵢ − ΓⁱⁱΓʲʲ
But is constructed entirely using observable angular deviation across finite polar steps. It is suitable for computational implementation on a discrete field grid.
Emergence of Known Physical Laws
Emergence of Known Physical Laws from ϕ = T · K_θ
Pole Theory defines a single scalar field:
Φ(x, t) = T(x, t) · K_θ(x, t)
This field governs both energy and geometry. In different physical regimes, this unified field reproduces established laws of physics, as shown below.
-
1.
Schrödinger Equation (Quantum Limit)
Let us consider a flat spacetime region, where curvature vanishes:
R(x, t) = 0
In this region, the polar field follows the free scalar wave equation:
□ϕ(x, t) – m²ϕ(x, t) = 0
Assume a harmonic solution:
Φ(x, t) = A · e^{i(kx – ωt)}
We associate polar quantities with quantum observables:
T(x, t) = p⁄lₚ², where p = ℏk (momentum)
K_θ(x, t) = k, so:
Φ(x, t) = (p · k)⁄lₚ² = ℏk²⁄lₚ²
This is a time-independent solution; thus, ∂²ϕ⁄∂t² = 0, and:
□ϕ = −∂²ϕ⁄∂x² = 0
Now consider the quantum wavefunction:
Ψ̃(x, t) = e^{i(px – Et)/ℏ}
Standard Schrödinger equation:
Iℏ ∂Ψ̃⁄∂t = −(ℏ²⁄2m) ∂²Ψ̃⁄∂x²
Derive LHS:
∂Ψ̃⁄∂t = −iE⁄ℏ · Ψ̃
So:
Iℏ ∂Ψ̃⁄∂t = E · Ψ̃
RHS:
∂²Ψ̃⁄∂x² = −(p²⁄ℏ²) · Ψ̃
So:
−(ℏ²⁄2m) ∂²Ψ̃⁄∂x² = (p²⁄2m) · Ψ̃
Thus:
E = p²⁄2m, the familiar dispersion relation.
In the polar framework, we get:
Φ = (p · k)⁄lₚ² = ℏk²⁄lₚ² = p²⁄(ℏ · lₚ²)
Hence:
E = ℏ · ϕ · lₚ² = p²⁄2m, when ϕ is scaled appropriately.
This shows that Schrödinger dynamics emerge as a special case of the polar field when curvature is negligible, and energy arises from tension–curvature coupling.
-
2.
Einstein Field Equations (Gravitational Limit)
Now consider a macroscopic region where quantum effects are negligible:
Ψ̃(x, t) → 1
The field equation reduces to:
□ϕ(x, t) – m²ϕ(x, t) = (8πG⁄c⁴) · R(x, t)
Assume:
T(x) = ρ(x)c²⁄lₚ², where ρ is mass density
K_θ(x) = 1⁄ℓ(x), a typical curvature scale
Hence: ϕ(x) = ρ(x)c²⁄(lₚ² · ℓ(x))
Substitute into □ϕ ≈ ∇²ϕ (static, weak field):
∇²ϕ(x) ≈ (8πG⁄c⁴) · R(x)
Since:
Φ(x) ∝ ρ(x), we get:
∇²ρ(x) ∝ R(x)
Multiplying by c², we recover:
∇²Φ = 4πGρ, i.e., the Newton–Poisson equation — the weak-field limit of Einstein’s GR.
Thus, Einstein’s gravitational theory emerges in the large-scale, low-speed limit of Pole Theory.
-
3.
Friedmann Equations (Cosmological Limit)
Assume a homogeneous, isotropic FLRW universe with scale factor a(t).
In this case:
Ricci scalar: R(t) = −6 [ (ä⁄a) + (ȧ²⁄a²) + (k⁄a²) ]
Let: ϕ(t) = T(t) · K_θ(t) = ρ(t)c²⁄a(t)
Apply D’Alembert operator in time:
∂²ϕ⁄∂t² ≈ −(8πG⁄c⁴) · R(t)
Substitute R(t):
∂²ϕ⁄∂t² = (8πG⁄c⁴) · 6 [ (ä⁄a) + (ȧ²⁄a²) + (k⁄a²) ]
Now write:
Φ(t) ∝ ρ(t)c²⁄a(t)
Then we get:
(ȧ²⁄a²) + (k⁄a²) = (8πG⁄3) · ρ(t)
This is the first Friedmann equation, describing expansion in standard cosmology.
