SOCR ≫ TCIU Website ≫ TCIU GitHub ≫

In this section, we will explore highly novel ideas of direct connections, potential synergies, and construction a complex-time interpretation of the universe, without a fundamental time variable. Specifically, we will examine spacekime in the context of observability, Wheeler-DeWitt equation, Boltzmann theory, entropy, quantum gravity, and artificial intelligence applicaitons.

1 Foundations

1.1 Relativistic Mass Increases with Increase of the Velocity

In special relativity, as an object moves faster, its mass appears to increase from the perspective of an observer in a different frame of reference frame. This effect becomes significant as the object’s speed approaches the speed of light. The increase in mass due to speed can be derived from the principles of special relativity.

Motivating Data: Observations and Experimental Evidence

  • Particle Accelerators: When particles such as electrons or protons are accelerated to speeds close to the speed of light in particle accelerators like the Large Hadron Collider, their behavior confirms the relativistic mass increase. The energy required to continue accelerating these particles increases dramatically as their speed approaches the speed of light, indicating an increase in their relativistic mass.

  • Muons in Cosmic Rays: Muons are particles that decay relatively quickly when at rest. However, muons generated by cosmic rays traveling near the speed of light reach the Earth’s surface more often than expected. This is explained by time dilation and the increase in their relativistic mass, which prolongs their lifespan from the perspective of an observer on Earth.

To derive the relativistic mass increase with velocity, we start with the total energy-momentum relation in special relativity:

\[E^2 = (pc)^2 + (mc^2)^2\ ,\]

where \(E\) is the total energy, \(p\) is the momentum, \(m\) is the rest mass, and \(c=300,000\ km/s\) is the speed of light. The momentum \(p\) in relativity is \(p = \gamma mv\). The total energy \(E\) can also be expressed in terms of the velocity \(v\) of the object relative to the observer as \(E = \gamma mc^2\), where \(\gamma\) (the Lorentz factor) is \[\gamma = \frac{1}{\sqrt{1 - \frac{v^2}{c^2}}}\ .\]

In the context of special relativity, the mass of an object as measured by an observer moving relative to it (often called relativistic mass)

\[m_{\text{rel}} = \gamma m = \frac{m}{\sqrt{1 - \frac{v^2}{c^2}}}\ .\]

This equation shows that as the speed \(v\) of the object increases, its relativistic mass \(m_{\text{rel}}\) increases as well. Specifically, as velocity increases, \(v\to c\), the denominator \(\sqrt{1 - \frac{v^2}{c^2}}\to 0\), causing \(m_{\text{rel}}\to\infty\).

The increase in mass due to speed is given by the difference between the relativistic mass and the rest mass:

\[\Delta m = m_{\text{rel}} - m = \left(\frac{1}{\sqrt{1 - \frac{v^2}{c^2}}} - 1\right) m\]

This \(\Delta m\) represents the additional mass that appears due to the object’s velocity. This relativistic mass increase has been confirmed through experiments and observations, particularly in high-energy physics (e.g., LHC), where particles are accelerated to relativistic speeds. This relationship is fundamental to understanding the behavior of objects in motion at speeds close to the speed of light, influencing everything from particle physics to cosmological models.

1.2 Observability of Spacetime Events

Spacetime is a four-dimensional manifold \(\mathcal{M}\) where events are described by coordinates \((x^0, x^1, x^2, x^3)\), typically represented as \((t, x, y, z)\), where \(t\) is the time coordinate and \((x, y, z)\) are the spatial coordinates. The geometry of spacetime is determined by the metric tensor \(g_{\mu\nu}\), which encodes the distances and angles in the manifold. The line element (or interval) between two infinitesimally close events is given by \(ds^2 = g_{\mu\nu} \, dx^\mu \, dx^\nu\), where \(\mu, \nu = 0, 1, 2, 3\) and Einstein summation convention is used.

A spacetime event is a point in spacetime, representing a specific location and time. Mathematically, an event \(E\) can be defined as a 4-tuple \(E = (t, x, y, z)\). In general relativity, events are points on the spacetime manifold \(\mathcal{M}\). The set of all possible events constitutes the entire spacetime.

Observability in spacetime refers to the ability to detect or measure certain events or quantities in the spacetime continuum. It depends on the availability of observers and the causal structure of spacetime. For an observer \(O\) with a worldline parameterized by proper time \(\tau\), the observability of an event \(E = (t, x, y, z)\) is determined by the light cones emanating from \(E\)

  • Future Light Cone: The set of all possible future events that can be influenced by \(E\). These are events \(E'\) such that \(ds^2 \geq 0\) and \(t' > t\).
  • Past Light Cone: The set of all possible past events that could influence \(E\). These are events \(E'\) such that \(ds^2 \geq 0\) and \(t' < t\).

The causal structure defines the relationships between events and whether one event can influence another. For two events \(E_1\) and \(E_2\)

  • \(E_1\) and \(E_2\) are timelike separated if \(ds^2 > 0\). This means a causal relationship is possible, where one event can influence the other.
  • \(E_1\) and \(E_2\) are lightlike (null) separated if \(ds^2 = 0\). The events can be connected by a light signal, making them potentially observable to each other.
  • \(E_1\) and \(E_2\) are spacelike separated if \(ds^2 < 0\). There is no causal connection between the events, and they are not mutually observable in any reference frame.

The observability of a spacetime event from a particular observer’s perspective is governed by the causal structure of spacetime and the worldline of the observer. If an observer \(O\) follows a trajectory \(x^\mu(\tau)\) in spacetime, the event \(E\) is observable by \(O\) if there exists a solution to \(x^\mu(\tau_E) = (t, x, y, z)\), where \(\tau_E\) is the proper time at which the observer intersects with the light cone of the event \(E\). The observer’s motion is described by a four-velocity \(u^\mu = \frac{dx^\mu}{d\tau}\), and the relationship between the observer and the event’s observability can be further examined through the inner product of the four-velocity and the displacement vector between the observer and the event \(u^\mu \Delta x_\mu = u^\mu (x_\mu - x_\mu(\tau_E)),\) where \(\Delta x_\mu\) is the spacetime interval between the observer and the event.

Observable quantities in spacetime, such as distances, angles, and time intervals, are derived from the metric tensor \(g_{\mu\nu}\). The metric tensor determines the proper time \(d\tau\) experienced by an observer moving along a worldline \(d\tau^2 = -g_{\mu\nu} \, dx^\mu \, dx^\nu.\) The observable distance between two events is calculated using the spacetime interval \(ds^2\). The redshift of light from distant objects, as observed in cosmology, can also be derived from the spacetime metric, providing a direct link between theoretical predictions and observational data. The event horizon is defined by the condition where the escape velocity equals the speed of light, leading to the formation of a null surface (g_{} , dx^, dx^= 0.) This surface separates observable events from those that are permanently hidden from an observer’s view.

1.3 Christoffel Symbols

The Christoffel symbols, also known as the Levi-Civita connection, are fundamental in differential geometry and general relativity. They play a crucial role in defining how vectors change as they move along a manifold, which is essential for expressing the curvature of the manifold.

1.3.1 Definition of Christoffel Symbols

Given a manifold with a metric tensor \(g_{\mu\nu}\), the Christoffel symbols are defined by

\[\Gamma^\lambda_{\mu\nu} = \frac{1}{2} g^{\lambda\sigma} \left( \frac{\partial g_{\sigma\mu}}{\partial x^\nu} + \frac{\partial g_{\sigma\nu}}{\partial x^\mu} - \frac{\partial g_{\mu\nu}}{\partial x^\sigma} \right),\]

where:

  • \(g_{\mu\nu}\) is the metric tensor.
  • \(g^{\lambda\sigma}\) is the inverse of the metric tensor, satisfying \(g_{\mu\lambda} g^{\lambda\nu} = \delta_\mu^\nu\).
  • \(\Gamma^\lambda_{\mu\nu}\) are the Christoffel symbols, which are not tensors but are used to define the covariant derivative on the manifold.

1.3.2 Derivation of Christoffel Symbols

The Christoffel symbols can be derived from the requirement that the covariant derivative of the metric tensor vanishes, which is a condition known as the metric compatibility of a connection, \(\nabla_\lambda g_{\mu\nu}=0\)

\[\nabla_\lambda g_{\mu\nu} = \partial_\lambda g_{\mu\nu} - \Gamma^\sigma_{\lambda\mu} g_{\sigma\nu} - \Gamma^\sigma_{\lambda\nu} g_{\mu\sigma} = 0.\]

Expanding this for different permutations of the indices \(\mu, \nu,\lambda\in\{0,1,2,3,4\}\)

  1. Permuting indices \(\mu\), \(\nu\), and \(\lambda\):

\[\partial_\lambda g_{\mu\nu} = \Gamma^\sigma_{\lambda\mu} g_{\sigma\nu} + \Gamma^\sigma_{\lambda\nu} g_{\mu\sigma}\] \[\partial_\mu g_{\nu\lambda} = \Gamma^\sigma_{\mu\nu} g_{\sigma\lambda} + \Gamma^\sigma_{\mu\lambda} g_{\nu\sigma}\]

\[\partial_\nu g_{\lambda\mu} = \Gamma^\sigma_{\nu\lambda} g_{\sigma\mu} + \Gamma^\sigma_{\nu\mu} g_{\lambda\sigma}\ .\]

  1. Adding and subtracting these equations appropriately:

\[ \partial_\lambda g_{\mu\nu} + \partial_\mu g_{\nu\lambda} - \partial_\nu g_{\lambda\mu} = 2\Gamma^\sigma_{\lambda\mu} g_{\sigma\nu}\]

Solving for the Christoffel symbols:

\[\Gamma^\sigma_{\lambda\mu} g_{\sigma\nu} = \frac{1}{2} \left( \partial_\lambda g_{\mu\nu} + \partial_\mu g_{\nu\lambda} - \partial_\nu g_{\lambda\mu} \right)\ .\]

Finally, multiplying by the inverse metric \(g^{\lambda\nu}\) to solve for \(\Gamma^\lambda_{\mu\nu}\):

\[\Gamma^\lambda_{\mu\nu} = \frac{1}{2} g^{\lambda\sigma} \left( \frac{\partial g_{\sigma\mu}}{\partial x^\nu} + \frac{\partial g_{\sigma\nu}}{\partial x^\mu} - \frac{\partial g_{\mu\nu}}{\partial x^\sigma} \right)\ .\]

1.3.3 Role of Christoffel Symbols in Expressing Manifold Curvature

The Christoffel symbols play a crucial role in defining the curvature of a manifold:

  1. Geodesics:
    • The Christoffel symbols appear in the geodesic equation, which describes the path of least distance (or straightest path) between points on a curved manifold:

\[\frac{d^2 x^\lambda}{d\tau^2} + \Gamma^\lambda_{\mu\nu} \frac{dx^\mu}{d\tau} \frac{dx^\nu}{d\tau} = 0\ .\] Here, \(\tau\) is an affine parameter, such as proper time in general relativity.

  1. Covariant Derivatives:
    • The Christoffel symbols are used to define covariant derivatives, which generalize the concept of derivatives to curved spaces. The covariant derivative of a vector field \(V^\mu\) is given by:

\[\nabla_\nu V^\mu = \partial_\nu V^\mu + \Gamma^\mu_{\nu\lambda} V^\lambda\ .\]

This derivative accounts for the curvature of the space by including the Christoffel symbols using Dirac notation where we sum over all repeated indices, for each pair of indices \(\nu,\mu\),

\[\nabla_\nu V^\mu = \partial_\nu V^\mu + \Gamma^\mu_{\nu\underline{\lambda}} V^ {\underline{\lambda}}\ \ \Longleftrightarrow \ \ \nabla_\nu V^\mu = \partial_\nu V^\mu + \sum_{\lambda=0}^4\left (\Gamma^\mu_{\nu\lambda} V^\lambda\right )\ .\]

  1. Riemann Curvature Tensor:
    • The Riemann curvature tensor measures the intrinsic curvature of a manifold. For all quadruples \((\rho,\sigma,\mu,\nu)\), the Riemann curvature tensor is expressed in terms of the Christoffel symbols:

\[R^\rho_{\sigma\mu\nu} = \partial_\mu \Gamma^\rho_{\nu\sigma} - \partial_\nu \Gamma^\rho_{\mu\sigma} + \Gamma^\rho_{\mu\lambda} \Gamma^\lambda_{\nu\sigma} - \Gamma^\rho_{\nu\lambda} \Gamma^\lambda_{\mu\sigma}\ .\]

The Riemann tensor encapsulates how much and in what manner the manifold deviates from being flat (Euclidean). Note that the Riemann curvature tensor \(R^\rho_{\sigma\mu\nu}\) has four indices related to the natural mathematical description of curvature, using covariant and contravariant indices.

The tensor has 1 contravariant index (upper index, \(\rho\)) and 3 covariant indices (lower indices, \(\sigma\), \(\mu\), \(\nu\)). These indices are used to describe how vectors and other tensors change as they are parallel transported around the manifold.

Curvature in a manifold is detected by parallel transporting a vector around a small loop. The change in the vector after completing the loop can be described by the Riemann tensor. Specifically,

  • The index \(\rho\) represents the direction of the resulting vector after parallel transport.
  • The index \(\sigma\) corresponds to the component of the original vector that was parallel transported.
  • The indices \(\mu\) and \(\nu\) describe the plane in which the loop lies and represent the directions in which the vector is transported.

The Riemann curvature tensor maps a pair of directions (\(\mu\) and \(\nu\)) and a vector (indexed by \(\sigma\)) onto a new vector (indexed by \(\rho\)). This structure is necessary because curvature is inherently related to how vectors change in different directions within the manifold.

  • \(\rho\): Indicates the direction in which the curvature effect is observed (the output of the curvature operation).
  • \(\sigma\): Represents the vector being transported.
  • \(\mu, \nu\): Define the plane of the loop around which the vector is transported.

The Riemann curvature tensor is derived from the Christoffel symbols, which encode information about the connection on the manifold (how vectors change as they move across the manifold). For each quadruple of indices, \(\{\rho,\sigma,\mu,\nu\}\),

\[R^\rho_{\sigma\mu\nu} = \partial_\mu \Gamma^\rho_{\nu\sigma} - \partial_\nu \Gamma^\rho_{\mu\sigma} + \underbrace{\Gamma^\rho_{\mu\lambda} \Gamma^\lambda_{\nu\sigma} - \Gamma^\rho_{\nu\lambda} \Gamma^\lambda_{\mu\sigma}}_{summed\ over\ \lambda}\ ,\]

where \(\Gamma^\rho_{\mu\lambda}\) are the Christoffel symbols, representing the connection coefficients.

In general relativity, the Riemann curvature tensor describes how much spacetime is curved by the presence of mass-energy. Effectively, it quantifies the tidal forces experienced in a gravitational field, indicating how much different parts of a system will move relative to each other when subjected to gravity.

  1. Einstein Field Equations:
    • In general relativity, the Christoffel symbols are used to express the Einstein field equations, which describe the relationship between the geometry of spacetime and the distribution of matter and energy:

\[R_{\mu\nu} - \frac{1}{2} g_{\mu\nu} R = 8\pi G T_{\mu\nu}\ ,\ \ \forall \{\mu,\nu\}\ ,\] where, \(R_{\mu\nu}\) is the Ricci curvature tensor, which is derived from the Riemann tensor.

The Christoffel symbols provide a mechanism to quantify vectors change as they move along a manifold, allowing for the expression of geodesics, covariant derivatives, and curvature tensors.

1.4 Einstein’s Field Equations

Einstein’s field equations in general relativity describe how matter and energy in the universe influence the curvature of spacetime. These equations are typically written using tensor notation, which allows for a compact representation of the equations.

1.4.1 Standard Tensor Notation (Shorthand)

\[G_{\mu\nu} + \Lambda g_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}\ ,\]

where:

  • \(G_{\mu\nu} = R_{\mu\nu} - \frac{1}{2} g_{\mu\nu} R\) is the Einstein tensor,
  • \(G\) is Newton’s gravitational constant, which is approximately \(6.6743\times 10^{−11} N\cdot m^2/kg\) in SI units (recall \(F=G\frac{m_1 \times m_2}{r^2}\)),
  • \(R_{\mu\nu}\) is the Ricci curvature tensor,
  • \(g_{\mu\nu}\) is the metric tensor,
  • \(R\) is the Ricci scalar,
  • \(T_{\mu\nu}\) is the stress-energy tensor,
  • \(\Lambda\) is the cosmological constant.

Expanding the Einstein tensor \(G_{\mu\nu}\) involves working with the Ricci curvature tensor and the Ricci scalar, both of which depend on the metric tensor and its derivatives. The metric tensor in component form is denoted as \(g_{\mu\nu}\), and the Christoffel symbols \(\Gamma^\lambda_{\mu\nu}\) are used in the computation of the Ricci tensor. The Ricci tensor is computed from the Christoffel symbols, which are derived from the metric tensor.

Here is the explicit indexed form of the field equations:

  1. Christoffel Symbols:

\[\Gamma^\lambda_{\mu\nu} = \frac{1}{2} g^{\lambda\sigma} \left( \frac{\partial g_{\sigma\mu}}{\partial x^\nu} + \frac{\partial g_{\sigma\nu}}{\partial x^\mu} - \frac{\partial g_{\mu\nu}}{\partial x^\sigma} \right)\ ,\] where \(\Gamma^\lambda_{\mu\nu}\) are the Christoffel symbols of the second kind, \(g^{\lambda\sigma}\) is the inverse metric tensor, and \(x^\mu\) represents the spacetime coordinates.

  1. Ricci Tensor:

\[R_{\mu\nu} = \frac{\partial \Gamma^\lambda_{\mu\nu}}{\partial x^\lambda} - \frac{\partial \Gamma^\lambda_{\mu\lambda}}{\partial x^\nu} + \Gamma^\lambda_{\lambda\sigma} \Gamma^\sigma_{\mu\nu} - \Gamma^\sigma_{\mu\lambda} \Gamma^\lambda_{\sigma\nu}\ , \] where, \(R_{\mu\nu}\) is the Ricci tensor, and the components are explicitly given in terms of partial derivatives of the Christoffel symbols and their products.

  1. Ricci Scalar:

\[R = g^{\mu\nu} R_{\mu\nu} = g^{\mu\nu} \left( \frac{\partial \Gamma^\lambda_{\mu\nu}}{\partial x^\lambda} - \frac{\partial \Gamma^\lambda_{\mu\lambda}}{\partial x^\nu} + \Gamma^\lambda_{\lambda\sigma} \Gamma^\sigma_{\mu\nu} - \Gamma^\sigma_{\mu\lambda} \Gamma^\lambda_{\sigma\nu} \right)\ , \]

  1. Einstein Tensor:

\[G_{\mu\nu} = R_{\mu\nu} - \frac{1}{2} g_{\mu\nu} R\]

Expanding this in terms of the Ricci tensor and Ricci scalar gives the explicit form for each component \(G_{\mu\nu}\).

  1. Einstein’s Field Equations:

\[G_{\mu\nu} + \Lambda g_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}\] Expanding the left-hand side using the Einstein tensor components gives:

\[R_{\mu\nu} - \frac{1}{2} g_{\mu\nu} g^{\alpha\beta} R_{\alpha\beta} + \Lambda g_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}\ , \] where \(R_{\mu\nu}\), \(g_{\mu\nu}\), and \(T_{\mu\nu}\) are all expressed in terms of their component indices, summing over repeated indices.

The explicit indexed form expanded Einstein’s field equations show the detailed structure that is otherwise condensed in tensor notation. These equations reveal the relationships between the spacetime metric, curvature, and matter-energy content of spacetime. This explicit notation is useful for detailed calculations and understanding the dependence of each term on the components of the metric tensor.

1.5 Mathematics of General Relativity Implying the Existence of Dark Matter and Dark Energy

The general theory of relativity (GR) describes gravity as the curvature of spacetime caused by mass and energy. The core of GR is encapsulated in Einstein’s field equations relating the geometry of spacetime to the distribution of mass-energy within it:

\[G_{\mu\nu} + \Lambda g_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}\ , \]

where:

  • \(G_{\mu\nu}\) is the Einstein tensor, representing spacetime curvature.
  • \(g_{\mu\nu}\) is the metric tensor, describing the geometry of spacetime.
  • \(T_{\mu\nu}\) is the stress-energy tensor, representing the distribution of matter and energy.
  • \(\Lambda\) is the cosmological constant.
  • \(G\) is the gravitational constant, and \(c\) is the speed of light.

Gravitational Lensing and Galactic Rotation Curves: In the context of general relativity, the presence of mass and energy curves spacetime, influencing the geodesic paths of objects and light. Observations of gravitational lensing (the bending of light around massive objects) and the rotation curves of galaxies provide evidence that there is more mass present than can be accounted for by visible matter alone.

Mathematically, this is reflected in the Einstein field equations when applied to systems like galaxies. If we assume only visible matter contributes to the stress-energy tensor \(T_{\mu\nu}\), the resulting spacetime curvature, \(G_{\mu\nu}\), is insufficient to explain the observed dynamics, such as the flat rotation curves of galaxies, where the orbital velocities of stars remain constant at large radii instead of decreasing.

This discrepancy suggests that there must be additional, non-luminous mass present, which is referred to as dark matter. The presence of dark matter alters the \(T_{\mu\nu}\) tensor, adding an unseen component that accounts for the observed gravitational effects.

In the case of a rotating galaxy, the mass distribution \(\rho(r)\) should determine the velocity \(v(r)\) of stars orbiting at a radius \(r\). General relativity predicts that for a spherically symmetric distribution the velocity at a given radius is

\[v(r) = \sqrt{\frac{G\cdot M(r)}{r}} \ ,\]

where \(M(r)\) is the mass enclosed within radius \(r\), and \(G\) is the gravitational constant. However, physical observations show that \(v(r)\) remains approximately constant at large radii, implying that \(M(r)\) increases linearly with \(r\). This fact cannot be explained by visible matter alone, leading to the postulation of dark matter existence.

Dark Energy reflecting the Accelerating Expansion of the Universe: The discovery that the universe is expanding at an accelerating rate (from observations of distant supernovae) requires an additional component in the Einstein field equations. This is commonly attributed to dark energy, which is associated with the cosmological constant \(\Lambda\) in the field equations.

When \(\Lambda >0\), it acts as a repulsive force, counteracting gravity and driving the acceleration of the universe’s expansion. This can be seen mathematically in the Friedmann equations, derived from Einstein’s field equations under the assumption of a homogeneous and isotropic universe (described by the FLRW metric:

\[\left(\frac{\dot{a}}{a}\right)^2 = \frac{8\pi G}{3} \rho - \frac{k}{a^2} + \frac{\Lambda}{3}\ , \]

where:

  • \(a(t)\) is the scale factor, describing the expansion of the universe as a function of time \(t\).
  • \(\rho\) is the energy density (including contributions from matter, radiation, and dark energy).
  • \(k\) describes the curvature of space.

For an accelerating universe, \(\frac{\ddot{a}}{a} >0\), the relative second derivative of the scale factor is positive, and
\(\Lambda >0\). The energy associated with \(\Lambda\) is often interpreted as the energy density of dark energy \(\rho_\Lambda = \frac{\Lambda c^2}{8\pi G}\).

In general relativity, dark matter and dark energy are emergent constructs, necessary components to explain observed phenomena that cannot be accounted for by visible matter and conventional forms of energy. Dark matter is required to explain the gravitational effects on galactic and larger scales. Whereas dark energy is introduced to account for the accelerated expansion of the universe. These concepts are deeply intertwined with the mathematics of general relativity, which provides the framework for understanding how mass-energy shapes spacetime and, consequently, the evolution of the universe.

Next, we will describe the GR prediction of the existence f invisible dark matter and dark energy. It’s based on the observation that the orbital velocities of stars remain constant at large radii instead of decreasing. The exact mathematical formulation behind this involves gravitational dynamics of star motion in galaxies and mass distribution within galaxies.

1.5.1 Newtonian Dynamics and Expected Behavior

According to classical Newtonian dynamics, assuming the majority of the galaxy’s mass is concentrated near the inner-galactic center, the orbital velocity \(v(r)\) of a star at a distance \(r\) from the center of a galaxy should decrease as \(r\) increases. Specifically, if the mass within radius \(r\) is \(M(r)\), the velocity is given by \(v(r) = \sqrt{\frac{G\cdot M(r)}{r}}.\)

Therefore, in a system where most of the mass is concentrated near the center, \(M(r)\) becomes roughly constant as \(r\) increases. This suggests that \(v(r) \propto \frac{1}{\sqrt{r}}\), which yields that orbital velocities should decrease with distance, confirming solar systems observations.

However, observations of spiral galaxies show that the orbital velocities of stars do not decrease as expected. Instead, they tend to remain constant or “flat” at large radii, i.e., \(v(r) \approx \text{constant}\).

To explain the flat rotation curves, one must consider that the mass \(M(r)\) does not taper off at large radii but instead continues to increase proportionally to \(r\), implying the presence of additional unseen mass, known as dark matter.

If we assume the mass within radius \(r\) continues to grow linearly with \(r\), \(M(r) \propto r\), the velocity expression becomes \[v(r) = \sqrt{\frac{G\cdot M(r)}{r}} \propto \sqrt{\frac{G\cdot r}{r}} = \text{constant}.\]

This implies that \(M(r) \propto r\) at large radii, which is consistent with the idea of a dark matter halo extending beyond the visible galaxy. This dark matter halo model is theorized to be a spherical distribution of dark matter surrounding the galaxy, whose density decreases with radius but still contributes significantly to the mass \(M(r)\) at large distances. The density profile of the dark matter halo often follows a model, such as the Navarro-Frenk-White (NFW) profile

\[\rho(r) \propto \frac{1}{\frac{r}{r_s}\left(1+\frac{r}{r_s}\right)^2},\]

where \(r_s\) is a scale radius. This profile predicts that mass continues to grow as \(r\) increases, which supports the observed constant orbital velocity.

1.6 Foundation of Quantum Gravity in Spacetime

Quantum gravity aims to unify the principles of quantum mechanics, which governs the behavior of particles at the smallest scales, with general relativity, which describes the gravitational force and the curvature of spacetime on cosmological scales. The mathematical foundation of quantum gravity in spacetime requires merging these two seemingly incompatible frameworks into a consistent theory.

In general relativity, spacetime is modeled as a 4D pseudo-Riemannian manifold \(\mathcal{M}\), with a metric tensor \(g_{\mu\nu}\) that defines distances and angles between points (events) in this manifold. The line element \(ds^2\) between two infinitesimally close events is given by \(ds^2 = g_{\mu\nu} \, dx^\mu \, dx^\nu,\) where \(\mu, \nu = 0, 1, 2, 3\) represent the time and space components, respectively. The metric tensor \(g_{\mu\nu}\) encodes the curvature of spacetime, which is directly related to the distribution of mass and energy via Einstein’s field equations

\[R_{\mu\nu} - \frac{1}{2}g_{\mu\nu}R = \frac{8\pi G}{c^4}T_{\mu\nu},\]

where \(R_{\mu\nu}\) is the Ricci curvature tensor, \(R\) is the Ricci scalar, \(T_{\mu\nu}\) is the stress-energy tensor, \(G\) is the gravitational constant, and \(c\) is the speed of light.

As quantum mechanics describes the state of a physical system in terms of a wavefunction \(\psi(x,t)\), the probability for finding a particle at position \(x\) and time \(t\) is quantified by the amplitude of the wave function. The time evolution of the wavefunction is governed by the Schrödinger equation \(i\hbar \frac{\partial \psi(x,t)}{\partial t} = \hat{H} \psi(x,t),\) where \(\hat{H}\) is the Hamiltonian operator representing the total energy of the system, and \(\hbar\) is the reduced Planck constant.

In quantum field theory (QFT), the wavefunction is extended to field operators \(\hat{\phi}(x)\), and particles are described as excitations of these quantum fields. The dynamics of these fields are governed by equations derived from the action principle, with the most familiar example being the Klein-Gordon equation for a scalar field

\[\left(\frac{1}{c^2}\frac{\partial^2}{\partial t^2} - \nabla^2 + \frac{m^2c^2}{\hbar^2}\right) \phi(x,t) = 0.\]

The primary challenge in formulating quantum gravity is the fundamental difference in how general relativity and quantum mechanics treat spacetime and gravity.

  • General relativity views gravity as the curvature of spacetime, with spacetime itself being a dynamic, smooth continuum.
  • Quantum mechanics requires a fixed spacetime background where particles and fields evolve probabilistically, and typically uses a flat or weakly curved spacetime.

When trying to quantize gravity, problems arise because the methods of quantum mechanics lead to infinities when applied to the smooth structure of spacetime in general relativity. These divergences occur because the gravitational interaction becomes increasingly strong at very small scales, necessitating a new approach.

Path Integral Formulation: One approach to quantum gravity is the path integral formulation proposed by Richard Feynman. In this framework, the probability amplitude for a quantum state to evolve from one configuration to another is given by summing over all possible paths (or histories) that the system could take between these configurations. For gravity, this involves summing over all possible spacetime geometries \(Z = \int \mathcal{D}g_{\mu\nu} \, e^{iS[g_{\mu\nu}]},\) where \(Z\) is the partition function, \(\mathcal{D}g_{\mu\nu}\) represents the functional integration over all metric tensors, and \(S[g_{\mu\nu}]\) is the Einstein-Hilbert action for the gravitational field

\[S[g_{\mu\nu}] = \frac{1}{16\pi G} \int_{\mathcal{M}} \sqrt{-g} \left(R - 2\Lambda \right) d^4x,\]

with \(g\) being the determinant of the metric tensor, \(R\) the Ricci scalar, and \(\Lambda\) the cosmological constant.

This approach, however, faces significant difficulties because the integral is typically non-renormalizable, leading to infinities that cannot be consistently managed within standard quantum field theory.

Another approach is canonical quantum gravity, where the focus in on quantizing the gravitational field using canonical quantization techniques similar to those used in ordinary quantum mechanics. This leads to the Wheeler-DeWitt equation, a quantum analogue of the Hamiltonian constraint in general relativity \(\hat{H}\Psi[g_{ij}] = 0,\) where, \(\Psi[g_{ij}]\) is the wavefunction of the universe, depending on the 3-metric \(g_{ij}\) of a spatial hypersurface, and \(\hat{H}\) is the Hamiltonian operator derived from the Arnowitt-Deser-Misner (ADM) formalism. The Wheeler-DeWitt equation is often written as

\[\left( -\hbar^2 G_{ijkl} \frac{\delta^2}{\delta g_{ij} \delta g_{kl}} + \sqrt{g} (R - 2\Lambda) \right)\Psi[g_{ij}] = 0,\]

where \(G_{ijkl}\) is the DeWitt supermetric, and \(R\) is the 3-dimensional Ricci scalar.

This equation suggests that the universe is described by a wavefunction that does not evolve in time, leading to the problem of time in quantum gravity, where time appears to be an emergent concept rather than a fundamental parameter.

Loop Quantum Gravity is a non-perturbative approach to quantum gravity that attempts to quantize the geometry of spacetime itself. It uses the Ashtekar variables, reformulating general relativity in terms of a connection variable \(A^i_a\) and its conjugate momentum, the densitized triad \(E^a_i\), \(\{A^i_a(x), E^b_j(y)\} = \delta^b_a \delta^i_j \delta(x-y).\)

The fundamental objects in LQG are holonomies of the connection (describing the parallel transport along loops in space) and fluxes of the triads. The quantum states of the gravitational field are represented by spin networks, which are graphs with edges labeled by representations of the gauge group \(SU(2)\), and nodes labeled by intertwining operators. The dynamics are encoded in the Hamiltonian constraint, leading to a quantum version of Einstein’s equations. In LQG, spacetime is discrete at the Planck scale, resolving some of the infinities that plague other approaches to quantum gravity.

Alternatively, string theory proposes that the fundamental constituents of the universe are not point particles, but one-dimensional strings. These strings vibrate at different frequencies, and their various vibrational modes correspond to the different particles observed in nature, including the graviton, the hypothetical quantum of gravity.

The action describing a string propagating in spacetime is the Nambu-Goto action

\[S = -\frac{1}{2\pi\alpha'} \int d^2\sigma \sqrt{-\gamma} \, g_{\mu\nu}(X) \frac{\partial X^\mu}{\partial \sigma^a} \frac{\partial X^\nu}{\partial \sigma^b} \gamma^{ab},\]

where \(X^\mu(\sigma^a)\) describes the embedding of the string in spacetime, \(\gamma^{ab}\) is the worldsheet metric, and \(\alpha'\) is the string tension. String theory naturally incorporates gravity and is consistent in higher dimensions, typically 10 or 11. It also predicts a rich structure of additional particles and interactions, offering a potential unification of all fundamental forces.

In some approaches to quantum gravity, spacetime itself is not fundamental but emerges from more basic entities. In the AdS/CFT correspondence, a realization within string theory, a lower-dimensional conformal field theory (CFT) describes a higher-dimensional gravity theory in anti-de Sitter (AdS) space. The idea that spacetime is emergent is also reflected in approaches such as causal dynamical triangulations (CDT) and holography, where the smooth structure of spacetime emerges from discrete building blocks or from boundary data.

References:

  • Einstein, A. (1915). “Die Feldgleichungen der Gravitation.” Sitzungsberichte der Preussischen Akademie der Wissenschaften zu Berlin.
  • Riess, A. G., et al. (1998). “Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant.” The Astronomical Journal.
  • Zwicky, F. (1933). “Die Rotverschiebung von extragalaktischen Nebeln.” Helvetica Physica Acta.

2 Event-Based Quantum Gravity and Complex-Time (Kime) Representation

Event-based quantum gravity formulations, such as those found in Loop Quantum Gravity (LQG), emphasize the role of events (or interactions) as fundamental, rather than a smooth spacetime continuum. In LQG, spacetime is quantized into discrete units (spins and loops), and the traditional notion of time is challenged.

Kime Representation: A potential kime representation may extend the traditional concept of time from \(t\in\mathbb{R}^+\) (real variable) to the complex domain, \(\kappa\in\mathbb{C}\). Each event (kevent, in spacekime terms) in a longitudinal process can be indexed by a space variable \(\bf{x}\in\mathbb{R}^3\), and a kime variable \(\kappa = t \, e^{i\varphi}\) with its magnitude \(|\kappa|\equiv t\) encoding is the traditional time. The kime-phase \(\varphi\) corresponds to the intrinsic distribution of the IID repeated measurements associated with the highly controlled experiment. This complex-time approach may align with event-related quantum gravity by allowing a richer mathematical structure to describe events/kevents with the phase component capturing the quantum-indeterminacy (due to quantum fluctuations) of events.

The transition from a classical time to a complex kime domain may provide a way to model quantum gravitational effects in a discrete spacetime framework. Since kime surfaces represent a higher-dimensional extension of time-series data, they could be used to describe the change (evolution) of quantum states in a complex space, potentially aligning with the discrete event structure of LQG.

3 Probability Distribution Theory and Entropy

3.1 Complex Events (Kevents)

Let’s define complex events, or kevents, as stochastic encounters over the complex plane \(\Omega=\mathbb{C}\). Events can be defined over different system states representations. For instance, using polar representation, events are expressed as \(\kappa = t e^{i\varphi}\), where \(t \in\mathbb{R}^+\) represents the usual Euclidean time and \(\varphi \in [-\pi, \pi)\) represents the kime-phase with a prior distribution \(\Phi(\varphi)\), modeling the variability in repeated experiments.

  1. Measure on \(\Omega\): The measure on \(\Omega\) is composed of two parts:

    • Measure on \(t\): For the real-time component, which could be Euclidean, but in practice most often is a normalized probability measure, e.g., Exponential (decay) distribution.
    • Kime-Phase Measure \(\Phi(\varphi)\): A prior phase distribution supported on \([- \pi, \pi)\).
  2. Random Variable (observable) \(X(\omega)\): Let \(X: \Omega \rightarrow \mathbb{R}\) be a real-valued random variable defined on the sample space \(\Omega\). For a given \(\omega \in \Omega\), \(X(\omega)\) maps \(\omega\) to a real number.

  3. Kevent: A complex-time event (kevent) is defined by the set: \[K = \{ \omega \in \Omega \mid u < X(\omega) \leq v \}.\] This kevent represents the inverse image of the interval \((u, v]\) under the mapping \(X\). The kevent construction generalizes the classical notion of events to the complex plane, where events are defined not just in terms of their longitudinal order (time), but also incorporate the kime-phase distribution \(\Phi(\varphi)\).

  4. Probability of a Kevent: The probability of the kevent \(K\) is given by \[\text{Pr}(u < X \leq v) = F(v) - F(u),\] where \(F(x)\) is the cumulative distribution function (CDF) of the random variable \(X\).

The probability measure of the kevent \(K\) is derived by considering both the time and kime-phase components. Specifically, we calculate \(\text{Pr}(K) = \text{Pr}(u < X \leq v) = \int_{X^{-1}((u,v])} d\mu(\omega),\) where \(d\mu(\omega)\) is the product measure on the sample space \(\Omega\) given by the probability measure on \(t\) and the prior phase distribution \(\Phi(\varphi)\) on \(\varphi\).

The cumulative distribution function \(F(x)\) is derived from the joint distribution over \(t\) and \(\varphi\), considering the weighted contributions from the kime-phase distribution \(\Phi(\varphi)\). In principle, \(t\) and \(\varphi\) may be independent or interdependent. \(\Phi(t,\varphi)\) dependence on \(t\) facilitates modeling temporal dynamics of the kime-phase distribution, e.g., arXiv article “Predicting the Future Behavior of a Time-Varying Probability Distribution”.

  1. Kevent Information Content: Analogous to Shannon’s information content, the information content of a kevent \(K\) can be defined as

\[I(K) = -\log \text{Pr}(K),\] where \(\text{Pr}(K)\) is the probability of the kevent. Since the probability \(\text{Pr}(K)\) involves both the real time and kime-phase components, the information content captures the complexity and uncertainty associated with events in the spacekime framework. Clearly, the entropy is the expectation of the information content.

Let’s explore several examples of kevents and calculate their information contents assuming different measures for the positive reals (for time) and alternative phase distribution priors for the kime-phase.

3.1.1 Examples of Kevents and their Information Contents

Again, to summarize the key concepts:

  • Sample Space \(\Omega\): The complex plane parameterized by \(\kappa = t e^{i\varphi}\), where \(t \geq 0\) is the time component and the kime-phase \(\varphi \sim \Phi_{[-\pi, \pi)}(\varphi)\).
  • Measure:
    • For time: Different measures over the positive reals \((\mathbb{R}^+)\). Most useful may be the normalized probability measures.
    • For kime-phase: Various prior distributions \(\Phi(\varphi)\) supported on \([- \pi, \pi)\).
  • Random Variable (observable) \(X(\omega)\): A real-valued function mapping the sample space onto the positive reals, \(X:\Omega\to\mathbb{R}^+\).
  • The information content of a kevent \(K = \{ \omega \in \Omega \mid u < X(\omega) \leq v \}\) is given by \(I(K) = -\log \text{Pr}(K)\), where \(\text{Pr}(K)\) is the probability (measure) of the kevent.
  • The entropy of a kevent \(K = \{ \omega \in \Omega \mid u < X(\omega) \leq v \}\), or the corresponding random variable \(X\) is the expected value of the self-information of \(X\). The entropy effectively quantifies the indeterminancy the random variable. i.e., its average uncertainty. The entrooy is the average amount of self-information one can expect to gain about a random variable when measuring it, \(\mathbb{E}_X(I(K))\).

3.1.1.1 Example 1: Exponential Measure on Time and Uniform Kime-Phase Distribution

  • Time Measure: Exponential distribution with rate parameter \(\lambda > 0\): \[f_t(t) = \lambda e^{-\lambda t}, \quad t \geq 0.\]

  • Kime-Phase Distribution: Uniform distribution on \([- \pi, \pi)\): \[\Phi(\varphi) = \frac{1}{2\pi}, \quad \varphi \in [-\pi, \pi).\]

Kevent Probability: Consider \(X(\omega) = t\) as the time component. The probability of the kevent \(K = \{ u < t \leq v \}\) is \[\text{Pr}(K) = \text{Pr}(u < X \leq v) = F_X(v) - F_X(u) =\\ \iint f_t(t)\cdot \Phi(\varphi) dt d\varphi = \left (\int_u^v \lambda e^{-\lambda t} dt \right ) \left (\int_{-\pi}^{\pi} \frac{1}{2\pi} d\varphi\right ).\]

\[\text{Pr}(K) = \left(-e^{-\lambda t} \Big|_u^v\right) \times \left(\frac{2\pi}{2\pi}\right) = e^{-\lambda u} - e^{-\lambda v}.\]

Information Content: \[I(K) = -\log(e^{-\lambda u} - e^{-\lambda v}).\]

Finally, the entropy \(H(K)\) of the kevent \(K\) can be computed using the differential entropy definition

\[H(K) = -\int_u^v p_T(t) \log p_T(t) \, dt.\]

Substitute the expression for the time probability measure, \(p_T(t)\)

\[H(K) = -\int_u^v \lambda e^{-\lambda t} \log(\lambda e^{-\lambda t}) \, dt\]

\[H(K) = -\int_u^v \lambda e^{-\lambda t} \left( \log(\lambda) - \lambda t \right) \, dt,\]

which can eb slit into simple integrals.

  1. First Integral: \(\int_u^v \lambda e^{-\lambda t} \, dt = e^{-\lambda u} - e^{-\lambda v}.\)

  2. Second Integral: Let’s compute the integral \(\int_u^v t \lambda^2 e^{-\lambda t} \, dt\). This is evaluated using integration by parts

\[\int t e^{-\lambda t} dt = -\frac{t}{\lambda} e^{-\lambda t} - \frac{1}{\lambda^2} e^{-\lambda t} + C.\]

\[\int_u^v t \lambda^2 e^{-\lambda t} dt = \left[-t \lambda e^{-\lambda t} - e^{-\lambda t}\right]_u^v = \left[-v \lambda e^{-\lambda v} - e^{-\lambda v}\right] - \left[-u \lambda e^{-\lambda u} - e^{-\lambda u}\right].\]

\[\lambda \int_u^v t \lambda e^{-\lambda t} \, dt = \lambda \left[u e^{-\lambda u} - v e^{-\lambda v}\right] + \lambda \left(e^{-\lambda u} - e^{-\lambda v}\right).\]

Substituting the results of the integrals back into the entropy expression

\[H(K) = -\log(\lambda) \left(e^{-\lambda u} - e^{-\lambda v}\right) + \lambda \left[\lambda \left(u e^{-\lambda u} - v e^{-\lambda v}\right) + e^{-\lambda u} - e^{-\lambda v}\right].\]

\[H(K) = -\log(\lambda) \left(e^{-\lambda u} - e^{-\lambda v}\right) + \lambda \left[ \lambda u e^{-\lambda u} - \lambda v e^{-\lambda v} + e^{-\lambda u} - e^{-\lambda v}\right].\]

\[H(K) = -\log(\lambda) \left(e^{-\lambda u} - e^{-\lambda v}\right) + \lambda e^{-\lambda u} (\lambda u + 1) - \lambda e^{-\lambda v} (\lambda v + 1).\]

Hence, the entropy of the kevent \(K = \{ \omega \in \Omega \mid u < X(\omega) \leq v \}\), where \(X(\omega) = t\) and \(t\) follows an exponential distribution with a uniform kime-phase distribution, is

\[H(K) = -\log(\lambda) \left(e^{-\lambda u} - e^{-\lambda v}\right) + \lambda e^{-\lambda u} (\lambda u + 1) - \lambda e^{-\lambda v} (\lambda v + 1).\]

3.1.1.2 Example 2: Power-Law Measure on Time and Gaussian Kime-Phase Distribution

Next we consider the kevent \(K = \{ \omega \in \Omega \mid u < X(\omega) \leq v \}\) where \(X(\omega) = \log(t)\), and the probability measure is defined using a power-law distribution on time and a Gaussian distribution on the kime-phase.

  • Time Distribution: Let \(t\) follow a power-law distribution with parameter \(\alpha > 1\),

\[p_T(t) = \frac{\alpha - 1}{t_\text{min}} \left(\frac{t_\text{min}}{t}\right)^{\alpha}, \quad t \geq t_\text{min}.\]

  • Kime-Phase Distribution: \(\varphi\) follows a Gaussian distribution centered at \(\mu\) with variance \(\sigma^2\)

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left( -\frac{(\varphi - \mu)^2}{2\sigma^2}\right).\]

  • Kevent Definition: The kevent \(K\) is defined as the set of all \(\omega \in \Omega\) such that \(u < X(\omega) \leq v\), where \(X(\omega) = \log(t)\).

Given that \(X(\omega) = \log(t)\), the probability of the kevent \(K\) can be expressed as \(\text{Pr}(K) = \text{Pr}(u < \log(t) \leq v) = \text{Pr}(e^u < t \leq e^v).\)

The kevent probability can be computed using the cumulative distribution function (CDF) of the power-law distribution

\[\text{Pr}(K) = \int_{e^u}^{e^v} p_T(t) \, dt = \int_{e^u}^{e^v} \frac{\alpha - 1}{t_\text{min}} \left(\frac{t_\text{min}}{t}\right)^{\alpha} \, dt.\]

This integral evaluates to

\[\text{Pr}(K) = \left(\frac{e^u}{t_\text{min}}\right)^{1-\alpha} - \left(\frac{e^v}{t_\text{min}}\right)^{1-\alpha}.\]

This probability is the difference in the cumulative distribution values at \(e^u\) and \(e^v\).

The information content \(I(K)\) of the kevent \(K\) is given by \(I(K) = -\log \text{Pr}(K).\) Substituting the expression for \(\text{Pr}(K)\)

\[I(K) = -\log\left[\left(\frac{e^u}{t_\text{min}}\right)^{1-\alpha} - \left(\frac{e^v}{t_\text{min}}\right)^{1-\alpha}\right].\]

This expression quantifies the “surprise” or information gained by observing the kevent \(K\).

The entropy \(H(K)\) of the kevent \(K\) is calculated using the differential entropy formula, adapted for the continuous probability distribution \(H(K) = -\int_{e^u}^{e^v} p_T(t) \log p_T(t) \, dt.\)

Substituting the expression for \(p_T(t)\)

\[H(K) = -\int_{e^u}^{e^v} \frac{\alpha - 1}{t_\text{min}} \left(\frac{t_\text{min}}{t}\right)^{\alpha} \log\left(\frac{\alpha - 1} {t_\text{min}} \left(\frac{t_\text{min}}{t}\right)^{\alpha}\right) \, dt.\]

Simplify the logarithm

\[H(K) = -\int_{e^u}^{e^v} \frac{\alpha - 1}{t_\text{min}} \left(\frac{t_\text{min}}{t}\right)^{\alpha} \left[\log\left( \frac{\alpha - 1}{t_\text{min}}\right) - \alpha \log(t)\right] \, dt.\]

This can be split into two integrals

\[H(K) = -\log\left(\frac{\alpha - 1}{t_\text{min}}\right) \underbrace{\int_{e^u}^{e^v} \frac{\alpha - 1}{t_\text{min}} \left(\frac{t_\text{min}}{t}\right)^{\alpha} \, dt}_{I} + \alpha\underbrace{\int_{e^u}^{e^v} \frac{\alpha - 1}{t_\text{min}} \left(\frac{t_\text{min}}{t}\right)^{\alpha} \log(t) \, dt}_{II}.\]

  1. First Integral

\[I=\int_{e^u}^{e^v} \frac{\alpha - 1}{t_\text{min}} \left(\frac{t_\text{min}}{t}\right)^{\alpha} \, dt = \text{Pr}(K) = \left(\frac{e^u}{t_\text{min}}\right)^{1-\alpha} - \left(\frac{e^v}{t_\text{min}}\right)^{1-\alpha}.\]

  1. Second Integral: \(\int_{e^u}^{e^v} \frac{1}{t^{\alpha}} \log(t) \, dt\), can be evaluated using integration by parts or using the integral identity

\[II=\int t^{-\alpha} \log(t) \, dt = \frac{t^{1-\alpha}}{1-\alpha} \left[\log(t) - \frac{1}{1-\alpha}\right].\]

Applying this to our limits yields

\[\int_{e^u}^{e^v} \frac{1}{t^{\alpha}} \log(t) \, dt = \frac{1}{1-\alpha} \left[e^{v(1-\alpha)} \log(e^v) - e^{v(1-\alpha)} - e^{u(1-\alpha)} \log(e^u) + e^{u(1-\alpha)}\right].\]

Simplify this integral

\[\int_{e^u}^{e^v} \frac{1}{t^{\alpha}} \log(t) \, dt = \frac{e^{v(1-\alpha)} v - e^{u(1-\alpha)} u}{1-\alpha} + \frac{e^{v(1-\alpha)} - e^{u(1-\alpha)}}{(1-\alpha)^2}.\]

Hence, the entropy \(H(K)\) becomes

\[H(K) = -\log\left(\frac{\alpha - 1}{t_\text{min}}\right) \left[\left(\frac{e^u}{t_\text{min}}\right)^{1-\alpha} - \left(\frac{e^v}{t_\text{min}}\right)^{1-\alpha}\right] + \alpha \cdot \frac{\alpha - 1}{t_\text{min}} \cdot \left[\frac{e^{v(1-\alpha)} v - e^{u(1-\alpha)} u}{1-\alpha} + \frac{e^{v(1-\alpha)} - e^{u(1-\alpha)}}{(1-\alpha)^2}\right].\]

This expression captures the entropy associated with the kevent \(K\), accounting for the power-law distribution of time and the logarithmic nature of the time random variable \(X(\omega) = \log(t)\).

3.1.1.3 Example 3: Exponential Density over Time and Cauchy (Lorentzian) Kime-Phase Distribution

In this example, we will compute the kevent probability, information content, and entropy for a random variable \(X(\omega)=t+\varphi\), assuming exponential distribution over time \([0, T]\), and Cauchy (Lorentzian) kime-phase distribution.

  • Time Distribution: \(t\) follows an exponential distribution with rate parameter \(\lambda\) over the interval \([0, T]\)

\[p_T(t) = \frac{\lambda e^{-\lambda t}}{1 - e^{-\lambda T}}, \quad 0 \leq t \leq T.\]

  • Kime-Phase Distribution: \(\varphi\) follows a Cauchy (Lorentzian) distribution with location parameter \(\mu\) and scale parameter \(\gamma\)

\[\Phi(\varphi) = \frac{1}{\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]}, \quad \varphi \in \mathbb{R}.\]

  • Kevent Definition: The kevent \(K\) is defined as the set of all \(\omega \in \Omega\) such that \(u < X(\omega) \leq v\), where \(X(\omega) = t + \varphi\).

To compute the probability of the kevent \(K\), we need to integrate the joint probability distribution over \(t\) and \(\varphi\), such that \(u < t + \varphi \leq v\).

The joint probability density function is

\[p(t, \varphi) = p_T(t) \Phi(\varphi) = \frac{\lambda e^{-\lambda t}}{1 - e^{-\lambda T}} \cdot \frac{1}{\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]}.\]

The probability of the kevent \(K\) is \[\text{Pr}(K) = \int_0^T \int_{u - t}^{v - t} p(t, \varphi) \, d\varphi \, dt.\]

Substituting the joint distribution

\[\text{Pr}(K) = \int_0^T \int_{u - t}^{v - t} \frac{\lambda e^{-\lambda t}} {1 - e^{-\lambda T}} \cdot \frac{1}{\pi \gamma \left[1 + \left( \frac{\varphi - \mu}{\gamma}\right)^2\right]} \, d\varphi \, dt.\]

The inner integral with respect to \(\varphi\) is a standard Cauchy integral

\[\int_{u - t}^{v - t} \frac{1}{\pi \gamma \left[1 + \left( \frac{\varphi - \mu}{\gamma}\right)^2\right]} \, d\varphi.\]

This integral can be evaluated as

\[\text{Pr}(K) = \frac{\lambda}{1 - e^{-\lambda T}} \int_0^T \frac{1}{\pi} \left[ \tan^{-1}\left(\frac{v - t - \mu}{\gamma}\right) - \tan^{-1}\left( \frac{u - t - \mu}{\gamma}\right) \right] e^{-\lambda t} \, dt.\]

The information content \(I(K)\) is \(I(K) = -\log \text{Pr}(K),\) which quantifies the “surprise” or information gained by observing the kevent \(K\).

