SOCR ≫ | TCIU Website ≫ | TCIU GitHub ≫ |
Note: This TCIU Section in Chapter 6 extends previous work on Chapter 3 work on Radon-Nikodym Derivatives, Kimemeasures, and Kime Operator.
Spacekime analytics is an emerging mathematical and computational framework that extends classical spacetime models by lifting the concept of time into the complex domain. Unlike traditional time representations, complex-time (kime) incorporates both magnitude (longitudinal event ordering, sequence or duration) and phase (characterizing the variability in repeated longitudinal experiments). This novel approach enhances our ability to represent, analyze, predict, and infer patterns within temporally dynamic systems. By blending statistical, quantum, and AI methodologies, spacekime analytics addresses fundamental questions in the modeling of longitudinal and multi-dimensional data, particularly through its applications in AI-driven inference and decision-making.
Complex time is denoted by \(\kappa = t e^{i\varphi}\), where the kime-magnitude \(t\) is the classical time (event ordering index) and the kime-phase \(\varphi\) reflects the random sampling from the repeated measurement distribution \(\Phi_{[-\pi, \pi)}\). This natural complex-time extension of time necessitates the reformulation of kime and spacekime events, spacekime metric tensor \(g_{\mu\nu}\), expected square interval \(\mathbb{E}[ds^2]\), and other classical concepts based on this kime-phase distribution.
A rigorous mathematical formulation of spacekime analytics begins with defining the probability space \((\Omega, \mathcal{F}, P)\) over a spacekime manifold. This space allows for integration over kime-events, with the following foundational elements.
Sample Space, \(\Omega\): The set of all possible outcomes, including all spatial and kime coordinates, \((\mathbf{x}, \kappa)\), where \(\mathbf{x} \in \mathbb{R}^n\) represents spatial dimensions, and \(\kappa \in \mathbb{C}\) represents complex time. The representation of kime as \(\kappa = r e^{i\phi}\) where \(r > 0\) and \(\phi\) is the kime-phase, defines each kime-coordinate in terms of event ordering and directional shifts.
Sigma-Algebra, \(\mathcal{F}\): The collection of subsets of \(\Omega\), representing all possible measurable events (kime-events). In this complex space, \(\mathcal{F}\) is generated by sets of points in both the real and imaginary components of kime, ensuring compatibility with probabilistic operations defined over kime-surfaces.
Probability Measure, \(P\): A probability measure assigning likelihoods to events in \(\mathcal{F}\), considering the topological and metric properties of spacekime. For a kime-event \(E \subset \mathcal{F}\), \(P(E)\) quantifies the probabilistic weighting of occurrences in both spatial and kime dimensions. This measure integrates over \((\mathbf{x}, \kappa)\), yielding probabilities influenced by both the magnitude and phase distributions of kime.
In the spacekime framework, probability densities must account for periodic behaviors and cyclic dependencies introduced by kime-phase. Standard probability density functions \(f(\mathbf{x}, r, \phi)\) are adjusted to include terms in both \(r\) (radial time component) and \(\phi\) (angular phase), allowing for more comprehensive modeling of time-varying dynamics.
Formalizing the geometry of the spacekime manifold requires introducing an appropriate metric \(d_{kime}\), defined over \((\mathbf{x}, \kappa) \in \mathbb{R}^n \times \mathbb{C}\), that respects the complex nature of kime and supports well-defined norms for distance calculations.
To rigorously define the spacekime metric and the corresponding square interval in terms of a kime-phase distribution \(\varphi \sim \Phi_{[-\pi, \pi)}\), we need to account for the variability in the kime-phase that reflects the intrinsic variability of repeated measurements in the observable process. The metric should encapsulate both real-time evolution and stochastic kime-phase behavior, while maintaining the fundamental properties of a metric: non-negativity, symmetry, and adherence to the triangle inequality.
Capturing both the real-time and phase variability effects, the spacekime square interval \(ds^2\) is defined in terms of expected squared phase difference \(\mathbb{E}[\theta^2]\), where \(\theta\) represents the kime-phase difference between two consecutive measurements \(\mathbb{E}[ds^2] = -c^2 t^2 - c^2 \mathbb{E}[\theta^2] + dx^2 + dy^2 + dz^2\), where \(t\) is the real-time component, \(\theta = \varphi_1 - \varphi_2\) is the phase difference, where \(\varphi_1, \varphi_2 \sim \Phi\), and \(\mathbb{E}[\theta^2]\) is the expected squared phase difference under \(\Phi\).