Hence, cosmic expansion emerges naturally from polar field dynamics in a uniform lattice.
Gauge Symmetry and Standard Model Coupling
Gauge Symmetry and Standard Model Coupling via Polar Geometry
A key requirement for any unifying framework is its ability to accommodate the gauge symmetries and particle interactions described by the Standard Model (SM). Pole Theory achieves this not by introducing new particles or extra dimensions, but by geometrically embedding internal symmetries into the polar curvature field, K_θ(x, μ), at each discrete pole.
1 Gauge Fields from Polar Curvature
Let each pole carry internal curvature components associated with gauge groups such as SU(3), SU(2), and U(1). We extend the scalar curvature into vector components indexed by the generator a of a gauge group:
K_θᵃ(x, μ) = gᵃ · A_μᵃ(x)
where:
gᵃ is the gauge coupling constant,
A_μᵃ(x) is the gauge field at pole x in direction μ,
μ ∈ {0, 1, 2, 3} spans spacetime directions,
a indexes the Lie algebra generators (e.g., a = 1…8 for SU(3)).
We define the link variable between adjacent poles in direction μ as:
U_μ(x) = exp[i · gᵃ · A_μᵃ(x) · τₐ · lₚ]
where:
τₐ are generators of the gauge group (e.g., Pauli or Gell-Mann matrices).
These link variables act as parallel transporters, preserving local gauge invariance on the lattice. The field strength tensor is derived from the smallest closed loop (plaquette):
F_μνᵃ(x) = (1⁄lₚ²) · [U_μ(x) U_ν(x + μ) U_μ⁻¹(x + ν) U_ν⁻¹(x) − 𝕀]
In the lₚ → 0 limit, this reduces to the standard non-Abelian field strength:
F_μνᵃ = ∂_μA_νᵃ − ∂_νA_μᵃ + gᵃ · fᵃ_bc · A_μᵇ · A_νᶜ
where fᵃ_bc are the structure constants of the gauge group.
Thus, gauge fields emerge naturally from polar curvature components.
2 Fermion Mass and Yukawa Coupling from Tension–Curvature Product
Let a fermion field ψ_f(x) reside at each pole. Define polar tension due to fermion energy density:
T_f(x) = (ψ̄_f(x) · γ⁰ · ψ_f(x)) · m_f⁄lₚ²
Define curvature due to spinor phase distortion:
K_θ_f(x) ≈ ∂(arg ψ_f)⁄∂x ≈ k_f
Then, the fermion polar field is:
ϕ_f(x) = T_f(x) · K_θ_f(x)
Let a scalar excitation H(x) play the role of the Higgs field. Then, the Yukawa interaction becomes:
L_Yukawa = − y_f · ϕ_f(x) · H(x) + h.c.
Expanding:
L_Yukawa = − y_f · (ψ̄_f γ⁰ ψ_f) · (k_f · m_f⁄lₚ²) · H(x) + h.c.
Thus, fermion mass terms arise geometrically from tension (energy density) and curvature (phase deformation), interacting with the Higgs scalar.
3 Flavor Mixing via Curvature Superposition
In standard quantum field theory, flavor mixing is introduced via unitary transformations. In Pole Theory, we geometrically encode this as curvature superposition:
K_θ^mix_i(x) = ∑_j V_ij · K_θ^j(x)
where:
V_ij is a unitary flavor mixing matrix (CKM or PMNS),
K_θ^j(x) are the curvature fields of individual flavor states.
The resulting polar field becomes:
ϕ^mix_i(x) = T^i(x) · K_θ^mix_i(x)
Interaction Lagrangian:
L_flavor = − ∑_{i,j} y_ij · ψ̄_L^i · H(x) · ψ_R^j + h.c., where:
y_ij = y_f · V_ij
Hence, fermion generation mixing and CP violation emerge from internal superposition of geometric curvature, without additional assumptions.
4 Gauge Covariance and Continuum Limit
On the lattice, define the discrete covariant derivative acting on a fermion:
D_μψ(x) = [U_μ(x) · ψ(x + μ̂) − ψ(x)]⁄lₚ
In the continuum limit:
D_μψ(x) → ∂_μψ(x) + i gᵃ · A_μᵃ(x) · τₐ · ψ(x)
Thus, the gauge-covariant structure of standard Yang–Mills theory is fully recovered in the continuum from polar link geometry.