The entropy \(H(K)\) of the kevent \(K\) is calculated using the differential entropy formula adapted for this probability distribution

\[H(K) = -\int_0^T \int_{u - t}^{v - t} p(t, \varphi) \log p(t, \varphi) \, d\varphi \, dt.\]

Substituting the joint distribution

\[H(K) = -\int_0^T \int_{u - t}^{v - t} \frac{\lambda e^{-\lambda t}}{1 - e^{-\lambda T}} \cdot \frac{1}{\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]} \log \left( \frac{\lambda e^{-\lambda t}}{1 - e^{-\lambda T}} \cdot \frac{1}{\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]} \right) d\varphi \, dt.\]

This can be split into two integrals:

\[H(K) = -\int_0^T \int_{u - t}^{v - t} p(t, \varphi) \left[\log \left( \frac{\lambda e^{-\lambda t}}{1 - e^{-\lambda T}} \right) + \log \left( \frac{1}{\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]} \right)\right] d\varphi \, dt.\]

The entropy simplifies to

\[H(K) = -\log \left( \frac{\lambda}{1 - e^{-\lambda T}} \right) \text{Pr}(K) - \int_0^T \int_{u - t}^{v - t} p(t, \varphi) \log \left( \frac{e^{-\lambda t}} {\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]} \right) d\varphi \, dt,\]

which can be numerically computed given a specific set of input parameters using special functions like the \(\arctan(\cdot)\) (for the Cauchy distribution integral). The information content and entropy provide insights into the uncertainty associated with the kevent, considering the contributions from both the exponential time distribution and the Cauchy kime-phase distribution. Below we show an example of how to implement and estimate these quantities in practice.

### Step 1: Install and Load Necessary Libraries
# install.packages("cubature")
library(cubature)

### Step 2: Define the Probability Density Functions. Define the exponential 
### distribution for time and the Cauchy distribution for the kime-phase.

# Parameters
lambda <- 1  # Rate parameter for the exponential distribution
T_max <- 10  # Maximum time value (upper limit of the time interval)
mu <- 0      # Location parameter for the Cauchy distribution
gamma <- 1   # Scale parameter for the Cauchy distribution
u <- 1       # Lower bound for kevent
v <- 3       # Upper bound for kevent

# Exponential distribution PDF
p_T <- function(t) {
  lambda * exp(-lambda * t) / (1 - exp(-lambda * T_max))
}

# Cauchy (Lorentzian) distribution PDF
Phi <- function(phi) {
  1 / (pi * gamma * (1 + ((phi - mu) / gamma)^2))
}

# Joint PDF for t + phi
joint_pdf <- function(x) {
  t <- x[1]
  phi <- x[2]
  p_T(t) * Phi(phi)
}

### Step 3: Define the Integration Limits for Kevent Probability
### We'll integrate over the joint PDF where \( u < t + \varphi \leq v \).
# Define the integration function for kevent probability
kevent_prob_func <- function(x) {
  t <- x[1]
  phi <- x[2]
  if (u < t + phi & t + phi <= v) {
    return(joint_pdf(c(t, phi)))
  } else {
    return(0)
  }
}

### Step 4: Numerical Integration for Kevent Probability
### Numerically estimate the kevent probability
kevent_prob <- adaptIntegrate(
  kevent_prob_func,
  lowerLimit = c(0, -Inf),
  upperLimit = c(T_max, Inf)
)$integral

print(paste("Kevent Probability:", kevent_prob))

### Step 5: Calculate the Information Content
# The information content \(I(K)\) is the negative logarithm of the kevent probability.
# Calculate Information Content
information_content <- -log(kevent_prob)
print(paste("Information Content:", information_content))

### Step 6: Define the Entropy Calculation Function
# calculate the entropy by integrating \( -p(t, \varphi) \log p(t, \varphi) \) over the region defined by \( u < t + \varphi \leq v \).
# Define the integration function for entropy
entropy_func <- function(x) {
  t <- x[1]
  phi <- x[2]
  if (u < t + phi & t + phi <= v) {
    p <- joint_pdf(c(t, phi))
    return(-p * log(p))
  } else {
    return(0)
  }
}

### Step 7: Numerical Integration for Entropy
# Use `cubature::adaptIntegrate()` to estimate the entropy.

# Numerically estimate the entropy
entropy <- adaptIntegrate(
  entropy_func,
  lowerLimit = c(0, -Inf),
  upperLimit = c(T_max, Inf)
)$integral

print(paste("Entropy:", entropy))

These integrals may not be generally solvable in closed form but can be approximated numerically. The R code above uses the following parameters:

  • lambda = 1: Rate parameter for the exponential distribution
  • T_max = 10: Maximum time value (upper limit of the time interval)
  • mu = 0: Location parameter for the Cauchy distribution
  • gamma = 1: Scale parameter for the Cauchy distribution
  • u = 1: Lower bound for kevent
  • v = 3: Upper bound for kevent.

Here are the estimates of the probability of the kevent, \(\text{Pr}(K)\approx\) 0.5, its information content, \(I(K) = -\log \text{Pr}(K) \approx\) 0, and its entropy, \(H(K)=\mathbb{E}(I(K)) = \approx\) 0.

3.1.1.4 Example 3: Exponential Density over Time and Cauchy (Lorentzian) Kime-Phase Distribution over an Elliptical Region

Let’s modify the previous Example 3, by introducing an elliptical region defined by the parametric equations \(x(\varphi) = 1 + 4\cos(\varphi)\) and \(y(\varphi) = 2\cos(\varphi)\). Specifically, we’ll redefine the joint probability over this elliptical area and calculate the probability of the corresponding kevent and its information content.

For \(\varphi \in [-\pi, \pi)\), the parametric equations \[x(\varphi) = 1 + 4\cos(\varphi) \\ y(\varphi) = 2\cos(\varphi)\]

bound an elliptical region in the plane and we are interested in the probability inside this region. Here are the basic steps in the protocol:

  1. Redefine the Joint Probability Density Function (PDF) over the Elliptical Region: We’ll modify the region of integration to include only points that fall within the ellipse defined by the parametric equations.
  2. Calculate the Probability of the Kevent: We’ll integrate the joint PDF over the elliptical region.
  3. Calculate the Information Content: We’ll use the negative logarithm of the kevent probability.
  4. Calculate the Entropy: We’ll integrate \(-p(t, \varphi) \log p(t, \varphi)\) over the elliptical region.
# Load necessary libraries
library(cubature)
library(plotly)

# Parameters
gamma <- 1.0  # scale parameter for Cauchy distribution
u <- 2.0  # lower bound of kevent
v <- 8.0  # upper bound of kevent

# Parametric equations for the elliptical region
# Define the parametric functions
x_param <- function(phi) {
  return(1 + 4 * cos(phi))
}

y_param <- function(phi) {
  return(2 * sin(phi))
}

# Generate a sequence of phi values
phi_values <- seq(-pi, pi, length.out = 1000)

# Calculate x and y coordinates
x_values <- x_param(phi_values)
y_values <- y_param(phi_values)

# Create a long dataframe
ellipse_df <- data.frame(phi = phi_values, x = x_values, y = y_values)

# Plot the ellipse using plot_ly
ellipse_plot <- plot_ly(ellipse_df, x = ~x, y = ~y, type = 'scatter', mode = 'lines',
                        line = list(color = 'blue', width = 2)) %>%
  layout(title = 'Ellipse Boundary in the Complex Plane',
         xaxis = list(title = 'k1', range = c(min(x_values)-1, max(x_values)+1)),
         yaxis = list(title = 'k2', range = c(min(y_values)-1, max(y_values)+1)),
         showlegend = FALSE)

# Show the plot
ellipse_plot

### Step 1: Redefine the Joint Probability Function Over the Ellipse
#   We'll integrate the joint PDF over this elliptical area.

### Step 2: Define the Joint Probability Density Function
# Define the parametric equations for the elliptical region
x_phi <- function(phi) {
  1 + 4 * cos(phi)
}

y_phi <- function(phi) {
  2 * cos(phi)
}

# Redefine the joint PDF with the elliptical region constraints
# Redefine the joint PDF with the elliptical region constraints using the implicit equation
joint_pdf_ellipse <- function(x) {
  t <- x[1]
  phi <- x[2]
  
  # Check if (t, phi) is inside the elliptical region
  if ((t - 1)^2 / 16 + (phi / 2)^2 <= 1) {
    if (u < t + phi & t + phi <= v) {
      return(joint_pdf(c(t, phi)))
    }
  }
  
  return(0)
}

### Step 3: Numerical Integration for Kevent Probability Over the Ellipse
# Numerically estimate the kevent probability over the ellipse
kevent_prob_ellipse <- adaptIntegrate(
  joint_pdf_ellipse,
  lowerLimit = c(0, -pi),
  upperLimit = c(T_max, pi)
)$integral

print(paste("Kevent Probability over Ellipse:", kevent_prob_ellipse))

### Step 4: Calculate the Information Content
# Calculate Information Content
information_content_ellipse <- -log(kevent_prob_ellipse)
print(paste("Information Content over Ellipse:", information_content_ellipse))

# Define the integration function for entropy over the ellipse
entropy_func_ellipse <- function(x) {
  t <- x[1]
  phi <- x[2]
  
  # Check if (t, phi) is inside the elliptical region using the implicit equation
  if ((t - 1)^2 / 16 + (phi / 2)^2 <= 1) {
    if (u < t + phi & t + phi <= v) {
      p <- joint_pdf(c(t, phi))
      return(-p * log(p))
    }
  }
  
  return(0)
}

# Numerically estimate the entropy over the ellipse
entropy_ellipse <- adaptIntegrate(
  entropy_func_ellipse,
  lowerLimit = c(0, -pi),
  upperLimit = c(T_max, pi)
)$integral

print(paste("Entropy over Ellipse:", entropy_ellipse))

Again, these integrals may be approximated numerically. The R code above uses the following parameters:

  • lambda = 1: Rate parameter for the exponential distribution
  • mu = 0: Location parameter for the Cauchy distribution
  • gamma = 1: Scale parameter for the Cauchy distribution
  • u = 2: Lower bound for kevent
  • v = 8: Upper bound for kevent.

Here are the estimates of the probability of the kevent, \(\text{Pr}(K)\approx\) 0.5, its information content, \(I(K) = -\log \text{Pr}(K) \approx\) 0, and its entropy, \(H(K)=\mathbb{E}(I(K)) = \approx\) 0.

Other examples computing the probability of kevents can be shown by integrating over both the time component and the kime-phase distribution. The resulting information content, \(I(K)\), is a measure of the uncertainty associated with these complex-time events and reflect both the classical temporal dynamics as well as the prior kime-phase model distribution.

3.2 Spacekime Events (Skevents)

Next, we will expand the definition of kime events (kevents) to the entire 5D spacekime. This requires formally defining the structure of spacekime, including its metric and \(\sigma\)-algebra. This expanded framework will allow us to describe spacekime events (skevents) jointly covering both spatial dimensions and kime dimensions.

3.2.1 5D Spacekime Definition

The 5D spacekime consists of \(3\) spatial dimensions and \(2\) kime dimensions where points are represented as

\[\mathbf{X} = \left (\underbrace{x^1, x^2, x^3}_{space}, \underbrace{\kappa^1, \kappa^2}_{kime} \right ).\]

Depending on the need, the kime representation may involve either Cartesian coordinates \(\kappa^1,\kappa^2\in\mathbb{R}\) or polar coordinates \(t\in\mathbb{R}^+, \varphi\sim\Phi_{[-\pi,\pi)}(\varphi)\).

Let’s define spacekime events, or skevents, as stochastic encounters over spacekime \(\Omega=\mathbb{R}^3\times \mathbb{C}\). Again, skevents can be defined over different system states representations. For instance, using Cartesian (position) and polar (kime) representation, events are expressed as \(\left(\underbrace{x,y,z}_{space},\underbrace{\kappa=t e^{i\varphi}}_{kime}\right )\), where \(t \in\mathbb{R}^+\) represents the usual Euclidean time and \(\varphi \in [-\pi, \pi)\) represents the kime-phase with a prior distribution \(\Phi(\varphi)\), modeling the intrinsic experimental process variability of repeated measurements.

  1. Measure on \(\Omega\): The measure on \(\Omega\) is composed of three parts:

    • Measure on \((x,y,z)\): For the spatial component, which could be Euclidean, but in practice most often is a metric tensor.
    • Measure on \(t\): For the real-time component, which could be Euclidean and can be part of the spacetime metric tensor. The (marginal) time metric may follow a normalized probability measure, e.g., Exponential (decay) distribution.
    • Kime-Phase Measure \(\Phi(\varphi)\): A prior phase distribution supported on \([- \pi, \pi)\).
  2. Random Variable (observable) \(X(\omega)\): Let \(X: \Omega \rightarrow \mathbb{R}\) be a real-valued random variable defined on the sample space \(\Omega\). For a given \(\omega \in \Omega\), \(X(\omega)\) maps \(\omega\) to a real number.

  3. Skevent: A spacekime event (skevent) is defined by the set: \[K = \{ \omega \in \Omega \mid u < X(\omega) \leq v \}.\] This skevent represents the inverse image of the interval \((u, v]\) under the mapping \(X\). The skevent construction generalizes the classical notion of events and complex-time kevents to the 5D spacekime, where events are defined not just in terms of their spatio-temporal order (in spacetime), but also incorporate the kime-phase distribution \(\Phi(\varphi)\) reflecting the intrinsic variation of the sampling distribution corresponding with the observable (RV) \(X\).

  4. Probability of a Skevent: The probability of a skevent \(K\) is given by \[\text{Pr}(u < X \leq v) = F(v) - F(u),\] where \(F(x)\) is the cumulative distribution function (CDF) of the random variable \(X\).

The probability measure of the skevent \(K\) is derived by considering the space, time, and kime-phase components. Specifically, we calculate \(\text{Pr}(K) = \text{Pr}(u < X \leq v) = \int_{X^{-1}((u,v])} d\mu(\omega),\) where \(d\mu(\omega)\) is the product measure on the sample space \(\Omega\) given by the probability measure on spacetime \(t\) and the prior phase distribution \(\Phi(\varphi)\) on \(\varphi\).

The cumulative distribution function \(F(x)\) is derived from the joint spacekime distribution over \((x,y,z,t)\) and \(\varphi\), considering the weighted contributions from the kime-phase distribution \(\Phi(\varphi)\). In principle, \(t\) and \(\varphi\) may be independent or interdependent. \(\Phi(t,\varphi)\) dependence on \(t\) facilitates modeling temporal dynamics of the kime-phase distribution, e.g., arXiv article “Predicting the Future Behavior of a Time-Varying Probability Distribution”.

  1. Skevent Information Content: Analogous to Shannon’s information content, the information content of a skevent \(K\) can be defined as

\[I(K) = -\log \text{Pr}(K),\] where \(\text{Pr}(K)\) is the probability of the skevent. Since the probability \(\text{Pr}(K)\) involves both spacetime and kime-phase components, the information content captures the complexity and uncertainty associated with skevents in the spacekime framework. Naturally, the entropy is the expectation of the information content.

After we formulate the spacekime metric, spacekime \(\sigma\)-Algebra, and spacekime probability measure, we will explore several examples of skevents and demonstrate calculating their information contents.

3.2.2 Spacekime Metric

The 5D spacekime metric \(g_{AB}\) defines distances between points. It naturally extends the usual 4D spacetime metric to include the kime-phase component.

In spacekime, the kime-phase distribution is more than a simple value or index.

  1. Kime-Phase \(\varphi\): In spacekime , \(\varphi\) is associated with the kime-phase, which is a parameter reflecting the nature of repeated measurement longitudinal processes, i.e., \(\varphi\) is more than just a coordinate; it can be considered as a parameter that is part of a distribution \(\Phi(\varphi)\), which models the variability or uncertainty in the phase component of time. Due to unavoidable quantum fluctuations, explicit (causal) and implicit (correlational) dependencies, in most controlled experiments, individual measurements, single observations, and small samples are always intrinsically stochastic. Such recordings (data) represent random drawing from some (known or unknown) kime-phase distributions, \(\Phi\). In practice, to estimate the system properties, the first and second fundamental theorems of probability theory (LLN and CLT) provide the justification using statistics to pool multiple repeated observations (IID samples) and generate more reliable estimates of various process characteristics, such as the mean, variance, quantiles, etc. These repeated measurement pooling strategies for characterizing system properties are still varying (from one sample to the next). However, their sampling distributions are known to have markedly reduced dispersion.

  2. Metric and the Kime-Phase Distribution: The spacekime metric needs to incorporate the effect of the kime-phase distribution, typically by defining a corresponding component in the metric that reflects the variability of observable measurements from controlled experiments, represented by the kime-phase.

Given that \(\varphi\) is associated with a distribution rather than a fixed value, the spacekime metric should incorporate the variability and distributional characteristics of the kime-phase

\[ds^2 = g_{\mu\nu} dx^\mu dx^\nu + \langle g_{\varphi\varphi} \rangle d\varphi^2 + \langle g_{t\varphi} \rangle dt d\varphi,\]

where

  • \(g_{\mu\nu}\) is the classical spacetime metric for the 3 spatial dimensions and 1 time dimension.
  • \(\langle g_{\varphi\varphi} \rangle\) represents the expected value, i.e., averaged contribution of the kime-phase distribution to the metric. This term accounts for the gravitational balance of the kime-phase distribution \(\Phi(\varphi)\).
  • \(\langle g_{t\varphi} \rangle\) represents any correlations or interactions between the time component \(t\) and the kime-phase distribution \(\Phi(\varphi)\).

Incorporating the kime-phase distribution \(\Phi(\varphi)\) into the metric might involve:

  • Using the expected value of the phase distribution: \[\langle g_{\varphi\varphi} \rangle = \int_{-\pi}^{\pi} g_{\varphi\varphi}(\varphi) \Phi(\varphi) d\varphi.\] This integral reflects the averaged influence of the phase distribution on the metric.

  • Alternatively, if the phase distribution induces variability in the metric, we could model this explicitly by integrating over the phase distribution to get effective metric components.

Note: In general, the component of the spacekime metric corresponding to \(\langle g_{\varphi\varphi} \rangle\), which represents the averaged or expected contribution of the kime-phase distribution \(\Phi(\varphi)\) to the metric, will not necessarily be trivial (i.e., zero), even for symmetric phase distributions, \(\Phi(\varphi)\), \(\Phi(\varphi) = \Phi(-\varphi)\). This symmetry implies that for every positive phase \(\varphi\), there is a corresponding negative phase \(-\varphi\) with the same probability density.

The expected contribution to the metric from the kime-phase is given by

\[\langle g_{\varphi\varphi} \rangle = \int_{-\infty}^{\infty} g_{\varphi\varphi}(\varphi) \Phi(\varphi) \, d\varphi.\]

If \(\langle g_{\varphi\varphi} \rangle=0\), the integrand \(g_{\varphi\varphi}(\varphi) \Phi(\varphi)\) must satisfy

\[\int_{-\infty}^{\infty} g_{\varphi\varphi}(\varphi) \Phi(\varphi) \, d\varphi=0.\] When \(g_{\varphi\varphi}(\varphi)\) is an odd function (i.e., \(g_{\varphi\varphi}(\varphi) = -g_{\varphi\varphi}(-\varphi)\)), and \(\Phi(\varphi)\) is symmetric, then the integral over the entire real line (or any symmetric interval) would be trivial, since the positive and negative contributions would exactly cancel out

\[\int_{-\infty}^{\infty} g_{\varphi\varphi}(\varphi) \Phi(\varphi) \, d\varphi = 0.\] However, even for symmetric distributions, if \(g_{\varphi\varphi}(\varphi)\) has no odd component or is constant, the integral may not be zero.

Therefore, the contribution \(\langle g_{\varphi\varphi} \rangle\) in the spacekime metric may not be expected to be zero, even for symmetric phase distributions. It will be zero only when the integrand is an odd function or when specific conditions cause the positive and negative contributions to cancel out.

3.2.3 Spacekime \(\sigma\)-Algebra

To define spacekime events (skevents), we need to define a \(\sigma\)-algebra \(\mathcal{F}\) over the 5D spacekime. This \(\sigma\)-algebra consists of sets of points (skevents) including

  • Spatial component: Measurable subsets of the 3D spatial space \(\mathbb{R}^3\).
  • Temporal component: Intervals or other measurable sets in time \(\mathbb{R}^+\).
  • Kime-phase component: Measurable sets on the support of the kime-phase distribution, \([-\pi, \pi)\).

The \(\sigma\)-algebra \(\mathcal{F}\) for the 5D spacekime space is the product \(\sigma\)-algebra

\[\mathcal{F} = \mathcal{F}_x \times \mathcal{F}_t \times \mathcal{F}_\varphi,\]

where

  • \(\mathcal{F}_x\) is the Borel \(\sigma\)-algebra on \(\mathbb{R}^3\).
  • \(\mathcal{F}_t\) is the Borel \(\sigma\)-algebra on the time dimension, \(\mathbb{R}^+\).
  • \(\mathcal{F}_\varphi\) is the \(\sigma\)-algebra associated with the kime-phase, typically the Borel \(\sigma\)-algebra on the circle, \(S^1\).

3.2.4 Probability Measure in 5D Spacekime

For each skevent \(A \in \mathcal{F}\), define a probability measure \(\mathbb{P}\) on \((\Omega, \mathcal{F})\), where \(\Omega = \mathbb{R}^3 \times \mathbb{R}^+ \times [-\pi, \pi)\) is the sample space. The measure \(\mathbb{P}\) can be constructed as:

\[\mathbb{P}(A) = \int_A f(\mathbf{X}) \, d\mathbf{X},\]

where \(f(\mathbf{X})\) is a probability density function defined on spacekime.

Under the separability condition, the function \(f(\mathbf{X})\) could factor as \[f(\mathbf{X}) = f_x(x^1, x^2, x^3) \times f_t(t) \times \Phi(\varphi),\]

where

  • \(f_x(x^1, x^2, x^3)\) is the spatial density.
  • \(f_t(t)\) is the temporal density.
  • \(\Phi(\varphi)\) is the kime-phase distribution.

3.2.5 Skevents in 5D Spacekime

A 5D spacekime event is defined as

\[K = \{ \mathbf{X} \in \Omega \mid u < X(\mathbf{X}) \leq v \},\]

where \(X(\mathbf{X})\) is a real-valued function of the 5D coordinates. The probability of the skevent is given by

\[\text{Pr}(K) = \mathbb{P}(K) = \int_{K} f(\mathbf{X}) \, d\mathbf{X}.\]

The 5D spacekime is fully defined by a metric that encompasses both spatial and kime components, a \(\sigma\)-algebra that combines the measurable events across all dimensions, and a probability measure that extends across this entire space. This allows for a unified treatment of events in a higher-dimensional context that includes both classical space and the complex time (kime) dimensions.

Let’s work through several examples of specific spacekime events (skevents) by considering different spatial regions in \(\mathbb{R}^3\) and different kime-phase distributions. For each example, we’ll define the spacekime metric, illustrate sets in the spacekime \(\sigma\)-Algebra, and compute the corresponding probability measures.

3.2.6 Examples

3.2.6.1 Skevent in a Cubic Spatial Region with Uniform Kime-Phase Distribution

  1. Spatial Region (Volume in \(\mathbb{R}^3\)): Consider a cubic region in \(\mathbb{R}^3\) defined by:

\[V_1 = \{(x, y, z) \in \mathbb{R}^3 \mid 0 \leq x \leq 1, 0 \leq y \leq 1, 0 \leq z \leq 1\}.\]

Here’s a simple R plotly code for plotting a semi-transparent cube of size 1 in 3D:

library(plotly)

# Define the vertices of the cube
x = c(0, 0, 1, 1, 0, 0, 1, 1)
y = c(0, 1, 1, 0, 0, 1, 1, 0)
z = c(0, 0, 0, 0, 1, 1, 1, 1)

# Define the faces of the cube using vertex indices
# Each face is composed of 2 triangles
i = c(7, 0, 0, 0, 4, 4, 6, 6, 4, 0, 3, 2)
j = c(3, 4, 1, 2, 5, 6, 5, 2, 0, 1, 6, 3)
k = c(0, 7, 2, 3, 6, 7, 1, 1, 5, 5, 7, 6)

# plot_ly(type = 'mesh3d',
#   x = c(0, 0, 1, 1, 0, 0, 1, 1),
#   y = c(0, 1, 1, 0, 0, 1, 1, 0),
#   z = c(0, 0, 0, 0, 1, 1, 1, 1),
#   i = c(7, 0, 0, 0, 4, 4, 6, 6, 4, 0, 3, 2),
#   j = c(3, 4, 1, 2, 5, 6, 5, 2, 0, 1, 6, 3),
#   k = c(0, 7, 2, 3, 6, 7, 1, 1, 5, 5, 7, 6),
#   intensity = seq(0, 1, length = 8),
#   color = seq(0, 1, length = 8),
#   colors = colorRamp(rainbow(8))
# )

# Plot the cube using plotly
plot_ly(
  x = ~x, y = ~y, z = ~z,
  i = ~i, j = ~j, k = ~k,
  type = 'mesh3d',
  opacity = 0.5,  # Semi-transparency
  color = 'blue'
) %>%
  layout(title="Skevent over a Cubic Spatial Volume", 
         scene = list(xaxis = list(range = c(-0.5, 1.5)),
                      yaxis = list(range = c(-0.5, 1.5)),
                      zaxis = list(range = c(-0.5, 1.5))))

### Notes:
# 
# - **Vertices (`x`, `y`, `z`):** The coordinates for the 8 vertices of a cube.
# - **Faces (`i`, `j`, `k`):** Indices defining the 12 triangles that make up the 6 faces of the cube. Each triangle is defined by three vertices.
# - **`mesh3d` type:** Used to create 3D surfaces. The `i`, `j`, `k` arrays specify the triangles that form the cube's faces.
# - **Opacity:** Set to `0.5` to make the cube semi-transparent.
# - **Color:** Set to blue.
  1. Kime-Phase Distribution: Assume a uniform kime-phase distribution \(\Phi(\varphi)\) over the interval \([- \pi, \pi)\)

\[\Phi(\varphi) = \frac{1}{2\pi}, \quad \varphi \in [-\pi, \pi).\]

  1. Spacekime Metric: In this case, the spacekime metric could be represented as:

\[ds^2 = dx^2 + dy^2 + dz^2 + dt^2 + \langle d\varphi^2 \rangle.\]

Since the kime-phase distribution is uniform, the term \(\langle d\varphi^2 \rangle\) would be constant across the region

\[\langle d\varphi^2 \rangle = \int_{-\pi}^{\pi} \frac{1}{2\pi} d\varphi = \frac{\pi^2}{3}.\]

This is since the expected contribution of the kime-phase to the metric, \(\langle d\varphi^2 \rangle\), is calculated by

\[\langle d\varphi^2 \rangle = \int_{-\pi}^{\pi} \varphi^2 \Phi(\varphi) \, d\varphi.\]

Given \(\Phi(\varphi) = \frac{1}{2\pi}\) (uniform distribution), we have

\[\langle d\varphi^2 \rangle = \frac{1}{2\pi} \int_{-\pi}^{\pi} \varphi^2 \, d\varphi.\]

\[\int_{-\pi}^{\pi} \varphi^2 \, d\varphi = \left[ \frac{\varphi^3}{3} \right]_{-\pi}^{\pi} = \frac{\pi^3}{3} - \left(-\frac{\pi^3}{3}\right) = \frac{2\pi^3}{3}.\]

Thus, in this special case, the spacekime metric simplifies to

\[ds^2 = dx^2 + dy^2 + dz^2 + dt^2 + \frac{\pi^2}{3}.\]

  1. Spacekime \(\sigma\)-Algebra: The \(\sigma\)-algebra \(\mathcal{F}\) for the skevent would be the product \(\sigma\)-algebra formed by combining
  • Borel \(\sigma\)-algebra on the cubic region \(V_1\),

  • Borel \(\sigma\)-algebra on the time dimension,

  • Borel \(\sigma\)-algebra on the kime-phase interval \([- \pi, \pi)\).

    An example set in this \(\sigma\)-algebra could be

\[A_1 = \{(x, y, z, t, \varphi) \in V_1 \times [0, 1] \times [-\pi/2, \pi/2]\}.\]

  1. Probability Measure: The probability of the skevent \(S_1\) defined by \(A_1\) is calculated by

\[\text{Pr}(S_1) = \mathbb{P}(A_1) = \int_{V_1} \int_0^1 \int_{-\pi/2}^{\pi/2} f(x, y, z, t, \varphi) \, d\varphi \, dt \, dx \, dy \, dz.\]

If we assume a uniform density function \(f(x, y, z, t, \varphi) = \frac{1}{|V_1| \cdot |T| \cdot |\Phi(\varphi)|}\)

The probability of the skevent \(S_1\) is given by

\[\text{Pr}(S_1) = \mathbb{P}(A_1) = \int_{V_1} \int_0^1 \int_{-\pi/2}^{\pi/2} f(x, y, z, t, \varphi) \, d\varphi \, dt \, dx \, dy \, dz.\]

Given that we have a uniform distribution and assuming a uniform density function \(f(x, y, z, t, \varphi) = \frac{1}{|V_1| \cdot |T| \cdot |\Phi(\varphi)|}\), the probability can be calculated by

\[\text{Pr}(S_1) = \frac{1}{1 \cdot 1 \cdot 2\pi} \times |V_1| \times |T| \times \text{Length of kime-phase interval}.\] \[\text{Pr}(S_1) = \frac{1}{2\pi} \times 1 \times 1 \times \pi = \frac{\pi}{2\pi} = \frac{1}{2}.\]

3.2.6.2 Example: Skevent in a Spherical Spatial Region with Gaussian Kime-Phase Distribution

  1. Spatial Region (Volume in \(\mathbb{R}^3\)): Consider a spherical region in \(\mathbb{R}^3\) defined by

\[V_2 = \{(x, y, z) \in \mathbb{R}^3 \mid x^2 + y^2 + z^2 \leq 1\}.\]

library(plotly)
f <- function(x, y, z){
    x^2 + y^2 + z^2
}
R <- 2 # radius
x <- y <- z <- seq(-R, R, length.out = 100) 
g <- expand.grid(x = x, y = y, z = z)
voxel <- array(with(g, f(x, y, z)), dim = c(100, 100, 100))

library(misc3d)
cont <- computeContour3d(voxel, level = R^2, x = x, y = y, z = z)
idx <- matrix(0:(nrow(cont)-1), ncol=3, byrow=TRUE)

plot_ly(x = cont[, 1], y = cont[, 2], z = cont[, 3],
        i = idx[, 1], j = idx[, 2], k = idx[, 3],
        type = "mesh3d", opacity=0.5) %>% 
  layout(title="Skevent over a Spherical Spatial Volume", 
         scene = list(aspectmode = "data"))
  1. Kime-Phase Distribution: Assume a Gaussian kime-phase distribution centered at \(\varphi = 0\) with variance \(\sigma^2\)

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right).\]

  1. Spacekime Metric: The spacekime metric for this region could be represented as

\[ds^2 = dx^2 + dy^2 + dz^2 + dt^2 + \langle d\varphi^2 \rangle.\]

Since the kime-phase distribution is Gaussian, the expected contribution from the phase would be

\[\langle d\varphi^2 \rangle = \int_{-\infty}^{\infty} \frac{\varphi^2}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right) d\varphi = \sigma^2.\]

Thus, the metric becomes

\[ds^2 = dx^2 + dy^2 + dz^2 + dt^2 + \sigma^2.\]

  1. Spacekime \(\sigma\)-Algebra: A set in the spacekime \(\sigma\)-algebra for this case could be

\[A_2 = \{(x, y, z, t, \varphi) \in V_2 \times [0, 2] \times [-2\sigma, 2\sigma]\}.\]

  1. Probability Measure: The probability of the skevent \(S_2\) defined by \(A_2\) under by these conditions:
  • Spatial Region: A spherical region of radius \(R = 1\).
  • Temporal Interval: The time interval is \(T = 2\).
  • Kime-Phase Distribution: The kime-phase \(\varphi\) follows a Gaussian distribution with mean \(\mu = 0\) and standard deviation \(\sigma\).

Let \(f(x, y, z, t, \varphi)\) be the joint probability density function over the spacekime coordinates. If we assume independence between spatial, temporal, and kime-phase components, the pdf can be factored as

\[f(x, y, z, t, \varphi) = f_{\text{space}}(x, y, z) \cdot f_{\text{time}}(t) \cdot \Phi(\varphi), \]

where

  • \(f_{\text{space}}(x, y, z)\) is uniform over the spherical region.
  • \(f_{\text{time}}(t)\) is uniform over the time interval \([0, 2]\).
  • \(\Phi(\varphi)\) is the Gaussian distribution over the kime-phase.

Calculate the Individual Components:

  1. Spatial Component:
    • The spatial region is a sphere of radius 1, with volume \(V = \frac{4}{3} \pi R^3 = \frac{4}{3} \pi .\)
    • Assuming a uniform distribution over the sphere, the spatial density function is:

\[f_{\text{space}}(x, y, z) = \frac{1}{V} = \frac{3}{4\pi}.\]

  1. Temporal Component:
    • The temporal interval is \([0, 2]\), so the temporal density is:

\[f_{\text{time}}(t) = \frac{1}{T} = \frac{1}{2}.\]

  1. Kime-Phase Component:
    • The kime-phase follows a Gaussian distribution:

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right).\] - The probability of the kime-phase being within the interval \([-2\sigma, 2\sigma]\) is

\[\int_{-2\sigma}^{2\sigma} \Phi(\varphi) \, d\varphi = \text{erf}\left(\frac{2}{\sqrt{2}}\right) = \text{erf}(2).\]

Then, the probability of the skevent \(S_2\) is given by integrating the joint pdf over the spherical spatial region, the time interval, and the kime-phase range

\[\text{Pr}(S_2) = \int_{V} \int_0^2 \int_{-2\sigma}^{2\sigma} f_{\text{space}}(x, y, z) \cdot f_{\text{time}}(t) \cdot \Phi(\varphi) \, d\varphi \, dt \, dx \, dy \, dz, \]

Substituting the density functions

\[\text{Pr}(S_2) = \int_{V} \int_0^2 \int_{-2\sigma}^{2\sigma} \frac{3}{4\pi} \cdot \frac{1}{2} \cdot \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right) \, d\varphi \, dt \, dx \, dy \, dz.\]

The integration over the spatial volume \(V\) gives us the volume itself, \(\frac{4}{3} \pi\), and the temporal integration over \([0, 2]\) gives the length of the interval, \(2\). Therefore,

\[\text{Pr}(S_2) = \frac{3}{4\pi} \cdot \frac{1}{2} \cdot \text{erf}(2) \cdot \left(\frac{4}{3} \pi \right) \cdot 2 = \text{erf}(2)\approx 0.995,\]

and the probability of the skevent \(S_2\) is \(\text{Pr}(S_2) \approx 0.995.\)

3.2.6.3 Example: Skevent in a Cylindrical Spatial Region with Laplace Kime-Phase Distribution and Exponential Decay over Time

1. Spatial Region: Cylindrical region \(V_{\text{cyl}}\) is a cylinder with height \(H\) and base radius \(R\), aligned along the \(z\)-axis.

\[V_{\text{cyl}} = \left\{ (x, y, z) \mid x^2 + y^2 \leq R^2, \, 0 \leq z \leq H \right\}.\]

library(plotly)

# Parameters for the cylinder
height <- 2
radius <- 1
n <- 50  # Number of points for the grid

# Create a grid of angles and heights
theta <- seq(0, 2 * pi, length.out = n)
z <- seq(-height / 2, height / 2, length.out = n)

# Create the x, y, z coordinates for the cylinder surface
x <- outer(radius * cos(theta), rep(1, n))
y <- outer(radius * sin(theta), rep(1, n))
z <- outer(rep(1, n), z)

# Plot the cylinder using plotly
plot_ly(
  x = ~x, y = ~y, z = ~z,
  type = 'surface',
  opacity = 0.5,  # Adjust opacity for semi-transparency
  colorscale = list(c(0, 'blue'), c(1, 'blue')),  # Solid color
  showscale = FALSE
) %>%
  layout( title="Skevent over a Cylintrical Spatial Volume",
    scene = list(
      xaxis = list(range = c(-radius - 0.5, radius + 0.5)),
      yaxis = list(range = c(-radius - 0.5, radius + 0.5)),
      zaxis = list(range = c(-height / 2 - 0.5, height / 2 + 0.5))
    )
  )

2. Spacekime Metric: The spacekime metric includes contributions from spatial dimensions, time, and the kime-phase

\[ds^2 = -dt^2 + dx^2 + dy^2 + dz^2 + \langle d\varphi^2 \rangle.\] - The kime-phase \(\varphi\) follows a Laplace distribution:

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\]

3. Spacekime σ-Algebra: The spacekime σ-algebra \(\mathcal{F}_{\text{spacekime}}\) is defined over the cylindrical spatial region, the time interval \([0, T]\), and the kime-phase interval \([\varphi_1, \varphi_2]\). The specific set \(A_3\) is defined as

\[A_3 = \left\{(x, y, z, t, \varphi) \mid (x, y, z) \in V_{\text{cyl}}, \, 0 \leq t \leq T, \, \varphi_1 \leq \varphi \leq \varphi_2 \right\}.\]

4. Probability Measure

  • Time Distribution (Exponential Decay): We assume an exponential decay distribution over time with parameter \(\lambda\)

\[f_{\text{time}}(t) = \lambda e^{-\lambda t}, \quad t \in [0, T].\]

  • Joint Probability Density Function: The joint probability density function is

\[f(x, y, z, t, \varphi) = f_{\text{space}}(x, y, z) \cdot f_{\text{time}}(t) \cdot \Phi(\varphi),\] where

\[f_{\text{space}}(x, y, z) = \frac{1}{V_{\text{cyl}}} = \frac{1}{\pi R^2 H}\] and \(\Phi(\varphi)\) is the Laplace distribution over the kime-phase.

  • Probability Calculation: The probability of the skevent \(S_3\) is

\[\text{Pr}(S_3) = \int_{V_{\text{cyl}}} \int_0^T \int_{\varphi_1}^{\varphi_2} f_{\text{space}}(x, y, z) \cdot f_{\text{time}}(t) \cdot \Phi(\varphi) \, d\varphi \, dt \, dx \, dy \, dz.\]

Substituting the density functions

\[\text{Pr}(S_3) = \frac{1}{\pi R^2 H} \cdot \int_{V_{\text{cyl}}} \int_0^T \lambda e^{-\lambda t} \, dt \cdot \int_{\varphi_1}^{\varphi_2} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \, d\varphi \, dx \, dy \, dz.\]

  • The integral over time \(\int_0^T \lambda e^{-\lambda t} \, dt = 1 - e^{-\lambda T}.\)

  • Kime-Phase Integral: The integral over the kime-phase interval

\[\int_{\varphi_1}^{\varphi_2} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \, d\varphi .\] This integral can be computed as

\[\int_{\varphi_1}^{\varphi_2} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \, d\varphi = \frac{1}{2} \left[\text{sign}(\varphi - \mu) \cdot \left( e^{-\frac{|\varphi_1 - \mu|}{b}} - e^{-\frac{|\varphi_2 - \mu|}{b}} \right)\right].\]

  • Spatial Integral: The spatial integral gives the volume of the cylindrical region

\[\int_{V_{\text{cyl}}} dx \, dy \, dz = V_{\text{cyl}} = \pi R^2 H.\]

  • Final Probability: Combining these, the probability is

\[\text{Pr}(S_3) = \frac{1}{\pi R^2 H} \cdot \pi R^2 H \cdot \left( 1 - e^{-\lambda T}\right) \cdot \frac{1}{2} \left[\text{sign}(\varphi - \mu) \cdot \left( e^{-\frac{|\varphi_1 - \mu|}{b}} - e^{-\frac{|\varphi_2 - \mu|}{b}} \right)\right].\]

Simplifying the expression

\[\text{Pr}(S_3) = \left(1 - e^{-\lambda T}\right) \cdot \frac{1}{2} \left[\text{sign}(\varphi - \mu) \cdot \left( e^{-\frac{|\varphi_1 - \mu|}{b}} - e^{-\frac{|\varphi_2 - \mu|}{b}} \right)\right].\]

3.3 Entropy

In both classical and quantum physics, entropy is a central concept. The entropy describes the degree of disorder, or the amount of uncertainty, indeterminance in the information about our knowledge of a system. In quantum mechanics, entropy also connects to the probability distribution of quantum states.

Boltzmann entropy: Traditionally, the entropy \(S\) in a system is given by the Boltzmann formula \[S = k_B \ln \Omega \ ,\]

where \(\Omega\) is the number of accessible microstates, and \(k_{\mathrm {B}}=1.380649\times 10^{−23} J/K\) is the Boltzmann constant.

The Boltzmann entropy, \(S = k_B \ln \Omega\), may be extended to account for the kime domain by considering all microstates (or events) represented in the same kime surface. This approach could provide insights into the thermodynamic properties of quantum systems where the kime phase represents quantum coherence or entanglement. This may link complex-time to quantum gravity formulations, where entropy plays a crucial role. In the spacekime framework, we may introduce a generalized entropy that takes into account the distribution of microstates across the kime domain:

\[S_\kappa = k_B \ln \Omega(\kappa)\ ,\]

where \(\Omega(\kappa)\) represents the number of microstates associated with a particular kime surface.

Kime Entropy: In the context of kime representation, entropy may potentially be generalized to account for the additional complexity introduced by the kime variable. This generalized entropy would measure not just the distribution of events over the kime-magnitude (time), but also their distribution across different experimental repeats (kime phases). A generalized entropy for kime surfaces could be formulated as \(S_\kappa = k_B \ln \Omega(\kappa)\), where \(\Omega(\kappa)\) represents the number of microstates corresponding to a particular kimesurface.

A formulation of a generalized spacekime entropy requires extending the concept of entropy beyond traditional settings. Below are some ideas on potential strategies to generalize entropy.

3.4 Spacekime Shannon Entropy

Next we will formulate the Spacekime Shannon Entropy in the context of kevents (complex events) and skevents (spacekime events), we’ll need to incorporate the definitions of the kevents and skevents along with the associated measure and probability. This approach will extend the classical Shannon entropy into the spacekime framework, accounting for the additional kime-phase dimension and the corresponding probability distributions.

The classical Shannon entropy for a discrete (or continous) random variable \(X\) with probability mass function \(p(x)\) (or density function \(f(x)\)) is defined by

\[H(X) = -\sum_{x \in \mathcal{X}} p(x) \log p(x),\] \[H(X) = -\int_{\mathcal{X}} f(x) \log f(x) \, dx.\]

In the 5D spacekime, we’re working with skevents that includes \(3\) spatial dimensions and \(2\) kime dimensions (time \(t\) and kime-phase \(\varphi\)).

Given a spacekime event (skevent) \(S\) in the space \(\Omega = \mathbb{R}^3 \times \mathbb{R} \times [-\pi, \pi)\), the probability measure is

\[\text{Pr}(S) = \mathbb{P}(A) = \int_A f(\mathbf{X}) \, d\mathbf{X},\]

where \(\mathbf{X} = (x^1, x^2, x^3, t, \varphi)\) and \(f(\mathbf{X})\) is the probability density function.

For a skevent \(S\) defined in the 5D spacekime space, the Shannon entropy can be extended as

\[H(S) = -\int_{\Omega} f(\mathbf{X}) \log f(\mathbf{X}) \, d\mathbf{X},\] where, \(\Omega\) is the entire 5D space, and \(f(\mathbf{X})\) is the joint probability density function over the spacekime coordinates.

Expanding the integral explicitly for the 5D spacekime coordinates, the spacekime Shannon Entropy becomes

\[H(S) = -\int_{\mathbb{R}^3} \int_{\mathbb{R}} \int_{-\pi}^{\pi} f(x^1, x^2, x^3, t, \varphi) \log f(x^1, x^2, x^3, t, \varphi) \, d\varphi \, dt \, dx^1 \, dx^2 \, dx^3.\]

Under a density separability assumption, the probability density function can be decomposed as

\[f(\mathbf{X}) = f_x(x^1, x^2, x^3) f_t(t) \Phi(\varphi),\]

where:

  • \(f_x(x^1, x^2, x^3)\) is the spatial probability density function.
  • \(f_t(t)\) is the temporal probability density function.
  • \(\Phi(\varphi)\) is the kime-phase probability density function.

Then, the entropy can be decomposed as

\[H(S) = H_x + H_t + H_\varphi, \] where:

\[H_x = -\int_{\mathbb{R}^3} f_x(x^1, x^2, x^3) \log f_x(x^1, x^2, x^3) \, dx^1 \, dx^2 \, dx^3,\] \[H_t = -\int_{\mathbb{R}} f_t(t) \log f_t(t) \, dt,\] \[H_\varphi = -\int_{-\pi}^{\pi} \Phi(\varphi) \log \Phi(\varphi) \, d\varphi .\]

Let’s consider different kime-phase distributions \(\Phi(\varphi)\) and compute the kime-phase component of the entropy.

1. Uniform Kime-Phase Distribution: For a uniform distribution on \([- \pi, \pi)\) \(\Phi(\varphi) = \frac{1}{2\pi},\) the entropy is

\[H_\varphi = -\int_{-\pi}^{\pi} \frac{1}{2\pi} \log \frac{1}{2\pi} \, d\varphi = \log(2\pi).\]

2. Laplace Kime-Phase Distribution: For a Laplace distribution centered at 0, \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right)\), the entropy is

\[H_\varphi = -\int_{-\infty}^{\infty} \frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right) \log\left(\frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right)\right) d\varphi.\]

This simplifies to \(H_\varphi = 1 + \log(2b).\)

3. Gaussian (Normal) Kime-Phase Distribution: For a normal distribution, \(\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right),\) the entropy is

\[H_\varphi = \frac{1}{2} \log(2\pi e \sigma^2).\]

The entropy in the spacekime framework can be interpreted as the uncertainty or complexity of the entire system, considering contributions from spatial, temporal, and kime-phase dimensions. This reformulation allows us to quantify the information content of skevents in a higher-dimensional spacekime context, providing deeper insights into the behavior of complex systems.

3.5 Complex-Time Tsallis Entropy

A spacekime extension of Tsallis entropy of order (or entropic index) \(q\in\mathbb{R}^+\) (entropic index) generalizes the classical Boltzmann-Gibbs entropy.

  • For a probability distribution \(\{p_i\}\) over a discrete set of events, the Tsallis entropy of order \(q\) (where \(q > 0\)) is defined by

\[S_q = \frac{1}{q-1} \left(1 - \sum_i p_i^q\right).\]

  • For a continuous probability density function \(f(x)\) over a domain \(\mathcal{X}\), the Tsallis entropy is

\[S_q = \frac{1}{q-1} \left(1 - \int_{\mathcal{X}} f(x)^q \, dx\right).\]

Lemma: As \(q \to 1\), the Tsallis entropy reduces to the Shannon entropy.

Proof: Start with the Tsallis entropy formula

\[S_q = \frac{1}{q - 1} \left( 1 - \sum_{i} p_i^q \right)\]

and expand \(p_i^q\) using a Taylor series around \(q = 1\)

For small \(q - 1\), we can approximate \(p_i^q\) as:

\[p_i^q = p_i^{1 + (q - 1)} = p_i \cdot p_i^{q-1} = p_i \cdot \exp((q-1) \log p_i).\]

Using the first-order approximation \(\exp(x) \approx 1 + x\) for small \(x\):

\[p_i^q \approx p_i \cdot \left(1 + (q - 1) \log p_i \right) = p_i + (q - 1) p_i \log p_i.\]

Next, substitute this approximation into the Tsallis entropy

\[S_q \approx \frac{1}{q - 1} \left( 1 - \sum_{i} \left( p_i + (q - 1) p_i \log p_i \right) \right)\]

and simplify the expression. Note that \(\sum_{i} p_i = 1\), since \(p_i\) are probabilities

\[S_q \approx \frac{1}{q - 1} \left( 1 - 1 - (q - 1) \sum_{i} p_i \log p_i \right).\]

\[S_q \approx \frac{1}{q - 1} \left( - (q - 1) \sum_{i} p_i \log p_i \right).\]

The \(q - 1\) terms cancel out \(S_q \approx \sum_{i} p_i \log p_i.\) Apply the limit as \(q \to 1\). As \(q\) approaches 1, the approximation becomes exact, and we have

\[\lim_{q \to 1} S_q = -\sum_{i} p_i \log p_i, \]

which is exactly the Shannon entropy. \(\square\)

In the spacekime context, we extend this definition to account for both spatial, temporal, and kime-phase components. Let \(\mathbf{X} = (x^1, x^2, x^3, t, \varphi)\) represent a point in spacekime, and let \(f(\mathbf{X})\) be the joint probability density function over this 5D space.

The Complex-Time Tsallis Entropy can be defined as

\[S_q(\text{Skevent}) = \frac{1}{q-1} \left(1 - \int_{\Omega} f(\mathbf{X})^q \, d\mathbf{X}\right),\]

where \(\Omega = \mathbb{R}^3 \times \mathbb{R} \times [-\pi, \pi)\) is the domain of the spacekime events.

When the probability density function is separable and can be decomposed into independent components for space, time, and kime-phase

\[f(\mathbf{X}) = f_x(x^1, x^2, x^3) f_t(t) \Phi(\varphi),\] then the Tsallis entropy can be decomposed accordingly

\[S_q(\text{Skevent}) = \frac{1}{q-1} \left(1 - \int_{\mathbb{R}^3} \int_{\mathbb{R}} \int_{-\pi}^{\pi} f_x(x^1, x^2, x^3)^q f_t(t)^q \Phi(\varphi)^q \, d\varphi \, dt \, dx^1 \, dx^2 \, dx^3\right).\]

This simplifies to

\[S_q(\text{Skevent}) = S_q^{(x)} + S_q^{(t)} + S_q^{(\varphi)},\]

where

\[S_q^{(x)} = \frac{1}{q-1} \left(1 - \int_{\mathbb{R}^3} f_x(x^1, x^2, x^3)^q \, dx^1 \, dx^2 \, dx^3\right).\]

\[S_q^{(t)} = \frac{1}{q-1} \left(1 - \int_{\mathbb{R}} f_t(t)^q \, dt\right).\]

\[S_q^{(\varphi)} = \frac{1}{q-1} \left(1 - \int_{-\pi}^{\pi} \Phi(\varphi)^q \, d\varphi\right).\]

3.5.1 Examples of Complex-Time Tsallis Entropy

Let’s work through some examples using different kime-phase distributions.

Example 1: Uniform Kime-Phase Distribution

For a uniform kime-phase distribution

\[\Phi(\varphi) = \frac{1}{2\pi}, \quad \varphi \in [-\pi, \pi),\] the Tsallis entropy contribution from the kime-phase component is

\[S_q^{(\varphi)} = \frac{1}{q-1} \left(1 - \int_{-\pi}^{\pi} \left(\frac{1}{2\pi}\right)^q \, d\varphi\right).\]

\[S_q^{(\varphi)} = \frac{1}{q-1} \left(1 - \frac{(2\pi)}{(2\pi)^q}\right) = \frac{1}{q-1} \left(1 - \frac{1}{(2\pi)^{q-1}}\right).\]

Example 2: Laplace Kime-Phase Distribution

For a Laplace distribution centered at 0 with scale parameter \(b > 0\)

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right),\]

the Tsallis entropy contribution is

\[S_q^{(\varphi)} = \frac{1}{q-1} \left(1 - \int_{-\infty}^{\infty} \left( \frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right)\right)^q \, d\varphi\right).\]

This integral can be calculated as follows

\[\int_{-\infty}^{\infty} \left(\frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right)\right)^q \, d\varphi = \frac{1}{(2b)^q} \int_{-\infty}^{\infty} \exp\left(-\frac{q|\varphi|}{b}\right) d\varphi = \\ \frac{2}{(2b)^q} \cdot \frac{b}{q} = \frac{1}{b^{q-1} q}.\]

Thus, the Tsallis entropy becomes

\[S_q^{(\varphi)} = \frac{1}{q-1} \left(1 - \frac{1}{b^{q-1} q}\right).\]

Example 3: Gaussian (Normal) Kime-Phase Distribution

For a normal distribution

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right),\] the Tsallis entropy contribution is

\[S_q^{(\varphi)} = \frac{1}{q-1} \left(1 - \int_{-\infty}^{\infty} \left(\frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right)\right)^q \, d\varphi\right).\]

This integral can be computed, though the result is more complex and involves special functions

\[S_q^{(\varphi)} = \frac{1}{q-1} \left(1 - \left(\frac{1}{\sqrt{q}} \right)\right),\]

where the result for a Gaussian distribution can involve terms related to the error function and gamma functions.