Assuming that the kime-phases \(\varphi_1\) and \(\varphi_2\) are independently sampled from a comon distribution \(\Phi\) with mean zero, we the expected squared phase difference is \(\mathbb{E}[\theta^2] = 2 \operatorname{Var}(\varphi),\) where \(\operatorname{Var}(\varphi)\) is the variance of the kime-phase under \(\Phi\). For a uniform distribution on \([-\pi, \pi)\), for instance, we have \(\operatorname{Var}(\varphi) = \frac{\pi^2}{3}\), yielding \(\mathbb{E}[\theta^2] = \frac{2\pi^2}{3}.\)
The metric tensor \(g_{\mu\nu}\) corresponding to this square interval encapsulates the classical spacetime and the kime-phase variability. Given that \(\mathbb{E}[\theta^2]\) represents phase variability as a constant scalar term for a stationary \(\Phi\), we can define \(g_{\mu\nu}\) as follows in the \((t, x, y, z)\) coordinates \[g_{\mu\nu} = \begin{pmatrix} -c^2 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \end{pmatrix} ,\] with an effective time component in the interval given by \(\mathbb{E}[ds^2] = -c^2 t^2 - c^2 \mathbb{E}[\theta^2] + dx^2 + dy^2 + dz^2.\) This metric reflects the additional length induced by the expected phase difference variability in complex time. This definition satisfies the mathematical conditions of a metric.
In the spacetime geometry, non-negativity doesn’t imply that the interval must be positive; rather, it reflects the pseudo-Riemannian nature of the metric. A time-like interval can be negative as long as it maintains consistency with the causal structure. In our spacekime context, \(\mathbb{E}[ds^2]\) being negative reflects a time-like separation, consistent with relativistic interval interpretations. This structure captures phase variability as an intrinsic source of time-like separation in repeated measurements, where the negative contributions from \(t\) and \(\mathbb{E}[\theta^2]\) reflect time-evolution and phase uncertainty, respectively.
Symmetry: The metric \(g_{\mu\nu}\) is symmetric, ensuring the symmetry of distances between any two points in the spacekime manifold.
Triangle Inequality: The metric tensor, by incorporating a constant \(\mathbb{E}[\theta^2]\) as a phase-difference variability term, preserves the triangle inequality by preventing negative distances, maintaining consistent metric properties.
In spacekime, the real-time evolution is mapped to complex time, where phase variability is encoded as \(\mathbb{E}[\theta^2]\). This implies that each time step is associated with a corresponding phase variation that scales with \(t\), leading to a spread in complex time that is observable as measurement variability. Whereas the inverse transform reconstructs the classical time behavior by averaging over the phase variability, effectively removing the stochastic component while retaining the real-time structure.
This spacekime metric enables advanced distance-based analyses, such as clustering and manifold learning, by leveraging both spatial proximity and kime-phase coherence. The phase information in particular allows for representing and modeling recurrent processes as intrinsic components of data analysis, thereby enhancing AI models’ ability to interpret periodic or cyclic temporal behaviors.
Spacekime analytics draws conceptual and operational parallels with quantum mechanics, statistical estimation, and AI. This interdisciplinary foundation allows the framework to exploit the structural advantages of complex time, with applications in higher-dimensional data analysis and quantum-inspired AI models:
Quantum Mechanics: Complex time (kime) representation directly translates quantum mechanics principles such as wave-particle duality, phase coherence, and operator eigenfunctions into statistical inference. In spacekime, each observable corresponds to a complex event or kime-phase, analogous to quantum observables. The transformation of longitudinal observations into kime-surfaces echoes the Heisenberg uncertainty principle, where time-phase uncertainties offer dual interpretations in data science, improving the predictive performance and interpretability of models.
Statistical Estimation and Bayesian Inference: Spacekime analytics provides a Bayesian framework where posterior distributions are conditioned on kime-phase aggregates or estimations. This approach enables inference on complex-valued processes, with posterior predictive distributions expanded as \(p(\gamma | X, \phi) \propto p(X | \gamma, \phi) \cdot p(\gamma | \phi),\) where \(\phi\) represents the kime-phase information aggregated across observations. This Bayesian structure allows for probabilistic estimation with reduced sample sizes, making spacekime analytics especially useful in high-dimensional, low-sample environments.
AI and Machine Learning Analytics: By representing time-dependent data as kime-surfaces, spacekime analytics unlocks novel AI applications. Machine learning models can exploit these complex embeddings, allowing for richer feature spaces that improve clustering, classification, and forecasting tasks. Techniques such as kime-surface tensor decomposition provide powerful tools for understanding multivariate dependencies, facilitating the development of robust, invariant, and sample-efficient AI systems.