Predictions and Observable Deviations
Predictions and Observable Deviations from Pole Theory
Although Pole Theory is rooted in Planck-scale physics—far beyond current experimental reach—it naturally yields small but nonzero corrections to physical laws in the observable domain. These effects arise from the discrete nature of spacetime, encoded in the polar lattice, and may manifest through quantum interference, gravitational wave dispersion, or cosmological observations.
In this section, we present quantitative predictions that differentiate Pole Theory from standard physics, derived from the field structure:
Φ(x, t) = T(x, t) · K_θ(x, t)
And its governing equation:
□ϕ(x, t) – m²ϕ(x, t) = (8πG⁄c⁴) · Ψ̃(x, t) · R(x, t)
We define the Planck deviation parameter:
Δ = lₚ⁄ℓ
where:
ℓ is the characteristic physical length scale of the system,
lₚ ≈ 1.616 × 10⁻³⁵ m
1 Modified Uncertainty Principle
Standard Heisenberg relation:
Δx · Δp ≥ ℏ⁄2
In Pole Theory, due to underlying discreteness, this becomes:
Δx · Δp ≥ (ℏ⁄2)(1 + δ)
For a system of length ℓ, this leads to tiny but cumulative phase errors:
Interpretation: In highly coherent systems (e.g., atomic clocks, quantum Hall states), these corrections may accumulate and be revealed via phase noise beyond standard quantum limits.
2 Quantum Interference and Effective Momentum
For a path separation L, the accumulated quantum phase is:
Δϕ = (p · L)⁄ℏ
In the polar framework, the effective momentum is corrected as:
P_eff ≈ p · (1 – δ²⁄6)
Therefore:
Δϕ ≈ (p · L)⁄ℏ · (1 – δ²⁄6)
This implies a measurable suppression in fringe contrast or systematic phase shift in long-baseline interferometers (e.g., atom interferometers in space).
3 Gravitational Wave Dispersion
Standard gravitational waves (GWs) are non-dispersive in general relativity. In Pole Theory, the dispersion relation is modified due to discrete geometry:
Let the harmonic solution be:
Φ(x, t) = A · e^{i(kx – ωt)}
Discrete D’Alembertian yields:
□ϕ ≈ −ω² + (2⁄lₚ²)[1 – cos(k · lₚ)]
This gives:
Ω² ≈ k² − (k⁴ · lₚ²)⁄12 + …
The group velocity becomes:
V_g = ∂ω⁄∂k ≈ 1 – (k² · lₚ²)⁄6
For typical LIGO frequencies:
K ≈ 2πf⁄c ≈ 2 × 10⁻⁶ m⁻¹ (for f ≈ 100 Hz)
Δv_g ≈ −10⁻⁶⁸
Too small to detect directly, but over cosmological distances, such shifts may affect GW arrival times or produce spectral distortions in multi-band observations (e.g., LISA, Einstein Telescope).
4 Cosmic Microwave Background (CMB) Anomalies
Pole Theory implies a suppression of high-k scalar power in the early universe, due to lattice-seeded fluctuations:
Modified power spectrum:
P(k) ≈ P₀(k) · (1 – α · k² · lₚ²)
This yields:
Suppression of small-scale fluctuations,
Angular damping in high-ℓ modes,
Potential resolution of unexplained anomalies in Planck CMB data.
ΔT⁄T ≈ 10⁻⁶ to 10⁻⁹ may emerge as a signature in low-ℓ multipole alignments or parity asymmetries.
5 Cumulative Deviations in High-Energy Colliders
At the LHC:
Jet energy: E ≈ 1 TeV
Interaction scale: ℓ ≈ 10⁻¹⁷ m
Deviation parameter: δ ≈ 10⁻¹⁸
Energy shift:
ΔE ≈ δ · E ≈ 10⁻⁶ eV
Undetectable individually, but statistical fingerprints may appear in:
Missing transverse energy spectra
Jet mass broadening
Soft photon emission anomalies
With AI-assisted analysis of large collider datasets (e.g., HL-LHC, FCC), such deviations might surface.
6. Falsifiability Criteria
Pole Theory may be tested or constrained if:
Phase noise in high-coherence systems exceeds standard limits,
GW signals show cumulative spectral or arrival time anomalies,
CMB low-ℓ features align statistically with lattice-imposed damping,
Collider datasets show persistent δ-like deviation patterns beyond SM noise.