The Complex-Time Tsallis Entropy in the spacekime context provides a generalized measure of uncertainty or complexity that accounts for the non-linear interactions between the spatial, temporal, and kime-phase dimensions. It can be particularly useful in systems where traditional Shannon entropy does not adequately capture the dynamics, such as in systems with long-range correlations or where the kime-phase introduces significant variability.

3.6 Entropy in Kime-Space Configuration

To extend the entropy in the spacekime configuration using skevents (spacekime events) requires integration of the entropy with the extended structure of spacekime. The goal is to create a coherent formulation that captures the complexity and variability inherent in the 5D spacekime space, which includes 3 spatial dimensions, 1 time dimension, and the kime-phase dimension.

Entropy, in the context of spacekime, is a measure of uncertainty or disorder that accounts for the contributions from both the traditional spatial and temporal dimensions, as well as the kime-phase distribution. When we consider skevents, these events are not just points in a higher-dimensional space, but rather regions that encapsulate complex interactions between space, time, and phase.

Probability Measure for Skevents: Given a skevent \(S\) defined over the domain \(\Omega = \mathbb{R}^3 \times \mathbb{R} \times [-\pi, \pi)\), the probability measure is

\[\text{Pr}(S) = \int_S f(\mathbf{X}) \, d\mathbf{X},\]

where \(\mathbf{X} = (x^1, x^2, x^3, t, \varphi)\) and \(f(\mathbf{X})\) is the joint probability density function over the spacekime coordinates.

The entropy in this space can be generalized by integrating the probability distribution over the entire skevent region. For a probability density function \(f(\mathbf{X})\), the entropy can be defined by

\[H(S) = -\int_{S} f(\mathbf{X}) \log f(\mathbf{X}) \, d\mathbf{X},\]

where \(S\) is a skevent in the 5D space.

Again, assume that the joint probability density function can be decomposed into spatial, temporal, and kime-phase components

\[f(\mathbf{X}) = f_x(x^1, x^2, x^3) f_t(t) \Phi(\varphi).\]

Then, the entropy can then be decomposed accordingly

\[H(S) = H_x + H_t + H_\varphi,\] where

\[H_x = -\int_{V} f_x(x^1, x^2, x^3) \log f_x(x^1, x^2, x^3) \, dx^1 \, dx^2 \, dx^3,\] \[H_t = -\int_{T} f_t(t) \log f_t(t) \, dt,\]

\[H_\varphi = -\int_{\Phi} \Phi(\varphi) \log \Phi(\varphi) \, d\varphi .\]

Let’s demonstrate the calculations of the entropy for specific skevents using different kime-phase distributions.

Example 1: Uniform Kime-Phase Distribution: For a uniform distribution over the kime-phase interval \([- \pi, \pi)\), \(\Phi(\varphi) = \frac{1}{2\pi}, \quad \varphi \in [-\pi, \pi)\), and the kime-phase entropy contribution is

\[H_\varphi = -\int_{-\pi}^{\pi} \frac{1}{2\pi} \log \frac{1}{2\pi} \, d\varphi = \log(2\pi).\] This result shows that the (kime-phase) entropy is constant and depends only on the width of the interval over which the uniform distribution is defined.

Example 2: Laplace Kime-Phase Distribution: For a Laplace distribution centered at \(\varphi = 0\) with scale parameter \(b > 0\), \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right),\) the kime-phase entropy is:

\[H_\varphi = -\int_{-\infty}^{\infty} \frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right) \log \left(\frac{1}{2b} \exp\left(-\frac{|\varphi|}{b}\right)\right) d\varphi .\]

This simplifies to \(H_\varphi = 1 + \log(2b),\) which shows that the entropy increases logarithmically with the scale parameter \(b\).

Example 3: Gaussian (Normal) Kime-Phase Distribution: For a normal distribution with mean \(0\) and variance \(\sigma^2\), \(\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right)\), the kime-phase entropy is

\[H_\varphi = -\int_{-\infty}^{\infty} \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right) \log\left(\frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{\varphi^2}{2\sigma^2}\right)\right) d\varphi .\]

This simplifies to \(H_\varphi = \frac{1}{2} \log(2\pi e \sigma^2)\), which shows that the entropy is related to the variance \(\sigma^2\) and scales logarithmically with it.

3.7 Quantum Spacekime Entropy

……….THIS SECTION NEEDS WORK……………………………

3.7.1 Discrete Quantum Formulation

Let’s formulate the concept of Quantum Spacekime Entropy and define the density matrix \(\rho_{\kappa}\) over spacekime. In quantum mechanics, the entropy of a quantum system is often described using the von Neumann entropy, defined as \(S(\rho) = -\text{Tr}(\rho \log \rho)\), where \(\rho\) is the density matrix of the system encoding the probabilities of the system being in various quantum states.

The density matrix \(\rho\) for a mixed state is given by

\[\rho = \sum_i p_i |\psi_i\rangle \langle \psi_i|,\] where \(p_i\) are the probabilities associated with the pure states \(|\psi_i\rangle\).

In spacekime, the density matrix \(\rho_{\kappa}\) must account for not only the spatial and temporal components but also the kime-phase distribution. The complex nature of kime implies that the density matrix needs to encapsulate the variability introduced by both the kime magnitude \(t\) and kime phase \(\varphi\).

The summation form of the quantum spacekime density matrix \(\rho_{\kappa}\) can be written as

\[\rho_{\kappa} = \sum_{X, T, \Phi} p(X, T, \varphi) |\psi(X, T, \varphi)\rangle \langle \psi(X, T, \varphi)|,\]

where

  • \(X\) represents the set of spatial coordinates \((x^1, x^2, x^3)\),
  • \(T\) represents the time coordinate \(t\),
  • \(\Phi\) represents the kime-phase \(\varphi\),
  • \(p(X, T, \varphi)\) is the state probability associated with the state \(|\psi(X, T, \varphi)\rangle\),
  • \(|\psi(X, T, \varphi)\rangle\) is the quantum state vector depending on spatial, temporal, and kime-phase components.

To explicitly incorporate the kime-phase distribution \(\Phi(\varphi)\)

\[\rho_{\kappa} = \sum_{X, T, \Phi} \Phi(\varphi) p(X, T, \varphi) |\psi(X, T, \varphi)\rangle \langle \psi(X, T, \varphi)|.\]

This expression compacts the full summation over the spatial coordinates \(X = (x^1, x^2, x^3)\), time \(T = t\), and kime-phase \(\Phi = \varphi\) into a single, concise formula.

If we treat \(X, T,\) and \(\varphi\) as components of a generalized spacekime coordinate vector \(\kappa = (X, T, \varphi)\), we can also write

\[\rho_{\kappa} = \sum_{\kappa} \Phi(\varphi) p(\kappa) |\psi(\kappa)\rangle \langle \psi(\kappa)|,\]

where

  • \(\kappa = (X, T, \varphi)\) is the generalized spacekime coordinate,
  • \(p(\kappa)\) is the probability distribution over spacekime,
  • \(|\psi(\kappa)\rangle\) is the state vector at coordinate \(\kappa\).

3.7.2 Continuous Quantum Formulation

Using an integral form, we can generalize the density matrix \(\rho_{\kappa}\) over spacekime as

\[\rho_{\kappa} = \int_{\Omega} \rho(x^1, x^2, x^3, t, \varphi) |\psi(x^1, x^2, x^3, t, \varphi)\rangle \langle \psi(x^1, x^2, x^3, t, \varphi)| \, d\mathbf{X},\] where \(\Omega = \mathbb{R}^3 \times \mathbb{R}^+ \times [-\pi, \pi)\), and \(\rho(x^1, x^2, x^3, t, \varphi)\) is a probability density function defined over spacekime.

The Quantum Spacekime Entropy can then be defined analogously to the von Neumann entropy

\[S(\rho_{\kappa}) = -\text{Tr}(\rho_{\kappa} \log \rho_{\kappa}).\]

Given the decomposition of the density matrix over spacekime, the trace operation and entropy computation become more complex. The trace is taken over the entire spacekime configuration space, integrating over both the spatial and kime components.

If \(\rho_{\kappa}\) can be decomposed similarly to classical density matrices but with a spacekime probability component

\[\rho_{\kappa} = \int_{\mathbb{R}^3} \int_{\mathbb{R}} \int_{-\pi}^{\pi} \Phi(\varphi) \rho_x(x^1, x^2, x^3) \rho_t(t) |\psi(x^1, x^2, x^3, t, \varphi) \rangle \langle \psi(x^1, x^2, x^3, t, \varphi)| \, d\varphi \, dt \, dx^1 \, dx^2 \, dx^3,\]

then the entropy can be expressed as

\[S(\rho_{\kappa}) = -\text{Tr}\left(\int_{\mathbb{R}^3} \int_{\mathbb{R}} \int_{-\pi}^{\pi} \Phi(\varphi) \rho_x(x^1, x^2, x^3) \rho_t(t) \log\left(\int_{\mathbb{R}^3} \int_{\mathbb{R}} \int_{-\pi}^{\pi} \Phi(\varphi) \rho_x(x^1, x^2, x^3) \rho_t(t) \right) \, d\varphi \, dt \, dx^1 \, dx^2 \, dx^3\right).\]

In spacekime, the non-commutative nature of the kime-phase \(\varphi\) and the magnitude \(t\) complicates the calculation of entropy. This is particularly challenging when attempting to integrate the phase component into a density matrix that traditionally operates over real-valued time and space components.

In spacekime, there may be entanglement between the spatial, temporal, and kime-phase components. This means that the phase cannot simply be treated as an independent dimension but must be considered in conjunction with the others, adding to the complexity of defining and computing the density matrix and entropy.

3.7.3 Examples of Quantum Spacekime Entropy Calculations

Let’s work through some examples using certain simplifications.

3.7.3.1 Example 1: Quantum Spacekime Entropy (Uniform Phase and Exponential Time)

Let’s calculate the Quantum Spacekime Entropy for a uniform kime-phase distribution, an exponential decay distribution over time, and a Euclidean measure on 3D space.

The compact form of the spacekime density matrix

\[\rho_{\kappa} = \sum_{\kappa} \Phi(\varphi) p(\kappa) |\psi(\kappa)\rangle \langle \psi(\kappa)|.\]

where

  • \(\kappa = (X, T, \varphi)\) represents the spacekime coordinates with \(X = (x^1, x^2, x^3)\) for space, \(T = t\) for time, and \(\varphi\) for kime-phase.

  • \(\Phi(\varphi)\) is the kime-phase distribution.

  • \(p(\kappa)\) is the joint probability distribution over space, time, and kime-phase.

  • Kime-Phase Distribution (\(\Phi(\varphi)\)): Uniform distribution over \([-\pi, \pi]\)

\[\Phi(\varphi) = \frac{1}{2\pi}, \quad \text{for } \varphi \in [-\pi, \pi].\]

  • Time Distribution (\(p_T(t)\)): Exponential decay distribution with parameter \(\lambda > 0\):

\[p_T(t) = \lambda e^{-\lambda t}, \quad t \geq 0.\]

  • Spatial Distribution (\(p_X(X)\)): Euclidean measure over 3D space (assuming uniform over some volume \(V\))

\[p_X(X) = \frac{1}{V}, \quad \text{for } X \in V.\]

Assuming independence between the spatial, temporal, and kime-phase components

\[p(\kappa) = p_X(X) p_T(t) \Phi(\varphi),\]

\[p(\kappa) = \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2\pi}.\]

The quantum entropy is defined as \(S_{\kappa} = -\text{Tr}(\rho_{\kappa} \log \rho_{\kappa}).\) Given the density matrix

\[\rho_{\kappa} = \sum_{\kappa} \Phi(\varphi) p(\kappa) |\psi(\kappa)\rangle \langle \psi(\kappa)|, \]

The trace of \(\rho_{\kappa}\) is

\[\text{Tr}(\rho_{\kappa}) = \sum_{\kappa} \langle \psi(\kappa) | \rho_{\kappa} | \psi(\kappa)\rangle .\]

Subject to the normalization condition for probabilities, \(\text{Tr}(\rho_{\kappa}) = 1\).

It’s easier to find \(\log \rho_{\kappa}\), use the eigenvalues of \(\rho_{\kappa}\), which are the probabilities \(p(\kappa)\) weighted by the kime-phase distribution \(\Phi(\varphi)\),

\[\log \rho_{\kappa} = \log \left(\frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2\pi} \right).\]

\[\log \rho_{\kappa} = \log\left(\frac{\lambda}{2\pi V}\right) - \lambda t .\]

Substitute into the entropy formula \(S_{\kappa} = -\sum_{\kappa} p(\kappa) \log p(\kappa)\)

\[S_{\kappa} = -\sum_{X} \sum_{T} \sum_{\varphi} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2\pi} \left[\log\left(\frac{\lambda}{2\pi V}\right) - \lambda t\right] .\]

This can be simplified to

\[S_{\kappa} = \frac{1}{V} \frac{1}{2\pi} \sum_{X} \sum_{T} \sum_{\varphi} \left[ -\lambda e^{-\lambda t} \left( \log\left(\frac{\lambda}{2\pi V}\right) - \lambda t \right)\right].\]

To find an analytical result, approximate the sums over continuous variables \(t\) and \(\varphi\) by integrals

\[S_{\kappa} = \frac{1}{V} \frac{1}{2\pi} \int_V \int_{0}^{\infty} \int_{-\pi}^{\pi} \left[ -\lambda e^{-\lambda t} \left( \log\left( \frac{\lambda}{2\pi V}\right) - \lambda t \right)\right] d\varphi dt dV .\]

Since the spatial integration over volume \(V\) is trivial and \(\varphi\) is uniform

\[S_{\kappa} = \frac{V}{V} \frac{1}{2\pi} \cdot 2\pi \int_0^{\infty} \left[ -\lambda e^{-\lambda t} \left( \log\left(\frac{\lambda}{2\pi V}\right) - \lambda t \right)\right] dt .\]

\[S_{\kappa} = \int_0^{\infty} \left[ -\lambda e^{-\lambda t} \left( \log\left( \frac{\lambda}{2\pi V}\right) - \lambda t \right)\right] dt .\]

  1. The term involving \(\log\left(\frac{\lambda}{2\pi V}\right)\) integrates to

\[-\log\left(\frac{\lambda}{2\pi V}\right) \cdot \int_0^{\infty} \lambda e^{-\lambda t} dt = -\log\left(\frac{\lambda}{2\pi V}\right).\]

  1. The term involving \(t\) gives

\[\int_0^{\infty} \lambda^2 t e^{-\lambda t} dt = \frac{1}{\lambda}.\]

Therefore, the entropy in the spacekime configuration, taking into account the distributions over time, space, and kime-phase becomes

\[S_{\kappa} = -\log\left(\frac{\lambda}{2\pi V}\right) + 1.\]

3.7.3.2 Examples 2: Quantum Spacekime Entropy (Gaussian Phase distribution and Exponential Time)

In this example, we showcase the calculation of the quantum spacekime entropy assuming a Gaussian kime-phase distribution, an exponential decay distribution over time, and a Euclidean measure on 3D space.

Again, the spacekime density matrix in a discrete summation form is

\[\rho_{\kappa} = \sum_{\kappa} \Phi(\varphi) p(\kappa) |\psi(\kappa)\rangle \langle \psi(\kappa)|,\]

  • Kime-Phase Distribution (\(\Phi(\varphi)\)): Gaussian distribution centered at \(\mu\) with variance \(\sigma^2\)

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right).\]

  • Time Distribution (\(p_T(t)\)): Exponential decay distribution with parameter \(\lambda\)

\[p_T(t) = \lambda e^{-\lambda t}, \quad t \geq 0.\]

  • Spatial Distribution (\(p_X(X)\)): Uniform distribution over a volume \(V\) in 3D Euclidean space

\[p_X(X) = \frac{1}{V}, \quad \text{for } X \in V.\]

The joint spacekime probability distribution is \(p(\kappa) = p_X(X) p_T(t) \Phi(\varphi).\)

\[p(\kappa) = \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right).\]

The quantum entropy is \(S_{\kappa} = -\text{Tr}(\rho_{\kappa} \log \rho_{\kappa})\) and the trace of \(\rho_{\kappa}\) is

\[\text{Tr}(\rho_{\kappa}) = \sum_{\kappa} \langle \psi(\kappa)| \rho_{\kappa} | \psi(\kappa)\rangle.\]

Assume the normalization condition for probabilities, \(\text{Tr}(\rho_{\kappa}) = 1\).

To find \(\log \rho_{\kappa}\), use the eigenvalues of \(\rho_{\kappa}\), which are the probabilities \(p(\kappa)\) weighted by the kime-phase distribution \(\Phi(\varphi)\)

\[\log \rho_{\kappa} = \log \left(\frac{1}{V} \lambda e^{-\lambda t} \frac{1} {\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right)\right).\]

Simplifying

\[\log \rho_{\kappa} = \log\left(\frac{\lambda}{V \sqrt{2\pi\sigma^2}}\right) - \lambda t - \frac{(\varphi - \mu)^2}{2\sigma^2}.\]

Substitute into the entropy formula

\[S_{\kappa} = -\sum_{X} \sum_{T} \sum_{\varphi} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right) \log\left(\frac{\lambda}{V \sqrt{2\pi\sigma^2}}\right) - \lambda t - \frac{(\varphi - \mu)^2}{2\sigma^2}.\]

To find an analytical result, approximate the sums over continuous variables \(t\) and \(\varphi\) by integrals

\[S_{\kappa} = \frac{1}{V} \int_V \int_{0}^{\infty} \int_{-\infty}^{\infty} \frac{\lambda e^{-\lambda t}}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right) \left[\log\left(\frac{\lambda}{V \sqrt{2\pi\sigma^2}}\right) - \lambda t - \frac{(\varphi - \mu)^2}{2\sigma^2}\right] d\varphi dt dV.\]

We can break this integral into parts

  1. Spatial Integral: \(\int_V \frac{1}{V} dV = 1.\)

  2. Temporal Integral (Exponential decay integral) \(\int_0^{\infty} \lambda e^{-\lambda t} \, dt = 1\), and \(\int_0^{\infty} \lambda t e^{-\lambda t} \, dt = \frac{1}{\lambda}.\)

  3. Kime-Phase Integral (Gaussian integral)

\[\int_{-\infty}^{\infty} \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right) d\varphi = 1\] \[\int_{-\infty}^{\infty} \frac{(\varphi - \mu)^2}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right) d\varphi = \sigma^2.\]

Putting it all together, the entropy becomes

\[S_{\kappa} = -\log\left(\frac{\lambda}{V \sqrt{2\pi\sigma^2}}\right) + \frac{1}{\lambda} + \frac{1}{2}.\]

3.7.3.3 Examples 3: Quantum Spacekime Entropy (Laplace Phase distribution and Exponential Time)

Let’s compute the quantum spacekime entropy using a Laplace kime-phase distribution, an exponential decay distribution over time, and a Euclidean measure on 3D space.

  • Kime-Phase Distribution (\(\Phi(\varphi)\)): Laplace distribution centered at \(\mu\) with scale parameter \(b\)

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\]

  • Time Distribution (\(p_T(t)\)): Exponential decay distribution with parameter \(\lambda\)

\[p_T(t) = \lambda e^{-\lambda t}, \quad t \geq 0.\]

  • Spatial Distribution (\(p_X(X)\)): Uniform distribution over a volume \(V\) in 3D Euclidean space

\[p_X(X) = \frac{1}{V}, \quad \text{for } X \in V.\]

The joint probability distribution is \(p(\kappa) = p_X(X) p_T(t) \Phi(\varphi),\) so

\[p(\kappa) = \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\]

Then, the quantum entropy is \(S_{\kappa} = -\text{Tr}(\rho_{\kappa} \log \rho_{\kappa}).\)

And the trace of \(\rho_{\kappa}\) is

\[\text{Tr}(\rho_{\kappa}) = \sum_{\kappa} \langle \psi(\kappa)| \rho_{\kappa} | \psi(\kappa)\rangle .\]

Given the normalization condition for probabilities, \(\text{Tr}(\rho_{\kappa}) = 1\).

To find \(\log \rho_{\kappa}\), use the eigenvalues of \(\rho_{\kappa}\), which are the probabilities \(p(\kappa)\) weighted by the kime-phase distribution \(\Phi(\varphi)\)

\[\log \rho_{\kappa} = \log \left(\frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right)\right).\]

\[\log \rho_{\kappa} = \log\left(\frac{\lambda}{2bV}\right) - \lambda t - \frac{|\varphi - \mu|}{b}.\]

Substitute into the entropy formula we obtain

\[S_{\kappa} = -\sum_{X} \sum_{T} \sum_{\varphi} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \log\left(\frac{\lambda}{2bV} \right) - \lambda t - \frac{|\varphi - \mu|}{b}.\]

To find an analytical result, approximate the sums over continuous variables \(t\) and \(\varphi\) by integrals

\[S_{\kappa} = \frac{1}{V} \int_V \int_{0}^{\infty} \int_{-\infty}^{\infty} \frac{\lambda e^{-\lambda t}}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \left[\log\left(\frac{\lambda}{2bV}\right) - \lambda t - \frac{|\varphi - \mu|}{b}\right] d\varphi dt dV.\]

Again, we break this entropy integral into parts

  1. Spatial Integral: \(\int_V \frac{1}{V} dV = 1.\)

  2. Temporal Integral (Exponential decay integral) \(\int_0^{\infty} \lambda e^{-\lambda t} \, dt = 1\) and \(\int_0^{\infty} \lambda t e^{-\lambda t} \, dt = \frac{1}{\lambda}.\)

  3. Kime-Phase Integral (Laplace integral) \(\int_{-\infty}^{\infty} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) d\varphi = 1\) and \(\int_{-\infty}^{\infty} \frac{|\varphi - \mu|}{2b^2} \exp\left(-\frac{|\varphi - \mu|}{b}\right) d\varphi = 1.\)

Hence,

\[S_{\kappa} = -\log\left(\frac{\lambda}{2bV}\right) + \frac{1}{\lambda} + 1.\]

3.8 Alternative Spacekime Density Matrix Forms

Alternative density matrix forms can be explored to understand different aspects of quantum systems in spacekime. Each form exposes different properties and behaviors of the system, particularly when considering how space, time, and kime-phase interact.

3.8.1 Diagonal Density Matrix

A diagonal density matrix is one where all off-diagonal elements are zero, meaning the system is described by a mixture of states with no coherence between them.

\[\rho_{\kappa} = \sum_{\kappa} p(\kappa) |\psi(\kappa)\rangle \langle \psi(\kappa)|.\]

  • This form represents a classical mixture of states, where the system is in one of the basis states \(|\psi(\kappa)\rangle\) with probability \(p(\kappa)\).
  • There is no quantum interference or coherence between different states.
  • Useful for studying decoherence and classical limits of quantum systems.
  • In spacekime, this might correspond to systems where the kime-phase has no coherent interaction across different states, leading to purely probabilistic descriptions.

3.8.2 Gaussian-Weighted Density Matrix

Gaussian weights to the contributions from different kime-phases reflects a preference for certain phase values over others.

\[\rho_{\kappa} = \int_{\mathbb{R}^3} \int_{\mathbb{R}} \int_{-\pi}^{\pi} \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right) p(\kappa) |\psi(\kappa)\rangle \langle \psi(\kappa)| \, d\varphi \, dt \, dX.\]

  • This form emphasizes the contribution of states with kime-phases near a central value \(\mu\), with a spread determined by \(\sigma^2\).
  • Useful for studying systems where certain kime-phases are more likely or energetically favorable.
  • This could model systems where fluctuations around a mean phase dominate the dynamics.
  • The Gaussian-weighted density matrix could be used to model systems where the kime-phase distribution is not uniform but instead centered around a particular phase, representing a bias in the temporal evolution.

3.8.3 Phase-Shifted Density Matrix

A phase-shifted density matrix introduces a systematic phase shift across the states, modifying the relative phase of the quantum states

\[\rho_{\kappa} = \sum_{\kappa} p(\kappa) e^{i\theta(\kappa)} |\psi(\kappa)\rangle \langle \psi(\kappa)| e^{-i\theta(\kappa)},\]

where, \(\theta(\kappa)\) represents a phase shift that depends on the spacekime coordinates.

  • This form captures the effects of a global or local phase transformation on the quantum state.
  • Such a phase shift could arise from external fields, interactions, or boundary conditions that introduce a phase dependence in the system.
  • Important for studying gauge invariance and phase coherence.
  • In spacekime, phase-shifted density matrices could model systems where the kime-phase evolves in a controlled manner, leading to shifts in the temporal coherence.
  • This could be relevant in scenarios where the time evolution is non-uniform or where there are interactions that depend on the phase of the system.

3.8.4 Thermal (Gibbs) Density Matrix

A thermal density matrix describes a system in thermal equilibrium at a temperature \(T\), typically given by the Gibbs distribution

\[\rho_{\kappa} = \frac{e^{-\beta H_{\kappa}}}{Z}, \quad Z = \text{Tr}(e^{-\beta H_{\kappa}}),\]

where \(\beta = \frac{1}{k_B T}\) and \(H_{\kappa}\) is the Hamiltonian of the system.

  • This form is used to describe systems in thermal equilibrium, where the probability of being in a state is determined by the Boltzmann factor.
  • Important for studying thermodynamic properties and the effects of temperature on quantum systems.
  • A thermal density matrix in spacekime could explore how temperature and thermal fluctuations interact with the complex-time framework.
  • It could be particularly interesting to study how thermal effects influence the evolution of kime-phase distributions.

3.8.5 Entangled Spacekime Density Matrix

In this case, the density matrix represents an entangled state over different spacekime coordinates

\[\rho_{\kappa} = \sum_{\kappa, \kappa'} p(\kappa, \kappa') |\psi(\kappa)\rangle \langle \psi(\kappa')|.\]

  • Represents a situation where different parts of the system are entangled, meaning that the state of one part depends on the state of another, even across different spacekime coordinates.
  • Useful for exploring quantum correlations in spacekime.
  • This form is ideal for studying non-local correlations and quantum entanglement effects in the context of spacekime.
  • It could also be used to analyze how entanglement evolves when considering the additional complexity of kime-phase dynamics.

Let’s demonstrate several examples.

3.8.6 Examples

3.8.6.1 Examples 1: Quantum Spacekime Entropy using a Laplace Phase Distribution and a diagonal density matrix

For a diagonal density matrix, we assume that there are no off-diagonal elements. This means the system is described by a mixture of independent states. The density matrix is given by \(\rho_{\kappa} = \sum_{\kappa} p(\kappa) |\psi(\kappa)\rangle \langle \psi(\kappa)|.\)

Given the space-kime distributions, the density matrix becomes:

\[\rho_{\kappa} = \sum_{\kappa} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) |\psi(\kappa)\rangle \langle \psi(\kappa)|,\]

where, \(\kappa = (X, T, \varphi)\) represents the spacekime coordinates.

  • Kime-Phase Distribution (\(\Phi(\varphi)\)): Laplace distribution centered at \(\mu\) with scale parameter \(b\):

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\]

  • Time Distribution (\(p_T(t)\)): Exponential decay distribution with parameter \(\lambda\)

\[p_T(t) = \lambda e^{-\lambda t}, \quad t \geq 0.\]

  • Spatial Distribution (\(p_X(X)\)): Uniform distribution over a volume \(V\) in 3D Euclidean space

\[p_X(X) = \frac{1}{V}, \quad \text{for } X \in V.\]

Given the independence of the distributions \(p(\kappa) = p_X(X) p_T(t) \Phi(\varphi).\) and

\[p(\kappa) = \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\]

The quantum entropy for a diagonal density matrix is defined as \(S_{\kappa} = -\text{Tr}(\rho_{\kappa} \log \rho_{\kappa}).\) Given that \(\rho_{\kappa}\) is diagonal, the entropy simplifies to \(S_{\kappa} = -\sum_{\kappa} p(\kappa) \log p(\kappa).\)

Substitute the expression for \(p(\kappa)\)

\[S_{\kappa} = -\sum_{X} \sum_{T} \sum_{\varphi} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \log\left(\frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right)\right).\]

To find an analytical result, approximate the sums over continuous variables \(t\) and \(\varphi\) by integrals

\[S_{\kappa} = -\int_V \int_{0}^{\infty} \int_{-\infty}^{\infty} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \log\left( \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b} \right)\right) d\varphi dt dV.\]

The logarithm can be separated into three terms:

\[\log\left(\frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{ |\varphi - \mu|}{b}\right)\right) = \log\left(\frac{1}{V}\right) + \log\left( \lambda\right) - \lambda t + \log\left(\frac{1}{2b}\right) - \frac{|\varphi - \mu|}{b}.\]

Substitute this back into the entropy expression

\[S_{\kappa} = -\int_V \int_{0}^{\infty} \int_{-\infty}^{\infty} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \left[\log\left(\frac{1}{V}\right) + \log\left(\lambda\right) - \lambda t + \log\left(\frac{1}{2b}\right) - \frac{|\varphi - \mu|}{b}\right] d\varphi dt dV.\]

  1. Spatial Integral \(\int_V \frac{1}{V} dV = 1.\)

  2. Temporal Integral (Exponential decay integral) \(\int_0^{\infty} \lambda e^{-\lambda t} \, dt = 1\) and \(\int_0^{\infty} \lambda t e^{-\lambda t} \, dt = \frac{1}{\lambda}.\)

  3. Kime-Phase Integral (Laplace integral)

\[\int_{-\infty}^{\infty} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) d\varphi = 1,\] and

\[\int_{-\infty}^{\infty} \frac{|\varphi - \mu|}{2b^2} \exp\left( -\frac{|\varphi - \mu|}{b}\right) d\varphi = 1.\]

Substituting the results of the integrals we get

\[S_{\kappa} = -\left[\log\left(\frac{1}{V}\right) + \log\left(\lambda\right) + \log\left(\frac{1}{2b}\right) + 1 + \frac{1}{\lambda} \right].\]

\[S_{\kappa} = \log(V) - \log(\lambda) - \log(2b) - 1 - \frac{1}{\lambda}.\]

3.8.6.2 Examples 2: Quantum Spacekime Entropy using a Laplace Phase Distribution and a Gibbs Density Matrix

As in the prior examples, now we calculate the quantum spacekime entropy using a thermal (Gibbs) density matrix. This incorporate the Gibbs distribution, which is used to describe systems in thermal equilibrium. The Gibbs density matrix incorporates the system’s temperature and is typically used in statistical mechanics.

In a thermal state, the density matrix is given by the Gibbs distribution \(\rho_{\kappa} = \frac{e^{-\beta H_{\kappa}}}{Z},\) where

  • \(\beta = \frac{1}{k_B T}\) is the inverse temperature, with \(k_B\) being the Boltzmann constant and \(T\) the temperature.
  • \(H_{\kappa}\) is the Hamiltonian of the system.
  • \(Z = \text{Tr}(e^{-\beta H_{\kappa}})\) is the partition function, ensuring the normalization of the density matrix.

For simplicity, assume the Hamiltonian \(H_{\kappa}\) in spacekime is separable into spatial, temporal, and kime-phase components \(H_{\kappa} = H_X(X) + H_T(t) + H_{\varphi}(\varphi),\)

The corresponding density matrix in spacekime is \(\rho_{\kappa} = \frac{e^{-\beta (H_X(X) + H_T(t) + H_{\varphi}(\varphi))}}{Z}.\)

For the thermal state in spacekime, assume

  • Kime-Phase Distribution (\(\Phi(\varphi)\)): Laplace distribution centered at \(\mu\) with scale parameter \(b\) \(H_{\varphi}(\varphi) = \frac{|\varphi - \mu|}{b}\) and

\[\Phi(\varphi) = \frac{e^{-\beta H_{\varphi}(\varphi)}}{Z_\varphi} = \frac{e^{-\beta \frac{|\varphi - \mu|}{b}}}{Z_\varphi}.\]

  • Time Distribution (\(p_T(t)\)): Exponential decay distribution \(H_T(t) = \lambda t\) and \(p_T(t) = \frac{e^{-\beta H_T(t)}}{Z_T} = \frac{e^{-\beta \lambda t}}{Z_T}.\)

  • Spatial Distribution (\(p_X(X)\)): Uniform over a volume \(V\), \(H_X(X) = \text{constant}\), \(p_X(X) = \frac{1}{V}.\)

Given these assumptions, the thermal density matrix becomes

\[\rho_{\kappa} = \frac{1}{Z} \frac{1}{V} \frac{e^{-\beta \lambda t}}{Z_T} \frac{e^{-\beta \frac{|\varphi - \mu|}{b}}}{Z_\varphi} |\psi(\kappa)\rangle \langle \psi(\kappa)|,\]

where, \(Z = Z_X Z_T Z_\varphi\) is the partition function.

Then, the quantum entropy is \(S_{\kappa} = -\text{Tr}(\rho_{\kappa} \log \rho_{\kappa})\) and given the density matrix \(\rho_{\kappa} = \frac{e^{-\beta H_{\kappa}}}{Z}\), the entropy is

\[S_{\kappa} = \beta \langle H_{\kappa} \rangle + \log Z,\]

Where \(\langle H_{\kappa} \rangle\) is the average energy in the system.

For the given Hamiltonian components

  • Kime-Phase Contribution:

\[Z_\varphi = \int_{-\infty}^{\infty} e^{-\beta \frac{|\varphi - \mu|}{b}} d\varphi = 2b e^{\beta \mu} \left(\frac{1 - e^{-2\beta \mu}}{\beta}\right).\]

  • Time Contribution: \(Z_T = \int_{0}^{\infty} e^{-\beta \lambda t} dt = \frac{1}{\beta \lambda}.\)

  • Spatial Contribution: \(Z_X = V \quad (\text{since } H_X(X) \text{ is constant}).\)

Then, the total partition function is

\[Z = V \cdot \frac{1}{\beta \lambda} \cdot 2b e^{\beta \mu} \left( \frac{1 - e^{-2\beta \mu}}{\beta}\right).\]

Therefore, the average energy \(\langle H_{\kappa} \rangle\) is

\[\langle H_{\kappa} \rangle = \langle H_X \rangle + \langle H_T \rangle + \langle H_{\varphi} \rangle = \text{constant} + \frac{1}{\beta} + \frac{b}{\beta}.\]

Substitute back into the entropy formula we get

\[S_{\kappa} = \beta \left(\text{constant} + \frac{1}{\beta} + \frac{b}{\beta}\right) + \log \left(V \cdot \frac{1}{\beta \lambda} \cdot 2b e^{\beta \mu} \left( \frac{1 - e^{-2\beta \mu}}{\beta}\right)\right).\]

\[S_{\kappa} = \beta (\text{constant} + \frac{1 + b}{\beta}) + \log \left(\frac{2b V e^{\beta \mu} (1 - e^{-2\beta \mu})}{\beta^3 \lambda}\right).\]

3.8.6.3 Examples 3: Quantum Spacekime Entropy using a Laplace Phase Distribution and a Phase-Shifted Density Matrix

The phase-shifted density matrix introduces a systematic phase shift across the states, modifying the relative phase of the quantum states. This can represent systems where external fields or boundary conditions induce a phase dependency.

In a phase-shifted density matrix, each quantum state is multiplied by a phase factor \(e^{i\theta(\kappa)}\), where \(\theta(\kappa)\) is a phase shift that depends on the spacekime coordinates \(\kappa = (X, T, \varphi)\).

The phase-shifted density matrix is

\[\rho_{\kappa} = \sum_{\kappa} p(\kappa) e^{i\theta(\kappa)} |\psi(\kappa)\rangle \langle \psi(\kappa)| e^{-i\theta(\kappa)}.\]

  • Kime-Phase Distribution (\(\Phi(\varphi)\)): Assume a Laplace distribution centered at \(\mu\) with scale parameter \(b\). Then, \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\)

  • Time Distribution (\(p_T(t)\)): Exponential decay distribution with parameter \(\lambda\), \(p_T(t) = \lambda e^{-\lambda t}, \quad t \geq 0.\)

  • Spatial Distribution (\(p_X(X)\)): Uniform over a volume \(V\) in 3D Euclidean space, \(p_X(X) = \frac{1}{V}, \quad \text{for } X \in V.\)

  • Phase Shift (\(\theta(\kappa)\)): Assume a linear phase shift in time and kime-phase, \(\theta(\kappa) = \alpha t + \beta \varphi.\)

The density matrix is

\[\rho_{\kappa} = \sum_{\kappa} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) e^{i(\alpha t + \beta \varphi)} |\psi(\kappa)\rangle \langle \psi(\kappa)| e^{-i(\alpha t + \beta \varphi)},\]

which can be simplified to

\[\rho_{\kappa} = \sum_{\kappa} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) |\psi(\kappa)\rangle \langle \psi(\kappa)|.\]

The phase factors \(e^{i(\alpha t + \beta \varphi)}\) cancel out in the final expression for \(\rho_{\kappa}\) since they are conjugates, leaving the density matrix similar to the non-phase-shifted case but with modified states.

The quantum entropy for a phase-shifted density matrix is still defined as \(S_{\kappa} = -\text{Tr}(\rho_{\kappa} \log \rho_{\kappa}).\) Given that the phase shift does not affect the eigenvalues of \(\rho_{\kappa}\), the entropy calculation proceeds similarly to the diagonal case. Substitute the density matrix expression into the entropy formula \(S_{\kappa} = -\sum_{\kappa} p(\kappa) \log p(\kappa).\)

\[S_{\kappa} = -\sum_{X} \sum_{T} \sum_{\varphi} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \log\left(\frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right)\right).\]

Approximate the sums over continuous variables \(t\) and \(\varphi\) by integrals

\[S_{\kappa} = -\int_V \int_{0}^{\infty} \int_{-\infty}^{\infty} \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \log\left( \frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b} \right)\right) d\varphi dt dV.\]

Separate the logarithm into components

\[\log\left(\frac{1}{V} \lambda e^{-\lambda t} \frac{1}{2b} \exp\left(-\frac{ |\varphi - \mu|}{b}\right)\right) = \log\left(\frac{1}{V}\right) + \log\left( \lambda\right) - \lambda t + \log\left(\frac{1}{2b}\right) - \frac{|\varphi - \mu|}{b}.\]

Perform the integrations similarly to the previous cases

  1. Spatial Integral: \(\int_V \frac{1}{V} dV = 1.\)

  2. Temporal Integral (Exponential decay integral): \(\int_0^{\infty} \lambda e^{-\lambda t} \, dt = 1\) and \(\int_0^{\infty} \lambda t e^{-\lambda t} \, dt = \frac{1}{\lambda}.\)

  3. Kime-Phase Integral (Laplace integral): \(\int_{-\infty}^{\infty} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) d\varphi = 1\) and \(\int_{-\infty}^{\infty} \frac{|\varphi - \mu|}{2b^2} \exp\left(-\frac{|\varphi - \mu|}{b}\right) d\varphi = 1.\)

Substituting the results of the integrals we get

\[S_{\kappa} = -\left[\log\left(\frac{1}{V}\right) + \log\left(\lambda\right) + \log\left(\frac{1}{2b}\right) + 1 + \frac{1}{\lambda} \right].\]

This simplifies to \(S_{\kappa} = \log(V) - \log(\lambda) - \log(2b) - 1 - \frac{1}{\lambda}.\)

3.9 Topological Kime Entropy

Definition: This entropy is defined based on the topological features of kimesurfaces, using persistent homology or other topological invariants

\[S_\text{topo}^\kappa = - \sum_{i} \mu_i^\kappa \log \mu_i^\kappa,\] where \(\mu_i^\kappa\) represents the persistence of a topological feature (e.g., a hole, loop) on the kimesurface. The topological kime entropy measures the complexity of the kimesurface topology, potentially reflecting the system’s robustness or stability.

Let’s consider several examples of the topological kime entropy using different topological features of kimesurfaces.

3.9.1 Example 1: Connected Components

Connected components represent distinct, disjoint subsets of a kimesurface that are not connected by any path. Imagine a kimesurface with several disconnected regions, each representing a different phase or region of spacetime that does not connect to the others. These could correspond to different, isolated spacetime regions in the complex-time framework.

The topological kime entropy is

\[S_{topo \choose components}^\kappa = - \sum_{i=1}^n \mu_i^\kappa \log \mu_i^\kappa,\] where, \(n\) is the number of connected components, and \(\mu_i^\kappa\) could represent the relative “size” (e.g., area or volume) of each component.

As an example, assume there are \(3\) connected components with equal size, \(\mu_1^\kappa = \mu_2^\kappa = \mu_3^\kappa = \frac{1}{3}\). Then

\[S_{topo \choose components}^\kappa = - \left(\frac{1}{3} \log \frac{1}{3} + \frac{1}{3} \log \frac{1}{3} + \frac{1}{3} \log \frac{1}{3}\right) = \log 3.\]

3.9.2 Example 2: Loops (1-Dimensional Homology)

Next consider a topological feature counting the loops, or 1-cycles, which correspond to circular paths on the kimesurface that do not bound any surface.

Consider a kimesurface where phase transitions, or time cycles, create circular features that represent non-trivial loops in the topology. Such loops might correspond to periodic phenomena or recurring processes in spacetime.

Then, the Topological Kime Entropy will be

\[S_{topo\choose loops}^\kappa = - \sum_{i=1}^m \nu_i^\kappa \log \nu_i^\kappa,\] where \(m\) is the number of loops and \(\nu_i^\kappa\) represents the persistence or significance of each loop.

In a more specific example, assume that there are \(2\) significant loops wit h persistence \(\nu_1^\kappa = 0.7\) and \(\nu_2^\kappa = 0.3\), respectively. Then,

\[S_{topo\choose loops}^\kappa = - (0.7 \log 0.7 + 0.3 \log 0.3) \approx 0.61.\]

3.9.3 Example 3: Voids (2-Dimensional Homology)

Let’s consider as topological feature the number of voids, or 2-cycles corresponding to hollow regions, or bubbles, on the kimesurface that are not filled. This corresponds to a kimesurface where complex interactions between space and kime may create voids, analogous to bubbles in spacetime. These could represent regions of phase space that are inaccessible or less probable, creating topological voids.

The corresponding Topological Kime Entropy will be

\[S_{topo\choose voids}^\kappa = - \sum_{i=1}^p \xi_i^\kappa \log \xi_i^\kappa,\] where \(p\) is the number of voids, and \(\xi_i^\kappa\) represents the persistence or significance of each void.

In the special case of a single void with \(\xi_1^\kappa = 1\), the entropy will be

\[S_{topo\choose voids}^\kappa = - (1 \log 1) = 0.\] This trivial topological entropy reflects the dominance of a single void with no significant competing structures.

Of course, these examples illustrate some simple Topological Kime Entropy calculations based on alternative kimesurface topological features. Other examples corresponding different kinds of topological structure—connected components may be move challenging to compute.

3.10 Spacekime Entropy of Fluctuations

The Gibbs’s view of thermodynamics states that the entropy is a static property of a stochastic phase space given by the functional \[H=-\sum_i (p_i \log p_i),\] where the equilibrium probability of outcome \(i\) is \(p_i\). For instance, using a fair die whose outcome probabilities are static in time yields \(pi=\frac{1}{6},\ 1\leq i\leq 6\), and the corresponding (static) entropy is

\[-\sum_{i=1}^6 \left(\frac{1}{6}\right)\log\frac{1}{6} = \log 6\approx 1.791759.\]

Entropy fluctuations are intrinsic when the underlying probability distribution itself fluctuates. In Bayesian inference, the equilibrium distribution is the most probable distribution (not unique), among all distributions that satisfy the system constraints. In spacekime, the kime-phase distribution may be time-dynamic, e.g, as in some of the examples we present, the time distribution over \(\mathbb{R}^+\) may be exponential, which would induce temporal dependence, i.e., entropy fluctuations.

Definition: The Entropy of Fluctuations tracks fluctuations in complex-time.

\[S_\text{fluct}^\kappa = \langle \delta \kappa \cdot \log \delta \kappa \rangle,\] where \(\delta \kappa = \kappa - \langle \kappa \rangle\) represents fluctuations around an average complex-time. This entropy measure could phonetically be used in dynamical systems with excessive complex-time fluctuations, e.g., in non-equilibrium thermodynamics. As before, we will consider several simplified examples.

3.10.0.1 Example 1: Cauchy (Lorentzian) Kime-Phase Distribution and Exponential Time Probability Measure

\[S_\text{fluct}^\kappa = \langle \delta \kappa \cdot \log \delta \kappa \rangle,\]

where \(\delta \kappa\) represents fluctuations in complex-time \(\kappa = t \, e^{i\varphi}\). Let’s work out the example using a Cauchy (Lorentzian) Kime-Phase Distribution with an exponential time distribution.

For calculating the Spacekime Entropy of Fluctuations in complex-time under a Cauchy (Lorentzian) Kime-Phase Distribution, the choice of time distribution matters as it directly influences the joint probability distribution that we integrate over to compute the entropy. Different time distributions weight differently on the complex-time fluctuations \(\delta \kappa = \delta t \cdot e^{i\varphi}\) and affect the computed entropy.

Using an exponential time probability distribution measure

\[p_T(t) = \lambda e^{-\lambda t}, \quad t \geq 0.\]

This distribution has a memoryless property, meaning that the probability of future time intervals is independent of the past. When combined with a Cauchy (Lorentzian) kime-phase distribution, it will affect the joint distribution over \(\delta \kappa\) and impact the calculation of the entropy.

\[S_\text{fluct}^\kappa = \langle \delta \kappa \cdot \log \delta \kappa \rangle.\]

The joint probability density function in this case is:

\[p(\delta t, \varphi) = p_T(\delta t) \cdot \Phi(\varphi),\]

where

  • \(p_T(\delta t) = \lambda e^{-\lambda \delta t}\) (Exponential Distribution)
  • \(\Phi(\varphi) = \frac{1}{\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]}\) (Cauchy Distribution)

So, the expected value can be rewritten as

\[S_\text{fluct}^\kappa = \int_{0}^{\infty} \int_{-\infty}^{\infty} \delta \kappa \cdot \log \delta \kappa \cdot p(\delta t, \varphi) \, d\varphi \, d(\delta t).\]

As \(\delta \kappa = \delta t \cdot e^{i\varphi}\),

\[S_\text{fluct}^\kappa = \int_{0}^{\infty} \int_{-\infty}^{\infty} \left( \delta t \cdot (\cos \varphi + i \sin \varphi) \right) \cdot \log \left( \delta t \cdot (\cos \varphi + i \sin \varphi) \right) \cdot \lambda e^{-\lambda \delta t} \cdot \frac{1}{\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]} \, d\varphi \, d(\delta t).\]

  1. Time Distribution Weighting: The exponential distribution gives higher probability to smaller \(\delta t\) values. This means the entropy calculation will be more influenced by the behavior of \(\delta \kappa\) at smaller time fluctuations.

  2. Kime-Phase Contribution: The Cauchy distribution, with its heavy tails, means that \(\varphi\) values far from the mean \(\mu\) still contribute significantly. The combination of these two distributions will lead to a different entropy value compared to using a uniform or Gaussian time distribution.

Given the exponential time distribution, the real part of the entropy calculation is

\[\int_{0}^{\infty} \int_{-\infty}^{\infty} \delta t \cdot \cos \varphi \cdot \log \delta t \cdot \lambda e^{-\lambda \delta t} \cdot \frac{1}{\pi \gamma \left[1 + \left(\frac{\varphi - \mu}{\gamma}\right)^2\right]} \, d\varphi \, d(\delta t).\]

library(cubature)

# Parameters
lambda <- 1    # Rate parameter for the exponential distribution
T_max <- 10    # Maximum time value (upper limit of the time interval)
mu <- 0        # Location parameter for the Cauchy distribution
gamma <- 1     # Scale parameter for the Cauchy distribution

# Exponential distribution for time fluctuations
p_T <- function(delta_t) {
  lambda * exp(-lambda * delta_t)
}

# Cauchy (Lorentzian) distribution for kime-phase
Phi <- function(phi) {
  1 / (pi * gamma * (1 + ((phi - mu) / gamma)^2))
}

# Joint PDF for fluctuations
joint_pdf_fluct <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  p_T(delta_t) * Phi(phi)
}

# Real part of the Spacekime Entropy of Fluctuations
S_fluct_real_func <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  return(delta_t * cos(phi) * log(delta_t) * joint_pdf_fluct(x))
}

# Imaginary part of the Spacekime Entropy of Fluctuations
S_fluct_imag_func <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  return(delta_t * sin(phi) * log(cos(phi) + 1i * sin(phi)) * joint_pdf_fluct(x))
}

# Numerically estimate the real part of the entropy fluctuation
S_fluct_real <- adaptIntegrate(S_fluct_real_func,
                               lowerLimit = c(0, -Inf),
                               upperLimit = c(T_max, Inf))$integral
print(paste("Real Part of Spacekime Entropy of Fluctuations:", S_fluct_real))

# Numerically estimate the imaginary part of the entropy fluctuation
S_fluct_imag <- adaptIntegrate(S_fluct_imag_func, maxEval=100000, tol = 1e-05,
                               lowerLimit = c(0, -Inf),
                               upperLimit = c(T_max, Inf))$integral
print(paste("Imaginary Part of Spacekime Entropy of Fluctuations:", S_fluct_imag))

We can split this (complex) integral into real and imaginary components

  • The real component of \(S_\text{fluct}^\kappa\) is \[\text{Re}(S_\text{fluct}^\kappa) = \int_{\delta t} \int_{\varphi} \delta t \cdot \cos \varphi \cdot \log \delta t \cdot p_T(\delta t) \cdot \Phi(\varphi) \, d\varphi \, d(\delta t).\]

  • The imaginary component of \(S_\text{fluct}^\kappa\) is

\[\text{Im}(S_\text{fluct}^\kappa) = \int_{\delta t} \int_{\varphi} \delta t \cdot \sin \varphi \cdot \log (\cos \varphi + i \sin \varphi) \cdot p_T(\delta t) \cdot \Phi(\varphi) \, d\varphi \, d(\delta t).\]

The R code above performs numerical integration for both the real and imaginary components using the cubature package, and assumes the following parameter settings:

  • lambda <- 1 # Rate parameter for the exponential distribution
  • T_max <- 10 # Maximum time value (upper limit of the time interval)
  • mu <- 0 # Location parameter for the Cauchy distribution
  • gamma <- 1 # Scale parameter for the Cauchy distribution

The Joint PDF for the time and kime-phase fluctuations uses exponential marginal for time and the Cauchy marginal for the kime-phase.

Real and Imaginary Parts of Spacekime Entropy of Fluctuations are

  • $(S_^) = $ 0, and
  • $(S_^) = $ 0.

3.10.0.2 Example 2: Beta Kime-Phase Distribution and Exponential Time Probability Measure

Next, we calculate the Spacekime Entropy of Fluctuations for a Beta Kime-Phase Distribution coupled with an Exponential Time Probability Measure, we’ll follow a similar process as before. The key difference is the Beta distribution for the kime-phase.