Domain-specific Applicaitons: Spacekime analytics holds significant promise across diverse scientific and applied domains. Three primary applications demonstrate its transformative potential. For instance, in neuroimaging, spacekime analytics enables the representation of fMRI or EEG time-series as kime-surfaces, capturing both the temporal and phase-specific dynamics of brain activity. Through kime-phase estimation, researchers can model patterns in brain connectivity that fluctuate over time, providing insights into cognitive processes and neurological conditions. This approach has proven invaluable in studying conditions such as Alzheimer’s disease, where the spacekime framework allows researchers to track disease progression and identify early biomarkers by modeling both the magnitude and phase of neural signals.
The application of spacekime analytics in biomedical signal processing allows for continuous, multi-dimensional monitoring of physiological signals, such as heart rate variability (HRV) and glucose levels. Kime-based metrics provide more accurate modeling of circadian rhythms and patient-specific variations, enabling real-time tracking of patient health and the early detection of anomalies. In personalized healthcare, AI-driven kime-based models support decision-making by integrating phase information, leading to dynamic adjustments in patient care plans based on nuanced temporal patterns.
In financial and economic forecasting, complex time is utilized to represent time-series data with high-frequency fluctuations and periodic trends. Spacekime analytics allows for enhanced modeling of cyclical trends, including economic cycles and seasonal market patterns. By leveraging kime-phase structures, these models capture complex correlations across assets and markets, providing better risk assessment and portfolio optimization. Quantum-inspired approaches allow for time-dependent probabilistic forecasts that are less sensitive to noise, supporting more robust decision-making in volatile markets.
The development of complex-time (kime) representation and spacekime analytics marks a milestone in data science, blending elements of quantum mechanics, advanced statistical inference, and AI. The spacekime framework, with its sigma-algebra structure, complex metric space, and robust AI applications, redefines the analytical landscape for high-dimensional, longitudinal data. Its interdisciplinary foundation and application versatility position spacekime analytics as a transformative tool for advancing scientific discovery and practical applications in diverse domains, from neuroscience to finance.
By facilitating complex temporal analyses, spacekime analytics not only enhances our understanding of dynamic systems but also opens up new research areas in AI and machine learning, ultimately paving the way for breakthroughs across multiple scientific and technological fields.
Following the notation in this description of General Relativity (GR) support for prediction of the precession of planitary orbits, we will explore the spacekime representation as another possible orbital prediciton strategy.
General Relativity (GR) predicts that the excess precession of the perihelion of Mercury’s orbit is approximately \(0.01^o\) per century. Mercury’s orbit precession had been perplexing for centuries, as Newton’s law of gravity can be used to model planetary orbits as closed ellipses with the Sun located at one focus of the corresponding ellipse orbit. Observations indicate that the major axes of these elliptical orbits are not stable and chance their direction with time. There are also gravitational interactions between planets also affect the precesional planetary orbits, complementing the primary solar gravitational force, which results in orbital motions that are not perfectly closed static ellipses but rather dynamic elliptical orbits within the solar orbital plane.
Ignoring the small interplanetary interactions, Newtonian gravity predicts that a planet’s orbit would be an idealized ellipse traversed by the planet at a distance from the Sun given by \(r = \frac{r_{min}}{1 + e \cos \phi}\), where \(r_{min}\) is the distance to the perihelion (the point of closest approach to the Sun) and \(e\) is the orbital eccentricity, representing the deviation from a circular orbit (\(e = 0\) for a circle).
The gravitational interactions between planets cause the perihelion to rotate. This perihelion precession results in a gradual rotation of the elliptical orbit’s major axis around the Sun.
Repeated astronomical observations of Mercury’s orbit suggest a perihelion precession of \(9.55\) arc minutes per century. Newtonian mechanics modeling that accounts for the gravitational interactions of other planets in the solar system predict a lower precession of \(8.85\) arc minutes per century, suggests a difference of \(\sim 0.7\) arc minutes, i.e., \(42\) arc seconds per century. However, observed astronomical measurements show a discrepancy of \(\sim 43.1\) arc seconds per century. General Relativity explains some of this inconsistency by modifying Newton’s gravitational model correcting the orbital elliptical equation by including the term \(\Delta\varphi\) (relativistic correction) \[r = \frac{r_{min}}{1 + e \cos(\phi - \Delta \phi)}.\]
According to general relativity \[\Delta \phi = \frac{6 \pi G M}{c^2 (1 - e^2) R},\] where \(G\) is the gravitational constant, \(M\) is the mass of the Sun, \(c\) is the speed of light, \(e\) is the orbital eccentricity, and \(R\) is the semi-major axis of the orbit.
Mercury’s small orbit is highly eccentric and hte planet has the largest relativistic precession among the 8 planets in the solar system. Plugging Mercury’s values into the GR-corrected equation model yields a precession of \(43.0\) arc seconds per century, which agrees with the observed value of \(43.1 \pm 0.5\) arc seconds.