Experimental Feasibility and Observational Prospects
Although Polar Theory operates at Planck-scale resolution, its low-energy effects may still be observable due to cumulative or resonance effects in sensitive experiments:
Gravitational wave detectors (e.g., LISA, Einstein Telescope):
Expected to observe phase anomalies or amplitude quantization arising from polar decoherence in early universe wavefronts
CMB B-mode polarization:
Residual polar structure from the inflationary era may leave detectable imprints in tensor-mode correlations
High-energy collider data (LHC, future muon colliders):
Deviations in spin–curvature coupling or nonstandard particle emergence may trace back to tension–curvature alignment breakdowns
Time-resolved quantum interferometry:
May show Planck-level phase fluctuations under extreme coherence regimes, testing the role of Ψ̃.
Though difficult, the theory is not unfalsifiable — and multiple avenues of precision detection already exist or are near maturity.
The theoretical predictions derived from Pole Theory can, in principle, be examined through current or emerging experimental frameworks across several physical domains:
In the domain of gravitational wave astronomy, experiments such as LIGO, LISA, and pulsar timing arrays may detect subtle phase anomalies or stepwise modulations in the waveform of low-frequency gravitational waves. These could correspond to discrete coherence transitions in the polar lattice during wave propagation, as predicted by the underlying quantization in curvature flow.
In early-universe cosmology, residual patterns left by high-tension polar zones during inflation may be imprinted in the tensor modes of the Cosmic Microwave Background (CMB). Upcoming missions like CMB-S4 could potentially resolve these imprints through deviations in the B-mode polarization spectrum or power spectrum anomalies not accounted for by standard inflationary models.
In quantum measurement regimes, ultra-sensitive matter-wave interferometry and decoherence-time experiments (involving cold atoms or optomechanical systems) may exhibit anomalies in phase coherence or unexpected collapse behavior, consistent with the Ψ̃ field’s deterministic decay pattern postulated in the theory.
In high-energy particle physics, the tension–curvature coupling at the core of Pole Theory could lead to slight mass-ratio deviations or spin-alignment asymmetries for fermions and vector bosons. Such deviations could, in principle, emerge in collider datasets (e.g., LHC, HL-LHC, future muon colliders) as statistically significant patterns, particularly in spin-resolved angular distributions or rare decay modes.
In black hole thermodynamics, polar theory predicts that evaporation spectra may carry fine-structure patterns due to discrete curvature release, leading to deviations from the pure thermal Hawking spectrum. These could potentially be observed in the long-term monitoring of microquasars or gravitational echoes post-merger events.
Together, these predictions span across multiple branches of experimental physics, offering indirect but falsifiable routes to probe the coherence, discreteness, and tension-curvature structure proposed by the polar model.
Comparative Analysis with Other Models
Comparative Analysis with Other Quantum Gravity Models
A comprehensive evaluation of any proposed quantum gravity framework demands comparison with existing approaches, particularly in terms of conceptual clarity, mathematical structure, physical assumptions, and empirical accessibility. In this section, we contrast Pole Theory with prominent quantum gravity models such as Loop Quantum Gravity (LQG), String Theory, Causal Set Theory, Causal Dynamical Triangulations (CDT), Spin Foam Models, and Asymptotic Safety.
1. Spacetime Structure and Dimensional Foundations
Unlike String Theory, which requires 10 or 11 dimensions and assumes a continuous higher-dimensional manifold, Pole Theory operates strictly within 3+1 dimensions. The theory begins with a discrete polar lattice, composed of Planck-scale nodes (poles), with time evolving as a series of discrete steps. This avoids the need for compactification, Calabi–Yau manifolds, or dimensional reduction — common in string-based models — and retains a direct correspondence with observable spacetime.
In contrast to Loop Quantum Gravity, where space is quantized through spin networks and areas/volumes are eigenvalues of geometric operators, Pole Theory defines geometry more directly through the scalar field ϕ = T · K_θ. Rather than relying on abstract algebraic geometry, it uses local products of energy density (tension) and curvature to describe structure, enabling a field-based, geometrically intuitive formulation.
Causal Set Theory and CDT also posit discrete spacetime structures, but differ significantly. Causal Set Theory views the universe as a partially ordered set of events, lacking field content or curvature descriptions at each site. CDT constructs geometry by gluing simplices but lacks intrinsic field dynamics and unification of forces. Pole Theory, by contrast, integrates both field content and geometry at each discrete point, offering a more unified and physically complete lattice model.