The Beta distribution for the kime-phase is given by

\[\Phi(\varphi) = \frac{\varphi^{\alpha-1} (1-\varphi)^{\beta-1}}{B(\alpha, \beta)}, \quad \varphi \in [0, 1],\]

where \(\alpha\) and \(\beta\) are shape parameters, and \(B(\alpha, \beta)\) is the Beta function

\[B(\alpha, \beta) = \int_0^1 t^{\alpha-1} (1-t)^{\beta-1} \, dt.\]

Given that \(\delta \kappa = \delta t \cdot e^{i\varphi}\), the integral setup for the entropy of fluctuation calculation is

\[S_\text{fluct}^\kappa = \int_{\delta t} \int_{\varphi} \left( \delta t \cdot (\cos \varphi + i \sin \varphi) \right) \cdot \log \left( \delta t \cdot (\cos \varphi + i \sin \varphi) \right) \cdot p_T(\delta t) \cdot \Phi(\varphi) \, d\varphi \, d(\delta t).\]

We will split this into real and imaginary components as before and specify the following parameter values:

  • lambda <- 1 # Rate parameter for the exponential distribution
  • alpha <- 2 # Shape parameter alpha for the Beta distribution
  • beta <- 5 # Shape parameter beta for the Beta distribution
  • T_max <- 10 # Maximum time value (upper limit of the time interval)
library(cubature)

# Parameters
lambda <- 1     # Rate parameter for the exponential distribution
alpha <- 2      # Shape parameter alpha for the Beta distribution
beta <- 5       # Shape parameter beta for the Beta distribution
T_max <- 10     # Maximum time value (upper limit of the time interval)

# Exponential distribution for time fluctuations
p_T <- function(delta_t) {
  lambda * exp(-lambda * delta_t)
}

# Beta distribution for kime-phase
Phi <- function(phi) {
  if (phi < 0 || phi > 1) return(0)  # Ensure phi is within [0, 1]
  return((phi^(alpha-1) * (1-phi)^(beta-1)) / beta(alpha, beta))
}

# Joint PDF for fluctuations
joint_pdf_fluct <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  p_T(delta_t) * Phi(phi)
}

# Real part of the Spacekime Entropy of Fluctuations
S_fluct_real_func <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  return(delta_t * cos(phi) * log(delta_t) * joint_pdf_fluct(x))
}

# Imaginary part of the Spacekime Entropy of Fluctuations
S_fluct_imag_func <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  return(delta_t * sin(phi) * log(cos(phi) + 1i * sin(phi)) * joint_pdf_fluct(x))
}

# Numerically estimate the real part of the entropy fluctuation
S_fluct_real <- adaptIntegrate(S_fluct_real_func,
                               lowerLimit = c(0, 0),
                               upperLimit = c(T_max, 1))$integral

print(paste("Real Part of Spacekime Entropy of Fluctuations:", S_fluct_real))

# Numerically estimate the imaginary part of the entropy fluctuation
S_fluct_imag <- adaptIntegrate(S_fluct_imag_func,
                               lowerLimit = c(0, 0),
                               upperLimit = c(T_max, 1))$integral

print(paste("Imaginary Part of Spacekime Entropy of Fluctuations:", S_fluct_imag))

The resulting (numerical estimates of the Real and Imaginary parts of the spacekime entropy of fluctuations are:

  • $(S_^) = $ 0, and
  • $(S_^) = $ 0.

3.10.0.3 Example 3: Laplace Kime-Phase Distribution and Exponential Time Probability Measure

To calculate the Spacekime Entropy of Fluctuations using a Laplace Kime-Phase Distribution and an Exponential Time Probability Measure, we’ll follow a similar process as in the previous examples. The Laplace distribution for the kime-phase \(\varphi\) is given by

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right),\]

where \(\mu\) is the location parameter and \(b\) is the scale parameter.

Given that \(\delta \kappa = \delta t \cdot e^{i\varphi}\), the integral for computing the entropy calculation

\[S_\text{fluct}^\kappa = \int_{\delta t} \int_{\varphi} \left( \delta t \cdot (\cos \varphi + i \sin \varphi) \right) \cdot \log \left( \delta t \cdot (\cos \varphi + i \sin \varphi) \right) \cdot p_T(\delta t) \cdot \Phi(\varphi) \, d\varphi \, d(\delta t).\]

Again, we split the integral into real and imaginary components.

library(cubature)

# Parameters
lambda <- 1      # Rate parameter for the exponential distribution
mu <- 0          # Location parameter for the Laplace distribution
b <- 1           # Scale parameter for the Laplace distribution
T_max <- 10      # Maximum time value (upper limit of the time interval)

# Exponential distribution for time fluctuations
p_T <- function(delta_t) {
  lambda * exp(-lambda * delta_t)
}

# Laplace distribution for kime-phase
Phi <- function(phi) {
  return((1 / (2 * b)) * exp(-abs(phi - mu) / b))
}

# Joint PDF for fluctuations
joint_pdf_fluct <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  p_T(delta_t) * Phi(phi)
}

# Real part of the Spacekime Entropy of Fluctuations
S_fluct_real_func <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  return(delta_t * cos(phi) * log(delta_t) * joint_pdf_fluct(x))
}

# Imaginary part of the Spacekime Entropy of Fluctuations
S_fluct_imag_func <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  return(delta_t * sin(phi) * log(cos(phi) + 1i * sin(phi)) * joint_pdf_fluct(x))
}

# Numerically estimate the real part of the entropy fluctuation
S_fluct_real <- adaptIntegrate(S_fluct_real_func,
                               lowerLimit = c(0, -Inf),
                               upperLimit = c(T_max, Inf))$integral

print(paste("Real Part of Spacekime Entropy of Fluctuations:", S_fluct_real))

# Numerically estimate the imaginary part of the entropy fluctuation
S_fluct_imag <- adaptIntegrate(S_fluct_imag_func,
                               lowerLimit = c(0, -Inf),
                               upperLimit = c(T_max, Inf))$integral

print(paste("Imaginary Part of Spacekime Entropy of Fluctuations:", S_fluct_imag))

In this case, the numerical estimates of the Real and Imaginary parts of the spacekime entropy of fluctuations are:

  • $(S_^) = $ 0, and
  • $(S_^) = $ 0.

3.10.0.4 Example 4: Weibull Kime-Phase Distribution

To calculate the Spacekime Entropy of Fluctuations using a Weibull Kime-Phase Distribution and an Exponential Time Probability Measure, we follow a similar approach. Here’s how we can proceed:

3.10.1 1. Weibull Kime-Phase Distribution

The Weibull distribution for the kime-phase \(\varphi\) is given by

\[\Phi(\varphi) = \frac{k}{\lambda}\left(\frac{\varphi}{\lambda}\right)^{k-1} \exp\left(-\left(\frac{\varphi}{\lambda}\right)^k\right),\]

where \(k\) is the shape parameter and \(\lambda\) is the scale parameter.

Given that \(\delta \kappa = \delta t \cdot e^{i\varphi}\), the integral for the entropy fluctuation calculation is

\[S_\text{fluct}^\kappa = \int_{\delta t} \int_{\varphi} \left( \delta t \cdot (\cos \varphi + i \sin \varphi) \right) \cdot \log \left( \delta t \cdot (\cos \varphi + i \sin \varphi) \right) \cdot p_T(\delta t) \cdot \Phi(\varphi) \, d\varphi \, d(\delta t).\]

Splitting this integral into its real and imaginary components and using the parameters listed below, and applying numerical integration yields the results below.

library(cubature)

# Parameters
lambda_time <- 1     # Rate parameter for the exponential distribution (time)
lambda_weibull <- 1  # Scale parameter for the Weibull distribution (kime-phase)
k_weibull <- 1.5     # Shape parameter for the Weibull distribution (kime-phase)
T_max <- 10          # Maximum time value (upper limit of the time interval)

# Exponential distribution for time fluctuations
p_T <- function(delta_t) {
  lambda_time * exp(-lambda_time * delta_t)
}

# Weibull distribution for kime-phase
Phi <- function(phi) {
  if (phi < 0) return(0)  # Ensure phi is non-negative, as Weibull is defined for phi >= 0
  return((k_weibull / lambda_weibull) * (phi / lambda_weibull)^(k_weibull - 1) * 
         exp(-(phi / lambda_weibull)^k_weibull))
}

# Joint PDF for fluctuations
joint_pdf_fluct <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  p_T(delta_t) * Phi(phi)
}

# Real part of the Spacekime Entropy of Fluctuations
S_fluct_real_func <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  return(delta_t * cos(phi) * log(delta_t) * joint_pdf_fluct(x))
}

# Imaginary part of the Spacekime Entropy of Fluctuations
S_fluct_imag_func <- function(x) {
  delta_t <- x[1]
  phi <- x[2]
  return(delta_t * sin(phi) * log(cos(phi) + 1i * sin(phi)) * joint_pdf_fluct(x))
}

# Numerically estimate the real part of the entropy fluctuation
S_fluct_real <- adaptIntegrate(S_fluct_real_func,
                               lowerLimit = c(0, 0),
                               upperLimit = c(T_max, Inf))$integral

print(paste("Real Part of Spacekime Entropy of Fluctuations:", S_fluct_real))

# Numerically estimate the imaginary part of the entropy fluctuation
S_fluct_imag <- adaptIntegrate(S_fluct_imag_func, maxEval=100000,
                               lowerLimit = c(0, 0),
                               upperLimit = c(T_max, Inf))$integral

print(paste("Imaginary Part of Spacekime Entropy of Fluctuations:", S_fluct_imag))

Note that the integration limits for \(\delta t \in [0, T_{\text{max}}]\) and \(\varphi >0\), non-negative for the Weibull distribution.

The numerical estimates of the Real and Imaginary parts of the spacekime entropy of fluctuations are:

  • $(S_^) = $ 0, and
  • $(S_^) = $ 0.

4 Loop Quantum Gravity and Kime Representation

Loop Quantum Gravity (LQG) is a theory that attempts to quantize general relativity and is characterized by the absence of a fundamental time variable. Instead, it focuses on the quantization of space itself, leading to a picture where time emerges from the evolution of quantum states.

Kime in LQG? In the kime framework, each “loop”, or “quantum of space” in LQG, could be associated with a kime surface, representing its evolution not just through its longitudinal course (classical time), but through complex time. This might provide a new way to represent quantum states of spacetime in LQG as evolving along complex trajectories in spacekime, rather than along a classical timeline (time-curves).

Loop Quantum Gravity (LQG) is a non-perturbative and background-independent approach to quantizing general relativity. In LQG, spacetime is discrete, and the fundamental structure is described by spin networks, which are quantum states of the gravitational field. Time in LQG is emergent and not fundamental, similar to the conceptualization of time in the spacekime framework.

Spacekime Representation extends the traditional 3D spatial and 1D temporal dimensions to a 3D spatial and 2D temporal (complex-time) framework, where the temporal dimension is described by a complex number \(\kappa = t e^{i\varphi}\). In this framework, time emerges as the magnitude of complex time \(t = |\kappa|\).

4.1 Connecting Complex-Time to LQG: Emergent Time

Both LQG and the spacekime framework share the concept of time as an emergent property rather than a fundamental entity. In LQG, time emerges from the dynamics of spin networks, while in the spacekime framework, time emerges from the computation of the magnitude of complex time.

  • Quantum States in Spacekime: In spacekime, quantum states could be defined on a spacekime manifold, where the evolution of these states gives rise to the perception of time. This is analogous to how spin networks evolve in LQG.
  • Spin Networks in Complex-Time: One could explore a formulation where the nodes and links of the spin networks in LQG are described by complex-time coordinates \(\kappa\). This could involve defining spin networks in a spacekime framework and studying how their evolution generates an emergent time dimension.

4.2 Quantization of the Spacekime Metric

In LQG, the quantization of spacetime geometry leads to a discrete spectrum of areas and volumes. A similar approach could be taken in spacekime by attempting to quantize the spacekime metric.

  • Quantization of Spacekime Geometry: Develop a quantum version of the spacekime metric, where the metric components are operators acting on quantum states in spacekime. This could involve defining a spacekime version of the Ashtekar variables used in LQG, which describe the gravitational field.
  • Spin Foam Models in Spacekime: Explore spin foam models in the context of spacekime, where the 2D surfaces in spin foams could be represented in the 5D spacekime manifold. This would involve studying the evolution of these surfaces and their role in generating an emergent temporal dimension.

4.3 Kime-Phase Distribution and Quantum States

In spacekime, the kime-phase \(\varphi\) plays a crucial role in defining the complex-time dimension. One could investigate how the kime-phase distribution affects the evolution of quantum states in a spacekime framework.

  • Kime-Phase as a Quantum Observable: Treat the kime-phase as a quantum observable that evolves according to a quantum dynamics governed by the spacekime field equations. The distribution of kime-phases across different quantum states could be linked to the structure of spin networks in LQG.
  • Entanglement in Spacekime: Investigate the role of quantum entanglement in spacekime, where entanglement might connect spatial and kime-phase dimensions. This could provide a mechanism for how time emerges from the quantum correlations in a spacekime manifold.

4.4 Discrete Spacekime and Quantum Geometry

LQG suggests that spacetime is discrete at the Planck scale. A similar discretization could be applied to the spacekime framework, where the complex-time and spatial dimensions are discretized.

  • Discrete Spacekime Manifold: Define a discrete version of the spacekime manifold, where both spatial and complex-time dimensions are quantized. This could involve defining a spacekime lattice or a spin network-like structure where the nodes and links are described by complex-time coordinates.
  • Emergence of Continuous Time: Study how continuous time could emerge from the evolution of quantum states on this discrete spacekime lattice. This approach could provide a deeper understanding of the relationship between discrete quantum geometry and emergent time in both LQG and spacekime.

4.5 Relating the Wheeler-DeWitt Equation to Spacekime

The Wheeler-DeWitt equation in quantum gravity is a time-independent equation that describes the quantum state of the universe. In spacekime, one could explore a complex-time version of the Wheeler-DeWitt equation.

  • Complex-Time Wheeler-DeWitt Equation: Develop a spacekime extension of the Wheeler-DeWitt equation, where the wavefunction of the universe is defined over a spacekime manifold. This would involve introducing complex-time coordinates into the equation and studying the resulting quantum dynamics.
  • Time Emergence from Wavefunction Evolution: Investigate how the evolution of the wavefunction in complex-time leads to the emergence of a classical time parameter. This could provide a new perspective on how time emerges in quantum gravity.

5 Wheeler-DeWitt equation (WDE)

In quantum gravity, the WDE describes the quantum state of the entire universe without reference to time, and can potentially be reinterpreted in the kime framework. The kime variable could serve as a tool to encode the evolution of the universe in a way that captures both its classical and quantum properties, thus offering a new perspective on time and space in quantum gravity.

Kime-Wheeler-DeWitt Equation: A generalized form of the Wheeler-DeWitt equation could be proposed where the wave function \(\Psi\) depends on the kime variable \(\kappa\) and the spatial geometry (represented by the 3-metric \(\gamma_{ij}\)), \(\Psi(\gamma_{ij}, \kappa)\). This equation would describe the state of the universe in a kime framework, integrating quantum and classical aspects.

The traditional Wheeler-DeWitt equation is given by: \[\mathcal{H} (x)\Psi = 0\ ,\]

where \(\mathcal{H}\) is the Hamiltonian constraint, and \(\Psi\) is the wavefunction of the universe. Rather that representing a classical complex-valued spatial wavefunction \(\Psi:\mathbb{R}^3\to\mathbb{C}\) with \(||\Psi||=1\), \(\Psi\) is a functional of field configurations on all of spacetime and contains the entire information about the geometry and matter content of the universe. Similarly, as an independent constraint in each spatial location \(x\), the Hamiltonian \(\mathcal{H} (x)\) is still an operator acting on the Hilbert space of wavefunctions, but it does not determines the evolution of the system, in this timeless system.

In a 5D spacekime framework, the wavefunction \(\Psi\) would depend on both the spatial metric \(\gamma_{ij}\) and the kime parameter \(\kappa\), perhaps on the distribution of spacekime configurations, or more directly, on the kime-phase distribution \(\Phi_{[-\pi,\pi)}(\varphi)\)

\[\mathcal{H}(x, \kappa) \Psi(\gamma_{ij}, \kappa) = 0\ .\]

To capture the dynamics in the kime domain, \(\mathcal{H}\) may need to be modified to include (distributional) derivatives with respect to \(\kappa\).

The Hamiltonian constraint in this extended space might take the form:

\[\mathcal{H} = G^{ijkl} \frac{\delta^2}{\delta \gamma_{ij} \delta \gamma_{kl}} - \sqrt{\gamma} \, R^{(3)} - \tilde{G} \frac{\partial^2}{\partial \kappa^2}\ ,\]

where

  • \(G^{ijkl}\) is the DeWitt supermetric,
  • \(R^{(3)}\) is the 3D Ricci scalar curvature of the spatial hypersurface,
  • \(\tilde{G}\) is a coefficient that would need to be determined based on the theory’s construction.

In this generalized equation, the wavefunction \(\Psi(\gamma_{ij}, \kappa)\) represents the quantum state of the universe in a combined space-kime framework, where \(\kappa\) incorporates both real and imaginary components, allowing for a richer description of quantum states.

Linking Entropy and the Wheeler-DeWitt Equation: Given that the Wheeler-DeWitt equation describes the wavefunction \(\Psi(\gamma_{ij}, \kappa)\), one could interpret \(\Psi\) as providing a probability amplitude for different configurations of space-kime. The entropy associated with these configurations could be derived from the distribution of these probabilities

\[S_\kappa = -k_B \int \Psi(\gamma_{ij}, \kappa) \ln \Psi(\gamma_{ij}, \kappa) \, d\gamma_{ij} \, d\kappa\]

Does this entropy quantify the uncertainty, or disorder, not just in the spatial configuration of the universe but also in its evolution through complex time?

5.1 Kime-Wheeler-DeWitt Equation Using Distributional Derivatives

Perhaps we can redefine the Kime-Wheeler-DeWitt equation using distributional derivatives with respect to the kime-phase \(\varphi\).

In the space-kime framework, we extend the wavefunction \(\Psi\) to depend on a complex-time (kime) variable \(\kappa\). Here, the kime-phase \(\varphi\) becomes a crucial component. We redefine the wavefunction as \(\Psi[\gamma_{ij}, \varphi]\).

In this context, distributional derivatives are employed to handle situations where the kime-phase \(\varphi\) is not smooth or well-behaved, which is common in quantum systems, where due to IID (random phase) sampling, the kime-phases may involve discontinuities or jumps. The distributional derivative of a function \(f(\varphi)\) is denoted by \(f'(\varphi)\) and is defined in the sense of distributions (generalized functions).

For a generalized function \(f(\varphi)\), its distributional derivative \(f'(\varphi)\) is defined by:

\[\langle f'(\varphi), \phi(\varphi) \rangle = - \langle f(\varphi), \phi'(\varphi) \rangle\ .\]

for any test function \(\phi(\varphi)\).

Let’s explore redefining the Wheeler-DeWitt equation to include distributional derivatives with respect to the kime-phase \(\varphi\):

\[\left( \mathcal{H} + \frac{\delta}{\delta \varphi} \right) \Psi[\gamma_{ij}, \varphi] = 0\ ,\]

where \(\frac{\delta}{\delta \varphi}\) represents the distributional derivative with respect to the kime-phase.

Physical Interpretation and Implications:

  • Quantum Phase Transitions: The use of distributional derivatives with respect to the kime-phase \(\varphi\) could model quantum phase transitions, where the wavefunction of the universe undergoes sudden changes in phase. These transitions are reflected in the non-smooth behavior of \(\varphi\), which is naturally handled by the distributional derivative.

  • Kime-Space Singularities: Distributional derivatives can also model singularities or discontinuities in the kime-phase, corresponding to physical situations where time itself may exhibit quantum fluctuations or discontinuities, as might be expected near the Planck scale or in early universe cosmology.

5.1.1 Example: Applying the Kime-Wheeler-DeWitt Equation

Consider a simple case where the wavefunction \(\Psi\) has a piecewise structure in \(\varphi\), such as:

\[\Psi[\gamma_{ij}, \varphi] = \begin{cases} A(\gamma_{ij}) & \text{if } 0 \leq \varphi < \varphi_0 \\ B(\gamma_{ij}) & \text{if } \varphi_0 \leq \varphi \leq 2\pi \end{cases}\ .\]

The distributional derivative \(\frac{\delta}{\delta \varphi} \Psi[\gamma_{ij}, \varphi]\) would involve Dirac delta functions at the points of discontinuity \(\varphi_0\):

\[\frac{\delta}{\delta \varphi} \Psi[\gamma_{ij}, \varphi] = (B(\gamma_{ij}) - A(\gamma_{ij})) \delta(\varphi - \varphi_0)\ .\]

This leads to a modified Wheeler-DeWitt equation that can capture the effects of sudden phase changes in the quantum state of the universe.

Distributional derivatives with respect to the kime-phase \(\varphi\) may offer a way to obtain a generalized form of the Wheeler-DeWitt equation that can handle non-smooth behaviors in quantum gravity. This formulation could be particularly useful in studying quantum cosmology, where the universe’s wavefunction may undergo rapid phase transitions or exhibit discontinuities in time-like variables. Further exploration of this approach could lead to new insights into the nature of time and the quantum state of the universe.

5.2 Using Averaging or Expectation

Alternatively, we could circumvent direct differentiation with respect to \(\varphi\) by working with the expected values or other statistical moments of the kime-phase distribution. This approach is akin to using moments in statistical physics or quantum mechanics. For instance, instead of differentiating directly, we can use the expectation of the distributional derivative

\[\left\langle \frac{\partial^2 \Psi}{\partial \varphi^2} \right\rangle = \int_{-\pi}^{\pi} \Phi(\varphi) \frac{\partial^2 \Psi(\kappa, x^1, x^2, x^3)} {\partial \varphi^2} \, d\varphi,\]

where \(\langle \cdot \rangle\) denotes an expectation value with respect to the distribution \(\Phi(\varphi)\).

Let’s redefine the Kime Hamiltonian Operator to avoid direct differentiation with respect to \(\varphi\).

A possible redefinition might involve using the distributional derivative with respect to the kime-phase distribution \(\Phi(\varphi)\), or considering the expected value of derivatives with respect to a parameterized function \(f(\theta)\).

5.2.1 Example: Using the Distributional Derivative Approach

\[\hat{H}_\kappa = -\hbar^2 \left( \frac{\partial^2}{\partial t^2} + \left\langle \frac{\partial^2 \Psi}{\partial \varphi^2} \right\rangle \right).\]

In this case, \(\langle \cdot \rangle\) could involve integrating over \(\varphi\) with respect to the kime-phase distribution

\[\left\langle \frac{\partial^2 \Psi}{\partial \varphi^2} \right\rangle = \int \Phi(\varphi) \frac{\partial^2 \Psi}{\partial \varphi^2} d\varphi.\]

Incorporating this into the generalized Kime Wheeler-DeWitt equation:

\[\left[ -\hbar^2 \left( \frac{\partial^2}{\partial t^2} + \left\langle \frac{\partial^2}{\partial \varphi^2} \right\rangle \right) + \hat{H}_{\text{space}} \right] \Psi(\kappa, x^1, x^2, x^3) = 0,\]

where

\[\left\langle \frac{\partial^2}{\partial \varphi^2} \right\rangle = \int \Phi(\varphi) \frac{\partial^2}{\partial \varphi^2} \, d\varphi .\]

This approach maintains the integrity of the equation without running into issues of differentiating with respect to a stochastic variable.

Consider the situation of the Kime Wheeler-DeWitt equation assuming a Laplace kime-phase distribution. We can explicate the wavefunction, the Hamiltonian, and the Wheeler-DeWitt equation in this context, and also explore possible constraints for solving the equation.

The Laplace distribution for the kime-phase \(\varphi\) is given by

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right),\]

where \(\mu\) is the location parameter and \(b\) is the scale parameter.

Consider a simple wavefunction that depends on both the kime variable \(\kappa = t e^{i\varphi}\) and the spatial coordinates \(x^i\)

\[\Psi(\kappa, x^1, x^2, x^3) = e^{-\alpha t} \cdot e^{i\beta \varphi} \cdot \Psi_{\text{space}}(x^1, x^2, x^3),\]

where \(\alpha\) and \(\beta\) are constants, and \(\Psi_{\text{space}}(x^1, x^2, x^3)\) is a function that describes the spatial part of the wavefunction.

In this situation, the spacekime Hamiltonian could be written as

\[\hat{H}_\kappa = -\hbar^2 \left( \frac{\partial^2}{\partial t^2} + \left\langle \frac{\partial^2}{\partial \varphi^2} \right\rangle \right).\]

The term \(\left\langle \frac{\partial^2}{\partial \varphi^2} \right\rangle\) will be computed using the Laplace kime-phase distribution \(\Phi(\varphi)\). Specifically, the I \(\left\langle \frac{\partial^2 \Psi}{\partial \varphi^2} \right\rangle\) using the Laplace distribution is

\[\left\langle \frac{\partial^2 \Psi}{\partial \varphi^2} \right\rangle = \int_{-\infty}^{\infty} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \frac{\partial^2 \Psi(\kappa, x^1, x^2, x^3)}{\partial \varphi^2} \, d\varphi .\]

For the wavefunction \(\Psi(\kappa, x^1, x^2, x^3) = e^{-\alpha t} \cdot e^{i\beta \varphi} \cdot \Psi_{\text{space}}(x^1, x^2, x^3)\), the second derivative with respect to \(\varphi\) is

\[\frac{\partial^2 \Psi}{\partial \varphi^2} = -\beta^2 e^{-\alpha t} \cdot e^{i\beta \varphi} \cdot \Psi_{\text{space}}(x^1, x^2, x^3),\]

and

\[\left\langle \frac{\partial^2 \Psi}{\partial \varphi^2} \right\rangle = -\beta^2 e^{-\alpha t} \cdot \Psi_{\text{space}}(x^1, x^2, x^3) \int_{-\infty}^{\infty} \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) e^{i\beta \varphi} \, d\varphi .\]

This integral can be computed explicitly, though it may be complex. For now, let’s assume the expectation value simplifies to a constant \(C_\varphi\)

\[\left\langle \frac{\partial^2 \Psi}{\partial \varphi^2} \right\rangle = C_\varphi \Psi(\kappa, x^1, x^2, x^3).\]

With the above components, the generalized Kime Wheeler-DeWitt equation becomes

\[\left[ -\hbar^2 \left( \frac{\partial^2}{\partial t^2} + C_\varphi \right) + \hat{H}_{\text{space}} \right] \Psi(\kappa, x^1, x^2, x^3) = 0.\]

Substituting the wavefunction \(\Psi(\kappa, x^1, x^2, x^3)\), we get

\[\left[ -\hbar^2 \left( \alpha^2 + C_\varphi \right) + \hat{H}_{\text{space}} \right] e^{-\alpha t} \cdot e^{i\beta \varphi} \cdot \Psi_{\text{space}}(x^1, x^2, x^3) = 0.\]

Additional constraints (or conditions) may need to be introduced to solve the Kime Wheeler-DeWitt equation. Here are some examples:

  • Boundary Conditions: Specify boundary conditions for the spatial part \(\Psi_{\text{space}}(x^1, x^2, x^3)\). For example, this could involve boundary conditions at infinity or on some compact spatial region.
  • Normalization: Ensure the wavefunction is normalizable, which imposes conditions on the parameters \(\alpha\) and \(\beta\).
  • Choice of \(C_\varphi\): The constant \(C_\varphi\) could be determined by the specific form of the kime-phase distribution and the integrals involved.

For instance, consider a simple case where \(\hat{H}_{\text{space}}\) is a free particle Hamiltonian (no potential), the equation reduces to

\[\hbar^2 \left( \alpha^2 + C_\varphi \right) \Psi(\kappa, x^1, x^2, x^3) = 0.\]

This suggests \(\alpha^2 + C_\varphi = 0.\) Given that \(C_\varphi\) is related to the Laplace distribution parameters, this equation could impose specific constraints on \(\alpha\) and \(\beta\).

5.2.2 Spatially constrained WDE

Let’s try to specifying boundary conditions on \(\Psi_{\text{space}}(x^1, x^2, x^3)\) to derive solutions to the Wheeler-DeWitt equation in spacekime.

We can impose appropriate boundary conditions on the spatial part of the wavefunction \(\Psi_{\text{space}}(x^1, x^2, x^3)\). Below, we explore two scenarios:

  1. constraining the wavefunction to a compact spatial region, such as a 3D box, and
  2. using boundary conditions that correspond to a free particle in a compact region.

Compact Spatial Region: 3D Box: Consider a compact spatial region defined by a 3D box with sides of length \(L_x\), \(L_y\), and \(L_z\). The spatial coordinates \(x^1\), \(x^2\), and \(x^3\) are constrained to this box

\[0 \leq x^1 \leq L_x, \quad 0 \leq x^2 \leq L_y, \quad 0 \leq x^3 \leq L_z .\]

For the wavefunction \(\Psi_{\text{space}}(x^1, x^2, x^3)\), we impose Dirichlet boundary conditions at the walls of the box

\[\Psi_{\text{space}}(x^1, x^2, x^3) = 0 \quad \text{at the boundaries} \quad x^1 = 0, L_x; \, x^2 = 0, L_y; \, x^3 = 0, L_z.\]

Wavefunction Solution in the 3D Box: The wavefunction that satisfies these boundary conditions can be written as

\[\Psi_{\text{space}}(x^1, x^2, x^3) = A \sin\left(\frac{n_1 \pi x^1}{L_x}\right) \sin\left(\frac{n_2 \pi x^2}{L_y}\right) \sin\left(\frac{n_3 \pi x^3}{L_z}\right),\]

where \(n_1\), \(n_2\), and \(n_3\) are positive integers, and \(A\) is a normalization constant.

Using this spatial wavefunction, we substitute into the Kime Wheeler-DeWitt equation

\[\left[ -\hbar^2 \left( \frac{\partial^2}{\partial t^2} + C_\varphi \right) + \hat{H}_{\text{space}} \right] \Psi(\kappa, x^1, x^2, x^3) = 0.\]

The spatial Hamiltonian operator \(\hat{H}_{\text{space}}\) for a free particle is given by the Laplacian

\[\hat{H}_{\text{space}} = -\frac{\hbar^2}{2m} \nabla^2.\]

Substituting \(\Psi_{\text{space}}(x^1, x^2, x^3)\) and applying the Laplacian

\[\nabla^2 \Psi_{\text{space}}(x^1, x^2, x^3) = -\left( \frac{n_1^2 \pi^2}{L_x^2} + \frac{n_2^2 \pi^2}{L_y^2} + \frac{n_3^2 \pi^2}{L_z^2} \right) \Psi_{\text{space}} (x^1, x^2, x^3).\]

Thus,

\[\hat{H}_{\text{space}} \Psi_{\text{space}}(x^1, x^2, x^3) = \frac{\hbar^2}{2m} \left( \frac{n_1^2 \pi^2}{L_x^2} + \frac{n_2^2 \pi^2}{L_y^2} + \frac{n_3^2 \pi^2}{L_z^2} \right) \Psi_{\text{space}}(x^1, x^2, x^3).\]

Substituting everything into the Kime Wheeler-DeWitt equation

\[\left[ -\hbar^2 \left( \alpha^2 + C_\varphi \right) + \frac{\hbar^2}{2m} \left( \frac{n_1^2 \pi^2}{L_x^2} + \frac{n_2^2 \pi^2}{L_y^2} + \frac{n_3^2 \pi^2}{L_z^2} \right) \right] e^{-\alpha t} \cdot e^{i\beta \varphi} \cdot \Psi_{\text{space}} (x^1, x^2, x^3) = 0.\]

This equation can be satisfied if the coefficients of the wavefunction are chosen such that

\[\alpha^2 + C_\varphi = \frac{1}{2m} \left( \frac{n_1^2 \pi^2}{L_x^2} + \frac{n_2^2 \pi^2}{L_y^2} + \frac{n_3^2 \pi^2}{L_z^2} \right).\]

Interpreting the Configuration and Solutions

  • Compact Spatial Region: The assumption of a compact spatial region (3D box) provides a discrete spectrum of solutions for the spatial wavefunction. The integers \(n_1\), \(n_2\), and \(n_3\) correspond to the quantum numbers associated with each spatial dimension.
  • Kime-Phase Contribution: The term \(C_\varphi\), which depends on the Laplace kime-phase distribution, influences the temporal part of the solution. This suggests that the kime-phase dynamics are tied to the spatial quantum numbers and the emergent time.
  • Emergent Time: In this setup, time \(t\) emerges as the magnitude of the complex-time variable \(\kappa\), and its behavior is influenced by the spatial configuration and kime-phase distribution. The fact that the energy eigenvalues (given by the right-hand side of the equation) must balance with the contributions from the kime-phase distribution underscores the interplay between space and kime.

6 Probability Distribution in Kime-Space

In quantum mechanics, the square of the wavefunction \(\Psi\) gives the probability distribution of outcomes. In space-kime, the wavefunction \(\Psi(\gamma_{ij}, \kappa)\) similarly provides a distribution over both space and kime.

The probability \(P\) of a particular spatial configuration \(\gamma_{ij}\) at a specific kime \(\kappa\) is:

\[P(\gamma_{ij}, \kappa) = |\Psi(\gamma_{ij}, \kappa)|^2 \]

Linking Probability and Entropy: This probability can be used to define an entropy that measures the uncertainty in the space-kime state

\[S_\kappa = -\int P(\gamma_{ij}, \kappa) \ln P(\gamma_{ij}, \kappa) \, d\gamma_{ij} \, d\kappa\]

Here, the entropy depends on both the spatial configuration and the kime parameter, reflecting the complex structure of the universe’s quantum state.

7 Bringing Kime and Space on the Same Level

Dimensional Equivalence: In traditional physics, space and time are treated differently, with time often being considered as a parameter rather than a dimension. In the spacekime framework, we treat the kime parameter \(\kappa\) on equal footing with spatial coordinates.

7.1 Spacekime Metric

The 5D metric in spacekime, including three spatial dimensions and one complex kime dimension, could be written as:

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, d\kappa^2 + \tilde{\beta}_a \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b \ ,\]

where \(A, B = 1, 2, 3, 4, 5\) and \(\kappa\) is treated as a dimension on par with the spatial coordinates \(X^a\).

In the space-kime framework, where \(\kappa = t \cdot e^{i\varphi}\) is a complex-valued quantity, defining \(d\kappa^2\) in the metric tensor requires careful consideration of how to handle the complex nature of \(\kappa\). Here’s how we can explicitly define \(d\kappa^2\):

7.2 How to interpret \(d\kappa\)?

Given \(\kappa = t \cdot e^{i\varphi}\), the differential of \(\kappa\) is:

\[d\kappa = e^{i\varphi} \, dt + i \, t \, e^{i\varphi} \, d\varphi\]

Expand the complex exponential \(e^{i\varphi}\) using Euler’s formula \(e^{i\varphi} = \cos\varphi + i\sin\varphi\).

This differential can be separated into its real and imaginary components

\[d\kappa = e^{i\varphi} \left( dt + i \, t \, d\varphi \right) =\\ (\cos\varphi + i\sin\varphi) \left( dt + i \, t \, d\varphi \right)=\\ (\cos\varphi \, dt - \sin\varphi \, t \, d\varphi) + i (\sin\varphi \, dt + \cos\varphi \, t \, d\varphi) \ .\]

7.2.1 Calculating \(d\kappa\) and \(d\kappa^2\) as real values?

If the spacekime metric needs to be real-valued, then \(d\kappa\) and \(d\kappa^2\) should also be real. This presents a challenge because \(\kappa = t \cdot e^{i\varphi} \in \mathbb{C}\) is inherently complex.

To reconcile this, we need to carefully redefine how we treat \(d\kappa\) in a way that maintains the real-valued nature of the metric. Here are two approaches:

Using the Modulus of \(d\kappa\): One way to ensure that the metric remains real is to use the modulus of \(d\kappa\), effectively treating the magnitude of the complex differential rather than the complex differential itself.

Given \(d\kappa = e^{i\varphi} \left( dt + i \, t \, d\varphi \right)\), the modulus (magnitude) of \(d\kappa\) is

\[|d\kappa| = \sqrt{\left(\cos\varphi \, dt - \sin\varphi \, t \, d\varphi\right)^2 + \left(\sin\varphi \, dt + \cos\varphi \, t \, d\varphi\right)^2}\ .\]

This simplifies to \(|d\kappa| = \sqrt{dt^2 + t^2 \, d\varphi^2}\). Using \(|d\kappa|\) ensures that the metric remains real-valued.

Alternatively, one could reinterpret the complex-time parameter \(\kappa\) such that only its real part (or an appropriate combination of real and imaginary parts) is used in the metric.

For instance, \(\kappa_{\text{real}} = t \cdot \cos(\varphi)\) and \(d\kappa_{\text{real}} = \cos(\varphi) \, dt - \sin(\varphi) \, t \, d\varphi\).

This real part of \(\kappa\) can then be used in the metric to ensure it remains real-valued.

To properly define \(d\kappa^2\), we take the square of the differential \(d\kappa\)

\[d\kappa^2 = \left(e^{i\varphi} \, dt + i \, t \, e^{i\varphi} \, d\varphi \right)^2 .\]

and expand it

\[d\kappa^2 = e^{2i\varphi} \, dt^2 + 2i \, t \, e^{2i\varphi} \, dt \, d\varphi - t^2 \, e^{2i\varphi} \, d\varphi^2\ (\in \mathbb{C})\ .\]

In the context of the space-kime metric, the term \(d\kappa^2\) must be interpreted carefully, as the metric should ultimately be a real-valued quantity. Therefore, this complex nature of \(d\kappa^2\) suggests that we must consider both the real and imaginary parts separately, or alternatively, consider the modulus squared of \(d\kappa\).

To ensure the metric remains real, one approach is to use the modulus squared of \(d\kappa\)

\[|d\kappa|^2 = d\kappa \, d\kappa^* = \left( e^{i\varphi} \, dt + i \, t \, e^{i\varphi} \, d\varphi \right) \left( e^{-i\varphi} \, dt - i \, t \, e^{-i\varphi} \, d\varphi \right) \ .\]

Expanding this \(|d\kappa|^2 = dt^2 + t^2 \, d\varphi^2\).

This real-valued expression can be interpreted as the complex-time contribution \(\kappa\) to the spacekime metric.

Given the above derivation, the differential \(d\kappa^2\) in the space-kime metric can be replaced with \(|d\kappa|^2 = dt^2 + t^2 \, d\varphi^2\) to maintain the reality of the metric

To integrate the concepts of complex-time (\(\kappa\)) and its real-valued differential into the formuation of the spacekime metric while maintaining the real-valued nature of the metric tensor, we need to carefully redefine how the metric interacts with \(d\kappa\).

The general form of the space-kime metric \(G_{AB}\) is:

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, d\kappa^2 + \tilde{\beta}_a \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]

where:

  • \(d\kappa\) is the differential associated with the complex-time \(\kappa = t \cdot e^{i\varphi}\).
  • \(dX^a\) refers to the differentials of the spatial coordinates.

Approach 1: Using the Modulus of \(d\kappa\): To ensure the metric remains real-valued, we can replace \(d\kappa\) with its modulus \(|d\kappa| = \sqrt{dt^2 + t^2 \, d\varphi^2}\).

This leads to a redefined metric:

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \left(dt^2 + t^2 \, d\varphi^2\right) + \tilde{\beta}_a \, |d\kappa| \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ .\]

This ensures that the metric remains real and consistent with the requirements of general relativity.

Approach 2: Incorporating the Real Part of \(d\kappa\): Alternatively, if we choose to use the real part of \(d\kappa\), the differential is \(\text{Re}(d\kappa) = \cos\varphi \, dt - \sin\varphi \, t \, d\varphi\), and the metric can then be redefined as

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \left(\cos\varphi \, dt - \sin\varphi \, t \, d\varphi\right)^2 + \tilde{\beta}_a \, \text{Re}(d\kappa) \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ .\]

This approach explicitly uses the real-valued differential while still incorporating the effects of the kime-phase \(\varphi\).

In either case, the spacekime metric takes the form:

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, |d\kappa|^2 + \tilde{\beta}_a \, \text{Real}(d\kappa) \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]

where:

  • \(|d\kappa|^2\) ensures the lapse term contributes positively to the metric, reflecting the proper “distance” in the space-kime framework.
  • \(\text{Real}(d\kappa)\) can be used if we want to keep the shift terms real and directly interpretable in the context of physical measurements.
  • \(\gamma_{ab}\) remains the spatial metric, which is unaffected by the kime phase directly but can be influenced through the dynamics dictated by \(\tilde{\beta}_a\).

With this redefined spacekime metric, the spacekime field equations can be derived analogously to the standard Einstein field equations but now include contributions from the kime-phase \(\varphi\). The dynamics of the spacekime manifold, including how energy and momentum are distributed, will be influenced by these additional terms, leading to potentially novel physical predictions.

7.2.2 Using Wirtinger derivatives

The formulation of the spacekime metric could also utilize Wirtinger derivatives. Here’s an alternative approach to define the spacekime metric.

Wirtinger derivatives are a method of differentiating functions of complex variables by using conjugate basis in teh complec plane, i.e., treating a complex variable and its conjugate as independent. For a complex variable \(\kappa = t \cdot e^{i\varphi}\), the Wirtinger derivatives with respect to a complex variable and its conjugate \(\kappa,\bar{\kappa}\in\mathbb{C}\) are defined as follows.

\[\frac{\partial}{\partial \kappa} = \frac{1}{2} \left(\frac{\partial}{\partial x} - i\frac{\partial}{\partial y}\right), \\ \frac{\partial}{\partial \bar{\kappa}} = \frac{1}{2} \left(\frac{\partial}{\partial x} + i\frac{\partial}{\partial y}\right)\ ,\]

where \(x\) and \(y\) are the real and imaginary parts of \(\kappa\).

Using Wirtinger derivatives allows us to treat \(\kappa\) and \(\bar{\kappa}\) independently in the spacekime metric, leading to a potentially different form of the metric components.

When differentiating with respect to \(\kappa\) and \(\bar{\kappa}\), the metric tensor \(G_{AB}\) would include terms involving both types of derivatives, potentially leading to a richer structure that can capture the complex interactions between the real and imaginary components of time. The metric could take a form similar to:

\[G_{AB} \, dX^A \, dX^B= -\tilde{N}^2 \, \left(d\kappa \, d\bar{\kappa}\right) + \tilde{\beta}_a \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]

where, \(d\kappa \, d\bar{\kappa}\) represents the mixed derivative term that inherently captures both real and imaginary parts through the Wirtinger formalism.

There may be advantages of using Wirtinger derivatives:

  • Handling Complex Time Directly: The use of Wirtinger derivatives allows for a direct and consistent method of differentiating with respect to complex time \(\kappa\). This approach automatically separates the contributions from the real and imaginary parts, making the equations more tractable.
  • Metric Symmetry: The resulting metric tensor may have symmetries that are not obvious when using standard derivatives, especially in the context of complex geometry.
  • Potential for New Physical Insights:

By leveraging the Wirtinger derivatives, one might uncover new physical effects related to the interaction of the real and imaginary components of time, which could be crucial for understanding phenomena in quantum gravity. The field equations derived from this metric would need to incorporate Wirtinger derivatives, potentially leading to new types of source terms or modifications to the standard Einstein equations. This could alter how we interpret energy, momentum, and curvature in the context of spacekime.

7.3 Curvature in Spacekime

The curvature of this space-kime manifold would involve both the usual spatial curvature and additional terms accounting for the kime dimension. The Ricci scalar for the 5D space-kime manifold could be expressed as:

\[R^{(5)} = R^{(4)} - \frac{1}{\tilde{N}} \nabla_a \nabla^a \tilde{N} + \text{(terms involving kime)}\ ,\]

where \(R^{(4)}\) is the Ricci scalar for the 4D spacetime slice.

An extended spacekime framework may generalize the Wheeler-DeWitt equation, entropy, and probability distribution theory to incorporate the complex-time domain. By treating the kime parameter on the same level as spatial dimensions, wemay develop a richer mathematical structure for describing the universe’s quantum state. Further exploration could lead to novel insights into the nature of time, entropy, and quantum gravity.

In the 5D spacekime, to bring the 2D kime (complex-time) on the same conceptual level as the 3D space in the context of quantum gravity and general relativity, we must properly define the space-kime metric. This involves not only defining the lapse function \(\tilde{N}\) but also incorporating the kime-phase distribution \(\Phi(\varphi)\) into the metric to fully describe the combined space-kime manifold.

Definition (Lapse Function \(\tilde{N}\)): In the ADM formalism of general relativity, the lapse function \(N\) controls how time progresses from one spacelike hypersurface to the next. In spacekime, we can define a separable generalized lapse function \(\tilde{N}\) as a function that incorporates both the real-time and kime-phase components:

\[\tilde{N}(\kappa) = N(t) \cdot \Phi(\varphi)\ , \]

where:

  • \(\kappa = t \cdot e^{i\varphi}\) is the complex-time (kime).
  • \(N(t)\) is the traditional lapse function that depends on the real-time \(t\).
  • \(\Phi(\varphi)\) is the kime-phase distribution, reflecting the probabilistic nature of the kime-phase.

This formulation allows the lapse function to vary not only with time but also with the kime-phase, capturing the dynamic and probabilistic nature of time in the space-kime framework.

7.4 Constructing the Spacekime Metric

The spacekime metric \(G_{AB}\) must incorporate the effects of the kime-phase distribution \(\Phi(\varphi)\) into both the spatial and kime components. The general form of the 5D metric in spacekime is given by

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, d\kappa^2 + \tilde{\beta}_a \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ .\]

Expanding this using the definitions:

\[G_{AB} \, dX^A \, dX^B = -N(t)^2 \, \Phi(\varphi)^2 \, d\kappa^2 + \tilde{\beta}_a \, \Phi(\varphi) \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]

where:

  • \(\tilde{\beta}_a\) is the shift vector in spacekime, which could also depend on \(\Phi(\varphi)\).
  • \(\gamma_{ab}\) is the three-metric on the 4D spacetime-like hypersurface.

7.5 The Role of \(\Phi(\varphi)\) in the Space-Kime Metric

The kime-phase distribution \(\Phi(\varphi)\) plays a critical role in determining how time and space are interconnected in spacekime. The generalized function \(\Phi(\varphi)\) modulates the lapse function and the shift vector, influencing how the complex-time dimension evolves and interacts with the spatial dimensions.

  • Probabilistic Modulation: \(\Phi(\varphi)\) can be chosen to reflect various probabilistic distributions, such as Gaussian, uniform, or other forms, depending on the physical scenario being modeled. This modulation affects the geometry of the space-kime manifold, potentially leading to new insights into how quantum fluctuations or phase transitions affect the structure of spacetime.

  • Physical Implications: The dependence of the lapse function and shift vector on \(\Phi(\varphi)\) implies that the geometry of spacetime could exhibit probabilistic or phase-dependent behaviors, especially in regimes where quantum gravitational effects are significant.

7.6 Spacekime Metric Tensor and Inverse Metric Tensor

In spacekime, we extend the usual 4D spacetime metric to a 5D structure that includes complex-time (kime) as an additional dimension. Here, we explicitly define both the metric tensor \(G_{AB}\) (lower indices) and its inverse \(G^{AB}\) (Upper indices).

7.6.1 Kime Metric Tensor \(G_{AB}\)

The spacekime metric tensor is defined to describe change, specifically the interval between two events in the 5D spacekime manifold. The general form of the metric tensor in this context is

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, d\kappa^2 + \tilde{\beta}_a \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]

where:

  • the 5D indices \(A,B\in(1, 2, 3, 4, 5)\), where the first four indices correspond to spacetime coordinates and the fifth index corresponds to the kime-phase (random sampling from teh phase distribution).
  • \(\tilde{N}(\kappa, \varphi)\) is the generalized lapse function, which may depend on the complex-time \(\kappa\) and kime-phase \(\varphi\).
  • \(\tilde{\beta}_a\) is the shift vector in the spacekime formalism, which might also depend on \(\varphi\).
  • \(\gamma_{ab}\) is the 4D metric tensor on the spacetime-like hypersurfaces.

Thus, the explicit form of the spacekime metric tensor in matrix form is

\[G_{AB} = \begin{pmatrix} \gamma_{11} & \gamma_{12} & \gamma_{13} & \gamma_{14} & \tilde{\beta}_1 \\ \gamma_{21} & \gamma_{22} & \gamma_{23} & \gamma_{24} & \tilde{\beta}_2 \\ \gamma_{31} & \gamma_{32} & \gamma_{33} & \gamma_{34} & \tilde{\beta}_3 \\ \gamma_{41} & \gamma_{42} & \gamma_{43} & \gamma_{44} & \tilde{\beta}_4 \\ \tilde{\beta}_1 & \tilde{\beta}_2 & \tilde{\beta}_3 & \tilde{\beta}_4 & -\tilde{N}^2 \end{pmatrix}\ ,\]

where:

  • The \(\gamma_{ab}\) components (\(a, b \in\{1, 2, 3, 4 \}\)) correspond to the 4D metric of the spacetime hypersurface.
  • The \(\tilde{\beta}_a\) components represent the interaction terms between the kime-phase and spacetime.
  • The last component \(-\tilde{N}^2\) represents the kime-phase.

7.6.2 Inverse Metric Tensor \(G^{AB}\)

Naturally, the inverse spacekime metric tensor \(G^{AB}\) is defined so that

\[G^{AB} G_{BC} = \delta^A_C= \begin{cases} 0 & A\not= C\\ 1 & A\equiv C \end{cases}\ .\]

Given the form of the metric tensor \(G_{AB}\), the inverse metric \(G^{AB}\) can be computed by matrix inversion. For the block matrix structure above, the inverse typically involves more complex relationships between the lapse function, shift vector, and spatial metric components. However, for illustrative purposes, the inverse metric generally takes the form:

\[G^{AB} = \begin{pmatrix} \gamma^{11} & \gamma^{12} & \gamma^{13} & \gamma^{14} & \tilde{\beta}^1 \\ \gamma^{21} & \gamma^{22} & \gamma^{23} & \gamma^{24} & \tilde{\beta}^2 \\ \gamma^{31} & \gamma^{32} & \gamma^{33} & \gamma^{34} & \tilde{\beta}^3 \\ \gamma^{41} & \gamma^{42} & \gamma^{43} & \gamma^{44} & \tilde{\beta}^4 \\ \tilde{\beta}^1 & \tilde{\beta}^2 & \tilde{\beta}^3 & \tilde{\beta}^4 & -\tilde{N}^{-2} \end{pmatrix}\ , \]

where:

  • \(\gamma^{ab}\) are the components of the inverse 4D metric tensor on the spacetime hypersurface.
  • \(\tilde{\beta}^a\) are components related to the shift vector, adjusted appropriately to satisfy the inverse metric conditions.
  • The \(-\tilde{N}^{-2}\) term is the inverse of the lapse function squared, associated with the kime dimension.

This inverse metric tensor will play a critical role in defining spacekime dynamical equations, energy conditions, and field equations.

The explicit definitions of the spacekime metric tensor \(G_{AB}\) and its inverse \(G^{AB}\) provide the foundation for the generalized treatment of general relativity with complex-time. These tensors encapsulate the geometry of the space-kime manifold, allowing for the derivation of the corresponding field equations, Hamiltonian formulation, and energy definitions in this extended setting.

7.7 Alternative Formulation of the Field Equations in Spacekime

With the metric now explicitly incorporating \(\Phi(\varphi)\), the spacekime field equations derived earlier can be re-expressed to include these effects. Specifically, the Ricci tensor \(R_{AB}\) and Ricci scalar \(R^{(5)}\) must now be computed considering the modulated lapse function \(\tilde{N}(\kappa, \varphi)\) and the spacekime metric \(G_{AB}\).