The following table shows the observed and predicted precession rates for several planets:
Planet | Orbits.per.century | Eccentricity | r_min.AU | GR.Predicted.arc.sec.per.century | Observed.arc.sec.per.century |
---|---|---|---|---|---|
Mercury | 415.2 | 0.2060 | 0.307 | 43.0 | \(43.1 \pm 0.5\) |
Venus | 162.5 | 0.0068 | 0.717 | 8.6 | \(8.4 \pm 4.8\) |
Earth | 100.0 | 0.0170 | 0.981 | 3.8 | \(5.0 \pm 1.2\) |
Icarus* | 89.3 | 0.8270 | 0.186 | 10.0 | \(9.8 \pm 0.8\) |
Note: The perihelion of the asteroid Icarus
,
which is closer to the Sun than any other asteroid, is within Mercury’s
orbit.
The precession of the perihelion of Mercury’s orbit is one of the classic tests of General Relativity (GR). In this derivation, we’ll show how GR predicts an additional precession of Mercury’s orbit that matches the observed anomaly not explained by Newtonian mechanics. Mercury’s orbit around the Sun exhibits a precession of its perihelion (the point of closest approach to the Sun) that cannot be fully explained by Newtonian mechanics and the gravitational perturbations from other planets.
The observed precession is about 43 arcseconds per century greater than what Newtonian physics predicts when accounting for perturbations from other planets. In Newtonian gravity, the orbit of a planet around the Sun is described by Kepler’s laws, resulting in elliptical orbits with no inherent precession (ignoring perturbations).
The equation of motion is \(\frac{d^2 \vec{r}}{dt^2} = -\frac{G M}{r^2} \hat{r}\), where conservation of Angular Momentum leads to a constant areal velocity. Einstein’s theory that describes gravity as the curvature of spacetime caused by mass and energy. Using the Schwarzschild Metric, the solution to Einstein’s field equations for the spacetime outside a spherical, non-rotating mass like the Sun is \[ds^2 = -\left(1 - \frac{2GM}{c^2 r}\right)c^2 dt^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} dr^2 + r^2 d\theta^2 + r^2 \sin^2 \theta \, d\phi^2 .\]
The derivation of the Equation of Motion in Schwarzschild Spacetime uses the assumptions that a test particle (Mercury) moving in the equatorial plane (\(\theta = \frac{\pi}{2}\)), so \(d\theta = 0\) with a simplified spacetime metric \[ds^2 = -\left(1 - \frac{2GM}{c^2 r}\right)c^2 dt^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} dr^2 + r^2 d\phi^2 . \] The equations of motion are derived from the geodesic equations for the Schwarzschild metric. The energy per Unit Mass (\(\tilde{E}\)) is \(\left(1 - \frac{2GM}{c^2 r}\right) c^2 \frac{dt}{d\tau} = \tilde{E}\), the angular momentum per Unit Mass (\(\tilde{L}\)) is \(r^2 \frac{d\phi}{d\tau} = \tilde{L},\) and the proper time (\(\tau\)) is measured by a clock moving with the particle.
The effective potential and radial equation of motion depend on the metric relation \[-\left(1 - \frac{2GM}{c^2 r}\right) c^2 \left( \frac{dt}{d\tau} \right)^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} \left( \frac{dr}{d\tau} \right)^2 + r^2 \left( \frac{d\phi}{d\tau} \right)^2 = -c^2.\]
Substituting the conserved quantities we obtain \[-\left(1 - \frac{2GM}{c^2 r}\right)^{-1} \left( \frac{\tilde{E}}{c} \right)^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} \left( \frac{dr}{d\tau} \right)^2 + \left( \frac{\tilde{L}}{r} \right)^2 = -c^2 .\]
Next, we can simplify the radial equation \[\left( \frac{dr}{d\tau} \right)^2 = \left( \frac{\tilde{E}}{c} \right)^2 - \left(1 - \frac{2GM}{c^2 r}\right) \left( c^2 + \left( \frac{\tilde{L}}{r} \right)^2 \right) .\] The effective potential (\(V_{\text{eff}}\)) is \[\left( \frac{dr}{d\tau} \right)^2 + V_{\text{eff}}(r) = \left( \frac{\tilde{E}}{c} \right)^2 ,\] where \[V_{\text{eff}}(r) = \left(1 - \frac{2GM}{c^2 r}\right) \left( c^2 + \left( \frac{\tilde{L}}{r} \right)^2 \right) .\]