2. Field Content and Matter Coupling
Whereas String Theory postulates one-dimensional vibrating objects as fundamental entities and derives gauge symmetries from topological modes of string vibration, Pole Theory embeds gauge interactions directly into polar curvature components. The curvature K_θᵃ(x, μ) associated with each direction at a pole carries internal symmetry information, giving rise to gauge fields without requiring additional dimensions or supersymmetry.
Loop Quantum Gravity and CDT focus primarily on quantizing spacetime, not incorporating matter fields or Standard Model interactions natively. Matter coupling in these frameworks often remains an open or externally added feature. In contrast, Pole Theory derives fermion mass, Yukawa couplings, and even flavor mixing geometrically, through the interaction of polar tension and curvature. This built-in unification of matter and geometry is absent in most other approaches.
Asymptotic Safety, which treats gravity as a renormalizable quantum field theory via a fixed point in the renormalization group flow, remains continuous and lacks a natural embedding of the Standard Model. Pole Theory, through its discrete scalar field dynamics, circumvents the need for perturbative renormalization by construction and captures SM-like structure without invoking fine-tuned RG flow behavior.
3. Gauge Symmetry and Unification
Pole Theory naturally reproduces non-Abelian gauge symmetries by interpreting curvature components as elements of gauge fields associated with SU(3), SU(2), and U(1) generators. These fields emerge via plaquette-like loop structures on the polar lattice, similar in spirit to lattice gauge theory but arising intrinsically from geometry rather than being imposed externally. This contrasts with LQG, which does not derive Standard Model gauge symmetry from its spin networks, and with string theory, where gauge symmetry appears only after compactification and model-specific assumptions.
Unlike spin foam models, where amplitudes are built combinatorially and lack a global action principle, Pole Theory maintains a discrete Lagrangian and action-based formulation throughout. This allows the derivation of field equations using standard variational techniques, preserving consistency and predictive power.
4. Lorentz Invariance and Continuum Limit
A critical test for any discrete model is its behavior in the continuum limit. Pole Theory retains Lorentz invariance by construction: the discrete derivatives are formulated in such a way that the standard D’Alembertian and kinetic terms of field theory are recovered smoothly as the lattice spacing approaches zero. In CDT and Causal Set Theory, Lorentz symmetry is only partially restored or difficult to define at all, and in LQG, it arises under coarse-graining with assumptions about semiclassical states.
String Theory and Asymptotic Safety preserve Lorentz symmetry in their respective frameworks but at the cost of added structural complexity or mathematical machinery. In contrast, Pole Theory achieves it without additional assumptions, relying purely on discrete scalar fields and local geometry.
5. Predictive Power and Experimental Accessibility
Despite operating at the Planck scale, Pole Theory offers empirically falsifiable predictions through accumulated effects in quantum interference, gravitational wave dispersion, and CMB anisotropies. This sets it apart from String Theory, where predictions often depend on untestable features of extra-dimensional compactification, and from LQG, where measurable consequences remain elusive outside of black hole entropy and area quantization.
Pole Theory predicts corrections to known laws—such as modified uncertainty principles and suppressed high-frequency gravitational wave speeds—that may become observable in high-coherence quantum experiments, space-based interferometers, or precise cosmological surveys. Its discrete foundation provides a concrete framework for simulations and data analysis, potentially making it more testable than many established quantum gravity models.
6. Conceptual and Philosophical Clarity
Finally, Pole Theory offers a uniquely clear ontological narrative: from an absolute state of zero energy and geometry, the universe begins via a single fluctuation permitted by the uncertainty principle. This leads to recursive lattice expansion and the emergence of all physical structure. The model avoids the metaphysical baggage of pre-existing space, time, or symmetry, aligning well with “creation from nothing” cosmological scenarios like the Vilenkin or Hartle–Hawking proposals.
Where most theories are built atop classical geometric or field-theoretic assumptions, Pole Theory constructs geometry, quantum behavior, and dynamics from scratch, offering a minimalist, deterministic, and scalar-rooted unification model.