The Einstein field equations in spacekime then take the form:

\[\underbrace{R_{AB} - \frac{1}{2} G_{AB} R^{(5)}}_{curvature\ tensors,\ incl. \choose second\ derivatives\ of\ the\ metric} = \underbrace{\overbrace{8\pi G \, T_{AB}}^{energy-momentum\ tensor} + \overbrace{\frac{1}{\Phi(\varphi)}\nabla_A \nabla_B \Phi(\varphi)}^{ kime-phase\ distribution\ impact\choose on\ spacetime\ curvature}}_{ matter\ \&\ sampling\ distribution}\ ,\]

where each term now includes contributions from \(\Phi(\varphi)\).

The term \(\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\) in the spacekime extension of the Einstein field equations introduces a modification that reflects the influence of the kime-phase distribution \(\Phi(\varphi)\) on spacetime curvature. This term consists of a few key components

  • \(\Phi(\varphi)\): This represents the kime-phase distribution, which describes the probabilistic distribution of the kime-phase \(\varphi\) within the complex-time framework. In essence, it encodes the fluctuations or variations in the phase component of complex time.
  • \(\nabla_A \nabla_B \Phi(\varphi)\): This denotes the covariant derivative of the kime-phase distribution \(\Phi(\varphi)\) with respect to the coordinates \(X^A\) and \(X^B\) in the 5D spacekime manifold. This second derivative measures how \(\Phi(\varphi)\) changes as we move through spacekime, capturing both the curvature of the spacekime manifold and the variation in the phase distribution.
  • \(\frac{1}{\Phi(\varphi)}\): This scaling factor normalizes the derivative term, allowing the expression to be independent of the magnitude of \(\Phi(\varphi)\) itself. It can be seen as introducing a weight that adjusts the influence of the kime-phase distribution based on its own value.

The term \(\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\) can be interpreted as follows.

  • Modification of Curvature: The term modifies the spacetime curvature based on the properties of the kime-phase distribution. Specifically, it reflects how the variations in the phase distribution influence the overall curvature of spacekime. When \(\Phi(\varphi)\) varies rapidly (i.e., has significant gradients), this term can lead to additional curvature effects that would not be present in a purely spacetime-based framework.
  • Coupling of Space and Kime Dynamics: The presence of this term indicates a coupling between the traditional spacetime dynamics and the kime-phase dynamics. The second derivative of \(\Phi(\varphi)\) effectively means that any non-trivial structure in the phase distribution directly contributes to the curvature of the spacekime manifold.

In a quantum gravity context, \(\Phi(\varphi)\) could represent quantum fluctuations or uncertainties in the phase of complex time. The term \(\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\) might then represent a correction to classical curvature due to these quantum effects. If \(\Phi(\varphi)\) varies spatially or temporally, this term indicates that the curvature (and thus the geometry of the universe) could be influenced by the underlying non-uniform evolution of time itself, as captured by the complex-time formalism. The term could also be seen as an effective source term in the Einstein field equations, similar to the way the energy-momentum tensor \(T_{AB}\) acts as a source for curvature. Here, instead of energy and momentum, the source is the structure of the kime-phase distribution.

Let’s consider two specific cases of \(\Phi(\varphi)\) to illustrate the role of this term.

Constant \(\Phi(\varphi)\): If \(\Phi(\varphi)\) is constant (i.e., it does not vary with \(\varphi\)), then \(\nabla_A \Phi(\varphi) = 0\), and this term vanishes. In this case, the spacekime curvature is not influenced by the kime-phase distribution, and the field equations reduce to their more familiar spacetime forms.

Non-Constant \(\Phi(\varphi)\) (e.g., Laplace Distribution): For a non-constant \(\Phi(\varphi)\), such as a Laplace distribution, this term does not vanish and instead introduces additional curvature based on the gradients and second derivatives of the distribution. This introduces a correction to the curvature based on the rate at which the kime-phase distribution varies. It might correspond to scenarios where the universe’s geometry is dynamically influenced by underlying quantum or phase fluctuations.

The additional term involves covariant derivatives, meaning it accounts for the underlying geometry of the spacekime manifold. This ensures that the contribution to curvature is compatible with the overall structure of spacekime, preserving the general covariance of the field equations. The scaling factor in ths new term captures the contribution of the term relative to the phase distribution itself. In regions where \(\Phi(\varphi)\) is small, the term could become significant, potentially indicating regions of high curvature or phase transition.

This, the extra term \(\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\) in the spacekime field equations acts as a source of curvature that depends on the kime-phase distribution \(\Phi(\varphi)\). It represents the influence of the underlying quantum or phase structure of time on the geometry of the universe, coupling the dynamics of space and kime. This term opens up possibilities for new physical phenomena, such as phase-induced curvature effects, that would not be present in a purely spacetime-based framework.

Let’s look at three examples of the spacekime field equations using Euclidean spacetime metric tensors and different kime-phase distributions (Uniform, Laplace, and Gaussian). In this case, the spacekime field equations are

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi),\]

where

  • \(R_{AB}\) is the 5D Ricci tensor.
  • \(G_{AB}\) is the 5D spacekime metric tensor.
  • \(R^{(5)}\) is the Ricci scalar.
  • \(T_{AB}\) is the energy-momentum tensor.
  • \(\Phi(\varphi)\) is the kime-phase distribution.
  • \(\nabla_A \nabla_B \Phi(\varphi)\) represents the covariant derivatives of the kime-phase distribution.

7.7.1 Example 1: Uniform Kime-Phase Distribution

Kime-Phase Distribution:

\[\Phi(\varphi) = \frac{1}{2\pi}, \quad \varphi \in [-\pi, \pi).\]

Spacekime Metric Tensor: Given the Euclidean spacetime metric, we have

\[G_{AB} = \begin{pmatrix} -1 & 0 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 1 \end{pmatrix}.\]

where the time component is \(-1\) and the spatial components are Euclidean.

Covariant Derivatives: For a uniform distribution \(\nabla_A \nabla_B \Phi(\varphi) = 0,\) because \(\Phi(\varphi)\) is constant.

Field Equations: Substituting into the field equations, the term involving \(\nabla_A \nabla_B \Phi(\varphi)\) vanishes, and we recover the usual Einstein field equations

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB}.\]

In this case, the kime-phase distribution does not affect the curvature, and the field equations reduce to those of a flat spacetime in 5D, with no additional phase-induced curvature.

7.7.2 Example 2: Laplace Kime-Phase Distribution

Kime-Phase Distribution:

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right),\]

where \(b\) is the scale parameter and \(\mu\) is the location parameter (mean).

Spacekime Metric Tensor: The same Euclidean metric is used

\[G_{AB} = \begin{pmatrix} -1 & 0 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 1 \end{pmatrix}.\]

Covariant Derivatives: The second derivatives of the Laplace distribution are

\[\nabla_A \nabla_B \Phi(\varphi) = \frac{\delta_{AB}}{b^2} \Phi(\varphi),\]

where \(\delta_{AB}\) is the Kronecker delta, indicating that the derivatives act primarily on the \(\varphi\) dimension.

Field Equations: Substituting into the field equations

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{\delta_{AB}}{b^2}.\]

The additional term \(\frac{\delta_{AB}}{b^2}\) represents a contribution to the curvature that is directly proportional to the inverse of the scale parameter \(b\). This implies that the smaller the scale of the kime-phase distribution, the larger the induced curvature, suggesting possible effects similar to those of a cosmological constant but varying with the phase distribution.

7.7.3 Example 3: Gaussian Kime-Phase Distribution

Kime-Phase Distribution:

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left( -\frac{(\varphi - \mu)^2}{2\sigma^2}\right),\]

where \(\sigma\) is the standard deviation.

Spacekime Metric Tensor: Again, we use the Euclidean metric

\[G_{AB} = \begin{pmatrix} -1 & 0 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 1 \end{pmatrix}.\]

Covariant Derivatives: For the Gaussian distribution

\[\nabla_A \nabla_B \Phi(\varphi) = \frac{(\varphi - \mu)}{\sigma^4} \delta_{AB} \Phi(\varphi).\]

This second derivative captures the curvature effects of the Gaussian distribution.

Substituting into the field equations we get

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{(\varphi - \mu)}{\sigma^4} \delta_{AB}.\]

Here, the additional term introduces curvature that is dependent on the displacement of \(\varphi\) from its mean \(\mu\). This could lead to phase-dependent gravitational effects, where different regions of spacekime experience different curvatures based on the underlying phase distribution. This might correspond to phase-induced inhomogeneities in the universe’s geometry.

In the next three examples, we will consider a non-trivial spacetime curvature, tensor can be computed from the first and second derivatives of the corresponding spacetime metric tensor.

7.8 Examples of Non-trivial Curvature Tensors

The curvature of spacetime is described by the Riemann curvature tensor, which can be computed from the first and second derivatives of the underlying spacetime metric tensor. Let’s explicate three examples of non-trivial spacetime curvature tensors: the Schwarzschild metric, the Friedmann-Lemaître-Robertson-Walker (FLRW) metric, and the Kerr metric.

7.8.1 Schwarzschild Metric

Metric Tensor:: The Schwarzschild metric describes the spacetime outside a non-rotating, spherically symmetric massive object. In Schwarzschild coordinates \((t, r, \theta, \phi)\), the metric tensor is

\[g_{\mu\nu} = \begin{pmatrix} -\left(1 - \frac{2GM}{r}\right) & 0 & 0 & 0 \\ 0 & \left(1 - \frac{2GM}{r}\right)^{-1} & 0 & 0 \\ 0 & 0 & r^2 & 0 \\ 0 & 0 & 0 & r^2 \sin^2\theta \end{pmatrix},\] where \(G\) is the gravitational constant and \(M\) is the mass of the object.

Riemann Curvature Tensor Calculation: The Riemann tensor \(R^\rho_{\sigma\mu\nu}\) is computed using the Christoffel symbols, which are functions of the first derivatives of the metric tensor

\[\Gamma^\rho_{\mu\nu} = \frac{1}{2} g^{\rho\lambda} \left( \partial_\mu g_{\nu\lambda} + \partial_\nu g_{\mu\lambda} - \partial_\lambda g_{\mu\nu} \right).\]

The Riemann curvature tensor is then calculated using

\[R^\rho_{\sigma\mu\nu} = \partial_\mu \Gamma^\rho_{\nu\sigma} - \partial_\nu \Gamma^\rho_{\mu\sigma} + \Gamma^\rho_{\mu\lambda} \Gamma^\lambda_{\nu\sigma} - \Gamma^\rho_{\nu\lambda} \Gamma^\lambda_{\mu\sigma}.\]

For the Schwarzschild metric, this calculation yields non-zero components

\[R^r_{\theta r \theta} = -\frac{GM}{r^2},\] which reflects the curvature due to the central mass. The non-trivial curvature tensor components indicate how spacetime is curved by the presence of mass, leading to phenomena such as gravitational time dilation and the bending of light.

7.8.2 Friedmann-Lemaître-Robertson-Walker (FLRW) Metric

Metric Tensor: The FLRW metric describes a homogeneous, isotropic expanding or contracting universe. In comoving coordinates \((t, r, \theta, \phi)\), the metric tensor is

\[g_{\mu\nu} = \text{diag}\left(-1, \frac{a(t)^2}{1 - kr^2}, a(t)^2 r^2, a(t)^2 r^2 \sin^2\theta\right),\]

where \(a(t)\) is the scale factor and \(k\) is the curvature parameter (\(k = 0\) for flat, \(k = 1\) for closed, \(k = -1\) for open universes).

Riemann Curvature Tensor Calculation: The Christoffel symbols for the FLRW metric can be derived similarly

\[\Gamma^t_{rr} = \frac{a(t) \dot{a}(t)}{1 - kr^2}, \quad \Gamma^r_{tr} = \frac{\dot{a}(t)}{a(t)}.\]

The Riemann tensor components are then computed using these Christoffel symbols. For the FLRW metric, a key component is \(R_{trtr} = -\frac{\ddot{a}(t)}{a(t)} g_{rr}\). This indicates how the scale factor \(a(t)\) influences the curvature of spacetime. The non-trivial components of the Riemann tensor describe the dynamics of the universe’s expansion or contraction, as governed by the scale factor \(a(t)\). The Ricci scalar \(R\) from the FLRW metric is directly related to the energy density and pressure of the universe.

7.8.3 Kerr Metric

Metric Tensor: The Kerr metric describes the spacetime around a rotating massive object. In Boyer-Lindquist coordinates \((t, r, \theta, \phi)\), the metric tensor is

\[g_{\mu\nu} = \begin{pmatrix} -\left(1 - \frac{2GMr}{\rho^2}\right) & 0 & 0 & -\frac{2GMar\sin^2\theta}{\rho^2} \\ 0 & \frac{\rho^2}{\Delta} & 0 & 0 \\ 0 & 0 & \rho^2 & 0 \\ -\frac{2GMar\sin^2\theta}{\rho^2} & 0 & 0 & \sin^2\theta \left( r^2 + a^2 + \frac{2GMr a^2 \sin^2\theta}{\rho^2} \right) \end{pmatrix},\] where \(\rho^2 = r^2 + a^2 \cos^2\theta\) and \(\Delta = r^2 - 2GMr + a^2\), with \(a\) being the spin parameter.

Riemann Curvature Tensor Calculation: The computation involves determining the Christoffel symbols, which are more complex due to the off-diagonal terms in the metric

\[\Gamma^\phi_{t\theta} = \frac{GMa\sin\theta}{\rho^2}, \quad \Gamma^r_{\theta\theta} = \frac{a^2\sin\theta\cos\theta}{\rho^2}.\]

The Riemann tensor components, \(R^r_{\phi r \phi} = \frac{GM \sin^2\theta (2r - M)}{\rho^4}\) indicate how the curvature is influenced by both mass and spin. The Kerr metric’s non-trivial curvature components reflect the frame-dragging effect and the ergosphere around a rotating black hole. These effects are key to understanding the dynamics near rotating black holes.

7.8.4 Spacekime Field Equations under Kerr Metric Tensor and Uniform Kime-Phase Distribution

We’ll use the Kerr metric in spacekime with a uniform kime-phase distributions to expand the spacekime field equations.

The Kerr metric describes the spacetime around a rotating black hole, and in Boyer-Lindquist coordinates \((t, r, \theta, \phi)\), the metric tensor \(g_{\mu\nu}\) is

\[g_{\mu\nu} = \begin{pmatrix} -\left(1 - \frac{2GMr}{\rho^2}\right) & 0 & 0 & -\frac{2GMar\sin^2\theta}{\rho^2} \\ 0 & \frac{\rho^2}{\Delta} & 0 & 0 \\ 0 & 0 & \rho^2 & 0 \\ -\frac{2GMar\sin^2\theta}{\rho^2} & 0 & 0 & \sin^2\theta \left( r^2 + a^2 + \frac{2GMr a^2 \sin^2\theta}{\rho^2} \right) \end{pmatrix},\]

where

  • \(\rho^2 = r^2 + a^2 \cos^2\theta\)
  • \(\Delta = r^2 - 2GMr + a^2\)
  • \(M\) is the mass of the black hole.
  • \(a\) is the spin parameter.

To express the Spacekime Metric Tensor, we extend the Kerr metric to 5D by introducing an additional dimension corresponding to the complex time \(\kappa = t e^{i\varphi}\), where \(\varphi\) is the kime-phase. The 5D spacekime metric tensor \(G_{AB}\) can be expressed as

\[G_{AB} = \begin{pmatrix} g_{\mu\nu} & 0 \\ 0 & \langle d\kappa^2 \rangle \end{pmatrix},\]

where \(\langle d\kappa^2 \rangle\) depends on the kime-phase distribution \(\Phi(\varphi)\).

Then, the generalized spacekime field equations are given by

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

For a uniform kime-phase distribution \(\Phi(\varphi) = \frac{1}{2\pi}, \quad \varphi \in [-\pi, \pi)\), the metric component \(\langle d\kappa^2 \rangle\) simplifies as \(\varphi\) is independent of the coordinates, leading to a constant additional term

\[\langle d\kappa^2 \rangle = dt^2 - t^2 \frac{d\varphi^2}{(2\pi)}.\]

Since \(\Phi(\varphi)\) is constant, the covariant derivatives vanish \(\nabla_A \nabla_B \Phi(\varphi) = 0.\) In this case, the spacekime field equations reduce to

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB}.\]

These are indeed the usual Einstein field equations, extended to include the additional \(\langle d\kappa^2 \rangle\) term. The uniform kime-phase distribution does not introduce additional curvature effects. Therefore, in the presence of a uniform kime-phase distribution, the spacetime curvature is not influenced by the kime-phase dynamics. This scenario is equivalent to the standard Kerr solution in 4D, but extended into the higher-dimensional spacekime.

7.8.5 Spacekime Field Equations under Kerr Metric Tensor and Laplace Kime-Phase Distribution

COnsider a Laplace distribution \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right),\) where \(b\) is a scale parameter and \(\mu\) is the mean. In this case, the metric component \(\langle d\kappa^2 \rangle\) is influenced by the Laplace distribution

\[\langle d\kappa^2 \rangle = \int_{-\infty}^{\infty} \left(dt^2 - t^2 d\varphi^2\right) \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) d\varphi.\]

The covariant derivatives of \(\Phi(\varphi)\) yield

\[\nabla_A \nabla_B \Phi(\varphi) = \frac{\delta_{AB}}{b^2} \Phi(\varphi).\]

Substituting these into the spacekime field equations

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{\delta_{AB}}{b^2}.\]

The Laplace kime-phase distribution introduces an additional curvature term proportional to \(\frac{1}{b^2}\). This term suggests that the curvature is sensitive to the scale of the kime-phase distribution, leading to potential modifications of the Kerr solution that vary with the kime-phase. For instance, smaller values of \(b\) would imply stronger curvature effects, potentially altering the nature of the black hole’s horizon or ergosphere depending on the kime-phase distribution’s properties.

The next tweo examples extend the Schwarzschild metric to spacekime and explore the effects of two different kime-phase distributions (Laplace and Gaussian) on the spacekime field equations.

7.8.6 Schwarzschild Metric Tensor

The Schwarzschild metric describes the spacetime around a non-rotating, spherically symmetric massive object. In Schwarzschild coordinates \((t, r, \theta, \phi)\), the metric tensor \(g_{\mu\nu}\) is

\[g_{\mu\nu} = \begin{pmatrix} -\left(1 - \frac{2GM}{r}\right) & 0 & 0 & 0 \\ 0 & \left(1 - \frac{2GM}{r}\right)^{-1} & 0 & 0 \\ 0 & 0 & r^2 & 0 \\ 0 & 0 & 0 & r^2 \sin^2\theta \end{pmatrix},\]

where \(G\) is the gravitational constant and \(M\) is the mass of the object.

The extended Schwarzschild metric to 5D spacekime relies on an additional dimension corresponding to the complex time \(\kappa = t e^{i\varphi}\), where \(\varphi\) is the kime-phase. The 5D spacekime metric tensor \(G_{AB}\) can be expressed as

\[G_{AB} = \begin{pmatrix} g_{\mu\nu} & 0 \\ 0 & \langle d\kappa^2 \rangle \end{pmatrix},\]

where \(\langle d\kappa^2 \rangle\) depends on the kime-phase distribution \(\Phi(\varphi)\).

Then, the generalized spacekime field equations are

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

7.8.7 Example 1: Schwarzschild spacetime metric and Laplace Kime-Phase Distribution

For a Laplace kime-phase distribution

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right),\]

where \(b\) is a scale parameter and \(\mu\) is the mean.

In this case, the metric component \(\langle d\kappa^2 \rangle\) is influenced by the Laplace distribution

\[\langle d\kappa^2 \rangle = \int_{-\infty}^{\infty} \left(dt^2 - t^2 d\varphi^2\right) \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) d\varphi.\]

The covariant derivatives of \(\Phi(\varphi)\) yield

\[\nabla_A \nabla_B \Phi(\varphi) = \frac{\delta_{AB}}{b^2} \Phi(\varphi).\]

Substituting these into the spacekime field equations

\[R_{AB}-\frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4}T_{AB} + \frac{\delta_{AB}}{b^2}.\]

In the presence of a Laplace kime-phase distribution, the additional term \(\frac{\delta_{AB}}{b^2}\) introduces a correction to the curvature. This term implies that the curvature is sensitive to the scale of the kime-phase distribution, leading to possible modifications in the Schwarzschild solution. For instance, smaller values of \(b\) would imply stronger curvature effects, potentially altering the nature of the gravitational field outside the black hole.

This example suggests that the Laplace distribution could lead to phase-dependent corrections to the classical Schwarzschild solution, possibly affecting the event horizon’s properties or the gravitational potential at various radii.

7.8.8 Example 2: Schwarzschild spacetime metric and Gaussian Kime-Phase Distribution

For a Gaussian kime-phase distribution

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right),\]

where \(\sigma\) is the standard deviation.

In this case, the metric component \(\langle d\kappa^2 \rangle\) is influenced by the Gaussian distribution

\[\langle d\kappa^2 \rangle = \int_{-\infty}^{\infty} \left(dt^2 - t^2 d\varphi^2\right) \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \mu)^2}{2\sigma^2}\right) d\varphi.\]

The covariant derivatives of \(\Phi(\varphi)\) yield

\[\nabla_A \nabla_B \Phi(\varphi) = \frac{(\varphi - \mu)}{\sigma^4} \delta_{AB} \Phi(\varphi).\]

Substituting these into the spacekime field equations

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{(\varphi - \mu)}{\sigma^4} \delta_{AB}.\]

The Gaussian kime-phase distribution introduces a term that depends on the displacement of \(\varphi\) from its mean \(\mu\). This term could lead to phase-dependent gravitational effects where different regions of spacekime experience different curvatures based on the underlying phase distribution.

For example, regions where \(\varphi\) is closer to the mean \(\mu\) would experience weaker corrections, while regions farther from the mean might experience stronger curvature effects. This could imply that the gravitational field outside a black hole might not be spherically symmetric when accounting for kime-phase fluctuations, leading to anisotropies in the gravitational potential.

7.8.9 Fields Equation Solutions using Schwarzschild spacetime metric and Laplace Kime-Phase Distribution

Let’s explore the conditions for solving the above field equations assuming Schwarzschild Metric tensor and Laplace kime-phase distribution. What constraints are necessary for solution consistency?

The spacekime field equations are given by

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

For the Laplace kime-phase distribution \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right)\), the covariant derivatives of \(\Phi(\varphi)\) yield

\[\nabla_A \nabla_B \Phi(\varphi) = \frac{\delta_{AB}}{b^2} \Phi(\varphi).\]

Substituting into the field equations

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{\delta_{AB}}{b^2}.\]

1. Consider the Schwarzschild Metric in 4D: The Schwarzschild metric is

\[g_{\mu\nu} = \begin{pmatrix} -\left(1 - \frac{2GM}{r}\right) & 0 & 0 & 0 \\ 0 & \left(1 - \frac{2GM}{r}\right)^{-1} & 0 & 0 \\ 0 & 0 & r^2 & 0 \\ 0 & 0 & 0 & r^2 \sin^2\theta \end{pmatrix}.\]

In 5D spacekime, the metric tensor \(G_{AB}\) will have an additional dimension corresponding to the kime-phase distribution

\[G_{AB} = \begin{pmatrix} g_{\mu\nu} & 0 \\ 0 & \langle d\kappa^2 \rangle \end{pmatrix},\]

where \(\langle d\kappa^2 \rangle\) represents the contribution from the kime-phase.

The Ricci tensor \(R_{AB}\) in 5D will have contributions from both the spacetime part and the kime-phase part. The additional term from the kime-phase is

\[\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi) = \frac{\delta_{AB}}{b^2}.\]

This term is a constant proportional to \(\delta_{AB}\), which means it adds a uniform curvature in all directions.

The spacekime field equations reduce to

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{\delta_{AB}}{b^2}.\]

Since \(R^{(5)}\) includes contributions from both spacetime curvature and the kime-phase term, we can express the Ricci scalar as \(R^{(5)} = R^{(4)} + \frac{4}{b^2},\) where \(R^{(4)}\) is the Ricci scalar in the 4D Schwarzschild spacetime.

Then, the equations simplify to

\[R_{AB} - \frac{1}{2} G_{AB} \left(R^{(4)} + \frac{4}{b^2}\right) = \frac{8\pi G}{c^4} T_{AB} + \frac{\delta_{AB}}{b^2}.\]

This equation suggests that the kime-phase distribution adds a term that mimics a cosmological constant but is dependent on the kime-phase distribution parameter \(b\).

Constraints Necessary for Solutions:

  • Matching Boundary Conditions: The solution must match the Schwarzschild solution at the boundary where the kime-phase effects become negligible (i.e., far from the central mass where the spacetime is asymptotically flat).
  • Stability Conditions: The solution should be stable under small perturbations. This may require conditions on the scale parameter \(b\) of the Laplace distribution to ensure that the additional curvature term does not lead to instability or singularities in the solution.
  • Energy Conditions: The energy-momentum tensor \(T_{AB}\) should satisfy the usual energy conditions (e.g., weak energy condition) to ensure physical plausibility.

The term \(\frac{\delta_{AB}}{b^2}\) acts like a phase-dependent curvature that modifies the Schwarzschild solution. This could be interpreted as a phase-induced cosmological constant, which might lead to effects similar to those observed in cosmological scenarios (e.g., accelerating expansion) but localized around the Schwarzschild mass.

To derive an explicit solution for the spacekime field equations using the Schwarzschild metric tensor and a Laplace kime-phase distribution, we will make the following assumptions:

  1. Matching Boundary Conditions: The solution should match the Schwarzschild solution at large distances from the central mass, where the effects of the kime-phase distribution become negligible. This implies that far from the source, the solution should asymptotically approach the standard Schwarzschild solution.

  2. Vacuum Solution: We’ll assume that the region we’re considering is a vacuum outside the central mass, so \(T_{AB} = 0\).

  3. Simplified Setup: We will focus on the time-time component and radial components of the field equations to illustrate the solution.

Working with Matching Boundary Conditions, the Schwarzschild metric, and the Laplace kime-phase distribution, the spacekime field equations are

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{\delta_{AB}}{b^2}.\]

For the Schwarzschild metric, the non-zero components of the Ricci tensor \(R_{\mu\nu}\) in the vacuum are typically zero, but due to the additional kime-phase term, we must modify the Ricci scalar \(R^{(5)}\) to account for the spacekime extension \(R^{(5)} = R^{(4)} + \frac{4}{b^2}.\)

Since \(R^{(4)} = 0\) for the Schwarzschild vacuum solution (outside the mass), the Ricci scalar simplifies to \(R^{(5)} = \frac{4}{b^2}.\)

Substituting \(R^{(5)}\) into the spacekime field equations, we get

\[R_{AB} - \frac{1}{2} G_{AB} \cdot \frac{4}{b^2} = \frac{\delta_{AB}}{b^2}.\]

We now examine the components of this equation.

  • The time-time component of the modified equation is

\[R_{tt} - \frac{1}{2} G_{tt} \cdot \frac{4}{b^2} = \frac{1}{b^2}.\]

For the Schwarzschild metric \(G_{tt} = -\left(1 - \frac{2GM}{r}\right)\) \(R_{tt}\) is typically zero in the Schwarzschild vacuum, and the equation simplifies to

\[-\frac{2}{b^2} \left(1 - \frac{2GM}{r}\right) = \frac{1}{b^2}\]

This simplifies further to \(1 - \frac{2GM}{r} = \frac{1}{2},\) which gives \(r = 4GM.\) This indicates a specific radial distance where the kime-phase effects balance out the Schwarzschild curvature.

Similarly, for the radial component, (\(A=B=r\))

\[R_{rr} - \frac{1}{2} G_{rr} \cdot \frac{4}{b^2} = \frac{1}{b^2},\]

where for the Schwarzschild metric \(G_{rr} = \left(1 - \frac{2GM}{r}\right)^{-1}.\)

Substituting in, we get

\[-\frac{2}{b^2} \left(1 - \frac{2GM}{r}\right)^{-1} = \frac{1}{b^2}.\]

This equation is more complex to solve directly, but it essentially implies a similar balance between the radial curvature and the kime-phase distribution at a specific radius.

As \(r\) increases (i.e., moving far from the central mass), the effects of the kime-phase distribution diminish, and the solution should approach the Schwarzschild solution

\[1 - \frac{2GM}{r} \underset{r \to \infty}{\longrightarrow 1} .\]

This matches the boundary condition that the solution asymptotically approaches flat spacetime.

Solution Interpretation

  • The solution we derived suggests that there exists a specific radial distance \(r = 4GM\) where the kime-phase effects are particularly significant, potentially leading to observable modifications of the gravitational field.
  • The Laplace distribution introduces a constant curvature correction to the Schwarzschild solution, acting like a cosmological constant that is phase-dependent.
  • This modification could affect the properties of the event horizon or lead to phase-dependent gravitational effects that vary with the scale parameter \(b\) of the Laplace distribution.

7.8.10 Energy Conditions for the Fields Equation Solutions based on Schwarzschild spacetime metric and Laplace Kime-Phase Distribution

To derive an explicit solution of the spacekime field equations using the Schwarzschild spacetime metric and a Laplace kime-phase distribution, we need to incorporate the energy conditions, which ensure that the solution is physically meaningful. The energy conditions typically include

  1. Weak Energy Condition (WEC): \(T_{AB} v^A v^B \geq 0\) for any timelike vector \(v^A\).
  2. Null Energy Condition (NEC): \(T_{AB} k^A k^B \geq 0\) for any null vector \(k^A\).
  3. Strong Energy Condition (SEC): \((T_{AB} - \frac{1}{2} G_{AB} T) v^A v^B \geq 0\) for any timelike vector \(v^A\).

The spacekime field equations, incorporating the kime-phase distribution, are

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

For a Laplace kime-phase distribution \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\)

The covariant derivatives of \(\Phi(\varphi)\) yield \(\nabla_A \nabla_B \Phi(\varphi) = \frac{\delta_{AB}}{b^2} \Phi(\varphi).\)

Substituting into the field equations

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{\delta_{AB}}{b^2}.\]

The Schwarzschild metric is

\[g_{\mu\nu} = \begin{pmatrix} -\left(1 - \frac{2GM}{r}\right) & 0 & 0 & 0 \\ 0 & \left(1 - \frac{2GM}{r}\right)^{-1} & 0 & 0 \\ 0 & 0 & r^2 & 0 \\ 0 & 0 & 0 & r^2 \sin^2\theta \end{pmatrix}.\]

The spacekime metric tensor in 5D will be

\[G_{AB} = \begin{pmatrix} g_{\mu\nu} & 0 \\ 0 & \langle d\kappa^2 \rangle \end{pmatrix},\]

where \(\langle d\kappa^2 \rangle\) is determined by the Laplace kime-phase distribution.

Given the Schwarzschild metric, the non-zero components of the Ricci tensor \(R_{\mu\nu}\) in vacuum are zero. However, due to the kime-phase term, the Ricci scalar \(R^{(5)}\) is modified as \(R^{(5)} = R^{(4)} + \frac{4}{b^2}.\) Since \(R^{(4)} = 0\) in the Schwarzschild vacuum, the Ricci scalar simplifies to \(R^{(5)} = \frac{4}{b^2}.\)

Assume the energy-momentum tensor \(T_{AB}\) satisfies the energy conditions (e.g., WEC, NEC, SEC). For simplicity, let’s assume \(T_{AB} = 0\) in the vacuum. Then, \(R_{AB} - \frac{1}{2} G_{AB} \frac{4}{b^2} = \frac{\delta_{AB}}{b^2}.\)

The trace of this equation gives the Ricci scalar \(R^{(5)}\), leading to \(R^{(5)} = \frac{4}{b^2}\) and hence \(T_{AB} = 0.\) This implies that the right-hand side of the field equations consists solely of the kime-phase-induced curvature term.

Explicit Solution:

  • For the time-time component \(A = B = t\) \[R_{tt} - \frac{1}{2} G_{tt} \frac{4}{b^2} = \frac{1}{b^2}.\]

Given the Schwarzschild metric \(G_{tt} = -\left(1 - \frac{2GM}{r}\right)\), this simplifies to \(-\frac{2}{b^2} \left(1 - \frac{2GM}{r}\right) = \frac{1}{b^2}.\)

This equation can be solved to find \(1 - \frac{2GM}{r} = \frac{1}{2}\), which implies \(r = 4GM.\)

  • For the radial component \(A = B = r\), \(R_{rr} - \frac{1}{2} G_{rr} \frac{4}{b^2} = \frac{1}{b^2}.\) Given the Schwarzschild metric \(G_{rr} = \left(1 - \frac{2GM}{r}\right)^{-1},\) the equation simplifies similarly, suggesting that the kime-phase distribution introduces a radius where the curvature changes significantly.

To satisfy the energy conditions and match boundary conditions:

  • Boundary Condition: The solution must asymptotically approach the Schwarzschild solution as \(r \rightarrow \infty\). This ensures that the spacekime effects diminish at large distances from the source, matching the classical Schwarzschild solution.
  • Energy Conditions: Ensure \(T_{AB} v^A v^B \geq 0\) for the WEC, which in this vacuum case implies ensuring that the curvature induced by the kime-phase distribution does not violate the physical constraints (e.g., no negative energy densities).

The solution suggests that the kime-phase distribution adds a constant curvature term, acting like a phase-dependent cosmological constant. This term modifies the Schwarzschild solution, leading to a modified radius of curvature that could impact the properties of the black hole’s event horizon or the gravitational potential at various radii.

7.8.11 Other Older Examples

7.8.11.1 Example 1: Simple Gaussian Distribution for \(\Phi(\varphi)\)

Consider a simple Gaussian distribution for \(\Phi(\varphi)\):

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} e^{-\frac{(\varphi - \varphi_0)^2}{2\sigma^2}}.\]

This distribution could represent the natural variability in the kime-phase due to quantum fluctuations. The lapse function in this case would be:

\[\tilde{N}(\kappa, \varphi) = N(t) \cdot \frac{1}{\sqrt{2\pi\sigma^2}} e^{-\frac{(\varphi - \varphi_0)^2}{2\sigma^2}}\ .\]

Substituting this into the spacekime metric gives us a framework where the spacetime curvature, and thus the gravitational dynamics, are influenced by the kime-phase prior, i.e., the assumed probabilistic nature of the kime-phase.

By defining the extended lapse function \(\tilde{N}\) in terms of the kime-phase distribution \(\Phi(\varphi)\), and integrating this into the spacekime metric, we can establish a framework where kime and space are treated on an equal footing.

7.8.11.2 Example 2: Using Laplace Distribution for \(\Phi(\varphi)\)

Let’s consider another example where the kime-phase distribution \(\Phi(\varphi)\) is modeled by a (symmetric) Laplace distribution, which is characterized by a sharp peak at the mean value and heavier tails compared to a Gaussian distribution. Laplace distribution is used to model situations where the kime-phase is likely to be close to a central value but with a significant probability of larger deviations.

The Laplace Distribution for \(\Phi(\varphi)\): The probability density function of a Laplace distribution is given by:

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right)\ ,\]

where:

  • \(\varphi_0\) is the mean (location parameter),
  • \(b\) is the scale parameter, which determines the width of the distribution.

This distribution is symmetric around \(\varphi_0\) and exhibits an exponential decay from the central value. In practice, the phase-distribution needs to be truncated on the region \([-\pi,\pi)\), or alternatively, periodically wrapped around using module \(2\pi\) arithmetic.

Lapse Function \(\tilde{N}(\kappa, \varphi)\) with Laplace Distribution: Using the Laplace distribution for \(\Phi(\varphi)\), we can define the generalized lapse function \(\tilde{N}(\kappa, \varphi)\) as:

\[\tilde{N}(\kappa, \varphi) = N(t) \cdot \frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right)\ .\]

This lapse function modulates the spacekime metric, allowing the kime-phase distribution to influence the progression of time in a complex manner.

Constructing the Space-Kime Metric: The space-kime metric, incorporating the Laplace-distributed kime-phase, is given by:

\[G_{AB} \, dX^A \, dX^B = \underbrace{-\left(N(t) \cdot \frac{1}{2b} \exp\left( -\frac{|\varphi - \varphi_0|}{b}\right)\right)^2 \, d\kappa^2}_{I} +\\ \underbrace{\tilde{\beta}_a \, \frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right) \, d\kappa \, dX^a}_{II} + \underbrace{\gamma_{ab} \, dX^a \, dX^b}_{III}\ .\]

In this expression:

  • The first term \(I\) represents the time-time (kime-magnitude - kime-phase) component of the metric, heavily influenced by the Laplace distribution of the kime-phase.
  • The second term \(II\) represents the interaction between the space-kime components, modulated by the same distribution.
  • The third term, \(III= \gamma_{ab} \, dX^a \, dX^b\), represents the purely spatial part of the metric, which remains unchanged.

The revised spacekime Field Equations assuming a Laplace-distribution kime-phase prior are

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \left(\frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right)\right)\ .\]

The (distributional) derivatives of \(\Phi(\varphi)\) with respect to \(\varphi\) will yield terms involving the exponential function and its gradient, leading to contributions that reflect the rapid decay of the kime-phase away from \(\varphi_0\).

Physical Interpretation:

  • Localized Time Dynamics: The Laplace distribution, with its sharp peak, implies that the dynamics of time as influenced by the kime-phase are highly localized around a central value. This could correspond to scenarios where quantum systems exhibit strong coherence around a particular phase but with a significant probability of deviations.

  • Enhanced Tail Effects: The heavier tails of the Laplace distribution imply that the influence of extreme kime-phase values is more pronounced than in a Gaussian distribution. This could model situations where rare but significant deviations in the kime-phase have substantial effects on the geometry of spacetime.

7.9 Alternative Formulation of the Field Equations in Spacekime

Let’s explore the derivation of an alternative formulation of the field equations in the spacekime framework. This derivation explicitly incorporates the kime-phase distribution \(\Phi(\varphi)\), distributional derivatives with respect to the kime-phase, and Wirtinger derivatives for the complex time \(\kappa\).

Again, starting with the spacekime metric \(G_{AB}\), which incorporates the kime-phase distribution \(\Phi(\varphi)\) and is defined in terms of the complex time \(\kappa = t \cdot e^{i\varphi}\)

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \left( |d\kappa|^2 \right) + \tilde{\beta}_a \, \text{Re}(d\kappa) \, dX^a + \gamma_{ab} \, dX^a \, dX^b \ ,\]

where:

  • \(|d\kappa|^2\) is the modulus squared of the differential, ensuring the metric remains real-valued.
  • \(\tilde{N}\) and \(\tilde{\beta}_a\) are generalized lapse functions and shift vectors, potentially dependent on \(\Phi(\varphi)\).
  • \(\gamma_{ab}\) represents the 4D metric on spacetime-like hypersurfaces.

The kime-phase distribution \(\Phi(\varphi)\) modulates the spacekime dynamics. We incorporate this by introducing distributional derivatives with respect to \(\varphi\):

\[\frac{\delta}{\delta \varphi} \Phi(\varphi) = \frac{\delta \Phi(\varphi)}{\delta \varphi}.\]

This allows us to express how the field equations depend on the kime-phase distribution, particularly when handling phase transitions or discontinuities.

Using Wirtinger derivatives, we treat \(\kappa\) and its conjugate \(\bar{\kappa}\) as independent variables:

\[\frac{\partial}{\partial \kappa} = \frac{1}{2} \left(\frac{\partial}{\partial t} - i\frac{\partial}{\partial \varphi}\right), \quad \frac{\partial}{\partial \bar{\kappa}} = \frac{1}{2} \left(\frac{\partial}{\partial t} + i\frac{\partial}{\partial \varphi}\right)\ .\]

These derivatives allow us to take into account both the real and imaginary parts of complex time, providing a more detailed description of the dynamics.

To derive the spacekime field equations, we start with the Einstein-Hilbert action generalized to spacekime

\[S = \int \left( R^{(5)} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi) \right) \sqrt{-G} \, d^5X\ ,\]

where, \(R^{(5)}\) is the Ricci scalar in 5D space-kime, and \(\sqrt{-G}\) is the determinant of the space-kime metric. Varying this action with respect to the metric \(G_{AB}\) gives the field equations

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi) - \frac{\partial G_{AB}}{\partial \kappa} \frac{\partial \kappa}{\partial X^A} - \frac{\partial G_{AB}}{\partial \bar{\kappa}} \frac{\partial \bar{\kappa}}{\partial X^A}\ .\]

This equation incorporates:

  • Distributional Derivatives: \(\frac{\delta \Phi(\varphi)}{\delta \varphi}\), capturing the effects of kime-phase transitions.
  • Wirtinger Derivatives: \(\frac{\partial}{\partial \kappa}\) and \(\frac{\partial}{\partial \bar{\kappa}}\), capturing the real and imaginary contributions of the complex time.

The spacekime field equations provide a richer structure than in traditional general relativity. They account for:

  • Kime-Phase Effects: The distributional derivatives allow for the modeling of phenomena where the kime-phase might jump or exhibit discontinuities, such as during quantum phase transitions.
  • Complex Time Dynamics: The Wirtinger derivatives enable the separation of real and imaginary components of time, which can lead to different physical interpretations of temporal evolution and its effects on the geometry of space-kime.

Physical Implications: These field equations suggest that the geometry of space-kime is influenced not only by the distribution of matter and energy (as in standard general relativity) but also by the kime-phase and the complex structure of time. This could lead to new predictions in quantum gravity, where the behavior of spacetime is sensitive to both the phase and magnitude of complex time.

This alternative formulation of the spacekime field equations using distributional derivatives for the kime-phase and Wirtinger derivatives for complex time, is not unique.

7.9.1 Spacekime Metric Tensor \(G_{AB}\)

Let’s try an alternative strategy to incorporates the kime-phase distribution \(\Phi(\varphi)\) in the spacekime metric tensor,

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, d\kappa^2 + \tilde{\beta}_a \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ .\]

Reformulating \(d\kappa^2\) in terms of the Kime-Phase distribution \(\Phi(\varphi)\) Instead of computing \(d\kappa^2\) directly, we compute the expected value of \(d\kappa^2\) under the distribution \(\Phi(\varphi)\)

\[\langle d\kappa^2 \rangle = \int_{-\pi}^{\pi} \left( e^{i\varphi} \, dt + i t e^{i\varphi} \, d\varphi \right)^2 \Phi(\varphi) \, d\varphi .\]

  • This integral can be expanded and simplified to provide a real-valued contribution to the metric tensor.
  • We need to calculate the expectation value of each component involving \(\varphi\).

\[\langle d\kappa^2 \rangle = \left\langle e^{2i\varphi} \right\rangle dt^2 + 2i t \left\langle e^{2i\varphi} \right\rangle dt \, d\varphi - t^2 \left\langle e^{2i\varphi} \right\rangle d\varphi^2 .\] - The expectation value \(\langle e^{2i\varphi} \rangle\) under the distribution \(\Phi(\varphi)\) is given by

\[\langle e^{2i\varphi} \rangle = \int_{-\pi}^{\pi} e^{2i\varphi} \Phi(\varphi) \, d\varphi .\] - The spacekime metric tensor incorporating the kime-phase distribution \(\Phi(\varphi)\) becomes

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, \langle d\kappa^2 \rangle + \tilde{\beta}_a \, \langle d\kappa \rangle \, dX^a + \gamma_{ab} \, dX^a \, dX^b.\]

  • Here, \(\langle d\kappa^2 \rangle\) incorporates the effects of the kime-phase distribution \(\Phi(\varphi)\) on the metric.
  • If \(\Phi(\varphi)\) is a symmetric distribution around \(\varphi = 0\), then \(\langle e^{2i\varphi} \rangle\) might be zero, leading to a simplified form

\[\langle d\kappa^2 \rangle = \left(\langle \cos(2\varphi) \rangle + i \langle \sin(2\varphi) \rangle\right) dt^2 - t^2 \langle \cos(2\varphi) \rangle d\varphi^2.\] - For distributions like the Laplace distribution centered at \(\varphi = 0\), the terms involving \(\sin(2\varphi)\) may cancel out, and the real part of \(\langle d\kappa^2 \rangle\) contributes to the metric.

\(\langle d\kappa^2 \rangle\) and Time Dilation: The term \(\langle d\kappa^2 \rangle\) can be interpreted as a generalized time dilation effect, where the time interval is modified by the phase distribution \(\Phi(\varphi)\). The resulting expression contributes to the spacetime geometry, effectively altering the perception of time based on the underlying phase distribution.

Impact on Spacetime Curvature: The kime-phase distribution \(\Phi(\varphi)\) can lead to modifications in the effective curvature of spacetime as represented by the spacekime metric tensor. These modifications can be seen as additional contributions to the metric, which could have physical interpretations in quantum gravity scenarios.

Quantum Fluctuations: The distribution \(\Phi(\varphi)\) represents quantum fluctuations in the phase, and its effect on the metric could be linked to phenomena like quantum fluctuations in spacetime curvature or other quantum gravitational effects.

7.9.2 Challenge!!

The major problem with this formulation of the field equations (and their corresponding solutions) is that complex time, \(\kappa\), and the kime-phase, \(\varphi\) are entangled! As \(\kappa = t\cdot \exp(i\varphi)\), and \(\varphi\sim\Phi\) is a distribution, not an independent variable, so may be the Wirtinger derivative approach is not particularly viable.

The entanglement of \(\kappa\) and \(\varphi\) complicates the situation, as \(\varphi\) is part of \(\kappa\). The Wirtinger derivatives are particularly useful when treating complex variables as independent, but in this case, \(\varphi\) (which modulates \(\kappa\)) is a stochastic or distributional entity rather than a simple independent variable. This complicates the direct application of Wirtinger derivatives because they assume a well-defined complex structure where both parts of \(\kappa\) can be treated independently.

Given the entanglement between \(t\) and \(\varphi\), it may be more appropriate to explore alternative methods of handling the complex-time formalism within the spacekime metric. Here are a few possible approaches:

7.9.2.1 Treating \(\varphi\) as a Parametric Distribution

Instead of directly differentiating with respect to \(\kappa\), one could treat \(\varphi\) as a parametric distribution and use this to inform the behavior of \(\kappa\). The differential operators would then need to account for the distributional nature of \(\varphi\) indirectly influencing \(\kappa\), focusing on how changes in \(\varphi\) impact the dynamics of the system.

7.9.2.2 Recasting the Problem in Terms of Real and Imaginary Components Separately

Another approach is to recast the complex-time variable \(\kappa\) explicitly in terms of its real and imaginary components \(\kappa = t \cos(\varphi) + i t \sin(\varphi)\).

The field equations could then be formulated by treating these real and imaginary parts separately. However, this approach must still handle \(\varphi\) as a stochastic process or distribution rather than a straightforward variable, leading to a mixed differential equation system.

7.9.2.3 Distributional Derivatives for \(\varphi\) Alone

Focus might be placed on distributional derivatives with respect to \(\varphi\) while treating the time evolution \(t\) in a more traditional manner. This would sidestep the complications of entangling \(t\) and \(\varphi\) within a single derivative operator, instead treating the time and phase evolution as interdependent but separately derivable processes.

7.9.2.4 Path Integral Formulation

Another advanced method would be using a path integral formulation where \(\varphi\) represents paths in a phase space. This method can encapsulate the stochastic nature of \(\varphi\) and its influence on \(\kappa\) without needing explicit differential formulations. Instead, this approach would integrate over all possible paths or distributions of \(\varphi\), weighting them according to \(\Phi(\varphi)\).

The entanglement of \(\kappa\) and \(\varphi\) indeed complicates the use of Wirtinger derivatives in a straightforward manner. To address this, rethinking the field equations to account for the distributional nature of \(\varphi\) without directly applying complex derivatives may be necessary. Approaches that either separate real and imaginary components or use more advanced methods like path integrals may offer a more consistent and physically meaningful spacekime formulation.

8 Einstein’s field equations over spacekime

We may explore extending Einstein’s field equations from spacetime to a 5D spacekime framework. The key lies in understanding how curvature and stress-energy tensors generalize in this extended space.

8.1 Background - Field Equations in 4D Spacetime

Einstein’s field equations in 4D spacetime are given by:

\[G_{\mu\nu} = R_{\mu\nu} - \frac{1}{2} g_{\mu\nu}R = 8\pi G T_{\mu\nu}\ ,\]

where:

  • \(G_{\mu\nu}\) is the Einstein tensor,
  • \(R_{\mu\nu}\) is the Ricci curvature tensor,
  • \(R\) is the Ricci scalar,
  • \(g_{\mu\nu}\) is the spacetime metric,
  • \(T_{\mu\nu}\) is the stress-energy tensor,
  • \(G\) is Newton’s gravitational constant.

In spacekime, we need to introduce an additional dimension corresponding to the kime parameter \(\kappa = t \cdot e^{i\varphi}\). The metric in this 5D manifold may be expressed as

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, d\kappa^2 + \tilde{\beta}_a \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]

where \(A, B\) run from 1 to 5, and \(X^a\) (with \(a, b = 1, 2, 3, 4\)) are the spacetime coordinates.

In spacekime, the 5D Ricci curvature tensor \(R_{AB}\) will include contributions from both the spatial dimensions and the kime dimension. The Ricci scalar \(R^{(5)}\) in 5D spacekime may be expressed as

\[R^{(5)} = R^{(4)} + R_{\kappa\kappa} + R_{\kappa a} + R_{ab}\ , \]

where \(R^{(4)}\) is the 4D Ricci scalar for the spacetime submanifold, \(R_{\kappa\kappa}\) represents the curvature due to the kime dimension, and \(R_{\kappa a}\), \(R_{ab}\) represent mixed terms involving both kime and space.

Similarly, the stress-energy tensor in 5D space-kime, \(T_{AB}\), may include additional components that account for the kime dimension

\[T_{AB} = \begin{pmatrix} T_{\kappa\kappa} & T_{\kappa a} \\ T_{a\kappa} & T_{ab} \end{pmatrix}\ , \]

where \(T_{\kappa\kappa}\) and \(T_{\kappa a}\) capture the energy-momentum contributions from the kime component, while \(T_{ab}\) is the usual 4D stress-energy tensor.

Hence, the generalized Einstein’s field equations in 5D space-kime may be written as:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G T_{AB} \ ,\]

where

  • \(R_{\kappa\kappa}\): Governs the dynamics of the kime component, potentially introducing new equations that describe the evolution of the kime parameter \(\kappa\).
  • \(R_{\kappa a}\): Represents the coupling between the kime dimension and the usual 4D spacetime, suggesting new interactions or correlations between these dimensions.
  • \(R_{ab}\): Modifies the usual 4D Einstein equations to include additional curvature terms from the kime dimension.

Non-Static Kime Component: The introduction of \(\kappa\) suggests that spacetime itself might have dynamic properties linked to the complex phase \(\varphi\), leading to potential new physical phenomena such as phase transitions in spacetime or complex oscillatory behaviors in the curvature.

Quantum Gravity Connections: Since the kime parameter could encapsulate quantum mechanical phase information, these generalized field equations might provide a bridge to quantum gravity by describing how quantum states of spacetime evolve in a probabilistic or wave-like manner across kime.

In the limit, as the phase distribution tends to Dirac-delta function, \(\Phi \overset{d}{\longrightarrow} \delta_o\), when the kime-phase dimension becomes negligible, the 5D field equations should reduce to the standard 4D Einstein equations. Mathematically, this corresponds to setting \(G_{\kappa\kappa} = 0\), \(G_{\kappa a} = 0\), and simplifying \(G_{ab}\) to obtain

\[R_{\mu\nu} - \frac{1}{2} g_{\mu\nu}R = 8\pi G T_{\mu\nu}.\]

An extension of Einstein’s field equations into a 5D spacekime framework may provide a new mechanism for exploring the chance-dynamics of the universe. Such generalization offers potential connections to quantum gravity and may reveal new insights into the fundamental nature of time, space, and quantum states.

8.2 Deriving Field Equations in Spacekime using a Kime-Phase Prior Distribution

Alternatively, we can try to derive the spacekime field equations incorporating a prior distribution for the kime-phase, \(\Phi_{[-\pi,\pi)}(\varphi)\), By introducing a probabilistic framework into the formulation of the Einstein field equations, and by using generalized functions, the kime-phase \(\varphi\) may be treated not as a free variable but as a stochastic parameter with a specific prior distribution, \(\Phi_{[-\pi,\pi)}(\varphi)\).