We can simplify the equations by changing the variables and using the inverse radius, \(u = \frac{1}{r}\). The derivatives transformation \(\frac{dr}{d\tau} = -\frac{1}{u^2} \frac{du}{d\tau} .\) Using \(\frac{d\phi}{d\tau}\) and the differentiation chain rule we get \[\frac{du}{d\tau} = \frac{du}{d\phi} \frac{d\phi}{d\tau} = \frac{du}{d\phi} \frac{\tilde{L}}{r^2} \] and plugging in \(r = 1/u\) yields \(\frac{du}{d\tau} = \tilde{L} u^2 \frac{du}{d\phi} .\)
Substitute this result into the radial equation and simplify \[\left( -\frac{1}{u^2} \tilde{L} u^2 \frac{du}{d\phi} \right)^2 + V_{\text{eff}}(r) = \left( \frac{\tilde{E}}{c} \right)^2\]
\[\tilde{L}^2 \left( \frac{du}{d\phi} \right)^2 + V_{\text{eff}}(r) = \left( \frac{\tilde{E}}{c} \right)^2 .\]
This equation can be expressed as \[\left( \frac{du}{d\phi} \right)^2 = \frac{1}{\tilde{L}^2} \left( \left( \frac{\tilde{E}}{c} \right)^2 - V_{\text{eff}}(r) \right) \]
Approximating \(V_{\text{eff}}(r)\) to obtain the weak gravitational fields \(\left ( \frac{GM}{c^2 r} \ll 1 \right )\)
\[V_{\text{eff}}(r) \approx c^2 + \left( \frac{\tilde{L}}{r} \right)^2 - \frac{2GM}{r} \left( c^2 + \left( \frac{\tilde{L}}{r} \right)^2 \right) .\]
Some simplification and differentiation w.r.t. \(\phi\) leads to \[\left( \frac{du}{d\phi} \right)^2 + u^2 = \frac{GM}{\tilde{L}^2} + 3 \frac{GM}{c^2} u^3 .\]
\[\frac{d^2 u}{d\phi^2} + u = \frac{GM}{\tilde{L}^2} + 3 \frac{GM}{c^2} u^2 .\]
The homogeneous solution to this differential equation is \(\frac{d^2 u}{d\phi^2} + u = 0\), and assuming a solution of the form \(u_p = \frac{GM}{\tilde{L}^2},\) its solution is \(u_h(\phi) = A \cos \phi + B \sin \phi.\)
Due to the nonlinearity of the quadratic term \(u^2\), we need to consider perturbation methods. Suppose \(u = u_0 + \delta u\), where \(u_0\) is the Newtonian solution and \(\delta u\) is a small relativistic correction. the Newtonial zero-order equation \(\frac{d^2 u_0}{d\phi^2} + u_0 = \frac{GM}{\tilde{L}^2}\) has a solution \(u_0(\phi) = \frac{GM}{\tilde{L}^2} \left(1 + e \cos \phi \right),\) where \(e\) is the orbital eccentricity.
A first-order correction uses \(u = u_0 + \delta u\) into the full equation to linearize \(\frac{d^2 \delta u}{d\phi^2} + \delta u = 3 \frac{GM}{c^2} u_0^2\) and compute \(u_0^2\). \[u_0^2 = \left( \frac{GM}{\tilde{L}^2} \right)^2 \left(1 + 2 e \cos \phi + e^2 \cos^2 \phi \right). \]
A partial solution for \(\delta u\) of the first-order differential equation \[\frac{d^2 \delta u}{d\phi^2} + \delta u = 3 \frac{GM}{c^2} \left( \frac{GM}{\tilde{L}^2} \right)^2 \left(1 + 2 e \cos \phi + e^2 \cos^2 \phi \right) \] can be expressed as \(\delta u = K_0 + K_1 \phi \sin \phi + K_2 \cos \phi.\) However, the term involving \(\cos^2 \phi\) leads to secular terms (terms that grow with \(\phi\)). To avoid these secular terms we can use the method of undetermined coefficients or variation of parameters to find a solution that remains bounded. Hence, \(\delta u\) contains a term proportional to \(\phi \sin \phi\) \[\delta u = \frac{3 (GM)^3}{c^2 \tilde{L}^4} e \phi \sin \phi + \text{bounded terms} .\]
the complete solution is \[u(\phi) = u_0(\phi) + \delta u(\phi) = \frac{GM}{\tilde{L}^2} \left(1 + e \cos \phi \right) + \delta u(\phi) ,\] where the term \(\delta u\) effectively causes the orbit to precess because it modifies the angular dependence. This solution suggests that the argument of the perihelion advances by a small angle \(\delta \phi\) per orbit.