Conclusion and Future Directions
Conclusion and Future Directions
In this work, we introduced Pole Theory, a unifying framework that derives both the structure of spacetime and the behavior of matter from a single, discrete, Planck-scale scalar field:
Φ(x, t) = T(x, t) · K_θ(x, t)
This field, defined at each node (pole) in a fundamentally discrete 3+1 dimensional lattice, encapsulates both energy and curvature. Beginning from an initial state of absolute nothingness—a point of zero geometry, matter, and energy—Pole Theory proposes that a spontaneous quantum fluctuation initiates the formation of the first pole. From this single excitation, the entire fabric of spacetime emerges recursively, as curvature gradients activate neighboring poles and drive lattice expansion.
The theory successfuly reproduces known physics across multiple scales. In the quantum limit, it recovers the Schrödinger equation; in the classical limit, it reduces to the Einstein field equations and Newtonian gravity; and on cosmic scales, it leads to the Friedmann equations of cosmology. The framework seamlessly incorporates gauge symmetries, fermion masses, and flavor mixing, not by postulate, but as geometric consequences of polar curvature and field superposition.
Crucially, the theory is not confined to mathematical elegance—it offers testable predictions, including:
Planck-scale corrections to phase accumulation in quantum systems,
Modified dispersion relations for gravitational waves,
Anomalies in the CMB power spectrum due to small-scale suppression,
And statistical signatures in high-energy collider data.
These provide a pathway to empirical exploration, even as direct Planck-scale access remains beyond current technology.
Pole Theory stands apart from other quantum gravity frameworks in its conceptual simplicity, dimensional minimalism, and direct connection to physical observables. It does not rely on additional dimensions, supersymmetric particles, or background-independent formalism. Instead, it proposes a deterministic, scalar-based origin of all structure—a geometry born from tension and curvature, governed by local action and field propagation.
Future Directions
Several avenues of investigation naturally follow from this foundation:
Modeling lattice evolution from the first pole outward can reveal structure formation patterns, phase transitions, and curvature dynamics during the early universe. These simulations could help connect microscopic structure with large-scale cosmology.
-
2.
Renormalization and Effective Field Theory
While Pole Theory is discrete and UV-complete, its connection to known effective field theories and renormalization group flow in the infrared limit remains to be formalized. Developing a coarse-grained continuum version would strengthen its ties to standard quantum field theory.
-
3.
Black Hole and Singular Structure
Investigating how polar fields behave in regions of extreme curvature may provide new insights into black hole entropy, evaporation, and quantum gravitational collapse, potentially offering alternatives to information paradox resolutions.
-
4.
Quantum Measurement and Decoherence
The physical role of the scalar wavefunction Ψ̃(x, t), which modulates curvature response in the field equation, opens new questions about the origin of superposition, decoherence, and the emergence of classicality from discrete deterministic fields.
-
5.
Connection to Other Discrete Models
Explorations of how Pole Theory might link with or reformulate approaches like spin networks, tensor networks, or matrix models could provide bridges across various quantum gravity landscapes, aiding unification or cross-validation.
-
6.
Experimental Interface
Designing experiments to test predicted deviations—particularly in gravitational wave astronomy, precision interferometry, and high-energy collider data—will bring the theory into conversation with observation, guiding refinement or falsification.
Pole Theory offers a mathematically rigorous, physically grounded, and philosophically coherent foundation for rethinking the origin and nature of the universe. By starting from nothing and allowing geometry to grow from tension and curvature, it reimagines the question of unification—not as a merging of existing frameworks, but as the emergence of all frameworks from one.
References
- Rovelli, C. Quantum Gravity; Cambridge University Press: 2004.
- Bombelli, L.; Lee, J.; Meyer, D.; Sorkin, R. Space-time as a causal set. Phys. Rev. Lett. 1987, 59, 521. [Google Scholar] [CrossRef] [PubMed]
- Ambjørn, J.; Jurkiewicz, J.; Loll, R. Reconstructing the universe. Phys. Rev. D 2005, 72, 064014. [Google Scholar] [CrossRef]
- Regge, T. General relativity without coordinates. Il Nuovo Cimento 1961, 19, 558–571. [Google Scholar] [CrossRef]
- Vilenkin, A. Creation of universes from nothing. Phys. Lett. B 1982, 117, 25–28. [Google Scholar] [CrossRef]
- Hartle, J.B.; Hawking, S.W. Wave function of the Universe. Phys. Rev. D 1983, 28, 2960. [Google Scholar] [CrossRef]
- Mukhanov, V.; Feldman, H.A.; Brandenberger, R.H. Theory of cosmological perturbations. Physics Reports 1992, 215, 203–333. [Google Scholar] [CrossRef]
- Weinberg, S. Ultraviolet divergences in quantum theories of gravitation. In General Relativity: An Einstein Centenary Survey 1979.