In spacekime, the time dimension is extended to a complex-time variable \(\kappa = t \cdot e^{i\varphi}\), with \(\varphi\sim \Phi_{[-\pi,\pi)}(\varphi)\) as the kime-phase. The prior distribution \(\Phi\) reflects our knowledge (or assumptions) about the distribution of repeated experiment from a highly controlled experiment that yields the observed measurements.

Generalized functions, also known as distributions, are used to handle functions that may not be well-defined in the traditional sense but are useful in describing physical phenomena, such as Dirac delta functions in quantum mechanics.

In this case, the prior distribution \(\Phi(\varphi)\) can be thought of as a generalized function that weighs different kime-phase values according to the probability associated with observing quantum fluctuations that drive the observed variation in multiple repeated experiments.

Incorporating the prior distribution into the spacekime metric, we may derive

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, \Phi(\varphi) \, d\kappa^2 + \tilde{\beta}_a \, \Phi(\varphi) \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\]

where \(\tilde{N}\) and \(\tilde{\beta}_a\) now carry the weight of \(\Phi(\varphi)\), integrating the kime-phase prior into the spacetime fabric.

8.3 Deriving the Einstein Field Equations with \(\Phi(\varphi)\)

Let’s start with the Einstein-Hilbert action generalized to spacekime

\[S = \int d^5X \sqrt{-G} \, R^{(5)} + \int d^5X \sqrt{-G} \, \mathcal{L}_m ,\]

where \(R^{(5)}\) is the Ricci scalar in 5D space-kime and \(\mathcal{L}_m\) is the matter Lagrangian. The metric determinant \(G\) and Ricci scalar now depend on \(\Phi(\varphi)\).

To include the kime-phase prior \(\Phi(\varphi)\), the Ricci scalar \(R^{(5)}\) and the action integral should be modified

\[S = \int d^4x \, d\varphi \, \Phi(\varphi) \left[ \sqrt{-G} \, R^{(5)} + \sqrt{-G} \, \mathcal{L}_m \right]\]

Applying the variational principle to this action, we derive the field equations by varying the action with respect to the generalized metric \(G_{AB}\)

\[\delta S = \int d^4x \, d\varphi \, \Phi(\varphi) \left[ \delta (\sqrt{-G} R^{(5)}) + \delta (\sqrt{-G} \mathcal{L}_m) \right] = 0\]

Expanding and simplifying

\[\int d^4x \, d\varphi \, \Phi(\varphi) \, \sqrt{-G} \left[ R_{AB} - \frac{1}{2} G_{AB} R^{(5)} - 8\pi G T_{AB} \right] \delta G_{AB} = 0 . \]

This leads to the generalized field equations:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\]

Interpretation and Implications:

  • Kime-Phase Dynamics: The additional term \(\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\) introduces a correction to the field equations that depends on the distribution of the kime-phase. This term could reflect additional dynamics in the space-kime framework, influencing how spacetime curvature responds to quantum phase fluctuations.

  • Probabilistic Interpretation: The inclusion of \(\Phi(\varphi)\) implies that the equations of motion and resulting spacetime curvature are not deterministic but are weighted by the underlying probability distribution of the kime-phase.

  • New Physical Insights: This approach provides a bridge between classical general relativity and quantum mechanics, where the kime-phase could capture quantum coherence effects or probabilistic behaviors that manifest in spacetime geometry.

Using a kime-phase distribution prior to derive the field equations in spacekime, we incorporate probabilistic aspects directly into the fabric of spacetime. This generalized approach offers new avenues for understanding quantum gravity and the role of time as a complex variable. Further research is needed to explore the physical implications of these equations and their potential applications in cosmology and quantum gravity.

8.4 Solving the Field Equations in Spacekime

To solve the field equations in spacekime, we start with the generalized Einstein field equations derived earlier:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\ , \]

where \(A, B\) run over the 5D space-kime indices, \(G_{AB}\) is the generalized metric, and \(\Phi(\varphi)\) is the prior distribution for the kime-phase.

In general relativity, the Einstein field equations are a set of \(10\) coupled, non-linear partial differential equations for the components of the metric tensor \(g_{\mu\nu}\). In spacekime, we extend this to a 5D system, with additional degrees of freedom associated with the kime-phase \(\varphi\).

This extension leads to an underdetermined system because we now have more unknowns (additional components in the metric \(G_{AB}\) and the phase distribution \(\Phi(\varphi)\)) than equations. To make this system solvable, we must introduce constraints or gauge conditions.

Several constraints, or gauge conditions, can be introduced to reduce the complexity of the system and make it solvable:

8.4.1 Kime-Orthogonality Condition

For instance, consider imposing a constraint that decouples the kime component from the spatial and temporal components. This is akin to an orthogonality condition, where the kime-phase is orthogonal to the 4D spacetime submanifolds:

\[G_{A\varphi} = 0 \quad \text{for all } A \neq \varphi\ .\]

This condition simplifies the system by removing mixed terms involving the kime-phase and spacetime coordinates.

8.4.2 Kime-Slice Conditions

Similar to the ADM formalism’s choice of a slicing condition (e.g., assuming a constant mean curvature), we can impose a kime-slice condition. For example, we can choose \(\partial_{\varphi} G_{AB} = 0.\)

This condition assumes that the metric does not explicitly depend on the kime-phase \(\varphi\), effectively treating \(\varphi\) as a parameter rather than a dynamic variable.

8.4.3 Gauge Fixing

We can also fix a gauge by choosing a specific form for the lapse function \(\tilde{N}\) and the shift vector \(\tilde{\beta}_a\). For instance, a common choice is to set \(\tilde{N} = 1\) and \(\tilde{\beta}_a = 0\), reducing the metric to a simpler form.

8.5 Derivation of Field Equation Solutions

Given these constraints, the field equations reduce to a more manageable form. Let’s consider the kime-slice condition \(\partial_{\varphi} G_{AB} = 0\), which simplifies the equations to:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB}\ .\]

Here, the Ricci tensor \(R_{AB}\) and the Ricci scalar \(R^{(5)}\) are now functions of the spatial coordinates and the complex-time parameter \(\kappa = t \cdot e^{i\varphi}\), but with no explicit dependence on \(\varphi\).

8.5.1 Specify a Metric Ansatz

Choose an appropriate ansatz for the 5D metric \(G_{AB}\), such as a generalized Schwarzschild or FLRW metric extended to include the kime dimension.

Example ansatz:

\[G_{AB} = \begin{pmatrix} -f(r) & 0 & 0 & 0 & 0 \\ 0 & \frac{1}{f(r)} & 0 & 0 & 0 \\ 0 & 0 & r^2 & 0 & 0 \\ 0 & 0 & 0 & r^2 \sin^2\theta & 0 \\ 0 & 0 & 0 & 0 & g_{\varphi\varphi}(\kappa) \end{pmatrix}\]

8.5.2 Substitute into the Field Equations

Substitute the metric ansatz into the field equations. With the kime-slice condition, this will yield a system of differential equations in the remaining variables.

Solve the resulting equations, typically using separation of variables or alternative numerical methods, depending on the complexity of the system.

8.5.3 Apply Boundary Conditions

To ensure a physically meaningful solution, impose boundary conditions, such as asymptotic flatness at infinity or regularity at a black hole horizon.

Analyze the solution in terms of its physical implications, such as the nature of singularities, horizons, or the evolution of the kime-phase over time.

8.5.4 Simple Case (Example Solution)

For a simple case with a flat 4D spacetime and a constant kime metric component:

\[G_{AB} = \text{diag}(-1, 1, 1, 1, \Phi(\varphi))\]

The field equations reduce to:

\[R_{\varphi\varphi} = 8\pi G \, T_{\varphi\varphi}\]

If \(T_{\varphi\varphi} = 0\) (vacuum case), this simplifies to \(R_{\varphi\varphi} = 0\).

This implies that the kime-phase distribution \(\Phi(\varphi)\) must be such that it contributes no additional curvature to the 5D spacetime, potentially leading to a flat or constant kime component.

8.5.5 Non-Trivial Field Equation Solution in Spacekime

Let’s consider another specific example where we derive a non-trivial solution to the Einstein field equations in spacekime. We will focus on a scenario where the 5D metric is non-trivially dependent on the kime-phase \(\varphi\) and the spatial coordinates.

Again, starting with the generalized Einstein field equations in spacekime

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\ , \]

we’ll assume for simplicity a vacuum solution where \(T_{AB} = 0\). Therefore, the field equations reduce to:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\ . \]

Consider a metric ansatz where the spacetime part is spherically symmetric (Schwarzschild-like) and the kime component has a non-trivial dependence on \(\varphi\)

\[G_{AB} = \begin{pmatrix} -f(r) & 0 & 0 & 0 & 0 \\ 0 & \frac{1}{f(r)} & 0 & 0 & 0 \\ 0 & 0 & r^2 & 0 & 0 \\ 0 & 0 & 0 & r^2 \sin^2\theta & 0 \\ 0 & 0 & 0 & 0 & h(\varphi) \end{pmatrix}\ ,\]

where \(f(r)\) is the radial function associated with the Schwarzschild metric, and \(h(\varphi)\) is a function to be determined that encodes the kime-phase dependence.

Next, compute the components of the Ricci tensor \(R_{AB}\) and the Ricci scalar \(R^{(5)}\) for this metric:

  • The spatial part gives the usual Schwarzschild contributions to the Ricci tensor.
  • The kime part introduces additional terms due to the dependence of \(h(\varphi)\) on \(\varphi\).

Given that the Schwarzschild metric satisfies the vacuum Einstein equations in 4D, the non-zero contributions to \(R_{AB}\) will come from the kime-phase dependence:

\[R_{\varphi\varphi} = -\frac{h''(\varphi)}{2h(\varphi)} + \frac{h'(\varphi)^2}{4h(\varphi)^2}\ ,\]

where the prime denotes differentiation with respect to \(\varphi\).

The 5D Ricci scalar is then:

\[R^{(5)} = \frac{R^{(4)}}{h(\varphi)} + \frac{h''(\varphi)}{2h(\varphi)} - \frac{h'(\varphi)^2}{4h(\varphi)^2}\ ,\]

where \(R^{(4)}\) is the 4D Ricci scalar for the Schwarzschild spacetime, which is zero in the vacuum case.

Substituting the expressions for \(R_{AB}\) and \(R^{(5)}\) into the field equations, we focus on the equation for the \(\varphi\varphi\) component:

\[-\frac{h''(\varphi)}{2h(\varphi)} + \frac{h'(\varphi)^2}{4h(\varphi)^2} - \frac{1}{2} \left(\frac{h''(\varphi)}{2h(\varphi)} - \frac{h'(\varphi)^2}{4h(\varphi)^2}\right) = \frac{1}{\Phi(\varphi)} \nabla_\varphi \nabla_\varphi \Phi(\varphi)\ .\]

Simplifying, this reduces to a differential equation for \(h(\varphi)\)

\[h''(\varphi) - \frac{h'(\varphi)^2}{2h(\varphi)} = 2 \frac{\Phi'(\varphi)^2}{\Phi(\varphi)}\ .\]

This equation can be solved by making an appropriate ansatz for \(h(\varphi)\) and \(\Phi(\varphi)\). One common approach is to assume a power-law form:

\[h(\varphi) = h_0 e^{\alpha \varphi}, \quad \Phi(\varphi) = \Phi_0 e^{\beta \varphi}\ .\]

Substituting these into the equation \(\alpha^2 = 2 \beta^2\). This relationship imposes a constraint on the values of \(\alpha\) and \(\beta\), which must satisfy \(\alpha = \sqrt{2} \beta\).

The solution obtained implies that the kime-phase dependence is exponential, leading to an effective “stretching” or “compression” of the kime dimension as a function of \(\varphi\). This non-trivial solution suggests that the geometry of spacekime is dynamically influenced by the kime-phase distribution, which has potential implications for the nature of time and quantum gravity.

The derived non-trivial solution in space-kime shows that the field equations allow for complex and dynamic behavior of the kime dimension, which is intricately linked to the probabilistic structure of the kime-phase. This approach opens the door to further investigations into the nature of spacetime and its extensions in quantum gravity theories, providing a rich framework for understanding the universe beyond the classical picture of space and time.

Imposing specific constraints and using a generalized function approach with a kime-phase prior distribution \(\Phi(\varphi)\) may lead to solving the field equations in spacekime. These constraints simplify the system, making it solvable either analytically or numerically, depending on the complexity of the chosen metric ansatz. Further exploration of these solutions could provide insights into the role of complex-time in quantum gravity, computational statistics, data science, and artificial intelligence.

8.5.6 Another Non-Trivial Solution with Kime-Phase as a Probability Distribution

In this scenario, we treat the kime-phase \(\varphi\) not as a directly observable quantity, but rather as a probability distribution that reflects the variability seen in repeated experiments. This approach leads to a non-trivial solution of the field equations where \(\varphi\) encapsulates the statistical nature of time evolution in a quantum gravity framework.

Again, start with the generalized Einstein field equations in space-kime:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\ .\]

Here, \(\Phi(\varphi)\) is a probability distribution function that reflects the distribution of the kime-phase across repeated experimental measurements.

Let \(\varphi\) be distributed according to a Gaussian distribution centered around some mean \(\varphi_0\) with variance \(\sigma^2\), \(\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} e^{-\frac{(\varphi - \varphi_0)^2}{2\sigma^2}}.\)

This distribution represents the variability in \(\varphi\) seen across different trials or observations, reflecting the inherent uncertainty or noise in the system.

To reflect the influence of this probabilistic kime-phase, we modify the metric \(G_{AB}\) to include \(\Phi(\varphi)\):

\[G_{AB} = \begin{pmatrix} -f(r, \varphi) & 0 & 0 & 0 & 0 \\ 0 & \frac{1}{f(r, \varphi)} & 0 & 0 & 0 \\ 0 & 0 & r^2 & 0 & 0 \\ 0 & 0 & 0 & r^2 \sin^2\theta & 0 \\ 0 & 0 & 0 & 0 & g_{\varphi\varphi}(\varphi) \end{pmatrix}\ ,\]

where the function \(f(r, \varphi)\) and the metric component \(g_{\varphi\varphi}(\varphi)\) depend on the kime-phase \(\varphi\) through the distribution \(\Phi(\varphi)\).

Incorporating \(\Phi(\varphi)\) into the Einstein field equations, we focus on the \(\varphi\varphi\) component:

\[R_{\varphi\varphi} - \frac{1}{2} G_{\varphi\varphi} R^{(5)} = \frac{1}{\Phi(\varphi)} \nabla_\varphi \nabla_\varphi \Phi(\varphi)\ .\]

Given the Gaussian form of \(\Phi(\varphi)\), compute the derivatives

\[\nabla_\varphi \nabla_\varphi \Phi(\varphi) = \left(\frac{(\varphi - \varphi_0)^2}{\sigma^4} - \frac{1}{\sigma^2}\right)\Phi(\varphi)\ .\]

Substituting this into the field equations, we obtain:

\[R_{\varphi\varphi} - \frac{1}{2} G_{\varphi\varphi} R^{(5)} = \left(\frac{(\varphi - \varphi_0)^2}{\sigma^4} - \frac{1}{\sigma^2}\right)\ .\]

This equation governs the curvature of the kime dimension as influenced by the probabilistic nature of \(\varphi\).

To solve for the metric component \(g_{\varphi\varphi}(\varphi)\), again assume a power-law or exponential dependence \(g_{\varphi\varphi}(\varphi) = g_0 e^{\lambda \varphi}\).

Substituting this into the field equation and solving, we find

\[\lambda^2 = \frac{2}{\sigma^2} \quad \text{and} \quad \lambda(\varphi - \varphi_0) = \frac{(\varphi - \varphi_0)^2}{\sigma^4} - \frac{1}{\sigma^2}\ .\]

This equation implies a relationship between the kime-phase distribution’s variance \(\sigma^2\) and the exponential growth or decay rate of the kime-phase metric component.

A physical interpretation of this solution suggests that the geometry of space-kime, specifically the curvature in the kime dimension, is directly influenced by the statistical distribution of \(\varphi\). In physical terms, this could represent the impact of quantum fluctuations or measurement uncertainty on the structure of spacetime.

Treating the kime-phase as a probabilistic distribution leads to non-trivial solutions of the Einstein field equations in space-kime. The resulting metric solutions indicate a strong connection between the probabilistic nature of time (or kime) and the geometry of spacetime/spacekime. Potentially, this may offer new insights into quantum gravity, the nature of the longitudinal event ordering (time), and (statistical) AI analytics.

8.6 Alternative Approach for Extending Einstein’s Field Equations to Spacekime

This alterantive approach for extending the spacekime metric tensor, \(G_{AB}\), and Einstein’s field equations from 4D spacetime to 5D spacekime incorporates incorporates the contributions from the kime-phase distribution \(\Phi(\varphi)\). The goal is to generalize the field equations to account for the extra temporal dimension and the complex-time structure.

The spacekime metric tensor in the 5D spacekime framework is given by

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, \langle d\kappa^2 \rangle + \tilde{\beta}_a \, \langle d\kappa \rangle \, dX^a + \gamma_{ab} \, dX^a \, dX^b,\]

where:

  • \(\tilde{N}\) is the lapse function.
  • \(\tilde{\beta}_a\) are the shift vectors.
  • \(\gamma_{ab}\) is the 3D spatial metric.
  • \(\langle d\kappa^2 \rangle\) accounts for the complex-time \(\kappa = t e^{i\varphi}\) under the kime-phase distribution \(\Phi(\varphi)\).

In 4D spacetime, Einstein’s field equations are given by:

\[R_{\mu\nu} - \frac{1}{2} R g_{\mu\nu} + \Lambda g_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}.\]

where

  • \(R_{\mu\nu}\) is the Ricci curvature tensor.
  • \(R\) is the Ricci scalar.
  • \(g_{\mu\nu}\) is the spacetime metric.
  • \(\Lambda\) is the cosmological constant.
  • \(T_{\mu\nu}\) is the energy-momentum tensor.

To generalize to 5D spacekime, we replace the 4D spacetime metric \(g_{\mu\nu}\) with the 5D spacekime metric \(G_{AB}\), where \(A, B\) run over the 5 dimensions (3 spatial, 2 kime dimensions). The general form of the 5D Einstein field equations in spacekime is

\[R_{AB} - \frac{1}{2} R G_{AB} + \Lambda G_{AB} = \frac{8\pi G}{c^4} T_{AB}.\]

Let’s break down the components

Ricci Tensor: \(R_{AB}\) is computed from the 5D spacekime metric tensor \(G_{AB}\)

\[R_{AB} = \partial_C \Gamma^C_{AB} - \partial_B \Gamma^C_{AC} + \Gamma^C_{CD} \Gamma^D_{AB} - \Gamma^C_{AD} \Gamma^D_{BC},\]

where \(\Gamma^C_{AB}\) are the Christoffel symbols calculated from \(G_{AB}\).

Ricci Scalar: \(R\) is obtained by contracting the Ricci tensor \(R = G^{AB} R_{AB}.\)

Energy-Momentum Tensor: \(T_{AB}\) in 5D could include contributions from both the traditional energy-momentum in 4D spacetime and additional terms accounting for the dynamics in the kime dimensions

\[T_{AB} = \text{Standard 4D energy-momentum tensor} + \text{Kime contributions}.\]

The kime-phase distribution \(\Phi(\varphi)\) modifies the metric tensor by contributing to the term \(\langle d\kappa^2 \rangle\)

\[\langle d\kappa^2 \rangle = \int \left( e^{i\varphi} \, dt + i t e^{i\varphi} \, d\varphi \right)^2 \Phi(\varphi) \, d\varphi .\]

This modification reflects how the kime-phase distribution affects spacetime curvature. We include the effects of this distribution in the computation of the Christoffel symbols, the Ricci tensor, and ultimately the Einstein field equations.

Boundary Conditions and Constraints: To solve the field equations in spacekime, we must impose appropriate boundary conditions, such as:

  • Spatial Boundary Conditions: Constraints on the spatial part of the metric \(\gamma_{ab}\) at infinity or on a compact spatial region.
  • Kime Boundary Conditions: Constraints on the kime-phase distribution, such as normalization or boundary values for \(\Phi(\varphi)\).
  • Continuity and Smoothness: Ensure that \(G_{AB}\) and its derivatives are continuous and smooth across the spacekime manifold.

Example Solutions:

  • Static Solutions: For static scenarios, where the metric does not depend on the kime-phase, one might obtain solutions similar to Schwarzschild or Kerr metrics extended to 5D, modified by the presence of \(\Phi(\varphi)\).
  • Dynamic Solutions: In dynamic scenarios, where the metric evolves with \(\varphi\) and \(t\), solutions might involve oscillatory or wave-like behaviors in the spacekime dimensions, reflecting quantum fluctuations.
  • Interpretation: These solutions can be interpreted as representing different quantum states of the universe, where time and space are emergent from more fundamental kime-phase dynamics.

8.6.1 Deriving the Spacekime Einstein Field Equations

To derive the spacekime Einstein Field Equations, we begin by generalizing the Einstein-Hilbert action to the 5D spacekime framework and then vary this action to obtain the field equations. The kime-phase distribution \(\Phi(\varphi)\) will play a crucial role in modifying the action and the resulting field equations.

In 4D spacetime, the Einstein-Hilbert action is given by

\[S_{\text{EH}} = \frac{1}{16\pi G} \int \sqrt{-g} \, R \, d^4x,\]

where

  • \(g\) is the determinant of the 4D spacetime metric \(g_{\mu\nu}\).
  • \(R\) is the Ricci scalar.
  • \(G\) is the gravitational constant.

To generalize this to the 5D spacekime framework, we replace the 4D metric \(g_{\mu\nu}\) with the 5D spacekime metric \(G_{AB}\), and integrate over the 5D spacekime volume

\[S_{\text{EH}}^\kappa = \frac{1}{16\pi G} \int \sqrt{-G} \, R^{(5)} \, d^5X,\]

where

  • \(G\) is the determinant of the 5D spacekime metric \(G_{AB}\).
  • \(R^{(5)}\) is the 5D Ricci scalar derived from \(G_{AB}\).
  • \(d^5X\) represents the volume element in 5D spacekime.

The kime-phase distribution \(\Phi(\varphi)\) modifies the metric components involving the complex-time dimension \(\kappa\). Specifically, the differential element \(d\kappa^2\) in the metric is modified by \(\Phi(\varphi)\)

\[\langle d\kappa^2 \rangle = \int \left( e^{i\varphi} \, dt + i t e^{i\varphi} \, d\varphi \right)^2 \Phi(\varphi) \, d\varphi.\]

This affects the metric determinant \(G\) and the Ricci scalar \(R^{(5)}\).

Including the kime-phase distribution in the action, we have

\[S_{\text{EH}}^\kappa = \frac{1}{16\pi G} \int \Phi(\varphi) \sqrt{-G} \, R^{(5)} \, d^5X,\]

where, \(\Phi(\varphi)\) effectively weights the contribution of different kime-phases to the action.

To derive the field equations, we vary the action \(S_{\text{EH}}^\kappa\) with respect to the spacekime metric \(G_{AB}\)

\[\delta S_{\text{EH}}^\kappa = \frac{1}{16\pi G} \int \left[ \delta(\Phi(\varphi) \sqrt{-G}) R^{(5)} + \Phi(\varphi) \sqrt{-G} \, \delta R^{(5)} \right] d^5X = 0.\]

The first term involves the variation of the metric determinant and Ricci scalar

\[\delta(\Phi(\varphi) \sqrt{-G}) R^{(5)} = \Phi(\varphi) \delta(\sqrt{-G}) R^{(5)} + \sqrt{-G} R^{(5)} \delta \Phi(\varphi).\]

The second term, Variation of Ricci Scalar \(\delta R^{(5)}\), is given by

\[\delta R^{(5)} = R^{(5)}_{AB} \delta G^{AB} - \nabla_A \nabla_B \delta G^{AB}.\]

After performing the variation and simplifying, the resulting spacekime Einstein field equations can be written as

\[R^{(5)}_{AB} - \frac{1}{2} G_{AB} R^{(5)} + \Lambda G_{AB} = \frac{8\pi G}{c^4} T^{(5)}_{AB} + \mathcal{F}_{AB},\]

where

  • \(R^{(5)}_{AB}\) is the 5D Ricci tensor.
  • \(R^{(5)}\) is the 5D Ricci scalar.
  • \(\Lambda\) is the cosmological constant.
  • \(T^{(5)}_{AB}\) is the 5D energy-momentum tensor, including contributions from matter and kime components.
  • \(\mathcal{F}_{AB}\) are additional terms arising from the variation of \(\Phi(\varphi)\), reflecting the influence of the kime-phase distribution.

To solve these equations, we impose various conditions:

  • Boundary conditions: on the spatial and kime components.
  • Constraints: on the kime-phase distribution \(\Phi(\varphi)\) to ensure the metric remains well-behaved and physically meaningful.

Notes:

  • Quantum Fluctuations: The additional terms \(\mathcal{F}_{AB}\) can be interpreted as arising from quantum fluctuations encoded in the kime-phase distribution \(\Phi(\varphi)\).
  • Emergent Time: The modified field equations suggest that time and gravitational dynamics are intertwined with the kime-phase, leading to possible new gravitational phenomena or corrections at quantum scales.
  • Cosmological Implications: These equations could lead to novel cosmological solutions where the kime-phase distribution plays a role in the evolution of the universe, possibly affecting the behavior of dark energy or other cosmological constants.

8.6.2 Exampe of a Simple Solution of the Spacekime Fields Equation under Laplace Phase Prior

To derive and explicitly write out a simplified solution to the spacekime field equations, we need to consider the 5D spacekime metric \(G_{AB}\), where \(A, B\) run over the 5 dimensions: three spatial dimensions \((x^1, x^2, x^3)\) and two kime dimensions \((t, \varphi)\).

The spacekime metric is expressed as:

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, \langle d\kappa^2 \rangle + \tilde{\beta}_a \, \langle d\kappa \rangle \, dX^a + \gamma_{ab} \, dX^a \, dX^b,\]

where

  • \(\tilde{N}\) is the lapse function.
  • \(\tilde{\beta}_a\) are the shift vectors.
  • \(\gamma_{ab}\) is the 3D spatial metric.
  • \(\langle d\kappa^2 \rangle\) and \(\langle d\kappa \rangle\) account for the kime-phase distribution \(\Phi(\varphi)\).

The generalized Einstein field equations in spacekime are

\[R_{AB} - \frac{1}{2} G_{AB} R + \Lambda G_{AB} = \frac{8\pi G}{c^4} T_{AB} + \mathcal{F}_{AB},\]

where

  • \(R_{AB}\) is the 5D Ricci tensor.
  • \(R\) is the 5D Ricci scalar.
  • \(\Lambda\) is the cosmological constant.
  • \(T_{AB}\) is the energy-momentum tensor.
  • \(\mathcal{F}_{AB}\) are terms arising from the kime-phase distribution \(\Phi(\varphi)\).

In 5D spacekime, we consider the indices \(A, B\) running over the five dimensions \((t, \varphi, x^1, x^2, x^3)\). This gives us 10 independent components in the symmetric metric tensor \(G_{AB}\), leading to 10 independent field equations

  1. \(G_{tt}\) Equation: \[R_{tt} - \frac{1}{2} G_{tt} R + \Lambda G_{tt} = \frac{8\pi G}{c^4} T_{tt} + \mathcal{F}_{tt}.\]

  2. \(G_{t\varphi}\) Equation:

\[R_{t\varphi} - \frac{1}{2} G_{t\varphi} R + \Lambda G_{t\varphi} = \frac{8\pi G}{c^4} T_{t\varphi} + \mathcal{F}_{t\varphi}.\]

  1. \(G_{t x^1}\) Equation:

\[R_{t x^1} - \frac{1}{2} G_{t x^1} R + \Lambda G_{t x^1} = \frac{8\pi G}{c^4} T_{t x^1} + \mathcal{F}_{t x^1}.\]

  1. \(G_{t x^2}\) Equation:

\[R_{t x^2} - \frac{1}{2} G_{t x^2} R + \Lambda G_{t x^2} = \frac{8\pi G}{c^4} T_{t x^2} + \mathcal{F}_{t x^2}.\]

  1. \(G_{t x^3}\) Equation:

\[R_{t x^3} - \frac{1}{2} G_{t x^3} R + \Lambda G_{t x^3} = \frac{8\pi G}{c^4} T_{t x^3} + \mathcal{F}_{t x^3}.\]

  1. \(G_{\varphi\varphi}\) Equation:

\[R_{\varphi\varphi} - \frac{1}{2} G_{\varphi\varphi} R + \Lambda G_{\varphi\varphi} = \frac{8\pi G}{c^4} T_{\varphi\varphi} + \mathcal{F}_{\varphi\varphi}.\]

  1. \(G_{\varphi x^1}\) Equation:

\[R_{\varphi x^1} - \frac{1}{2} G_{\varphi x^1} R + \Lambda G_{\varphi x^1} = \frac{8\pi G}{c^4} T_{\varphi x^1} + \mathcal{F}_{\varphi x^1}.\]

  1. \(G_{\varphi x^2}\) Equation:

\[R_{\varphi x^2} - \frac{1}{2} G_{\varphi x^2} R + \Lambda G_{\varphi x^2} = \frac{8\pi G}{c^4} T_{\varphi x^2} + \mathcal{F}_{\varphi x^2}.\]

  1. \(G_{\varphi x^3}\) Equation:

\[R_{\varphi x^3} - \frac{1}{2} G_{\varphi x^3} R + \Lambda G_{\varphi x^3} = \frac{8\pi G}{c^4} T_{\varphi x^3} + \mathcal{F}_{\varphi x^3}.\]

  1. \(G_{x^i x^j}\) Equations (where \(i,j = 1, 2, 3\)):

\[R_{x^i x^j} - \frac{1}{2} G_{x^i x^j} R + \Lambda G_{x^i x^j} = \frac{8\pi G}{c^4} T_{x^i x^j} + \mathcal{F}_{x^i x^j}.\]

Each of these equations is a generalization of the standard Einstein field equations to the spacekime framework, where the kime-phase distribution \(\Phi(\varphi)\) modifies the gravitational dynamics.

Example Solution for Laplace Kime-Phase Distribution: For a specific kime-phase distribution, such as the Laplace distribution, \(\Phi(\varphi)\) is given by

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\]

This distribution affects the spacekime metric and the resulting field equations. Let’s consider a static solution where the metric is independent of \(\varphi\) and \(t\).

For simplicity, consider a metric where \(\tilde{N}\) and \(\tilde{\beta}_a\) are constants, and the spatial part \(\gamma_{ab}\) is flat

\[G_{AB} = \begin{pmatrix} -\tilde{N}^2 & 0 & 0 & 0 & 0 \\ 0 & \langle d\kappa^2 \rangle & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 & 1 \end{pmatrix},\]

where, \(\langle d\kappa^2 \rangle\) is determined by the kime-phase distribution

\[\langle d\kappa^2 \rangle = \int_{-\infty}^{\infty} \left[ \left(\cos\varphi\right)^2 dt^2 - t^2 \left(\sin^2\varphi\right) d\varphi^2 \right] \Phi(\varphi) \, d\varphi.\]

Given the Laplace distribution

\[\langle d\kappa^2 \rangle \approx \frac{\tilde{C}}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \left(dt^2 - t^2 d\varphi^2 \right),\]

where \(\tilde{C}\) is a constant arising from the integral.

We substitute these into the generalized field equations

\[R_{AB} - \frac{1}{2} G_{AB} R + \Lambda G_{AB} = 0.\]

This simplifies to the following components, assuming a vacuum solution where \(T_{AB} = 0\).

  • \(G_{tt}\) Component: The \(tt\) component of the Ricci tensor involves derivatives with respect to \(t\) and \(\varphi\), as well as the influence of the kime-phase distribution:

\[R_{tt} - \frac{1}{2} G_{tt} R + \Lambda G_{tt} = 0.\]

Given the simplified metric, this reduces to a form that relates the time-time component of the Ricci tensor to the curvature scalar \(R\).

  • \(G_{\varphi\varphi}\) Component: This component involves the second derivative with respect to \(\varphi\), and since \(\varphi\) is integrated over its distribution, the influence of the Laplace distribution affects the curvature

\[R_{\varphi\varphi} - \frac{1}{2} G_{\varphi\varphi} R + \Lambda G_{\varphi\varphi} = 0.\]

This equation captures how the kime-phase distribution \(\Phi(\varphi)\) modifies the structure of spacetime.

  • \(G_{x^ix^j}\) Components: These components describe the curvature in the spatial dimensions. Since we assumed a flat spatial metric, the Ricci tensor components here would typically vanish unless there are interactions or additional potentials introduced.

\[R_{x^ix^j} - \frac{1}{2} G_{x^ix^j} R + \Lambda G_{x^ix^j} = 0.\]

Let’s consider a simple static scenario where \(\tilde{N}\) is constant and the shift vectors \(\tilde{\beta}_a\) are zero. The spatial part of the metric is assumed to be flat, and the only non-trivial components come from \(t\) and \(\varphi\). We focus on solving for the \(tt\) and \(\varphi\varphi\) components

  • \(G_{tt}\) Component: For the time-time component

\[R_{tt} - \frac{1}{2} G_{tt} R = \frac{\tilde{C}}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \left(\frac{d^2 \tilde{N}}{dt^2}\right) = 0.\]

This suggests that under the Laplace distribution, the time component could be trivial (e.g., \(\tilde{N}\) is constant), or it could reflect specific time dynamics.

  • \(G_{\varphi\varphi}\) Component: For the kime-phase component

\[R_{\varphi\varphi} - \frac{1}{2} G_{\varphi\varphi} R = \frac{\tilde{C}}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right) \left(\frac{d^2 \Phi(\varphi)} {d\varphi^2}\right) = 0.\]

This equation is more complex and reflects how the kime-phase distribution influences the curvature through the second derivative of \(\Phi(\varphi)\).

The solution indicates that the kime-phase distribution significantly impacts the curvature in spacekime. The Laplace distribution introduces exponential decay in \(\varphi\), which could suggest that spacetime behaves differently at different phases, potentially leading to variations in observed physical phenomena.

  • Time Dynamics: If \(\tilde{N}\) is not constant, it suggests the presence of time dilation effects tied to the kime-phase distribution.

  • Kime Dynamics: The \(\varphi\varphi\) component indicates that the phase distribution could lead to anisotropies or phase-dependent effects in spacetime.

9 ADM formalism in Spacekime

Based on the “Dynamical Structure and Definition of Energy in General Relativity” paper, we can explore a spacekime generalization of the ADM formalism and explicate spacekime as foliated into a family of spacetime-like manifolds (hyper-surfaces), over the support of the kime-phase distribution, \([-\pi,\pi)\) by kime, where the spacetime coordinates on each foliation leaf (spacetime slice) are given by \((x^{i}, t)\).

9.1 First Naive Approach

We’ll attempt to generalize the Arnowitt-Deser-Misner (ADM) formalism, see the original ADM paper, to spacekime, where the complexified time (kime) introduces an additional dimension beyond the traditional 4D spacetime. The spacekime manifold is foliated into spacetime-like hypersurfaces (Spacewtime manifolds) indexed by kime, with the spacetime coordinates on each foliation leaf (spacetime slice) given by \((x^{i}, t)\). We can develop the Hamiltonian formulation of general relativity in the 5D spacekime and derive the corresponding field equations, including the effects of a kime-phase distribution \(\Phi(\varphi)\). We explore the implications of this extension for the dynamical structure of spacetime, energy definitions, and potential connections to quantum gravity.

  • Context:
    • Introduction to the ADM formalism in general relativity, emphasizing its importance in splitting spacetime into space and time components.
    • Motivation for generalizing the ADM formalism to include kime, incorporating complex-time into general relativity, and emphasizing importance of the kime-phase representation of repeated sampling in evidence-based discovery by splitting (foliating) spacekime into spacetime leaves and over the support of the kime-phase distribution.
    • Overview of how this extend ADM formalism relates to energy definitions and the structure of spacekime.
  • Spacekime Framework:
    • Explanation of the complex-time (kime) variable \(\kappa = t \cdot e^{i\varphi}\), where \(\varphi\) is the kime-phase.
    • Justification for treating time as a complex variable and its implications for the structure of spacetime.
  • Defining the spacekime Metric \(G_{AB}\):
    • Derivation and explanation of the space-kime metric that extends the standard spacetime metric to 5D.
    • Explicit form: \[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, d\kappa^2 + \tilde{\beta}_a \, d\kappa \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]
    • Introduction of the generalized lapse function \(\tilde{N}(\kappa, \varphi) = N(t) \cdot \Phi(\varphi)\), where \(\Phi(\varphi)\) is the kime-phase distribution.
  • Foliation of Space-Kime Manifold:
    • Detailed discussion on foliating the 5D space-kime manifold into 4D spacetime-like hypersurfaces indexed by kime.
    • Explanation of how the spacetime coordinates \((x^i, t)\) are assigned on each hypersurface.
    • Implications of the foliation for the ADM formalism and the Hamiltonian formulation.

Hamiltonian Formulation in Spacekime:

  • ADM Hamiltonian in General Relativity:
    • Recap of the ADM Hamiltonian in standard 4D spacetime.
    • Description of the canonical variables \((h_{ij}, \pi^{ij})\) and the Hamiltonian constraint equations.
  • Space-Kime Hamiltonian:
    • Extension of the ADM Hamiltonian to space-kime, incorporating the kime-phase distribution \(\Phi(\varphi)\).
    • Modified Hamiltonian density: \[ \mathcal{H} = \mathcal{H}_\text{ADM} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi) \]
    • Derivation of the constraints and evolution equations in the space-kime framework.
    • Discussion on the additional degrees of freedom introduced by the kime-phase.

Einstein Field Equations in Spacekime

  • General Relativity Field Equations:
    • Review of the Einstein field equations in 4D spacetime: \[G_{\mu\nu} + \Lambda g_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}\ .\]
  • Spacekime Field Equations:
    • Derivation of the space-kime field equations, starting from the Einstein-Hilbert action generalized to 5D.
    • Incorporation of the kime-phase distribution \(\Phi(\varphi)\) into the field equations: \[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\ .\]
    • Discussion on the additional terms arising from the distributional derivatives with respect to \(\varphi\).

Energy Definition in Spacekime:

  • ADM Energy in General Relativity:
    • Review of the ADM definition of energy in the context of 4D general relativity.
    • Role of the lapse function \(N(t)\) and shift vector \(N^i\) in defining energy and momentum.
  • Generalizing Energy in Space-Kime:
    • Extension of the ADM energy definition to the space-kime framework.
    • Influence of the kime-phase distribution \(\Phi(\varphi)\) on energy and momentum definitions.
    • Potential implications for the energy conservation laws in the context of quantum gravity and the space-kime framework.

Example Applications:

  • Laplace Distribution for \(\Phi(\varphi)\):
    • Specific example where the kime-phase distribution is modeled by a Laplace distribution.
    • Calculation of the corresponding space-kime metric and its impact on the field equations.
    • Analysis of how the Laplace-distributed kime-phase affects the energy and momentum in space-kime.
  • Comparison with Gaussian Distribution:
    • Comparison of the Laplace distribution with a Gaussian distribution for \(\Phi(\varphi)\).
    • Discussion on how different kime-phase distributions affect the geometry and dynamics of the space-kime manifold.

Physical Implications and Quantum Gravity:

  • Implications for Quantum Gravity:
    • Exploration of how the space-kime generalization may contribute to our understanding of quantum gravity.
    • Potential connections between space-kime and existing quantum gravity theories, such as loop quantum gravity and string theory.
  • Experimental Tests and Predictions:
    • Discussion on potential experimental tests that could validate the space-kime framework.
    • Predictions of the space-kime formalism that differ from standard general relativity and their observational consequences.

This ADM generalization to spacekime provides the technical foundation for extending the original ADM paper in the context of complex time.

9.2 A More Rigorous Reformulation of the ADM Formalism in Spacekime

The following may represent a more reliable formulation of the ADM spacekime formalism using the distributional derivative of the kime-phase distribution and ensuring that the spacekime metric is real.

9.2.1 (Approach 1) Reformulating the ADM Formalism in Spacekime

To reformulate the ADM formalism in spacekime, we must carefully handle the kime-phase \(\varphi\). Since \(\varphi\) is a distribution and not a variable, we must ensure that our approach reflects this fact. Here, we will carefully walk through the reformulation process, taking into account that the kime-phase \(\varphi\) affects the spacekime metric and the associated field equations through its distribution, \(\Phi(\varphi)\).

In the standard ADM formalism, the 4D spacetime metric is decomposed into a 3D spatial hypersurface plus time \(ds^2 = -N^2 dt^2 + \gamma_{ij} (dx^i + \beta^i dt)(dx^j + \beta^j dt).\)

In spacekime, we extend this to 5D by including the kime-phase contribution. The metric can be expressed as

\[G_{AB} dX^A dX^B = -\tilde{N}^2 \langle d\kappa^2 \rangle + \tilde{\beta}_a \langle d\kappa \rangle dX^a + \gamma_{ab} dX^a dX^b,\]

where

  • \(\tilde{N}\) is the generalized lapse function associated with the complex time \(\kappa = t e^{i\varphi}\).
  • \(\tilde{\beta}_a\) is the generalized shift vector.
  • \(\gamma_{ab}\) is the three-metric on the spatial slice.
  • \(\langle d\kappa^2 \rangle\) and \(\langle d\kappa \rangle\) are the expected values of the corresponding differential quantities, integrated over the kime-phase distribution \(\Phi(\varphi)\).

Given that \(\varphi\) is distributed according to \(\Phi(\varphi)\), we must calculate the expected values of \(d\kappa\) and \(d\kappa^2\).

\[\langle d\kappa \rangle = \int_{-\infty}^{\infty} e^{i\varphi} \left( dt + i t d\varphi \right) \Phi(\varphi) d\varphi.\]

Given that \(\Phi(\varphi)\) is a probability distribution, such as a Laplace distribution \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right).\)

The expected value \(\langle d\kappa \rangle\) should account for the average phase

\[\langle d\kappa \rangle = \left\langle e^{i\varphi} \right\rangle dt + i t \left\langle e^{i\varphi} \right\rangle d\varphi,\]

where for Laplace phase distribution \(\left\langle e^{i\varphi} \right\rangle = \int_{-\infty}^{\infty} e^{i\varphi} \Phi(\varphi) d\varphi = e^{i\mu} \frac{b}{b^2 + 1}.\)

This implies \(\langle d\kappa \rangle = e^{i\mu} \frac{b}{b^2 + 1} \left(dt + i t d\varphi \right).\)

The expected value of \(d\kappa^2\) is \[\langle d\kappa^2 \rangle = \int_{-\infty}^{\infty} e^{2i\varphi} \left(dt^2 - t^2 d\varphi^2\right) \Phi(\varphi) d\varphi,\]

where: \(\left\langle e^{2i\varphi} \right\rangle = \int_{-\infty}^{\infty} e^{2i\varphi} \Phi(\varphi) d\varphi = e^{2i\mu} \frac{b^2}{b^2 + 4}.\)

Thus, \(\langle d\kappa^2 \rangle = e^{2i\mu} \frac{b^2}{b^2 + 4} \left(dt^2 - t^2 d\varphi^2 \right).\)

Given that \(\varphi\) is a distribution and not a variable, the correct way to account for the kime-phase is through these expected values, rather than treating \(\varphi\) as an independent variable.

Thus, the spacekime metric tensor should be

\[G_{AB} dX^A dX^B = -\tilde{N}^2 \langle d\kappa^2 \rangle + \tilde{\beta}_a \langle d\kappa \rangle dX^a + \gamma_{ab} dX^a dX^b.\]

In the ADM Formalism in spacekime, the Hamiltonian and momentum constraints must be adapted to account for the contributions from the kime-phase distribution.

The Hamiltonian constraint is now

\[\mathcal{H} = \int \left(R^{(3)} + \frac{K^2 - K_{ab} K^{ab}}{\tilde{N}^2 \langle d\kappa^2 \rangle}\right) \sqrt{\gamma} \, d^3x \, \Phi(\varphi) d\varphi,\]

where \(K_{ab}\) is the extrinsic curvature, and \(R^{(3)}\) is the scalar curvature of the 3D spatial hypersurface.

The momentum constraint becomes \(\mathcal{M}_a = -2 \nabla_b \left( \frac{K^b_a}{\langle d\kappa \rangle} \right).\)

This explicitly incorporates the effects of the kime-phase distribution, rather than treating \(\varphi\) as a simple variable.

To obtain solutions to the spacekime ADM equations, consider:

  • Boundary Conditions: Solutions must match classical general relativity (e.g., Schwarzschild solution) at large distances where spacekime effects diminish.
  • Energy Conditions: Solutions should satisfy standard energy conditions when integrated over the kime-phase distribution.
  • Stability Analysis: Ensure that the inclusion of the kime-phase distribution does not lead to instabilities or unphysical solutions.

9.2.2 (Approach 2) Reformulating the ADM Formalism in Spacekime

To reformulate the ADM spacekime formalism, we need to account for the complex nature of time (\(\kappa = t e^{i\varphi}\)) while ensuring that the spacekime metric remains real-valued. We will also incorporate the kime-phase distribution, \(\Phi(\varphi)\), using distributional derivatives to capture the stochastic or probabilistic behavior of the phase.

We begin by defining the spacekime metric \(G_{AB}\) that remains real-valued despite the complex nature of \(\kappa\). To achieve this, we use the modulus of \(d\kappa\) to ensure that the time-related terms in the metric do not introduce imaginary components. The space-kime metric can be written as

\[ G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \left(|d\kappa|^2\right) + \tilde{\beta}_a \, \text{Re}(d\kappa) \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]

where:

  • \(|d\kappa|^2 = dt^2 + t^2 \, d\varphi^2\) is the modulus squared of the differential, ensuring the metric’s reality.
  • \(\text{Re}(d\kappa) = \cos\varphi \, dt - \sin\varphi \, t \, d\varphi\) ensures the shift vector interaction terms remain real.
  • \(\gamma_{ab}\) represents the 4D spatial metric, unaffected by the kime-phase directly but influenced through the dynamics dictated by the lapse function \(\tilde{N}(\kappa, \varphi)\) and shift vector \(\tilde{\beta}_a\).

The 5D spacekime manifold is foliated into spacetime-like hypersurfaces indexed by \(\phi\). Each leaf of this foliation is a 4D spacetime hypersurface, characterized by the coordinates \((x^i, t)\).

The kime-phase \(\varphi\) modulates the progression from one hypersurface to another, governed by the distribution \(\varphi\sim \Phi(\varphi)\).

To incorporate the probabilistic nature of \(\varphi\), we introduce distributional derivatives of the kime-phase distribution \(\Phi(\varphi)\). This allows us to capture the effects of phase transitions or other stochastic processes influencing the kime-phase.

If \(\Phi(\varphi)\) is a smooth distribution, we define \(\frac{\delta \Phi(\varphi)}{\delta \varphi}\).

These derivatives play a crucial role in the dynamics, particularly when \(\Phi(\varphi)\) is not a simple function but instead reflects more complex, stochastic behavior.

We extend the ADM formalism by incorporating the kime-phase distribution. The Hamiltonian constraint equation in spacekime is modified to include terms involving the distributional derivatives:

\[\mathcal{H} = \mathcal{H}_\text{ADM} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\ ,\]

where:

  • \(\mathcal{H}_\text{ADM}\) is the standard ADM Hamiltonian.
  • The term \(\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\) accounts for the influence of the kime-phase distribution on the geometry and dynamics of space-kime.

The Einstein field equations in spacekime are now derived by varying the generalized Einstein-Hilbert action with respect to the metric tensor \(G_{AB}\):

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\ ,\]

where:

  • \(R_{AB}\) is the Ricci tensor for the 5D space-kime manifold.
  • \(R^{(5)}\) is the Ricci scalar.
  • \(T_{AB}\) is the energy-momentum tensor, describing the distribution of matter and energy.

9.2.3 Example Solutions: Laplace and Normal Kime-Phase Distributions

To solve the field equations, we consider specific forms for \(\Phi(\varphi)\).

Laplace Kime-Phase Distribution: \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right)\).

  • The distributional derivative introduces a term in the field equations that concentrates curvature around \(\varphi_0\), leading to localized effects in the space-kime manifold.

Gaussian (Normal) Kime-Phase Distribution: \(\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \varphi_0)^2}{2\sigma^2}\right)\).

  • A smoother distribution like the Gaussian spreads curvature effects more uniformly across the space-kime manifold, resulting in smoother variations in the geometry.

To solve these equations:

  • Laplace Distribution: The solution would show sharp curvature spikes at \(\varphi_0\), suggesting regions of concentrated gravitational effects, akin to phase transitions.
  • Gaussian Distribution: The solution would result in a smoother curvature profile, indicating more gradual changes in the space-kime structure.

By reformulating the ADM formalism in space-kime, we capture the effects of a complexified time and its associated kime-phase distribution while ensuring the metric remains real. This framework opens up new possibilities for understanding the dynamics of the universe through the lens of complex time and its probabilistic influences.

9.3 Linking the Spacekime Metric, Field Equations, and ADM Formalism

To fully integrate the spacekime metric, the field equations, and the ADM formalism, we need to explicitly detail how each component influences the overall structure of spacekime general relativity. Let’s try to solve the equations and show examples using different lapse functions and distributions (e.g., Gaussian and Laplace kime-phase distributions).

The spacekime metric extends the standard 4D spacetime metric to a 5D manifold where the extra dimension corresponds to the complex time (\(\kappa = t \cdot e^{i\varphi}\)). The metric is given by:

\[G_{AB} \, dX^A \, dX^B = -\tilde{N}^2 \, |d\kappa|^2 + \tilde{\beta}_a \, \text{Re}(d\kappa) \, dX^a + \gamma_{ab} \, dX^a \, dX^b\ ,\]

where:

  • \(|d\kappa|^2 = dt^2 + t^2 \, d\varphi^2\) ensures the metric is real-valued.
  • \(\text{Re}(d\kappa) = \cos\varphi \, dt - \sin\varphi \, t \, d\varphi\) ensures the reality of the metric components.
  • \(\tilde{N}\) is the generalized lapse function, which modulates the relationship between space and complex time.
  • \(\tilde{\beta}_a\) represents the shift vector, and \(\gamma_{ab}\) is the 4D spatial metric.

The field equations are derived from varying the Einstein-Hilbert action in the spacekime framework

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\ ,\]

where:

  • \(R_{AB}\) is the Ricci tensor for the 5D space-kime manifold.
  • \(T_{AB}\) is the energy-momentum tensor, describing the matter content.
  • The term involving \(\Phi(\varphi)\) introduces contributions from the kime-phase distribution.

The ADM formalism provides a Hamiltonian formulation of general relativity, adapted here for spacekime. The key equations are the Hamiltonian constraint and the momentum constraints

\[\mathcal{H} = \mathcal{H}_\text{ADM} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\]

\[\mathcal{H}_\text{ADM} = \sqrt{\gamma} \left( R^{(3)} + \frac{1}{\tilde{N}^2} (K_{ij}K^{ij} - K^2) \right)\ ,\]

where:

  • \(R^{(3)}\) is the Ricci scalar for the 3D spatial hypersurfaces.
  • \(K_{ij}\) is the extrinsic curvature, capturing how the 3D hypersurfaces are embedded in the 5D space-kime manifold.

Let’s consider three example lapse functions and solve the corresponding field equations.

1. Constant Lapse Function: \(\tilde{N}(\kappa) = 1\). For this simplest case, the spacekime metric reduces to:

\[G_{AB} \, dX^A \, dX^B = -|d\kappa|^2 + \gamma_{ab} \, dX^a \, dX^b\ .\] The field equations simplify to \(R_{AB} = 8\pi G \, T_{AB}\). This corresponds to a static spacekime manifold where the complex time has no additional modulation effects beyond those of standard general relativity.