The presence of the term \(\phi\) in \(\delta u\) indicates that the orbit is not exactly periodic with period \(2\pi\). The new period \(\Phi\) satisfies \(\Phi = 2\pi \left(1 + \Delta \right)\), where \(\Delta\) is a small correction. From the perturbation, the shift per revolution is \(\delta \phi = 2\pi \Delta = \frac{6\pi GM}{c^2 a (1 - e^2)}\), where \(a\) is the semi-major axis of Mercury’s orbit and \(e\) is the orbital eccentricity.
The numerical precision calculation for Mercury uses the following parameters Gravitational constant \(G\); Mass of the Sun \(M\); Speed of light \(c\); Semi-major axis \(a\); and Eccentricity \(e\). \[\delta \phi = \frac{6\pi GM}{c^2 a (1 - e^2)}\]
We can convert to arcseconds per century. Mercury completes approximately \(415\) revolutions per century and we multiply \(\delta \phi\) by \(415\) to get the total precession per century. Therefore, \(\delta \phi_{\text{total}} = 43'' \text{ (arcseconds per century)},\) which is a good prediction of the observed excess precession of Mercury’s perihelion.
Similarly, the Earth’s precession perihelion which is predicted by General Relativity (GR) to be \(3.8''\), whereas repeated observations estimate it to be \(5.0'' \pm 1.2''\). We follow a similar approach we used for Mercury, but with using Earth’s specific orbital parameters. The point of closest approach of a planet to the Sun (perihelion) shifts over time, resulting in a precession of the orbit. Newtonian Mechanics predicts some precession due to perturbations from other planets and General Relativity provides an additional correction to the precession not accounted for by Newtonian physics.
The relativistic correction to the perihelion precession per orbit is given by \(\delta\phi = \frac{6\pi G M}{c^2 a (1 - e^2)}\), where \(\delta\phi\) is Additional precession per orbit in radians; \(G\) is the gravitational constant; \(M\) is the Mass of the Sun; \(c\) is the Speed of light; \(a\) is the Semi-major axis of Earth’s orbit; and \(e\) is the orbital eccentricity of Earth.
These Earth’s orbital parameters have the following values:
We start by calculating the numerator \(N = 6\pi G M\) \[\begin{align*} N &= 6\pi \times (6.67430 \times 10^{-11}) \times (1.98847 \times 10^{30}) \\ &= 6\pi \times 1.3271244 \times 10^{20} \\ &= 6 \times 3.1415926536 \times 1.3271244 \times 10^{20} \\ &= 18.84955592 \times 1.3271244 \times 10^{20} \\ &= 25.0226386 \times 10^{20} \, \text{m}^3\,\text{s}^{-2} \end{align*} .\]
Next, we compute the denominator \(D = c^2 a (1 - e^2)\). First, compute \(c^2\), \(c^2 = (299,792,458)^2 = 8.98755179 \times 10^{16} \, \text{m}^2\,\text{s}^{-2}\), and \(1 - e^2\), \(1 - e^2 = 1 - (0.0167086)^2 = 1 - 0.000279 \approx 0.999721.\)
\[\begin{align*} D &= (8.98755179 \times 10^{16}) \times (1.495978707 \times 10^{11}) \times 0.999721 \\ &= 1.34562455 \times 10^{28} \times 0.999721 \\ &\approx 1.34524933 \times 10^{28} \, \text{m}^3\,\text{s}^{-2} \end{align*} .\]
\[\delta\phi = \frac{N}{D} = \frac{25.0226386 \times 10^{20}}{1.34524933 \times 10^{28}} = 1.860695 \times 10^{-7} \, \text{radians per orbit} .\]
Again, convert the precession to arcseconds per century, using 1 radian \(= 206,264.806247\) arcseconds
\[\delta\phi_{\text{arcsec/orbit}} = \delta\phi \times 206,264.806247 = (1.860695 \times 10^{-7}) \times 206,264.806247 \approx 0.038389 \, \text{arcseconds per orbit} .\]
The number of Earth’s orbits around the Sun in a century is approximately \(100\) (by definition of a century) \[\delta\phi_{\text{arcsec/century}} = \delta\phi_{\text{arcsec/orbit}} \times 100 = 0.038389 \times 100 = 3.8389 \, \text{arcseconds per century} .\]
Let’s compare the GR predictions against real observed data
The observed precession has an uncertainty of \(\pm 1.2''\), which means the observed value ranges from \(3.8''\) to \(6.2''\) arcseconds per century. The lower bound of the observed value (\(3.8''\)) coincides with the GR prediction (\(3.84''\)), suggesting that the discrepancy may be within observational uncertainties.