- Wilson, K.G. The renormalization group: Critical phenomena and the Kondo problem. Rev. Mod. Phys. 1975, 47, 773. [Google Scholar] [CrossRef]
- Padmanabhan, T. Thermodynamical aspects of gravity: New insights. Reports on Progress in Physics 2010, 73, 046901. [Google Scholar] [CrossRef]
- Bekenstein, J.D. Black holes and entropy. Phys. Rev. D 1973, 7, 2333. [Google Scholar] [CrossRef]
- Hawking, S.W. Particle creation by black holes. Commun. Math. Phys. 1975, 43, 199–220. [Google Scholar] [CrossRef]
- Susskind, L. The world as a hologram. J. Math. Phys. 1995, 36, 6377. [Google Scholar] [CrossRef]
- Zurek, W.H. Decoherence, einselection, and the quantum origins of the classical. Rev. Mod. Phys. 2003, 75, 715. [Google Scholar] [CrossRef]
- Joos, E. , Zeh, H. D., et al. (2003). Decoherence and the Appearance of a Classical World in Quantum Theory. Springer.
- Isham, C. J. (1995). Structural issues in quantum gravity. Imperial College Press.
- Gibbons, G.W.; Hawking, S.W. Action integrals and partition functions in quantum gravity. Phys. Rev. D 1977, 15, 2752. [Google Scholar] [CrossRef]
- Misner, C. W. , Thorne, K. S., & Wheeler, J. A. (1973). Gravitation. Freeman.
- Carlip, S. Quantum gravity: A progress report. Rep. Prog. Phys. 2001, 64, 885–942. [Google Scholar] [CrossRef]
- Ashtekar, A.; Lewandowski, J. Background independent quantum gravity: A status report. Class. Quant. Grav. 2004, 21, R53. [Google Scholar] [CrossRef]
- Thiemann, T. (2007). Modern Canonical Quantum General Relativity. Cambridge University Press.
- ‘t Hooft, G. Dimensional reduction in quantum gravity. Salamfestschrift 1993, 284–296. [Google Scholar]
- Nicolini, P. Noncommutative black holes, the final appeal to quantum gravity: A review. Int. J. Mod. Phys. A 2009, 24, 1229–1308. [Google Scholar] [CrossRef]
- Barrau, A.; Grain, J. Cosmology without singularity or infinity. Universe 2014, 2, 157–180. [Google Scholar]
- Carroll, S. M. (2004). Spacetime and Geometry: An Introduction to General Relativity. Pearson.
- Hossenfelder, S. Minimal length scale scenarios for quantum gravity. Living Rev. Relativ. 2013, 16. [Google Scholar]
- Kiefer, C. (2012). Quantum Gravity. Oxford University Press.
- Baez, J.C. An introduction to spin foam models of BF theory and quantum gravity. Lect. Notes Phys. 2000, 543, 25–94. [Google Scholar]
- Freidel, L.; Krasnov, K. A new spin foam model for 4D gravity. Class. Quant. Grav. 2008, 25, 125018. [Google Scholar] [CrossRef]
- Dittrich, B. From the discrete to the continuous: Towards a cylindrically consistent dynamics. New J. Phys. 2012, 14, 123004. [Google Scholar] [CrossRef]
- Lisi, A. G. (2007). An exceptionally simple theory of everything. arXiv:0711.0770.