2. Exponential Lapse Function: \(\tilde{N}(\kappa) = e^{\lambda t}\). In this case, \(\lambda\) is a constant that modulates the lapse function over time. The metric is

\[G_{AB} \, dX^A \, dX^B = -e^{2\lambda t} \, |d\kappa|^2 + \gamma_{ab} \, dX^a \, dX^b.\]

The field equations in this case suggest an accelerating expansion in the kime direction, similar to inflationary models in cosmology but extended into complex time.

3. Power-Law Lapse Function: \(\tilde{N}(\kappa) = t^\alpha\). For a power-law lapse function

\[G_{AB} \, dX^A \, dX^B = -t^{2\alpha} \, |d\kappa|^2 + \gamma_{ab} \, dX^a \, dX^b.\]

This metric describes a space-kime manifold where the influence of time grows or decays polynomially, depending on the sign and magnitude of \(\alpha\).

Gaussian Kime-Phase Distribution: For a Gaussian distribution of the kime-phase:

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left(-\frac{(\varphi - \varphi_0)^2}{2\sigma^2}\right)\]

The corresponding field equations include terms that smoothly vary with \(\varphi\), leading to a spacekime manifold where curvature is distributed smoothly, with the most significant effects centered around \(\varphi_0\).

C. Laplace Kime-Phase Distribution: For a Laplace distribution:

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right).\]

The field equations now involve sharp curvature effects centered at \(\varphi_0\), resulting in concentrated gravitational effects analogous to rapid phase transitions or localized “shocks” in the space-kime manifold.

9.4 Examples of spacekime field equation solutions

To derive and explicate the solutions to the spacekime field equations, we need to incorporate the specific lapse functions and kime-phase distributions discussed earlier. This process involves identifying any necessary constraints and then solving the equations under these conditions.

Again, the generalized space-kime field equations are:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

To solve these field equations, we impose the following constraints:

  1. Metric Compatibility: The metric \(G_{AB}\) must remain real and consistent with the structure of space-kime. This requires that the lapse function \(\tilde{N}(\kappa)\) and the shift vector \(\tilde{\beta}_a\) do not introduce any imaginary components.

  2. Energy-Momentum Conservation: The solutions must satisfy generalized energy-momentum conservation laws, which may include additional terms from the kime-phase distribution.

  3. Boundary Conditions: Appropriate boundary conditions must be specified, particularly for the kime-phase \(\varphi\) at critical points such as infinity or phase transitions.

Example 1: Constant Lapse Function: \(\tilde{N}(\kappa) = 1\). With a constant lapse function, the metric simplifies, and the field equations reduce to

\[R_{AB} = 8\pi G \, T_{AB}.\]

This is equivalent to the standard Einstein field equations in 4D general relativity but extended into the space-kime framework. No additional constraints are necessary beyond those typical for solving Einstein’s equations.

The solutions correspond to standard GR solutions (e.g., Schwarzschild, FLRW metrics), extended trivially into the complex-time domain.

Example 2: Exponential Lapse Function: \(\tilde{N}(\kappa) = e^{\lambda t}\).

The metric now incorporates an exponentially growing or decaying lapse function, leading to:

\[R_{AB} = 8\pi G \, T_{AB} + \lambda^2 e^{2\lambda t} \gamma_{AB}.\]

The exponential term introduces additional curvature in the space-kime manifold, necessitating a constraint on \(\lambda\) to maintain stability and prevent singularities.

This setup resembles inflationary cosmological models but extended into the spacekime domain. The exponential growth in the lapse function suggests rapid expansion or contraction in the kime direction, influencing how spacetime evolves with \(\varphi\).

To avoid singularities or uncontrolled growth, constraints on the parameter \(\lambda\) may be necessary, such as \(\lambda < 0\) for contraction, or specific initial conditions to stabilize expansion.

Example 3: Power-Law Lapse Function: \(\tilde{N}(\kappa) = t^\alpha\). The field equations with a power-law lapse function become

\[R_{AB} = 8\pi G \, T_{AB} + \alpha^2 t^{2\alpha - 2} \gamma_{AB}.\]

The power-law introduces a polynomial dependence on time, affecting how curvature evolves.

For positive \(\alpha\), curvature grows or decays as a function of \(t\). For \(\alpha = 1\), this reduces to linear time evolution, similar to the standard cosmological expansion.

The value of \(\alpha\) must be chosen to ensure that the curvature remains finite and the evolution is physically meaningful. For example, \(\alpha > 1\) might lead to accelerated expansion, requiring boundary conditions at infinity.

9.4.1 Gaussian Kime-Phase Distribution

\[\Phi(\varphi) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp\left( -\frac{(\varphi - \varphi_0)^2}{2\sigma^2}\right)\ .\]

The smooth Gaussian distribution leads to field equations of the form

\[R_{AB} = 8\pi G \, T_{AB} + \frac{1}{\sigma^2} \left( \varphi - \varphi_0 \right) \gamma_{AB}\] This equation reflects a smooth curvature distribution centered around \(\varphi_0\).

The solutions indicate that curvature is most significant around \(\varphi_0\), decaying smoothly as \(\varphi\) deviates from \(\varphi_0\).

To maintain finite curvature, the standard deviation \(\sigma\) must be appropriately scaled relative to the spacetime scales involved.

9.4.2 Laplace Kime-Phase Distribution

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right).\]

With the Laplace distribution, the field equations become

\[R_{AB} = 8\pi G \, T_{AB} + \frac{1}{b^2} \text{sign}(\varphi - \varphi_0) \gamma_{AB}.\]

This introduces a sharp transition in curvature at \(\varphi = \varphi_0\).

Curvature is concentrated at \(\varphi_0\), with rapid decay away from this point. The solution might represent localized gravitational phenomena, such as phase transitions or abrupt changes in spacetime properties.

To prevent singularities at \(\varphi = \varphi_0\), constraints on \(b\) and boundary conditions around \(\varphi_0\) are necessary.

Solving the spacekime field equations with different lapse functions and kime-phase distributions illustrate how complex time and its phase distribution influence the curvature and dynamics of spacetime.

9.4.3 Spacekime Metric Signature

In standard general relativity, the metric signature often takes the form \((-, +, +, +)\) for a 4D spacetime, reflecting one time-like dimension and three space-like dimensions. In spacekime, where time is complexified (\(\kappa = t e^{i\varphi}\)), the introduction of an additional kime dimension complicates the definition of a metric signature.

Given that \(\varphi\) is a distribution rather than a traditional variable, the spacekime metric must reflect how this distribution affects the geometry. However, defining a signature like \((- - + + +)\) might be misleading because it implies two “time-like” dimensions, one of which is not a variable but a distribution influencing the overall structure.

A more appropriate approach could be to consider the metric as fundamentally 5D, but with one of the dimensions influenced by a stochastic or distributional process. The signature might still be \((- \cdot + + +)\), depending on how the kime-phase distribution affects the dynamics, but it would need careful interpretation to reflect the fact that the second “time-like” dimension is not observable.

The Einstein-Hilbert action is given by

\[S = \int \sqrt{-g} \, R \, d^4x\ ,\]

where \(R\) is the Ricci scalar, \(g\) is the determinant of the metric tensor, and \(d^4x\) represents the spacetime volume element.

To generalize the Einstein-Hilbert action for complex time

\[S = \int \sqrt{-G} \, R^{(5)} \, d^5X\ ,\]

where:

  • \(G\) is the determinant of the space-kime metric tensor.
  • \(R^{(5)}\) is the Ricci scalar for the 5D space-kime manifold.
  • \(d^5X\) represents the volume element in the 5D spacekime manifold.

To find the corresponding field equations, we vary the generalized action with respect to the space-kime metric \(G_{AB}\)

\[\delta S = \delta \int \sqrt{-G} \, R^{(5)} \, d^5X = 0.\]

The variation gives us the generalized Einstein field equations in spacekime

\[R_{AB}^{(5)} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \text{(additional terms)}.\]

Applying the stationary-action principle (\(\delta S = 0\)) to this generalized action should yield the field equations directly. However, the additional complexity introduced by the kime-phase distribution \(\Phi(\varphi)\) must be handled carefully.

Key Points:

  • Metric Reality: Ensure that the space-kime metric remains real-valued despite the presence of complex-time.
  • Kime-Phase Distribution: The additional terms arising from the distributional nature of \(\varphi\) might introduce source terms in the field equations, affecting curvature in a non-trivial way.
  • Constraints: The solutions to these equations might require additional constraints related to the behavior of \(\Phi(\varphi)\) to ensure physical relevance and stability.

9.4.4 Additional Terms

To explicitly determine the additional terms in the above generalized spacekime Einstein field equations, we must carefully examine the variation of the Einstein-Hilbert action when complex-time and the kime-phase distribution are introduced.

The generalized action for a space-kime manifold is:

\[S = \int \sqrt{-G} \left( R^{(5)} + \mathcal{L}_{\text{matter}} \right) d^5X\ ,\]

where:

  • \(G\) is the determinant of the space-kime metric tensor \(G_{AB}\).
  • \(R^{(5)}\) is the 5D Ricci scalar.
  • \(\mathcal{L}_{\text{matter}}\) is the Lagrangian density for matter fields.

The variation of the action with respect to the metric tensor \(G_{AB}\) yields:

\[\delta S = \delta \int \sqrt{-G} \, R^{(5)} \, d^5X = \int \left( \delta \sqrt{-G} \, R^{(5)} + \sqrt{-G} \, \delta R^{(5)} \right) d^5X.\]

The first term varies the metric determinant, while the second varies the Ricci scalar. The variation of the Ricci scalar in higher dimensions, particularly when the metric is complexified, is complex and introduces additional terms, especially when \(\kappa\) and \(\Phi(\varphi)\) are involved.

Additional Terms due to Complex Time and Kime-Phase Distribution: When considering complex-time \(\kappa = t \cdot e^{i\varphi}\) and a distributional kime-phase \(\Phi(\varphi)\), the field equations pick up additional terms from the following sources:

  • A. Distributional Derivatives of \(\Phi(\varphi)\). When \(\varphi\) is treated as a distribution, its derivatives introduce source-like terms. Specifically:

\[\frac{\delta S}{\delta \varphi} \propto \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

These terms reflect how variations in the kime-phase distribution influence the spacetime curvature. In the field equations, this translates to additional contributions that can be interpreted as effective stress-energy components, contributing to the overall curvature.

  • B. Coupling Between Real and Imaginary Parts of Complex Time. The complex nature of time, represented by \(\kappa = t \cdot e^{i\varphi}\), involves both real and imaginary components. When considering variations of the action, the coupling between these components introduces mixed derivative terms in the field equations. Specifically:
    • Real Part Contributions: Derivatives with respect to \(t\) (the real part) contribute to the usual time-evolution terms.
    • Imaginary Part Contributions: Derivatives with respect to \(\varphi\) (the kime-phase) introduce new terms reflecting the phase-dependent behavior. Since \(\varphi\) is distributional, these terms can manifest as sources that affect the curvature in a non-standard way.

The additional terms involving \(\varphi\) and its distributional derivatives can be seen as analogous to source terms in the Einstein field equations, except they are linked to the probabilistic nature of the kime-phase rather than traditional matter or energy distributions.

When all contributions are considered, the modified spacekime field equations take the form:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \underbrace{\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)}_{\text{Distributional Kime-Phase Term}} + \underbrace{\text{Coupling Terms Between } t \text{ and } \varphi}_{\text{Complex-Time Contributions}}.\]

Here, the additional terms can be detailed as:

  1. Distributional Kime-Phase Term: \(\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\)
  • This term represents the influence of the kime-phase distribution on the curvature. It acts similarly to an effective stress-energy tensor but is linked to the stochastic properties of \(\varphi\).
  1. Complex-Time Coupling Terms:
  • These arise from the interaction between the real and imaginary parts of \(\kappa\). They include mixed derivatives of \(t\) and \(\varphi\), which contribute to the overall dynamics in a way that can be interpreted as phase-modulated forces or curvatures.

The presence of these additional terms implies several significant modifications to the usual Einstein field equations:

  • Modified Curvature Dynamics: The distributional nature of \(\varphi\) leads to a non-uniform curvature distribution that could, for example, describe regions of spacetime where quantum effects or phase transitions play a significant role.
  • Phase-Dependent Gravity: The coupling terms suggest that gravity (or more generally, spacetime curvature) could have a phase-dependent component, which might manifest as variations in gravitational strength or spacetime structure in different regions of the universe.

Let’s fully expand and explore explicitly the field equations under a Laplace distribution prior for the kime-phase. The Laplace distribution for the kime-phase \(\varphi\) is given by:

\[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right)\ ,\]

where \(b\) is a scale parameter, and \(\varphi_0\) is the location parameter (mean of the distribution). The distribution is characterized by a sharp peak at \(\varphi_0\) and exponential decay away from \(\varphi_0\). The generalized Einstein field equations in the space-kime framework are:

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

Given the Laplace distribution for \(\Phi(\varphi)\), we wil expand the additional term \(\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi)\).

  1. First Derivative of \(\Phi(\varphi)\): The first derivative with respect to \(\varphi\) is

\[\frac{d\Phi(\varphi)}{d\varphi} = -\frac{\text{sign}(\varphi - \varphi_0)}{2b^2} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right),\]

where, \(\text{sign}(\varphi - \varphi_0)\) indicates the sign of the difference \(\varphi - \varphi_0\), which is \(+1\) if \(\varphi > \varphi_0\) and \(-1\) if \(\varphi < \varphi_0\).

  1. Second Derivative of \(\Phi(\varphi)\): The second derivative is

\[\frac{d^2\Phi(\varphi)}{d\varphi^2} = \frac{1}{2b^3} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right) \delta(\varphi - \varphi_0),\]

where, \(\delta(\varphi - \varphi_0)\) is the Dirac delta function, which is non-zero only at \(\varphi = \varphi_0\).

  1. Additional Term in Field Equations: Now, substitute these derivatives back into the additional term of the field equations:

\[\frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi) = \frac{1}{2b} \left( \frac{1}{b} \delta(\varphi - \varphi_0) \gamma_{AB} \right).\]

Thus, the additional term simplifies to \(\frac{1}{b^2} \delta(\varphi - \varphi_0) \gamma_{AB}\).

Then, the spacekime Field Equations with Laplace Distribution are obtained by substituting this into the generalized field equations

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{b^2} \delta(\varphi - \varphi_0) \gamma_{AB}.\]

Notes:

  1. Localized Curvature:
  • The delta function \(\delta(\varphi - \varphi_0)\) indicates that the additional curvature contribution is highly localized at \(\varphi = \varphi_0\). This suggests that the space-kime manifold experiences a concentrated curvature effect specifically at the kime-phase \(\varphi_0\), corresponding to a phase transition or a sharp change in the spacetime structure.
  1. Effective Stress-Energy Term:
  • The term \(\frac{1}{b^2} \gamma_{AB}\) can be interpreted as an effective stress-energy contribution localized at \(\varphi_0\), which could represent, for example, the effects of a localized energy distribution or a quantum phase transition in the space-kime framework.
  1. Geometric Interpretation:
  • The space-kime metric is modified by this additional term, which could lead to new geometric structures or solutions in general relativity when extended to complex time. This might manifest as regions of spacetime that have different properties depending on their kime-phase, potentially influencing how we understand spacetime at a fundamental level.

We can also expand the spacekime field equations for all indices \(A,B\in\{0,1,2,3,4\}\). Considering a 5D spacekime metric \(G_{AB}\), corresponding to the 4 spacetime dimensions and the additional kime-phase dimension.

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{b^2} \delta(\varphi - \varphi_0) \gamma_{AB}.\]

Let’s explicitly write these out for all indices \(A, B in\{0, 1, 2, 3, 4\}\), where: - \(A, B = 0\) corresponds to the kime-phase dimension. - \(A, B = 1, 2, 3\) corresponds to the spatial dimensions. - \(A, B = 4\) corresponds to the time dimension.

For the metric \(G_{AB}\), the Ricci tensor \(R_{AB}\) in 5 dimensions is

\[R_{AB} = \partial_C \Gamma^C_{AB} - \partial_B \Gamma^C_{AC} + \Gamma^C_{AB} \Gamma^D_{CD} - \Gamma^C_{AD} \Gamma^D_{BC},\]

where \(\Gamma^C_{AB}\) are the Christoffel symbols defined by

\[\Gamma^C_{AB} = \frac{1}{2} G^{CD} \left( \partial_A G_{BD} + \partial_B G_{AD} - \partial_D G_{AB} \right).\]

We compute the components of \(R_{AB}\) for the different cases of indices:

(a) \(A, B = 0\): The Kime-Kime Component

\[R_{00} = \partial_C \Gamma^C_{00} - \partial_0 \Gamma^C_{0C} + \Gamma^C_{00} \Gamma^D_{CD} - \Gamma^C_{0D} \Gamma^D_{0C}\]

(b) \(A = 0, B = i\) and \(A = i, B = 0\): The Kime-Spatial Components

\[R_{0i} = \partial_C \Gamma^C_{0i} - \partial_i \Gamma^C_{0C} + \Gamma^C_{0i} \Gamma^D_{CD} - \Gamma^C_{0D} \Gamma^D_{iC},\]

and similarly,

\[R_{i0} = \partial_C \Gamma^C_{i0} - \partial_0 \Gamma^C_{iC} + \Gamma^C_{i0} \Gamma^D_{CD} - \Gamma^C_{iD} \Gamma^D_{0C}.\]

(c) \(A, B = i, j\): The Spatial Components

\[R_{ij} = \partial_C \Gamma^C_{ij} - \partial_j \Gamma^C_{iC} + \Gamma^C_{ij} \Gamma^D_{CD} - \Gamma^C_{iD} \Gamma^D_{jC}.\]

(d) \(A = 0, B = 4\) and \(A = 4, B = 0\): The Kime-Time Components

\[R_{04} = \partial_C \Gamma^C_{04} - \partial_4 \Gamma^C_{0C} + \Gamma^C_{04} \Gamma^D_{CD} - \Gamma^C_{0D} \Gamma^D_{4C},\]

and analogously,

\[R_{40} = \partial_C \Gamma^C_{40} - \partial_0 \Gamma^C_{4C} + \Gamma^C_{40} \Gamma^D_{CD} - \Gamma^C_{4D} \Gamma^D_{0C}.\]

(e) \(A = i, B = 4\) and \(A = 4, B = i\): The Spatial-Time Components

\[R_{i4} = \partial_C \Gamma^C_{i4} - \partial_4 \Gamma^C_{iC} + \Gamma^C_{i4} \Gamma^D_{CD} - \Gamma^C_{iD} \Gamma^D_{4C},\]

and similarly,

\[R_{4i} = \partial_C \Gamma^C_{4i} - \partial_i \Gamma^C_{4C} + \Gamma^C_{4i} \Gamma^D_{CD} - \Gamma^C_{4D} \Gamma^D_{iC}.\]

(f) \(A, B = 4\): The Time-Time Component

\[R_{44} = \partial_C \Gamma^C_{44} - \partial_4 \Gamma^C_{4C} + \Gamma^C_{44} \Gamma^D_{CD} - \Gamma^C_{4D} \Gamma^D_{4C}.\]

The Ricci scalar in 5D is \(R^{(5)} = G^{AB} R_{AB}\). Given the complexity of the Ricci scalar and the corresponding calculations, the term \(G_{AB} R^{(5)}\) must also be expanded for each pair of indices \(A\) and \(B\). This will interact with the stress-energy tensor \(T_{AB}\) and the additional term \(\frac{1}{b^2} \delta(\varphi - \varphi_0) \gamma_{AB}\).

Therefore, the fully expanded equations take the form

\[R_{AB} - \frac{1}{2} G_{AB} \left( G^{CD} R_{CD} \right) = 8\pi G \, T_{AB} + \frac{1}{b^2} \delta(\varphi - \varphi_0) \gamma_{AB}\]

For Specific Indices:

  • \(A = 0, B = 0\): The kime-kime component.
  • \(A = 0, B = i\): The kime-spatial component.
  • \(A = i, B = j\): The spatial-spatial component.
  • \(A = 4, B = 4\): The time-time component.
  • \(A = 4, B = i\): The time-spatial component.

Each equation will involve explicit forms of \(R_{AB}\) derived from the Christoffel symbols and their derivatives, considering the specific kime-phase prior distribution and its localization effects.

9.4.5 Hands-on Example: Solving the Field Equations in Spacekime under Laplace Prior

Let’s work out an explicit solution to the spacekime field equations, using the Laplace distribution prior for the kime-phase and considering a simplified case with specific assumptions.

Given:

  1. Metric Signature: We assume a metric signature of \((- + + + \cdot)\) for the 5D space-kime manifold.
  2. Laplace Distribution: The kime-phase \(\varphi\) follows the distribution \[\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \varphi_0|}{b}\right).\]
  3. Field Equations: \[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = 8\pi G \, T_{AB} + \frac{1}{b^2} \delta(\varphi - \varphi_0) \gamma_{AB}.\]

We aim to solve for \(R_{00}\), the kime-kime component of the Ricci tensor, and the corresponding curvature in the space-kime manifold. We will assume a simplified, static spacetime scenario with \(T_{AB} = 0\) (vacuum solution).

Step 1: Christoffel Symbols Calculation

For simplicity, let’s assume a diagonal spacekime metric \[G_{AB} = \text{diag}(-\tilde{N}^2, \gamma_{11}, \gamma_{22}, \gamma_{33}, \gamma_{44}),\] where \(\tilde{N}\) is the lapse function, and \(\gamma_{ii}\) are the spatial metric components. The Christoffel symbols \(\Gamma^C_{AB}\) are calculated by

\[\Gamma^0_{00} = \frac{1}{2} G^{00} \partial_0 G_{00},\] \[\Gamma^i_{00} = \frac{1}{2} G^{ii} \partial_0 G_{00},\] \[\Gamma^0_{ij} = \frac{1}{2} G^{00} (\partial_i G_{0j} + \partial_j G_{0i} - \partial_0 G_{ij}).\]

Step 2: Calculate the Ricci Tensor Component \(R_{00}\)

Given the Christoffel symbols, the Ricci tensor component \(R_{00}\) is

\[R_{00} = \partial_C \Gamma^C_{00} - \partial_0 \Gamma^C_{0C} + \Gamma^C_{00} \Gamma^D_{CD} - \Gamma^C_{0D} \Gamma^D_{0C}.\]

In this simplified static case \[R_{00} = \partial_0 \Gamma^0_{00} - \Gamma^i_{00} \Gamma^0_{i0}.\]

Step 3: Solve the Field Equation for \(R_{00}\)

Using the Laplace distribution, the field equation for \(R_{00}\) becomes

\[R_{00} - \frac{1}{2} G_{00} R^{(5)} = \frac{1}{b^2} \delta(\varphi - \varphi_0) \gamma_{00}.\]

Given that \(G_{00} = -\tilde{N}^2\) and assuming a constant lapse function \(\tilde{N} = 1\)

\[R_{00} - \frac{1}{2} (-1) R^{(5)} = \frac{1}{b^2} \delta(\varphi - \varphi_0).\]

The Ricci scalar in 5D, \(R^{(5)}\), must be computed considering the contributions from both the kime and spatial dimensions. However, in this simplified case, assuming the Ricci scalar contribution is small, we approximate

\[R_{00} \approx \frac{1}{b^2} \delta(\varphi - \varphi_0).\]

This solution of the spacekime field equations indicates that:

  • Localized Curvature: The curvature \(R_{00}\) in the kime-kime direction is highly localized at \(\varphi = \varphi_0\), with magnitude inversely proportional to \(b^2\).
  • Physical Interpretation: This could correspond to a “shock” or rapid change in the kime dimension, possibly related to a phase transition or abrupt change in the spacetime structure at \(\varphi_0\).

This example solved a simplified version of the spacekime field equations under the influence of a Laplace-distributed kime-phase. The resulting solution suggests that significant curvature effects are localized at specific kime-phase values, potentially representing regions of spacetime with unusual properties or rapid phase transitions.

References:

  • Arnowitt, R., Deser, S., Misner, C. (1959). “Dynamical Structure and Definition of Energy in General Relativity” (PDF). Physical Review. 116 (5): 1322–1330. Bibcode:1959PhRv..116.1322A. doi:10.1103/PhysRev.116.1322.

10 Experimental Tests to Evaluate and Potentially Falsify the Complex-Time Representation and the Spacekime Framework

To evaluate the complex-time representation of repeated measurement longitudinal models and the spacekime framework for representing quantum gravity and field equations, it is essential to design experiments that can test the fundamental predictions made by these models. Below are several proposed experimental tests, categorized into physical experiments, data science experiments, and mathematical derivations, aimed at assessing the validity of these theoretical frameworks.

10.1 Physical Experiments

10.1.1 Time Dilation Experiments with Quantum Systems

Objective: Test the predictions of the spacekime framework concerning time dilation effects in quantum systems, especially those under different kime-phase distributions.

Approach:

  • Utilize high-precision atomic clocks or quantum systems (e.g., trapped ions, superconducting qubits) in varying gravitational fields or under acceleration.
  • Measure the time dilation effects as predicted by standard general relativity and compare them with predictions from the spacekime framework, where time is complex and the kime-phase distribution influences the observed time intervals.
  • Any deviation from the standard relativistic predictions could indicate the presence of complex-time effects or kime-phase influences, potentially validating or falsifying the spacekime framework.

Falsifiability Criterion: If the experimental results match standard general relativity without any detectable deviation, it would challenge the necessity of introducing complex-time or kime-phase distributions.

10.1.2 Testing Quantum Superpositions of Time Intervals

Objective: Investigate the impact of the complex-time representation on quantum superpositions, specifically regarding the measurement of time intervals in superposed states.

Approach:

  • Design an experiment where quantum systems are prepared in superpositions of different time intervals, e.g., using quantum clocks or entangled particles with different time evolution paths.
  • Measure the outcomes of these superpositions and analyze the results to determine whether they exhibit the interference patterns predicted by the complex-time framework.
  • The complex-time representation suggests that time intervals may interfere differently compared to classical expectations.

Falsifiability Criterion: The absence of predicted interference patterns or the presence of patterns that can be fully explained by classical or conventional quantum mechanics would challenge the validity of the complex-time model.

10.2 Artificial Intelligence (AI) and Data Science Experiments

10.2.1 Analysis of Longitudinal Biomedical Data

Objective: Evaluate whether the complex-time representation of longitudinal data offers better predictive accuracy or insights compared to traditional methods.

Approach:

  • Use real-world longitudinal biomedical datasets (e.g., patient health records, clinical trial data) and apply both the complex-time representation and traditional longitudinal models.
  • Compare the models’ performance in predicting future outcomes, accounting for temporal correlations, and identifying underlying trends.
  • The spacekime framework should theoretically capture more nuanced temporal dynamics, potentially leading to improved predictions.

Falsifiability Criterion: If the complex-time model does not outperform traditional models or fails to provide additional insights, it would question the practical utility of the spacekime representation in real-world data analysis.

10.2.2 Simulation of Kime-Phase Influences on Longitudinal Data

Objective: Simulate data using the spacekime framework and compare it with observed data from quantum systems or other time-sensitive processes.

Approach:

  • Generate synthetic datasets under different kime-phase distributions (e.g., Gaussian, Laplace) and simulate how these distributions affect the evolution of a system over time.
  • Compare the statistical properties of the simulated data with real-world datasets to see if the spacekime framework can accurately replicate observed phenomena.
  • Analyze whether the phase-distribution influences manifest as predicted in real-world data.

Falsifiability Criterion: Inability to replicate observed data characteristics using the spacekime framework, or if real-world data fits traditional models better, would weaken the argument for the complex-time approach.

10.3 Mathematical Derivations and Theoretical Tests

10.3.1 Derivation of Consistent Field Equations

Objective: Examine whether the field equations derived in the spacekime framework are consistent with known physical laws and lead to sensible physical predictions.

Approach:

  • Derive explicit solutions to the spacekime field equations under different kime-phase distributions and compare them with known solutions in general relativity and quantum field theory.
  • Analyze whether these solutions are stable, causal, and lead to physically meaningful predictions.
  • Investigate whether the introduction of a kime-phase distribution creates new conservation laws or breaks existing ones.

Falsifiability Criterion: If the derived solutions are unstable, non-causal, or violate known physical laws (e.g., energy conservation), it would suggest that the spacekime framework may not be a valid extension of current theories.

10.3.2 Consistency with Loop Quantum Gravity (LQG)

Objective: Test whether the spacekime framework can be integrated with or remains consistent with the principles of Loop Quantum Gravity (LQG).

Approach:

  • Analyze the spacekime formalism to see if it can be formulated within the existing LQG framework, particularly in terms of the quantization of spacetime and the absence of a fundamental time variable.
  • Investigate whether spacekime introduces additional degrees of freedom that are incompatible with LQG or whether it provides a natural extension to LQG’s spin network and spin foam models.
  • Determine if the complex-time representation can be used to explain phenomena that LQG currently cannot address.

Falsifiability Criterion: If spacekime cannot be integrated with LQG or leads to contradictions within the established LQG framework, this would be a significant challenge to its validity as a quantum gravity theory.

10.3.3 Example

Objective: Design an direct comparison contrasting the explicit solutions to the spacekime field equations under Laplace kime-phase distribution and compare them with known spacetime solutions in general relativity and quantum field theory using the same time of constraints.

10.3.4 Direct Comparison: Spacekime Field Equations under Laplace Kime-Phase Distribution vs. Known Spacetime Solutions in General Relativity and Quantum Field Theory

To design a direct comparison between explicit solutions to the spacekime field equations under a Laplace kime-phase distribution and known solutions in general relativity (GR) and quantum field theory (QFT), we can follow a systematic approach that highlights the differences and similarities in the resulting geometries, field dynamics, and physical predictions.

The spacekime field equations generalize Einstein’s field equations to a 5D framework, incorporating complex time (\(\kappa = t e^{i\varphi}\)) and a kime-phase distribution (\(\Phi(\varphi)\)). The generalized field equations are

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

where

  • the Laplace kime-phase distribution \(\Phi(\varphi) = \frac{1}{2b} \exp\left(-\frac{|\varphi - \mu|}{b}\right),\)
  • \(R_{AB}\) is the Ricci tensor in 5D spacekime.
  • \(G_{AB}\) is the spacekime metric tensor.
  • \(T_{AB}\) is the energy-momentum tensor.
  • \(\nabla_A\) are the covariant derivatives.

For comparison, we will use known solutions from GR and QFT, such as:

  • Schwarzschild Solution in general relativity, describing the spacetime around a non-rotating spherical mass.
  • Kerr Solution in general relativity, describing spacetime around a rotating mass.
  • Quantum Field Theory (QFT) solutions, particularly in the context of curved spacetime, where quantum fields are treated on a fixed classical background.

For both the spacekime framework and the traditional GR/QFT frameworks, we assume the following boundary conditions:

  • Vacuum Solutions: We consider solutions in vacuum, where \(T_{AB} = 0\).
  • Asymptotic Flatness: Solutions should asymptotically approach flat spacetime at large distances from the source.
  • Boundary Conditions: The solutions should satisfy physical boundary conditions such as the regularity of the metric and continuity across boundaries.

The spacekime field equations under a Laplace kime-phase distribution lead to solutions that include corrections to the traditional spacetime metric. For example, the Schwarzschild metric in spacekime might take the form

\[G_{tt} = -\left(1 - \frac{2GM}{r}\right) - \frac{2}{b^2} \left(1 - \frac{2GM}{r}\right),\]

where the second term represents the contribution from the kime-phase distribution.

Similarly, for the radial component

\[G_{rr} = \left(1 - \frac{2GM}{r}\right)^{-1} + \frac{2}{b^2} \left(1 - \frac{2GM}{r}\right)^{-1}.\]

Known spacetime solutions in GR include

Schwarzschild Metric: \(g_{tt} = -\left(1 - \frac{2GM}{r}\right),\) and \(g_{rr} = \left(1 - \frac{2GM}{r}\right)^{-1}.\)

Kerr Metric: For a rotating mass, the metric components depend on both \(r\) and the angular momentum \(J\) of the mass. The exact form of the metric is more complex, but it retains a similar structure to the Schwarzschild solution, with additional terms for the rotational effects.

We compare the metric components \(G_{tt}, G_{rr}\) from the spacekime field equations under the Laplace distribution with \(g_{tt}, g_{rr}\) from the Schwarzschild solution. The key difference is the additional phase-dependent curvature term in the spacekime framework, which modifies the traditional metric.

Time Component: The spacekime correction adds a phase-dependent term that modifies the gravitational potential. This can lead to observable differences in time dilation effects, especially close to the event horizon.

Radial Component: The radial component in spacekime includes an additional term that could modify the effective radius of gravitational effects, potentially altering the location of the event horizon or the gravitational redshift.

We calculate the spacetime Ricci scalar \(R^{(4)}\) and the spacekime Ricci scalar in \(R^{(5)}\). For example,

\[R^{(5)} = R^{(4)} + \frac{4}{b^2}.\] However, comparing the Spacetime Ricci Scalar \(R^{(4)}\) and the Spacekime Ricci Scalar \(R^{(5)}\) are approximately equal, when the Laplace scale parameter \(b\) is large, or under a Uniform Kime-Phase distribution.

Both, the spacetime Ricci scalar \(R^{(4)}\) and the spacekime Ricci scalar \(R^{(5)}\) are quantities that describe the curvature of spacetime and spacekime, respectively. The spacekime Ricci scalar is expected to include additional contributions from the kime-phase distribution, \(\Phi(\varphi)\). When the kime-phase distribution is uniform, we need to evaluate whether these additional contributions significantly alter the Ricci scalar, and under what conditions \(R^{(4)}\) and \(R^{(5)}\) might be similar.

In standard general relativity, the Ricci scalar \(R^{(4)}\) is a 4D quantity computed from the spacetime metric \(g_{\mu\nu}\), \(R^{(4)} = g^{\mu\nu} R_{\mu\nu},\) where \(g^{\mu\nu}\) is the inverse spacetime metric, and \(R_{\mu\nu}\) is the Ricci tensor derived from the Riemann curvature tensor, describing the curvature of spacetime.

In vacuum solutions like the Schwarzschild or Kerr metrics, \(R^{(4)} = 0\) because these solutions are solutions to Einstein’s vacuum field equations.

In the spacekime framework, the Ricci scalar \(R^{(5)}\) is computed from the 5D metric \(G_{AB}\), which incorporates the complex time \(\kappa = t e^{i\varphi}\) and the kime-phase distribution \(\Phi(\varphi)\), \(R^{(5)} = G^{AB} R_{AB},\) where \(G^{AB}\) is the inverse spacekime metric, and \(R_{AB}\) is the Ricci tensor in 5D, which now includes contributions from the kime-phase distribution.

A uniform kime-phase distribution implies that \(\Phi(\varphi)\) is constant over the phase interval, \(\Phi(\varphi) = \frac{1}{2\pi} \quad \text{for} \, \varphi \in [-\pi, \pi).\) This uniform distribution means that the kime-phase contributes equally at all values of \(\varphi\), and its effect on the spacekime metric is averaged across the entire phase range.

When the kime-phase distribution is uniform, the spacekime metric may average out the contributions of the kime-phase in a way that minimizes its impact on the overall curvature. To see if \(R^{(4)}\) and \(R^{(5)}\) would be similar, let’s consider the structure of the spacekime Ricci scalar \[R^{(5)} = R^{(4)} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

For a uniform distribution, \(\Phi(\varphi)\) is constant and \(\nabla_A \nabla_B \Phi(\varphi) = 0.\) This means the term involving the derivatives of \(\Phi(\varphi)\) drops out, and the spacekime Ricci scalar reduces to

\[R^{(5)} = R^{(4)} + \text{(possible additional terms from kime contributions to metric)}.\]

However, if the kime contributions to the metric do not introduce additional curvature effects (e.g., if the spacekime metric is designed such that the kime-phase effects cancel out in the Ricci tensor calculation), then \(R^{(5)} \approx R^{(4)}.\)

Under a uniform kime-phase distribution, the spacekime Ricci scalar \(R^{(5)}\) and the spacetime Ricci scalar \(R^{(4)}\) would be very similar, particularly if the uniform distribution leads to an averaging out of the kime-phase contributions. In such a scenario, the additional phase dimension does not introduce significant curvature effects beyond those already present in the 4D spacetime. This result suggests that for a uniform kime-phase distribution, the spacekime framework may reduce to a form that closely resembles traditional general relativity, with minimal deviation in the curvature characteristics. However, this similarity depends on the specific form of the spacekime metric and the boundary conditions applied. In more general cases with non-uniform kime-phase distributions, \(R^{(5)}\) would differ from \(R^{(4)}\) due to the additional curvature contributions from the kime-phase distribution.

In GR, \(R^{(4)}\) is zero in vacuum, but in spacekime, the additional term introduces a non-zero curvature even in the vacuum solution, reflecting the influence of the kime-phase distribution.

Physical Predictions and Gravitational Effects: The spacekime corrections could lead to observable differences in gravitational lensing, time dilation, and the behavior of objects near strong gravitational sources (e.g., near black holes).

Quantum Field Behavior: In QFT, fields in curved spacetime may experience different vacuum states and particle creation rates when the background spacetime is modified by spacekime effects.

Experimental Falsifiability: If experiments (e.g., gravitational wave observations, black hole shadow measurements) detect deviations from predictions based on standard GR, it could indicate spacekime effects. Conversely, if no deviations are found, this could falsify the spacekime approach.

11 Spacekime Framework Representatio of Quantum Mechanics and Gravity

11.1 Schrödinger equation in spacekime

11.1.1 Approach 1

The Schrödinger equation traditionally governs the time evolution of a quantum state \(\psi(x, t)\)

\[i\hbar \frac{\partial \psi(x, t)}{\partial t} = \hat{H} \psi(x, t).\]

In spacekime, time \(t\) is replaced by complex time \(\kappa\). Since \(\kappa\) is complex and its phase is a distribution, the formulation must account for the distribution of \(\varphi\).

The corrected spacekime Schrödinger equation should be

\[i\hbar \frac{\partial \psi(x, \kappa)}{\partial \kappa} = \hat{H}^\kappa \psi(x, \kappa),\]

where

  • \(\frac{\partial \psi(x, \kappa)}{\partial \kappa}\) should be interpreted in the sense of distributional derivatives when \(\kappa\) involves a phase distribution.
  • \(\hat{H}^\kappa\) is the Hamiltonian operator in spacekime, which may include contributions from both the spatial dimensions and the complex kime component.

However, since \(\kappa\) itself is not a simple variable but includes a phase \(\varphi\) that is distributed, the differential with respect to \(\kappa\) needs careful handling. The proper approach involves considering the expected values and treating the evolution as an averaged effect over the distribution \(\Phi(\varphi)\).

Given the complex nature of \(\kappa\), we interpret the derivative with respect to \(\kappa\) as

\[\frac{\partial \psi(x, \kappa)}{\partial \kappa} \equiv \left\langle \frac{\partial \psi(x, \kappa)}{\partial t} + i \frac{\partial \psi(x, \kappa)}{\partial \varphi} \right\rangle_{\Phi(\varphi)},\]

where the angle brackets denote the expectation value taken over the kime-phase distribution \(\Phi(\varphi)\).

This leads to a modified Schrödinger equation in spacekime

\[i\hbar \left\langle \frac{\partial \psi(x, t, \varphi)}{\partial t} + i \frac{\partial \psi(x, t, \varphi)}{\partial \varphi} \right\rangle_{\Phi(\varphi)} = \hat{H}^\kappa \psi(x, t, \varphi).\]

When \(\Phi(\varphi)\) is sharply peaked (e.g., a delta function), the equation reduces to something closer to the traditional Schrödinger equation, with a well-defined time evolution. For broader distributions (e.g., Gaussian, Laplace), the phase distribution introduces additional terms that affect the evolution of \(\psi(x, \kappa)\). These terms can be interpreted as introducing an effective “quantum time” that reflects the uncertainty or variability in the phase \(\varphi\).

This formulation suggests that the quantum evolution of states in the spacekime framework is influenced by the underlying kime-phase distribution, which adds a new layer of complexity to the time evolution. This approach effectively integrates quantum uncertainty into the time dimension, potentially offering new insights into the relationship between quantum mechanics and spacetime geometry.

11.1.2 Approach 2

Let’s re-examine the Schrödinger equation in spacekime clarifying the differentiation with respect to a (random) phase \(\varphi\sim \Phi(\varphi)\). In spacekime, time is extended to a complex dimension \(\kappa = t e^{i\varphi}\), where\(t\) is real time and \(\varphi\sim \Phi(\varphi)\).

Our goal is to derive a version of the Schrödinger equation that is consistent with this complex-time representation, without improperly differentiating with respect to the distributed phase \(\varphi\).

In quantum mechanics, the time evolution of the wavefunction \(\psi(x, t)\) is governed by the spacetime Schrödinger equation

\[i\hbar \frac{\partial \psi(x, t)}{\partial t} = \hat{H} \psi(x, t).\]

To extend this to spacekime, we consider how \(\psi(x, \kappa)\) evolves with respect to \(\kappa\). Instead of differentiating with respect to \(\varphi\), we treat the kime-phase distribution as influencing the evolution of the wavefunction in an averaged or probabilistic manner.

Expected Value Formulation: To incorporate the kime-phase distribution into the Schrödinger equation consider the expected evolution of the wavefunction with respect to the distribution \(\Phi(\varphi)\). One approach to formulate the evolution of \(\psi(x, \kappa)\) can be expressed as

\[\frac{\partial \psi(x, \kappa)}{\partial \kappa} = \left\langle \frac{\partial \psi(x, t)}{\partial t} \cdot \frac{\partial t}{\partial \kappa} \right\rangle_{\Phi(\varphi)},\]

where, we only differentiate with respect to the real time \(t\), and then account for how this derivative affects the complex time \(\kappa\). Since \(\kappa = t e^{i\varphi}\), the derivative \(\frac{\partial t}{\partial \kappa}\) in complex terms is

\[\frac{\partial t}{\partial \kappa} = \frac{1}{e^{i\varphi}} = e^{-i\varphi}.\]

So the evolution equation becomes

\[\frac{\partial \psi(x, \kappa)}{\partial \kappa} = \left\langle e^{-i\varphi} \frac{\partial \psi(x, t)}{\partial t} \right\rangle_{\Phi(\varphi)}.\]

In this case, incorporating the Hamiltonian operator \(\hat{H}\), the Schrödinger equation in spacekime becomes

\[i\hbar \frac{\partial \psi(x, \kappa)}{\partial \kappa} = \hat{H}^\kappa \psi(x, \kappa),\]

where \[\frac{\partial \psi(x, \kappa)}{\partial \kappa} = \left\langle e^{-i\varphi} \frac{\partial \psi(x, t)}{\partial t} \right\rangle_{\Phi(\varphi)}.\]

Thus,

\[i\hbar \left\langle e^{-i\varphi} \frac{\partial \psi(x, t)}{\partial t} \right\rangle_{\Phi(\varphi)} = \hat{H}^\kappa \psi(x, \kappa).\]

This formulation suggests that the evolution of the wavefunction in spacekime is an averaged effect, influenced by the kime-phase distribution. The expected value accounts for the distribution of \(\varphi\) rather than treating \(\varphi\) as a variable to be differentiated directly.

To validate this formulation:

  • Consistency with Standard Quantum Mechanics: When \(\Phi(\varphi)\) is sharply peaked (e.g., at a single phase value), the expected value reduces to a standard derivative with respect to real time \(t\), recovering the traditional Schrödinger equation.
  • Incorporation of Quantum Effects: For broader distributions, this equation captures the additional phase uncertainty, which could be interpreted as introducing quantum corrections to the evolution.

11.2 Alternative Spacekime Evolutions of the Wavefunction

There are alternative ways to formulate the evolution of the wavefunction with respect to a phase distribution. In the context of quantum mechanics, the phase may be interpreted probabilistically or statistically rather than as a simple deterministic variable.

11.2.1 Path Integral Formulation

The path integral formulation of quantum mechanics, introduced by Richard Feynman, provides a natural framework to incorporate phase distributions. In this approach, the evolution of the wavefunction is understood as a sum over all possible paths, each weighted by a phase factor.

In the spacekime context, we could generalize the path integral to include contributions from the kime-phase distribution \(\Phi(\varphi)\):

\[\psi(x, \kappa) = \int \mathcal{D}[x(t)] \, e^{\frac{i}{\hbar} S[x(t)]} \Phi(\varphi),\]

where, \(S[x(t)]\) is the action along a path \(x(t)\), and \(\Phi(\varphi)\) modulates the contribution of each path according to the kime-phase.

For example, let’s work out an example with a path integral formulation under Laplace kime-phase distribution. Consider a simple quantum system where the action \(S[x(t)]\) is known, and the spacekime wavefunction \(\psi(x, \kappa)\) is to be computed. In the spacekime framework, the time parameter is complex, \(\kappa = t e^{i\varphi}\), where \(\varphi\sim\Phi(\varphi)\) is Laplace distributed.

The path integral formulation states that the wavefunction is obtained by summing over all possible paths, each weighted by the phase factor \(e^{iS/\hbar}\) and the kime-phase distribution

\[\psi(x, \kappa) = \int \mathcal{D}[x(t)] \, e^{\frac{i}{\hbar} S[x(t)]} \Phi(\varphi).\]

The Laplace distribution for the phase \(\varphi\) is \(\Phi(\varphi) = \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}},\) where \(b\) is the scale parameter, and \(\varphi_0\) is the location parameter (the mean of the distribution).

We are interested in computing the wavefunction \(\psi(x, \kappa)\), where \(\kappa = t e^{i\varphi}\) and \(\varphi\) follows the Laplace distribution.

\[\psi(x, \kappa) = \int \mathcal{D}[x(t)] \int_{-\infty}^{\infty} \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}} \, e^{\frac{i}{\hbar} S[x(t)]} \, d\varphi.\]

To simplify the integral, we focus on the contribution from the kime-phase distribution

\[\psi(x, \kappa) =\int_{-\infty}^{\infty} \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}} \, e^{i\varphi} \, d\varphi.\]

This integral can be split into two parts (for \(\varphi \geq \varphi_0\) and \(\varphi < \varphi_0\)):

\[\psi(x, \kappa) =\underbrace{\int_{\varphi_0}^{\infty} \frac{1}{2b} e^{-\frac{\varphi - \varphi_0}{b}} \, e^{i\varphi} \, d\varphi}_{I} + \underbrace{\int_{-\infty}^{\varphi_0} \frac{1}{2b} e^{\frac{\varphi - \varphi_0}{b}} \, e^{i\varphi} \, d\varphi}_{II}.\]

These integrals can be individually evaluated, leading to

\[I=\int_{\varphi_0}^{\infty} \frac{1}{2b} e^{-\frac{\varphi - \varphi_0}{b}} \, e^{i\varphi} \, d\varphi = \frac{1}{2b} \cdot \frac{e^{i\varphi_0}}{1 - i b}.\]

\[II=\int_{-\infty}^{\varphi_0} \frac{1}{2b} e^{\frac{\varphi - \varphi_0}{b}} \, e^{i\varphi} \, d\varphi = \frac{1}{2b} \cdot \frac{e^{i\varphi_0}}{1 + i b}.\]

Adding these, the total contribution from the kime-phase distribution is:

\[II =\int_{-\infty}^{\infty} \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}} \, e^{i\varphi} \, d\varphi = \frac{e^{i\varphi_0}}{b^2 + 1}.\]

Substituting this back into the path integral expression for the wavefunction

\[\psi(x, \kappa) = \frac{e^{i\varphi_0}}{b^2 + 1} \underbrace{\int \mathcal{D}[x(t)] \,e^{\frac{i}{\hbar} S[x(t)]}}_{spacetime\ evolution}.\]

This result indicates that the spacekime wavefunction is modulated by the kime-phase distribution and includes a phase shift \(\varphi_0\) and a damping factor \(b^2 + 1\).

  • Phase Modulation: The Laplace kime-phase distribution introduces a phase shift \(\varphi_0\) in the wavefunction. This shift depends on the location parameter of the distribution.
  • Damping: The factor \(b^2 + 1\) reflects a damping effect introduced by the distribution’s scale parameter \(b\). For larger \(b\), the contribution to the wavefunction is spread out more, leading to a more pronounced damping effect.

This approach doesn’t require explicit differentiation with respect to the phase. Instead, the phase distribution modifies the weight of each path in the integral.

11.2.2 Density Matrix Formalism

The density matrix formalism is another approach, especially when dealing with mixed states or statistical ensembles. Instead of evolving a pure wavefunction, we can describe the system’s state using a density matrix \(\rho(t)\), which evolves according to the Liouville-von Neumann equation.

In the spacekime framework, the density matrix can be generalized to include phase distribution

\[\rho(\kappa) = \int_{-\pi}^{\pi} \Phi(\varphi) \, \rho(t, \varphi) \, d\varphi.\]

The evolution of the density matrix would then be expressed as

\[\frac{d\rho(\kappa)}{d\kappa} = \left\langle \frac{d\rho(t, \varphi)}{dt} \cdot \frac{\partial t}{\partial \kappa} \right\rangle_{\Phi(\varphi)},\]

where, the density matrix evolution integrates over the phase distribution, allowing for a more statistical treatment of the phase.

Let’s show another example of spacekime evolution using the Density Matrix Formalism under Laplace Kime-Phase Distribution in the Spacekime Framework. The density matrix formalism is particularly useful in quantum mechanics for describing mixed states and statistical ensembles. In the spacekime framework, where time is extended into a complex dimension \(\kappa = t e^{i\varphi}\) and \(\varphi\) follows a probability distribution \(\Phi(\varphi)\), the density matrix can be generalized to include the effects of this kime-phase distribution.

In standard quantum mechanics, the density matrix \(\rho(t)\) is given by

\[\rho(t) = \sum_i p_i |\psi_i(t)\rangle \langle \psi_i(t)|,\]

where \(p_i\) are the probabilities associated with different pure states \(|\psi_i(t)\rangle\). The evolution of the density matrix is governed by the Liouville-von Neumann equation

\[i\hbar \frac{d\rho(t)}{dt} = [\hat{H}, \rho(t)].\]

In spacekime, the time parameter \(t\) is replaced by the complex time \(\kappa = t e^{i\varphi}\), and the phase \(\varphi\) is distributed according to a Laplace distribution \(\Phi(\varphi)\). The density matrix in this context can be defined as

\[\rho(\kappa) = \int_{-\infty}^{\infty} \Phi(\varphi) \, \rho(t, \varphi) \, d\varphi,\]

where \(\Phi(\varphi) = \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}}\) is the Laplace distribution, with \(b\) as the scale parameter and \(\varphi_0\) as the location parameter.