The calculation demonstrates that General Relativity predicts a perihelion precession for Earth’s orbit that is consistent with observational data when uncertainties are considered. The slight discrepancy can be attributed to observational errors and unmodeled perturbations, and does not indicate a failure of GR in explaining planetary motion.
A kime-representation model can be applied to the perihelion precession of Earth’s orbit to potentially improve upon the General Relativity (GR) prediction by incorporating a specific kime-phase distribution, such as the Laplace distribution. We’ll consider the observed measurements as repeated samples from a sampling distribution and see if this approach can account for the discrepancy between the GR prediction and the observed value.
A kime-phase difference (\(\Delta \phi\)) with a probability distribution \(P(\Delta \phi)\) represents the uncertainty, or fluctuations, in time measurements across repeated experiments. To see if incorporating a specific kime-phase distribution (e.g., Laplace or Normal distribution) into the perihelion precession calculation can adjust the GR prediction to better match the observed value.
Knowledge about the Earth’s Perihelion Precession include
Observed Precession: \(5.0'' \pm 1.2''\) arcseconds per century.
GR Prediction: Approximately \(3.84''\) arcseconds per century.
Discrepancy: The observed value is higher than the GR prediction, with an uncertainty that partially overlaps the predicted value.
The standard GR approach provides a deterministic prediction for perihelion precession based on the curvature of spacetime.
Kime-Representation Perspective: Introduces stochasticity into the time evolution, potentially affecting orbital dynamics.
Hypothesis: The discrepancies between GR predictions and observations might be accounted for by fluctuations in the kime-phase, modeled by a probability distribution.
The standard equation of motion is \(\frac{d^2 u}{d\phi^2} + u = \frac{GM}{\tilde{L}^2} + 3 \frac{GM}{c^2} u^2\), and the kime-adjusted equation includes a kime-phase correction term \(\delta \phi_k\) to the angular coordinate \(\phi\), \(\phi \rightarrow \phi + \delta \phi_k,\) where \(\delta \phi_k\) is a random variable with a probability distribution \(P(\delta \phi_k)\).
Consider modeling the kime-phase difference with a Laplace Distribution \(P(\delta \phi_k) = \frac{1}{2b} \exp\left( -\frac{|\delta \phi_k - \mu|}{b} \right),\) where \(\mu\): Location parameter (mean of the distribution) and \(b\): Scale parameter (related to the variance). Let’s try to adjust the calculation of the Earth’s perihelion precession by averaging over the Kime-Phase Distribution. The modified solution for \(u(\phi)\) is \(u(\phi) = u_0(\phi + \delta \phi_k)\), where \(u_0(\phi)\) is the solution without the kime-phase correction.
The expectation value of \(u(\phi)\) is \[\langle u(\phi) \rangle = \int_{-\infty}^{\infty} u_0(\phi + \delta \phi_k) P(\delta \phi_k) \, d(\delta \phi_k) .\]
For small \(\delta \phi_k\), expand \(u_0(\phi + \delta \phi_k)\) \[u_0(\phi + \delta \phi_k) \approx u_0(\phi) + \delta \phi_k \frac{du_0}{d\phi}\]
The expectation value is \[\langle u(\phi) \rangle = u_0(\phi) + \langle \delta \phi_k \rangle \frac{du_0}{d\phi}\]
Since the Laplace distribution is symmetric around \(\mu\), if we set \(\mu = 0\), then \(\langle \delta \phi_k \rangle = 0\).* The variance \(\sigma^2\) of the Laplace distribution is \(2b^2\).
As \(\langle \delta \phi_k \rangle = 0\), the first-order term vanishes. The second-order term in the expansion is \[u_0(\phi + \delta \phi_k) \approx u_0(\phi) + \delta \phi_k \frac{du_0}{d\phi} + \frac{1}{2} (\delta \phi_k)^2 \frac{d^2 u_0}{d\phi^2} .\]
Taking the expectation value \[\langle u(\phi) \rangle = u_0(\phi) + \frac{1}{2} \langle (\delta \phi_k)^2 \rangle \frac{d^2 u_0}{d\phi^2} .\]
Since \(\langle (\delta \phi_k)^2 \rangle = 2b^2\) \(\frac{d^2 u_0}{d\phi^2} + u_0 = \frac{GM}{\tilde{L}^2}.\)
\[\langle u(\phi) \rangle = u_0(\phi) + b^2 \left( \frac{GM}{\tilde{L}^2} - u_0 \right)\]
\[\langle u(\phi) \rangle = u_0(\phi) + b^2 \left( \frac{GM}{\tilde{L}^2} - u_0(\phi) \right)\]
\[\langle u(\phi) \rangle = (1 - b^2) u_0(\phi) + b^2 \frac{GM}{\tilde{L}^2}\]
This suggests that the effect of the kime-phase distribution is to scale the Newtonian solution and add a constant term. The precession is related to the angular shift in the orbit due to deviations in \(u(\phi)\). Let’s consider how the scaling affects the precession term in GR where the extra precession per orbit is due to the \(3GM u^2 / c^2\) term in the equation of motion. With the kime-phase correction, this term becomes modified due to the change in \(u(\phi)\).