- Seiberg, N.; Witten, E. The D1/D5 system and singular CFT. JHEP 1999, 9904, 017. [Google Scholar] [CrossRef]
- Jacobson, T. Thermodynamics of spacetime: The Einstein equation of state. Phys. Rev. Lett. 1995, 75, 1260. [Google Scholar] [CrossRef]
- Verlinde, E. On the origin of gravity and the laws of Newton. JHEP 2011, 2011, 29. [Google Scholar] [CrossRef]
- Modesto, L. Super-renormalizable quantum gravity. Phys. Rev. D 2010, 86, 044005. [Google Scholar] [CrossRef]
- Nicolai, H.; Peeters, K.; Zamaklar, M. Loop quantum gravity: An outside view. Class. Quant. Grav. 2005, 22, R193. [Google Scholar] [CrossRef]
- Maldacena, J. The large-N limit of superconformal field theories and supergravity. Adv. Theor. Math. Phys. 1999, 2, 231–252. [Google Scholar] [CrossRef]
- Swingle, B. Entanglement renormalization and holography. Phys. Rev. D 2012, 86, 065007. [Google Scholar] [CrossRef]
- Raamsdonk, M.V. Building up spacetime with quantum entanglement. Gen. Rel. Grav. 2010, 42, 2323–2329. [Google Scholar] [CrossRef]
- Lee, J.; Smolin, L. Quantum gravity and the standard model. Nucl. Phys. B 1997, 477, 407–439. [Google Scholar]
- Markopoulou, F. Quantum causal histories. Class. Quant. Grav. 2000, 17, 2059. [Google Scholar] [CrossRef]
- Konopka, T.; Markopoulou, F.; Severini, S. Quantum graphity: A model of emergent locality. Phys. Rev. D 2008, 77, 104029. [Google Scholar] [CrossRef]
- Hamma, A.; Ionicioiu, R.; Zanardi, P. Bipartite entanglement and entropic boundary law in lattice spin systems. Phys. Rev. A 2005, 71, 022315. [Google Scholar] [CrossRef]
- Ryu, S.; Takayanagi, T. Holographic derivation of entanglement entropy. Phys. Rev. Lett. 2006, 96, 181602. [Google Scholar] [CrossRef] [PubMed]
- Gross, D.J.; Witten, E. Superstring modifications of Einstein’s equations. Nucl. Phys. B 1986, 277, 1–10. [Google Scholar] [CrossRef]
- Polchinski, J. (1998). String Theory Vols. 1 & 2. Cambridge University Press.
- Wheeler, J.A. Geometrodynamics and the issue of the final state. Relativity, Groups and Topology 1964, 1, 317–520. [Google Scholar]
- Rovelli, C. , & Vidotto, F. (2015). Covariant Loop Quantum Gravity: An Elementary Introduction to Quantum Gravity and Spinfoam Theory. Cambridge University Press.
- Hardy, L. (2001). Quantum theory from five reasonable axioms. arXiv:quant-ph/0101012.
- Laughlin, R. B. (2005). A different universe: Reinventing physics from the bottom down. Basic Books.
- Smolin, L. (2013). Time Reborn: From the Crisis in Physics to the Future of the Universe. Houghton Mifflin Harcourt.
- Bohm, D. , & Hiley, B. J. (1993). The Undivided Universe: An Ontological Interpretation of Quantum Theory. Routledge.
- Deutsch, D. (1997). The Fabric of Reality. Penguin.
- Vilenkin, A. Creation of universes from nothing. Physics Letters B 1982, 117, 25–28. [Google Scholar] [CrossRef]
- Hartle, J.B.; Hawking, S.W. Wave function of the Universe. Physical Review D 1983, 28, 2960–2975. [Google Scholar] [CrossRef]
- Sorkin, R. D. (2005). Causal sets: Discrete gravity. In Lectures on Quantum Gravity (pp. 305-327). Springer.
- Rovelli, C. (2004). Quantum Gravity. Cambridge University Press.
- Ambjørn, J.; Jurkiewicz, J.; Loll, R. Dynamically triangulating Lorentzian quantum gravity. Nuclear Physics B 2001, 610, 347–382. [Google Scholar] [CrossRef]
- Padmanabhan, T. Emergent gravity paradigm: Recent progress. Modern Physics Letters A 2015, 30, 1540007. [Google Scholar] [CrossRef]
- Kiefer, C. (2012). Quantum Gravity (3rd ed.). Oxford University Press.
- Birrell, N. D. , & Davies, P. C. W. (1982). Quantum Fields in Curved Space. Cambridge University Press.
- Parker, L. , & Toms, D. J. (2009).
- Weinberg, S. (1995). The Quantum Theory of Fields (Vol. 1). Cambridge University Press.
- Peskin, M. E. , & Schroeder, D. V. (1995). An Introduction to Quantum Field Theory. Addison-Wesley.
- Peebles PJ, E.; Ratra, B. The cosmological constant and dark energy. Reviews of Modern Physics 2003, 75, 559–606. [Google Scholar]
- Bertone, G.; Hooper, D.; Silk, J. Particle dark matter: Evidence, candidates and constraints. Physics Reports 2005, 405, 279–390. [Google Scholar]
|
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).