Assume the system can be described by a pure state \(|\psi(\kappa)\rangle\) that evolves over complex time. The corresponding density matrix is

\[\rho(\kappa) = \int_{-\infty}^{\infty} \Phi(\varphi) |\psi(\kappa)\rangle \langle \psi(\kappa)| \, d\varphi.\]

Substituting the expression for \(\kappa\) and the Laplace distribution

\[\rho(\kappa) = \int_{-\infty}^{\infty} \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}} |\psi(t e^{i\varphi})\rangle \langle \psi(t e^{i\varphi})| \, d\varphi.\]

To evaluate this integral, we can split it into two parts for \(\varphi \geq \varphi_0\) and \(\varphi < \varphi_0\), similarly to the path integral case

\[\rho(\kappa) = \frac{1}{2b} \left[\int_{\varphi_0}^{\infty} e^{-\frac{\varphi - \varphi_0}{b}} |\psi(t e^{i\varphi})\rangle \langle \psi(t e^{i\varphi})| \, d\varphi + \int_{-\infty}^{\varphi_0} e^{\frac{\varphi - \varphi_0}{b}} |\psi(t e^{i\varphi})\rangle \langle \psi(t e^{i\varphi})| \, d\varphi \right].\]

For simplicity, assume that the wavefunction \(|\psi(t e^{i\varphi})\rangle\) can be expanded as a power series in \(\varphi\), allowing us to approximate the integral

\[|\psi(t e^{i\varphi})\rangle = \sum_{n=0}^{\infty} \frac{(i\varphi)^n}{n!} |\psi_n(t)\rangle.\]

Inserting this into the integral, and assuming the leading order terms dominate

\[\rho(\kappa) \approx \frac{1}{2b} \left[\frac{|\psi(t e^{i\varphi_0})\rangle \langle \psi(t e^{i\varphi_0})|}{1 - \frac{i}{b}} + \frac{|\psi(t e^{i\varphi_0}) \rangle \langle \psi(t e^{i\varphi_0})|}{1 + \frac{i}{b}} \right].\]

After simplifying, the density matrix becomes

\[\rho(\kappa) \approx \frac{|\psi(t e^{i\varphi_0})\rangle \langle \psi(t e^{i\varphi_0})|}{b^2 + 1}.\]

This result shows that the density matrix is modulated by the kime-phase distribution, similarly to the path integral formulation. The phase \(\varphi_0\) introduces a shift, and the scale parameter \(b\) introduces a damping factor. To determine the evolution of the density matrix \(\rho(\kappa)\), we use the modified Liouville-von Neumann equation

\[i\hbar \frac{d\rho(\kappa)}{d\kappa} = [\hat{H}^\kappa, \rho(\kappa)],\]

where, \(\hat{H}^\kappa\) is the Hamiltonian in the spacekime framework, which includes the effects of the kime-phase distribution. Some notes:

  • Mixed State: Even if the original state \(|\psi(t)\rangle\) was pure, the integration over the phase distribution \(\Phi(\varphi)\) introduces a mixed state character to the system.
  • Phase Modulation: The phase \(\varphi_0\) modulates the quantum state, potentially leading to new observable effects.
  • Damping: The factor \(b^2 + 1\) indicates how the phase distribution smooths or damps the density matrix, which could be interpreted as introducing decoherence effects into the system.

11.2.3 Stochastic Schrödinger Equation

Another approach is to model the phase \(\varphi\) as a stochastic process rather than a fixed distribution. This leads to the formulation of a stochastic Schrödinger equation, where the wavefunction evolves according to a differential equation with stochastic terms.

If \(\varphi\) is treated as a stochastic variable with dynamics described by some stochastic differential equation (e.g., a Wiener process), the evolution of the wavefunction could be described by

\[i\hbar \frac{d\psi(x, t)}{dt} = \hat{H} \psi(x, t) + \eta(t, \varphi),\]

where \(\eta(t, \varphi)\) is a noise term that depends on the phase and time, representing the stochastic nature of the kime-phase. In this case, the wavefunction evolution is influenced by random fluctuations in the kime-phase, leading to an ensemble of possible states.

Next, we consider an example of using the Stochastic Schrödinger Equation Formalism under Laplace Kime-Phase Distribution. The Stochastic Schrödinger Equation (SSE) formalism extends the standard Schrödinger equation to account for random processes, often used to describe systems where quantum states are influenced by noise or other stochastic effects. In the context of the spacekime framework, where time is extended to a complex dimension \(\kappa = t e^{i\varphi}\) with the phase \(\varphi\) distributed according to a Laplace distribution \(\Phi(\varphi)\), we can model the evolution of the wavefunction with a stochastic Schrödinger equation.

The traditional Schrödinger equation is

\[i\hbar \frac{\partial \psi(x, t)}{\partial t} = \hat{H} \psi(x, t).\]

Its stochastic extension includes a noise term that influences the evolution of the wavefunction

\[i\hbar \frac{d\psi(x, t)}{dt} = \hat{H} \psi(x, t) + \eta(t),\]

where \(\eta(t)\) represents a stochastic noise term.

In spacekime, the complex time \(\kappa = t e^{i\varphi}\) introduces an additional stochastic component due to the kime-phase distribution \(\Phi(\varphi)\). The phase \(\varphi\) can be modeled as a stochastic process, and the evolution of the wavefunction can be described by a stochastic differential equation (SDE).

Given that \(\varphi\) follows a Laplace distribution, we need to incorporate this into the noise term. To incorporate the Laplace-distributed phase \(\varphi\) into the Schrödinger equation, we assume that the noise term \(\eta(t, \varphi)\) has contributions from both the real time \(t\) and the stochastic phase \(\varphi\).

The modified stochastic Schrödinger equation in spacekime is

\[i\hbar \frac{d\psi(x, \kappa)}{d\kappa} = \hat{H}^\kappa \psi(x, \kappa) + \eta(\kappa),\]

Where the noise term \(\eta(\kappa)\) is influenced by the kime-phase

\[\eta(\kappa) = \eta_0(t) e^{i\varphi} + \eta_\varphi(t),\]

where \(\eta_0(t)\) is a standard noise term depending on time, and \(\eta_\varphi(t)\) is a stochastic term representing the influence of the kime-phase \(\varphi\). Given that \(\varphi\) is Laplace-distributed, we model \(\eta_\varphi(t)\) accordingly \(\eta_\varphi(t) \sim \text{Laplace}(0, b),\) where \(b\) is the scale parameter of the Laplace distribution.

Then, the equation for the wavefunction evolution becomes

\[i\hbar \frac{d\psi(x, \kappa)}{d\kappa} = \hat{H}^\kappa \psi(x, \kappa) + \eta_0(t) e^{i\varphi} + \eta_\varphi(t).\]

This is a stochastic differential equation where the noise term \(\eta_\varphi(t)\) introduces randomness based on the Laplace distribution of \(\varphi\).

To solve this equation, we treat it as an SDE and use stochastic calculus techniques. For simplicity, we assume \(\hat{H}^\kappa\) is independent of \(\kappa\) for the initial case

\[\psi(x, \kappa) = \psi(x, \kappa_0) e^{-\frac{i}{\hbar} \hat{H}^\kappa (\kappa - \kappa_0)} + \int_{\kappa_0}^\kappa G(\kappa', \kappa_0) [\eta_0(t') e^{i\varphi} + \eta_\varphi(t')] d\kappa',\]

where, \(G(\kappa', \kappa_0)\) is the Green’s function associated with the Hamiltonian \(\hat{H}^\kappa\).

Given the Laplace distribution of \(\varphi\) \(\Phi(\varphi) = \frac{1}{2b} e^{-\frac{|\varphi|}{b}}\), we can calculate the expected value of the wavefunction by integrating over \(\varphi\)

\[\langle \psi(x, \kappa) \rangle_{\Phi(\varphi)} = \int_{-\infty}^{\infty} \psi(x, \kappa) \Phi(\varphi) \, d\varphi.\]

The solution involves averaging the Green’s function and the noise terms over the Laplace-distributed phase.

Let’s look at an example calculation for a simple harmonic oscillator with Laplace Noise. Consider the Hamiltonian of a simple harmonic oscillator

\[\hat{H}^\kappa = \frac{p^2}{2m} + \frac{1}{2}m\omega^2 x^2.\]

The noise term is given by \(\eta_\varphi(t) \sim \text{Laplace}(0, b)\) and the wavefunction evolution is described by

\[\langle \psi(x, \kappa) \rangle_{\Phi(\varphi)} = \psi(x, \kappa_0) e^{-\frac{i}{\hbar} \hat{H}^\kappa (\kappa - \kappa_0)} + \int_{\kappa_0}^\kappa G(\kappa', \kappa_0) \left\langle \eta_0(t') e^{i\varphi} + \eta_\varphi(t') \right\rangle_{\Phi(\varphi)} d\kappa',\]

where \[\left\langle \eta_0(t') e^{i\varphi} \right\rangle_{\Phi(\varphi)} = \eta_0(t') \frac{1}{b^2 + 1}.\]

Some notes:

  • Stochastic Dynamics: The phase noise introduces randomness into the evolution of the wavefunction, simulating effects like decoherence or quantum noise.
  • Phase Averaging: The phase \(\varphi\) being Laplace-distributed introduces a non-Gaussian stochastic process, which may lead to heavy-tailed distributions of the wavefunction’s behavior.
  • Decoherence: The averaging process can be interpreted as leading to decoherence, where certain quantum states become more “classical” due to the influence of the noise.

11.2.4 Phase Space Representation

Quantum mechanics can also be formulated in phase space using the Wigner function or other quasi-probability distributions. In this framework, the evolution of the quantum state is described in terms of phase space variables, and the kime-phase distribution could be integrated into this representation.

The Wigner function \(W(x, p, t)\) could be generalized to include a kime-phase distribution

\[W(x, p, \kappa) = \int_{-\pi}^{\pi} \Phi(\varphi) \, W(x, p, t, \varphi) \, d\varphi,\]

The evolution of this function would be governed by a modified quantum Liouville equation that incorporates the kime-phase effects.

The phase space representation of quantum mechanics offers a different approach to understanding quantum states, where states are described in terms of position and momentum (or other conjugate variables). One of the most common phase space representations is the Wigner function, which is a quasi-probability distribution function. In the context of the spacekime framework, where time is extended to a complex dimension \(\kappa = t e^{i\varphi}\) with the kime-phase \(\varphi\) distributed according to a Laplace distribution, we can explore how this affects the Wigner function and the dynamics of the system.

In standard quantum mechanics, the Wigner function \(W(x, p, t)\) for a state \(|\psi(t)\rangle\) is defined as

\[W(x, p, t) = \frac{1}{\pi \hbar} \int_{-\infty}^{\infty} \psi^*\left( x - \frac{y}{2}, t\right) \psi\left(x + \frac{y}{2}, t\right) e^{ipy/\hbar} \, dy.\]

The Wigner function provides a way to calculate quantum expectations and is often used to bridge the gap between quantum and classical mechanics.

In the spacekime framework, we extend the time parameter to a complex dimension \(\kappa = t e^{i\varphi}\). The corresponding Wigner function in spacekime can be defined as

\[W(x, p, \kappa) = \frac{1}{\pi \hbar} \int_{-\infty}^{\infty} \psi^*\left(x - \frac{y}{2}, \kappa\right) \psi\left(x + \frac{y}{2}, \kappa\right) e^{ipy/\hbar} \, dy.\]

Given that \(\varphi\sim\Phi(\varphi) = \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}}\) is a stochastic variable with a Laplace distribution, the Wigner function must account for this distribution. Let’s incorporate this into the Wigner function by averaging over the kime-phase distribution

\[\langle W(x, p, \kappa) \rangle_{\Phi(\varphi)} = \int_{-\infty}^{\infty} \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}} W(x, p, t e^{i\varphi}) \, d\varphi.\]

Again, let’s consider a simple harmonic oscillator with the Hamiltonian

\[\hat{H}^\kappa = \frac{p^2}{2m} + \frac{1}{2}m\omega^2 x^2.\]

For a pure state \(|\psi(t)\rangle\), the Wigner function at time \(t\) is well known. We extend this to the spacekime Wigner function \(W(x, p, \kappa) = W(x, p, t e^{i\varphi}).\)

Consider the initial Gaussian wave packet: \[\psi(x, 0) = \left(\frac{m\omega}{\pi\hbar}\right)^{1/4} e^{-\frac{m\omega x^2}{2\hbar}}.\]

The Wigner function at time \(t = 0\) is

\[W(x, p, 0) = \frac{1}{\pi\hbar} \exp\left(-\frac{2m\omega x^2}{\hbar} - \frac{p^2}{2m\hbar\omega}\right).\]

To evaluate the spacekime Wigner function, we consider the evolution of the Wigner function over complex time \(\kappa = t e^{i\varphi}\), \(W(x, p, \kappa) = W\left(x, p, t e^{i\varphi}\right).\) This Wigner function is then averaged over the Laplace distribution of \(\varphi\)

\[\langle W(x, p, \kappa) \rangle_{\Phi(\varphi)} = \int_{-\infty}^{\infty} \frac{1}{2b} e^{-\frac{|\varphi - \varphi_0|}{b}} W\left(x, p, t e^{i\varphi}\right) \, d\varphi.\]

For the simple harmonic oscillator, we assume that the phase shift \(e^{i\varphi}\) modifies the oscillation frequency

\[W\left(x, p, t e^{i\varphi}\right) = \frac{1}{\pi\hbar} \exp\left(-\frac{2m\omega x^2} {\hbar e^{i\varphi}} - \frac{p^2 e^{i\varphi}}{2m\hbar\omega}\right).\]

The integral can be challenging due to the complex exponential, but can be approximated numerically or by using series expansion methods. We expand the exponentials to the first order around \(\varphi = \varphi_0\)

\[\langle W(x, p, \kappa) \rangle_{\Phi(\varphi)} \approx \frac{1}{\pi\hbar} \exp\left(-\frac{2m\omega x^2}{\hbar e^{i\varphi_0}} - \frac{p^2 e^{i\varphi_0}}{2m\hbar\omega}\right) \cdot \underbrace{\frac{1}{b^2 + 1}}_{damping}.\]

This result shows how the Laplace distribution modulates the phase space distribution function, introducing a damping factor and modifying the frequency of oscillation.

Some notes:

  • Phase Space Evolution: The evolution of the quantum state in phase space is influenced by the kime-phase distribution, introducing modifications to the Wigner function that reflect the stochastic nature of the kime-phase.
  • Decoherence: The damping factor introduced by the Laplace distribution suggests that the kime-phase distribution can lead to decoherence effects, where the quantum state becomes more classical in behavior over time.
  • Modulated Oscillations: The phase shift affects the frequency of the oscillations in phase space, potentially leading to new observable phenomena in quantum systems described by the spacekime framework.

Each of these approaches offers a way to address the challenge of phase distributions in quantum mechanics, particularly in the context of the spacekime framework.

11.3 Gravity in Spacekime

In general relativity (GR), gravity is understood as the curvature of 4D spacetime, caused by the presence of mass and energy. The Einstein field equations describe how this curvature relates to the energy-momentum tensor, which encodes the distribution of matter and energy.

The unification of quantum mechanics and gravity is one of the most significant challenges in theoretical physics. The spacekime framework, which introduces complex time (\(\kappa = t e^{i\varphi}\)) and extends the traditional 4D spacetime to 5D spacekime, offers a novel approach to addressing this challenge. Below, I outline a conceptual and mathematical strategy for unifying quantum mechanics and gravity within the spacekime framework.

In standard physics, quantum mechanics and gravity are described by fundamentally different frameworks

  • Quantum Mechanics: Governed by wave functions and operators, typically formulated in a fixed spacetime background.
  • General Relativity: Describes gravity as the curvature of spacetime, without an intrinsic role for quantum uncertainty or superposition.

The spacekime framework may bridge these by extending spacetime to include a complex time dimension, where time is represented as \(\kappa = t e^{i\varphi}\), with \(\varphi\) being a distributed kime-phase. This extension allows for a unified description where quantum states evolve not just over time but over spacekime, incorporating both spatial and kime-phase dynamics.

The spacekime metric tensor \(G_{AB}\) is a 5D extension of the 4D spacetime metric, incorporating the complex time dimension:

\[G_{AB} dX^A dX^B = -\tilde{N}^2 \langle d\kappa^2 \rangle + \tilde{\beta}_a \langle d\kappa \rangle dX^a + \gamma_{ab} dX^a dX^b,\]

where

  • \(\tilde{N}\) is the lapse function in spacekime.
  • \(\tilde{\beta}_a\) is the shift vector.
  • \(\gamma_{ab}\) is the 3D spatial metric.
  • The expected values \(\langle d\kappa \rangle\) and \(\langle d\kappa^2 \rangle\) account for the kime-phase distribution.

The field equations in the spacekime framework extend Einstein’s field equations to include the effects of the kime-phase distribution

\[R_{AB} - \frac{1}{2} G_{AB} R^{(5)} = \frac{8\pi G}{c^4} T_{AB} + \frac{1}{\Phi(\varphi)} \nabla_A \nabla_B \Phi(\varphi).\]

These equations govern the curvature of spacekime and include contributions from both classical energy-momentum (via \(T_{AB}\)) and quantum effects through the kime-phase distribution.

In quantum mechanics, the state of a system is described by a wave function \(\psi(x, t)\), evolving according to the Schrödinger equation. In spacekime, this is generalized to

\[\hat{H}^\kappa \psi(x, \kappa) = i\hbar \frac{\partial \psi(x, \kappa)}{\partial \kappa},\]

where - \(\hat{H}^\kappa\) is the Hamiltonian operator in spacekime, which may include both spatial and kime-phase components. - The evolution of \(\psi(x, \kappa)\) is governed by the spacekime extension of the Schrödinger equation, which incorporates the complex time variable \(\kappa\).

The key to unifying quantum mechanics and gravity in spacekime lies in integrating the quantum state dynamics with the curvature of spacekime.

In this framework, the quantum state \(\psi(x, \kappa)\) is not just a function but can be viewed as a geometric object (a manifold) embedded in spacekime. The evolution of this manifold is influenced by the underlying spacekime geometry, which is determined by the field equations.

To incorporate quantum gravity, the spacekime field equations must be quantized. This involves promoting the metric \(G_{AB}\) to an operator \(\hat{G}_{AB}\) and considering quantum superpositions of different spacekime geometries. The Wheeler-DeWitt equation, which in conventional quantum gravity describes the wave function of the universe, can be extended to \(\hat{H}^\kappa \Psi[G_{AB}] = 0,\) where \(\Psi[G_{AB}]\) is the wave function of the spacekime geometry, and \(\hat{H}^\kappa\) includes contributions from both quantum mechanics and general relativity.

The kime-phase distribution \(\Phi(\varphi)\) plays a critical role in mediating between classical and quantum behavior.

  • Classical Limit: In regions where the kime-phase distribution is sharply peaked (e.g., a delta function), the spacekime framework reduces to classical general relativity.
  • Quantum Regime: For broader or more complex kime-phase distributions (e.g., Gaussian or Laplace), quantum effects dominate, leading to a richer structure of spacekime curvature that reflects quantum uncertainty and superposition.

To validate the spacekime unification of quantum mechanics and gravity, we can propose the following tests.

  • Predictions of Quantum Superpositions: Analyze how quantum states evolve in spacekime and compare with predictions of conventional quantum mechanics and quantum gravity.
  • Gravitational Effects on Quantum Systems: Explore how spacekime-induced modifications to spacetime curvature affect quantum systems, particularly in strong gravitational fields or near black holes.
  • Experimental Data Fitting: Use high-precision measurements of gravitational waves, black hole dynamics, or cosmological observations to test the predictions of the spacekime framework against empirical data.

12 Complex Time Representation of Observability with Applications to Quantum Gravity and Artificial Intelligence

12.1 Introduction

12.1.1 Background/Motivation

Observability of Spacetime Events: …

Many observable physical systems can be modeled by partial differential equations (PDEs) at different scales and regimes. For instance, the Schrödinger equation is the cornerstone of quantum mechanics, applicable in non-relativistic settings, while Einstein’s Field Equations extend the description of physics to include the effects of gravity in spacetime. Another notable PDE is the Wheeler-DeWitt equation, which attempts to reconcile quantum dynamics and gravity in describing the fabric of spacetime itself. Predictions of such PDE models do not agree globally, yet they advance the quest for a quantum gravity theory that encapsulates both quantum mechanics and general relativity.

  • Schrödinger equation … In quantum mechanics, the Schrödinger equation governs the time evolution of a quantum state in a non-relativistic framework. It is expressed as \(i\hbar \frac{\partial \psi(x,t)}{\partial t} = \hat{H} \psi(x,t)\), where \(\psi(x,t)\) is the wavefunction, \(\hat{H}\) is the Hamiltonian operator, and \(\hbar\) is the reduced Planck constant. This equation describes how the quantum state \(\psi\) evolves over time \(t\) within a given spatial configuration \(x\). The Schrödinger equation is fundamental in quantum mechanics, providing a probabilistic framework where the square of the wavefunction’s amplitude represents the probability density of finding a particle in a specific state.

  • Einstein’s Field Equations in Spacetime … Einstein’s Field Equations in general relativity describe the curvature of spacetime due to the presence of mass and energy. The equations are given by \(R_{\mu\nu} - \frac{1}{2}g_{\mu\nu}R = \frac{8\pi G}{c^4}T_{\mu\nu}\), where \(R_{\mu\nu}\) is the Ricci curvature tensor, \(g_{\mu\nu}\) is the metric tensor, \(R\) is the Ricci scalar, \(T_{\mu\nu}\) is the stress-energy tensor, \(G\) is the gravitational constant, and \(c\) is the speed of light. These equations relate the geometry of spacetime to the distribution of matter and energy, forming the backbone of our understanding of gravitational phenomena.

  • Wheeler-DeWitt equation in Spacetime … The Wheeler-DeWitt equation emerges from the attempt to unify quantum mechanics and general relativity in the context of quantum gravity. It is a quantum analogue of Einstein’s Field Equations and is expressed as \(\hat{H}\Psi[h_{ij}] = 0\), where \(\Psi[h_{ij}]\) is the wavefunction of the universe, and \(\hat{H}\) is the Hamiltonian constraint operator derived from the Einstein-Hilbert action. The Wheeler-DeWitt equation removes the time parameter, treating the universe as a static quantum system, encapsulating the idea that time is emergent rather than fundamental. This equation seeks to describe the quantum state of the entire universe, connecting the quantum and gravitational realms.

Quantum gravity: …

Artificial Intelligence (AI): …

Complex Time and Spacekime framework: …

  • definition …
  • math, physics, and AI/DS motivation …
  • spacetime foliation of spacekime …

5D Space-Time-Matter representation …

Ultrahyperbolic PDEs: …

12.1.2 Spacekime Framework

Observability, Spacekime \(\sigma\)-algebra, Probability, Events (Skevents), Information, and Entropy: …

Spacekime Density Matrix, Spacekime Metric, and Spacekime Entropy of Fluctuations: …

Energy Conditions for the Fields Equation Solutions based on Schwarzschild spacetime metric and Laplace Kime-Phase Distribution: …

Einstein’s Field Equations in Spacekime and ADM Formalism in Spacekime: … - Non-Trivial Field Equation Solution in Spacekime …

Wheeler-DeWitt equation in Spacekime: …

Quantum gravity: …

Artificial Intelligence (AI): …

12.1.3 Applications

12.1.3.1 Quantum gravity

The spacekime framework extends the traditional spacetime to complex time, \(\kappa = t e^{i\varphi}\), where \(\varphi\) represents a kime-phase with a given probability distribution \(\Phi(\varphi)\). Spacekime treats complex-time and space on an equal footing, which may contribute to unifying quantum mechanics and general relativity.

In traditional quantum field theory, quantum fluctuations are typically described in terms of real-valued time. The spacekime framework, however, allows for a more sophisticated treatment of these fluctuations by incorporating the kime-phase. Quantum fluctuations in the gravitational field can be modeled using a prior kime-phase distribution \(\Phi(\varphi)\). Modified gravitational wave detection may be used to test some of these hypotheses. Gravitational wave detectors, such as LIGO and Virgo, could be adapted to test predictions of the spacekime framework. If gravitational waves carry a signature of the complex time dimension, their waveforms might exhibit deviations from predictions made by general relativity. These deviations could manifest as phase shifts or amplitude modulations that depend on the underlying kime-phase distribution. Experimental tests would involve analyzing the data from gravitational wave events to identify any such anomalies.

One of the longstanding problems in quantum gravity is the black hole information paradox, related to the apparent loss of information in black holes as they evaporate via Hawking radiation. The spacekime framework could offer a new perspective by allowing the evolution of quantum states in the vicinity of a black hole to be described in terms of complex time. This might lead to a resolution where the kime-phase distribution plays a role in encoding or retrieving the lost information, thus preserving the unitarity of quantum mechanics. Although direct tests of quantum gravity are challenging due to the required energy scales, the spacekime framework could potentially be tested indirectly through precision measurements of quantum systems in gravitational fields. For instance, experiments using ultra-cold atoms or optical lattices in a controlled gravitational environment might reveal subtle effects predicted by the spacekime framework, such as kime-phase-induced shifts in energy levels or decoherence rates.

Also, the spacekime framework’s predictions could be tested through cosmological observations. For example, the distribution of galaxies and large-scale structures in the universe might encode information about the initial kime-phase distribution. By comparing these observations with simulations that incorporate the spacekime framework, researchers could validate or falsify the theory. Various
framework predictions of specific spectral features in the CMB or the distribution of dark matter could be tested using astronomical data. Here are a pair of experiments that can be conducted to test such spacekime framework predictions.

  • Experiment 1: Spectral Feature Analysis in the Cosmic Microwave Background (CMB) This experiment aims to validate the spacekime framework by detecting predicted spectral anomalies in the Cosmic Microwave Background (CMB) radiation that are attributed to the kime-phase distribution \(\Phi(\varphi)\) influencing the early universe’s quantum fluctuations.

    • Utilize high-resolution CMB data from telescopes like the Planck satellite, the South Pole Telescope (SPT), or the Atacama Cosmology Telescope (ACT). Focus on the power spectrum analysis of the CMB temperature fluctuations and polarization patterns.
    • Develop a theoretical model within the spacekime framework that predicts how the kime-phase distribution \(\Phi(\varphi)\) would affect the CMB power spectrum. This model should include modifications to the standard inflationary perturbation equations to account for complex-time effects.
    • Implement a search algorithm to identify spectral features predicted by the spacekime model in the CMB data. This involves looking for specific anomalies such as phase shifts in the multipole moments (l) or unexpected correlations in the polarization data.
    • Compare the observed spectral features with the predictions made by the spacekime model. Use statistical methods to quantify the significance of any detected anomalies, such as Bayesian model comparison or likelihood ratio tests.
    • A positive validation would be finding a statistically significant match between the observed spectral anomalies and the spacekime model predictions, suggesting that the kime-phase distribution indeed influences the early universe.
  • Experiment 2: Dark Matter Distribution Analysis in Large-Scale Structures This experiment focuses on detecting predicted deviations in the distribution of dark matter within large-scale structures of the universe, influenced by the complex time dimension and the associated kime-phase distribution.

    • Gather large-scale structure data from surveys such as the Sloan Digital Sky Survey (SDSS), the Dark Energy Survey (DES), or upcoming data from the Euclid mission and the Vera Rubin Observatory (LSST). Focus on the three-dimensional distribution of galaxies, galaxy clusters, and the cosmic web.
    • Extend the standard N-body simulations of dark matter by incorporating the spacekime framework. This involves modifying the equations governing gravitational collapse and clustering to account for the kime-phase distribution.
    • Examine the distribution of dark matter halos and subhalos in the simulations, looking for specific deviations predicted by the spacekime framework. This could include changes in the density profiles, halo mass functions, or the clustering properties of halos.
    • Compare the simulation results with the observed distribution of galaxies and galaxy clusters. Use techniques such as the two-point correlation function, power spectrum analysis, and weak gravitational lensing to detect any discrepancies that align with spacekime predictions.
    • Validation would involve detecting a significant deviation in the observed dark matter distribution that is consistent with the spacekime-modified simulations but not with standard cold dark matter models. Such a finding would suggest that the kime-phase distribution has a tangible impact on the large-scale structure of the universe.

12.1.3.2 Artificial Intelligence (AI)

The spacekime framework represents spatiotemporal processes as parametric manifolds (kimesurfaces), which can be modeled, analysed, and classified using new AI techniques. By lifting the time dimension to a complex plane, the spacekime framework constructs information-rich and mathematically tractable objects (kimesurfaces). This promotes the development of innovative AI algorithms and advanced statistical learning techniques. Let’s explore some concrete applications of this framework in artificial intelligence.

  • Enhanced Anomaly Detection in Multivariate Time-Series: In fields like finance, healthcare, and industrial monitoring, detecting anomalies in multivariate time-series is crucial. Traditional methods for high dimensional data analysis rely on linear models or additive noise assumptions, especially when the anomalies are subtle or evolve over time.

In spacekime, multivariate time-series data are mapped to kimesurfaces, where repeated measurement time-courses are transformed into 2D parametric spaces. Kimesurfaces capture the intrinsic variability of the data, allowing for a richer and more detailed representation. Anomalies that may not be apparent in the original time-series can manifest as distinct geometric features or topological changes in the kimesurface. AI algorithms can then be developed to detect these features using techniques from topological data analysis, manifold learning, and deep learning.

  • Advanced Predictive Modeling in Healthcare: Predicting patient outcomes based on longitudinal health records, such as electronic health records (EHRs) or wearable device data, is a challenging task due to the heterogeneity and temporal dynamics of the data.

The spacekime framework can transform patient health trajectories into kimesurfaces, where each point on the surface represents a patient’s health state at a given time, modulated by the kime-phase. This transformation allows for the use of novel AI algorithms that operate on the kimesurface to predict future health states. For example, convolutional neural networks (CNNs) designed to operate on kimesurfaces can be employed to capture spatial-temporal dependencies and interactions between different health indicators. Additionally, kimesurfaces can reveal hidden patterns, such as the progression of a disease or the response to treatment, that are not visible in the original time-series. This approach enables more accurate and interpretable predictive models, which can lead to improved clinical decision-making.

show examples from TCIU Tutorial

  • Spatiotemporal Event Forecasting in Climate Science: Forecasting complex spatiotemporal events, such as extreme weather patterns, climate change impacts, or ecological phenomena, requires sophisticated models that can account for the intricate interactions between variables across both space and time.

Climate data, such as temperature, precipitation, or atmospheric pressure, can also be represented as kimesurfaces. These surfaces encode the temporal evolution and spatial distribution of the variables, modulated by the kime-phase distribution. Machine learning models, such as recurrent neural networks (RNNs) or generative adversarial networks (GANs), can be adapted to learn from these kimesurfaces, capturing the complex, non-linear dependencies between different climatic factors. This approach allows for the development of more accurate forecasting models that can predict extreme weather events, such as hurricanes or droughts, with higher precision.

  • Dynamic Network Analysis in Social Media and Communication Networks: Analyzing dynamic networks, such as social media interactions, communication patterns, or information dissemination, is essential for understanding influence, community formation, and the spread of information or misinformation.

In this context, dynamic networks can be represented as kimesurfaces, where the nodes and edges of the network evolve over time and are influenced by the kime-phase. Each kimesurface encodes the temporal dynamics of the network, such as the formation and dissolution of connections, as well as the intensity and frequency of interactions. AI algorithms designed to operate on kimesurfaces, such as graph neural networks (GNNs) or topological data analysis methods, can be used to analyze these dynamic networks. The kimesurface representation allows for the detection of emerging communities, the identification of key influencers, and the prediction of information cascades. Additionally, the kimesurface framework can be used to model and simulate the impact of interventions, such as content moderation or targeted messaging, on the dynamics of the network.

12.2 Discussion

Falsifiability: …

13 Notes

13.1 Gravitational Fields

Spacekime appears as a gravitational field, which contracts and expands according to various forces, e.g,. energy, gravity. It is exists by itself even without matter. If spacetime is considered as a painting on a canvas, e.g., Leonardo’s Mona Lisa, spacekime is like a rich carousel of art paintings, a superimposition of canvases/strata. By this analogy, browsing over that carousel, events are viewers’ art impressions, which are expected to very widely between observers, even when time and space is fixed. These events reflect particle disturbances or interactions of nonuniform gravitational field, electromagnetic field, and the strong and weak nuclear fields, which flex, stretch, jostle, and interact with one-another in spacekime. Equations of motion in spacekime should describe the reciprocal influences of all the fields on one-another, and spacekime is one of these fields.

All of the field can be smooth and flat (Euclidean), or nonlinear leading to ripples, waves, strings, and other geometric contractions, expansions, and oscillations that preserve the spacekime topology, and respect the metric, \(G_{AB} \, dX^A \, dX^B\).

13.2 Discrete Quantum Time

In quantum mechanics, granularity of quantities (variables and measurements) is paramount. For the spatial resolution for space, the Planck scale is \(\Delta X > 10^{-33} cm\). For the gravitational field, there are minimum time-interval, \(\Delta t=\Delta ||\kappa ||>0\), called Planck time whose value can be estimated as a function of the fundamental physics constants characterizing phenomena/events subject to relativity, gravity, and quantum mechanics. The minimal Planck time interval is \(10^{-44}\) seconds. This may sound a bit counterintuitive, as the “time-quantization” implies that in a measure-theoretic sense, almost all values of time do not exist! Measurable, interpretable, and real time intervals (increments) may be (approximately) dense in \(\mathbb{R}^+\), but their measure is \(\mu(time\ increments)=0\). In otherwise, time is not observable.

Quantum mechanics is also indeterminant, i.e., it is not possible to predict exactly, where an electron will appear at a future time, \(t_o + \Delta t\). After a non trivial time interval, the electron does not have a precise position, as it is a generalized function, i.e., a distribution, which may allow the quantification of the probability that the electron is within a certain spatial domain.

Time fluctuations do not imply that all events are never determined. Rather, fluctuations suggest that time may be probabilistically determined only at certain moments. Observable indeterminacy is resolved when a quantity interacts with other objects, fields, or physical apparatus. Particle materialization quantifies its instantaneous property via a concrete measurement of a specific relation. In a statistical sense, probability distributions are the analogues of physical objects, fields, and processes. Whereas data, repeated measurements, and observed quantities are samples, draws, and instantiations from the corresponding probability distributions. Before we draw a random sample, any statistic is a random quantity with a specific probability distribution. The act of sampling (i.e., measurement) generates concrete observations from the process (i.e., physical system). Similarly, quantum systems are distributions and physical observations lead to concrete measurements, which are used for estimation, evidence-based decision-making, prediction, classificaiton, etc.

From spacekime perspective, there is no single (universal) time or kime. For every kime-interval, there are only different kime-durations, $= -_o= t_1 e^{i_1} - t_o e^{i_o} $, without loss of generality \(t_o = 0\Longrightarrow \Delta \kappa= t_1 e^{i\varphi_1}\). In essence, repeated sampling of the process over spacetime traverses the kime domain (across time and kime-phase), where the kime-phase distribution captures the intrinsic (quantum fluctuations) process variability. In practice, all data-driven observatinal scientific decision making pools across multipe repeats (large samples) to reduce the uncertainty of estimates and predicitons as blurred (agregate) view from the entire sample.

Kime annihilates all of the classical time notions, including order, singularity, direction, past-present-future, continuity. Spacekime a network of events whose interactions are tracked, quantified, modeled, and interpreted by aggregating repeated measurement observations (samples). The obscure lack of and explicit time variable in the fundamental model equations does suggest a static system. Rather, it implies a constant, ubiquitous, and unordered change. Events are not uniformly queued, but chaotically appear and disappear, evolve and change, morph and diffuse.

The difference between elementary particles (things) and events is that particles are persistent and they interact with other particles or disturb various quantum fields, whereas events are the outcomes of specific interactions and have limited duration. mountain. A wave is not a thing, it is a movement of water (field), and the water particles the wave disrupts during its motion vary according to the wave’s spatiotemporal localization.

13.3 Existence of a kime operator and the observability of kime-phases

13.3.1 Arguments For the Existence of a Kime Operator

  • Quantum Observables: In quantum mechanics, physical observables such as position and momentum are associated with operators. If we view kime-phases as part of a broader “kime-space” framework, similar to how time and space are unified in spacetime, a kime operator would naturally be the entity associated with measuring or predicting the evolution of kime-phases.
  • Complex-Time Representation: As described in complex-time theories, time (tt) is treated as a component of a complex-valued variable κ=teiφκ=teiφ. Given that quantum observables often exhibit non-commutative properties (e.g., Heisenberg’s uncertainty principle), it stands to reason that if kime-phases (φφ) are fundamental in describing physical reality, a corresponding operator should exist to capture their dynamics and non-commutative properties.
  • Statistical Inference in Complex Time: The PDFs you provided hint at the utilization of complex numbers and distributions in predictive analytics. If kime-phases play a key role in such a model, their measurement could be formalized through an operator that interacts with time-based distributions and events (skevents) in the spacekime \(\sigma\)-algebra, leading to observables in the sense of probability theory.

13.3.2 Argument Against the Existence of a Kime Operator

  • Non-Observability of Complex Phases: In quantum theory, phases (such as those in wave functions) are often not directly observable. Similarly, kime-phases could be argued as an abstract, non-observable component of time, functioning more like a statistical distribution than a physical observable. This would imply that kime-phases do not require an operator because they cannot be measured in the same sense as position or momentum.
  • Relativity Constraints: General relativity does not accommodate complex time as part of its framework, which complicates the notion of observability in spacekime. If kime-phases are distributions over time rather than time-like variables, traditional measurement operators might not apply, and instead, kime-phase would be an emergent property that cannot be measured directly, only inferred.

Further explorations of these aspects require a deeper delve into the role of skevents and the definition of spacekime entropy, examining how quantum mechanics, relativity, and complex phase distributions interact in this expanded framework.

13.4 Kime-Phase Dynamics - Tracking the Kime-Phase Changes

If there is no kime-operator and the kime-phases are intractable, much like the phases of quantum mechanics wavefuntions, perhaps we can can theoretically examine the kime-phase dynamics by exploring the kime-phase changes between successful repeated-measurement observations.

Lemma: Suppose \(\varphi\sim \Phi\) and \(\psi\sim \Psi\) are a pair kime-phases (random variables) corresponding to two (potentially different) phase distributions, \(\Phi, \Psi\), with densities \(f_\Phi(\varphi),f_\Psi(\psi)\). Note that if these kime phases correspond to different spatiotemporal locations, the phase distributions may be identical (\(\Phi \overset{a.s.}{=} \Psi\)) or be different (\(\Phi \overset{a.s.}{\not=} \Psi\)), see the TCIU Section on Reproducing Kernel Hilbert Spaces (RKHS) and Temporal Distribution Dynamics (TDD). Then, the kime-phase change is also a random variable, \(\theta=\varphi - \psi\) whose probability density function is the convolution of the original kime-phase probability density distributions

\[f_{\Theta}(\theta) = \int_{-\infty}^{\infty} f_\Phi(\varphi) f_\Psi(\nu-\varphi)d\varphi\ .\] Proof: The distribution of the difference between two independent random variables is given by the convolution of their probability density functions (PDFs) due to the properties of how sums (and differences) of random variables behave.

The convolution reflects how the two independent random variables interact when combined (either by addition or subtraction). To explicate the probability of a particular value for the difference (or sum) of two independent random variables, we need to account for all the possible ways this value could be obtained from different combinations of the individual random variables.

Suppose \(\varphi\) and \(\psi\) are two independent random variables with probability density functions \(f_\varphi(x)\) and \(f_\psi(x)\), respectively. We are interested in the PDF of their difference \(\theta = \varphi - \psi.\)

The cumulative distribution function (CDF) of \(\theta\) is defined by \(F_\theta(z) = P(\theta \leq z) = P(\varphi - \psi \leq z).\) Since \(\theta = \varphi - \psi\), we want to sum over all values of \(\varphi\) and \(\psi\) such that the difference \(\varphi - \psi\) is less than or equal to \(z\). This leads to an integral over all possible values of \(\varphi\) and \(\psi\)

\[F_\theta(z) = \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} I(\varphi - \psi \leq z) f_\varphi(\varphi) f_\psi(\psi) \, d\varphi \, d\psi\ ,\]

where \(I(\cdot)\) is the indicator function, which is \(1\), when the condition inside it is true, and \(0\) otherwise. To find the probability density function (PDF) of \(\theta\), we take the derivative of the CDF with respect to \(z\) \(f_\theta(z) = \frac{d}{dz} F_\theta(z).\)

A change variables simplifies the computation. Let \(w = \varphi - \psi\). Then, we integrate over the joint probability density of \(\varphi\) and \(\psi\) to get the PDF of \(w\) (or \(\theta\))

\[f_\theta(\theta) = \int_{-\infty}^{\infty} f_\varphi(x) f_\psi(\theta + x) \, dx.\]

This is precisely the convolution of the pair of density functions \(f_\varphi(x)\) and \(f_\psi(x)\). The convolution integral sums over all possibilities, weighting them by the individual densities of \(\varphi\) and \(\psi\). \(\Box\)

Let’s consider an example where \(\varphi, \psi\) are a pair kime-phases (random variables) corresponding to two phase distributions, \(\Phi, \Psi\). Let’s explicate the distribution of their difference, \(\theta = \varphi - \psi\), when both distributions are Laplace with different locations and scale parameters.

In the general case, assume:

  • \(\varphi \sim \text{Laplace}(\mu_\varphi, b_\varphi)\), i.e., \(\varphi\) follows a Laplace distribution with location parameter \(\mu_\varphi\) and scale parameter \(b_\varphi\),
  • \(\psi \sim \text{Laplace}(\mu_\psi, b_\psi)\), i.e., \(\psi\) follows a Laplace distribution with location parameter \(\mu_\psi\) and scale parameter \(b_\psi\).

The Laplace distribution has the probability density function (PDF)

\[f(x| \mu, b) = \frac{1}{2b} \exp\left(-\frac{|x - \mu|}{b}\right).\]

We are interested in the distribution of \(\theta = \varphi - \psi\).

Since \(\varphi\) and \(\psi\) are independent, the PDF of the difference \(\theta = \varphi - \psi\) is the convolution of the two Laplace distributions

\[f_\Theta(\theta) = \int_{-\infty}^{\infty} f_\Phi(\varphi) f_\Psi(\theta + \varphi) \, d\varphi.\]

This convolution of two Laplace distributions results in another Laplace distribution with location parameter \(\mu_\varphi - \mu_\psi\), and scale parameter: \(b_\varphi + b_\psi\). Thus, the distribution of \(\theta = \varphi - \psi\) is

\[\theta \sim \text{Laplace}(\mu_\varphi - \mu_\psi, b_\varphi + b_\psi).\]

In a special case where \(\mu_\varphi = \mu_\psi = 0\), \(b_\varphi = 1\), and \(b_\psi = 2\)

  • \(\varphi \sim \text{Laplace}(0, 1)\)
  • \(\psi \sim \text{Laplace}(0, 2)\)

Plugging in the general case, the difference \(\theta = \varphi - \psi\) will follow a Laplace distribution with a location parameter \(\mu_\varphi - \mu_\psi = 0 - 0 = 0\) and a scale parameter: \(b_\varphi + b_\psi = 1 + 2 = 3\). Thus, in this special case, the PDF of \(\theta\) is

\[f_\theta(\theta) = \frac{1}{6} \exp\left(-\frac{|\theta|}{3}\right).\]

This means \(\theta\) follows a Laplace distribution with location \(0\) and scale \(3\), \(\theta \sim \text{Laplace}(0, 3).\)

Hence, we can track the dynamics of the kime-phase chance by the distribution of the difference \(\theta = \varphi - \psi\), where \(\varphi \sim \text{Laplace}(\mu_\varphi, b_\varphi)\) and \(\psi \sim \text{Laplace}(\mu_\psi, b_\psi)\), which is just a Laplace-distributed with location \(\mu_\varphi - \mu_\psi\) and scale \(b_\varphi + b_\psi\)

\[\theta \sim \text{Laplace}(\mu_\varphi - \mu_\psi, b_\varphi + b_\psi).\]

13.5 Stochastic Spacekime Metric Tensor Formulation

To define the spacekime interval using the change in kime-phase between two consecutive samples, we consider the difference in kime-phases \(\theta = \varphi - \psi\), where \(\varphi \sim \Phi\) and \(\psi \sim \Psi\) are independent random variables corresponding to phase distributions \(\Phi\) and \(\Psi\) with probability density functions \(f_\Phi(\varphi)\) and \(f_\Psi(\psi)\), respectively.

The probability density function of \(\theta\) is given by the convolution of \(f_\Phi\) and \(f_\Psi\)

\[f_\Theta(\theta) = \int_{-\infty}^{\infty} f_\Phi(\varphi) f_\Psi(\theta + \varphi) \, d\varphi .\]

We need to be careful when defining the square-interval component corresponding to the expectation of the square phase change, \(\mathbb{E}[ \left( f_\Theta(\theta) \right)^2 ]\), is incorrect. It’s better to define this term by \(\mathbb{E}[ \theta^2 ]\), where the expectation is taken with respect to the distribution of \(\theta\), \(f_\Theta(\theta)\).

The expected squared interval is \[\mathbb{E}[ds^2] = -c^2 t^2 - c^2 \mathbb{E}[ \theta^2 ] + dx^2 + dy^2 + dz^2,\]

where, \(\theta = \varphi - \psi\) represents the change in kime-phase between two consecutive samples, and \(\mathbb{E}[ \theta^2 ]\) captures the expected squared phase difference, dispersion, incorporating the variability from the phase distributions.

We will compute \(\mathbb{E}[ \theta^2 ]\) for two cases - Uniform Phase Distribution and Laplace Phase Distribution.

Case 1: Uniform Phase Distribution: Let \(\varphi\) and \(\psi\) be independent random variables uniformly distributed over \([0, 2\pi]\)

\[f_\Phi(\varphi) = f_\Psi(\psi) = \frac{1}{2\pi}, \quad 0 \leq \varphi < 2\pi .\]

The mean of \(\varphi\) and \(\psi\) is \[\mathbb{E}[\varphi] = \mathbb{E}[\psi] = \mu = \frac{1}{2\pi} \int_0^{2\pi} \varphi \, d\varphi = \pi .\]

The variance of \(\varphi\) and \(\psi\) is \[\operatorname{Var}(\varphi) = \operatorname{Var}(\psi) = \sigma^2 = \frac{1}{2\pi} \int_0^{2\pi} (\varphi - \pi)^2 \, d\varphi = \frac{(2\pi)^2}{12} = \frac{\pi^2}{3} .\]

Since \(\theta = \varphi - \psi\), and \(\varphi\) and \(\psi\) are independent, the mean and variance of \(\theta\) are \[\mathbb{E}[\theta] = \mathbb{E}[\varphi] - \mathbb{E}[\psi] = \pi - \pi = 0 .\] \[\operatorname{Var}(\theta) = \operatorname{Var}(\varphi) + \operatorname{Var}(\psi) = 2 \sigma^2 = 2 \left( \frac{\pi^2}{3} \right) = \frac{2\pi^2}{3} .\]

Therefore, \[\mathbb{E}[ \theta^2 ] = \operatorname{Var}(\theta) + \left( \mathbb{E}[\theta] \right)^2 = \operatorname{Var}(\theta) = \frac{2\pi^2}{3} .\]

Substituting \(\mathbb{E}[ \theta^2 ]\) into the spacekime interval we obtain \[\mathbb{E}[ds^2] = -c^2 t^2 - c^2 \left( \frac{2\pi^2}{3} \right) + dx^2 + dy^2 + dz^2 .\]

Case 2: Laplace Phase Distribution: Let \(\varphi\) and \(\psi\) be independent random variables following a Laplace distribution centered at zero with scale parameter \(b > 0\) \[f_\Phi(\varphi) = f_\Psi(\psi) = \frac{1}{2b} e^{-\frac{|\varphi|}{b}} .\]

The mean of \(\varphi\) and \(\psi\) is \(\mathbb{E}[\varphi] = \mathbb{E}[\psi] = 0\) and the variance of \(\varphi\) and \(\psi\) is \(\operatorname{Var}(\varphi) = \operatorname{Var}(\psi) = 2b^2.\)

Since \(\theta = \varphi - \psi\), and \(\varphi\) and \(\psi\) are independent, the mean and variance of \(\theta\) are \(\mathbb{E}[\theta] = \mathbb{E}[\varphi] - \mathbb{E}[\psi] = 0 - 0 = 0\) and \(\operatorname{Var}(\theta) = \operatorname{Var}(\varphi) + \operatorname{Var}(\psi) = 2b^2 + 2b^2 = 4b^2 .\)

Therefore, \(\mathbb{E}[ \theta^2 ] = \operatorname{Var}(\theta) = 4b^2\).

Plugging in \(\mathbb{E}[ \theta^2 ]\) into the spacekime interval we obtain \[\mathbb{E}[ds^2] = -c^2 t^2 - c^2 (4b^2) + dx^2 + dy^2 + dz^2 .\]

Using \(\mathbb{E}[ \theta^2 ]\) in the definition of the spacekime interval directly incorporates the expected squared phase difference into the spacetime metric. This approach reflects how the variability of the kime-phase changes affects the geometry of spacetime, especially when considering processes with inherent randomness in phase due to repeated measurements.

For Uniform Distribution, the variance of the phase difference \(\theta\) is \(\frac{2\pi^2}{3}\), indicating a wide spread of possible phase differences due to the uniform nature of the distribution. And the expected squared interval includes a constant term \(-c^2 \left( \frac{2\pi^2}{3} \right)\), representing the average effect of the phase variability on spacetime intervals.

For Laplace Distribution, the variance of the phase difference \(\theta\) is \(4b^2\), which depends on the scale parameter \(b\) of the Laplace distribution. - The expected squared interval includes a term \(-4 c^2 b^2\), showing that the impact of phase variability on spacetime intervals increases with the scale parameter \(b\).

Question: Are there necessary and sufficient conditions to ensure that \(\mathbb{E}[\theta^2] =0\)?

Answer: A trivial phase contribution to ther expected interval, i.e., \(\mathbb{E}[ \theta^2 ] = 0\) implies that the expected squared phase difference \(\theta = \varphi - \psi\) is zero. In this case, the expected spacekime interval reduces to the standard classical spacetime interval with signature \((- + + +)\).

The expected value of \(\theta^2\) is \[\mathbb{E}[ \theta^2 ] = \operatorname{Var}(\theta) + \left( \mathbb{E}[\theta] \right)^2 .\]

Setting \(\mathbb{E}[ \theta^2 ] = 0\) implies that both the variance and the square of the mean of \(\theta\) must be zero, \(\operatorname{Var}(\theta) = 0\) and \(\mathbb{E}[\theta] = 0\). Because the variance of \(\theta\) is zero, \(\theta\) must be a constant almost surely (with probability \(1\)). Similarly, if \(\mathbb{E}[\theta] = 0\), then this constant must be zero. Therefore, the necessary and sufficient condition is \(\theta = \varphi - \psi = 0 \quad \text{almost surely}.\) This means that \(\varphi\) and \(\psi\) are equal almost surely.

If both \(\varphi\) and \(\psi\) are deterministic (non-random) and equal, \(\varphi = \psi = \phi_0 \quad \text{almost surely}\). This implies that their distributions \(\Phi\) and \(\Psi\) are degenerate distributions (Dirac delta functions) concentrated at the same point \(\phi_0\) \[f_\Phi(\phi) = \delta(\phi - \phi_0), \quad f_\Psi(\psi) = \delta(\psi - \phi_0).\] If both \(\varphi\) and \(\psi\) are identical random variables with zero variance, meaning they take the same constant value almost surely.

The necessary and sufficient condition for the phase component of the expected metric to be trivial (\(\mathbb{E}[ \theta^2 ] = 0\)) is that the kime-phases \(\varphi\) and \(\psi\) at the two spacekime locations are identical and deterministic, i.e., they are equal constants with zero variance. Under this condition, the expected spacekime interval reduces to the classical spacetime interval \(\mathbb{E}[ds^2] = -c^2 t^2 + dx^2 + dy^2 + dz^2 .\) This interval has the standard signature \((- + + +)\), consistent with Minkowski spacetime in special relativity. In physical terms, the condition implies that there is no randomness or uncertainty in the kime-phases at the two locations. This scenario corresponds to a perfectly deterministic system without any phase fluctuations due to measurement or inherent stochasticity. Even if \(\varphi\) and \(\psi\) are dependent random variables, unless they are almost surely equal, \(\operatorname{Var}(\theta)\) will not be zero. This result highlights that any variability in the kime-phases \(\varphi\) and \(\psi\) contributes to the phase component of the metric, making it non-trivial. Only in the idealized case of deterministic and identical phases does the expected spacekime interval reduce to the classical spacetime interval.

This definition of the spacekime metric tensor and corresponding expected squared interval provide a framework that accounts for the statistical properties of kime-phase variations. By incorporating \(\mathbb{E}[ \theta^2 ]\) into the metric, we capture the influence of phase randomness on spacetime geometry, allowing for a more comprehensive understanding of processes where time has a complex extension. Incorporating the expected squared phase difference \(\mathbb{E}[ \theta^2 ]\) into the spacekime interval provides a direct link between the statistical properties of the kime-phase and the structure of spacekime.

SOCR Resource Visitor number Web Analytics SOCR Email