Assuming the kime-phase correction effectively scales the precession term, we can adjust the GR precession term \(\delta \phi_{\text{total}} = \delta \phi_{\text{GR}} \times f(b) ,\) where \(f(b)\) is a function depending on the scale parameter \(b\) of the Laplace distribution.
Can we determine an appropriate value for the scale parameter \(b\) that may link the observed Precession, \(\delta \phi_{\text{obs}} = 5.0''\) arcseconds per century with the GR prediction, \(\delta \phi_{\text{GR}} = 3.84''\) arcseconds per century? \[f(b) = \frac{\delta \phi_{\text{obs}}}{\delta \phi_{\text{GR}}} = \frac{5.0}{3.84} \approx 1.3021 \]
From our earlier expression, the precession term is proportional to \(u_0^2\). With the kime-phase correction, \(u_0\) is effectively scaled by \((1 - b^2)\). Thus, the precession term becomes \[\delta \phi_{\text{kime}} \propto \left( (1 - b^2) u_0 \right)^2 = (1 - b^2)^2 u_0^2 .\] To match the observed precession, we set \[f(b) = \frac{\delta \phi_{\text{kime}}}{\delta \phi_{\text{GR}}} = (1 - b^2)^2 .\] However, this leads to \(f(b) \leq 1\), which cannot increase the precession beyond the GR prediction.
Are there alternative approaches? If the assumption is that the kime-phase correction scales the precession term as \((1 - b^2)^2\), this may not increase in the precession value. Perhaps the kime-phase correction introduces an additive term to the precession? Is there is an additional precession \(\delta \phi_{\text{kime}}\) due to the kime-phase fluctuations? Let’s suppose the total precession is \[\delta \phi_{\text{total}} = \delta \phi_{\text{GR}} + \delta \phi_{\text{kime}}.\] To estimate \(\delta \phi_{\text{kime}}\) we can use a theoretical expression for \(\delta \phi_{\text{kime}}\) in terms of \(b\). Suppose the second-order term in \(\langle u(\phi) \rangle\) introduces an effective potential perturbation. Using perturbation theory, the additional precession per orbit due to a small potential perturbation \(\delta V(r)\) is given by \[\delta \phi_{\text{kime}} = -\frac{\partial}{\partial L} \left( \frac{1}{T} \int_0^T \delta V(r) \, dt \right) .\] However, this approach may be over-complexifies and may not yield a simple expression in terms of \(b\).
Yet another approach may be to directly modify the precession formula. Suppose the kime-phase fluctuations effectively modify the gravitational constant \(G\) to \(G_{\text{eff}}\). Consider a modified precession formula \(\delta \phi_{\text{total}} = \frac{6\pi G_{\text{eff}} M}{c^2 a (1 - e^2)}.\) Assuming \(G_{\text{eff}} = G (1 + \delta G)\), where \(\delta G\) is a small correction due to kime-phase fluctuations. If \(\delta G\) is proportional to the variance of the kime-phase distribution \(\delta G = k b^2\), where \(k\) is a proportionality constant. To find \(b\) such that \[\delta \phi_{\text{total}} = \delta \phi_{\text{GR}} (1 + \delta G) = \delta \phi_{\text{GR}} (1 + k b^2) \] we can set \(\delta \phi_{\text{total}} = 5.0''\), \(\delta \phi_{\text{GR}} = 3.84''\), i.e., \(5.0 = 3.84 (1 + k b^2) .\) Solving for \(k b^2\), yields \(1 + k b^2 = \frac{5.0}{3.84} \approx 1.3021.\), \(k b^2 = 0.3021\). For simplicity, assume \(k = 1\). Then, \(b = \sqrt{0.3021} \approx 0.5496.\) This suggests that the scale parameter \(b \approx 0.55\) radians (about 31.5 degrees). Given that \(b\) is relatively large, this may not be physically reasonable. A large value of \(b\) implies significant fluctuations in the kime-phase, leading to substantial deviations in orbital dynamics. Such large fluctuations are not supported by observational data of planetary motion, which is highly precise. Introducing significant kime-phase fluctuations would likely affect other aspects of Earth’s orbit, which are not observed. Hence, modifications to \(G\) would have wide-ranging effects beyond the perihelion precession.
The discrepancy between the GR prediction and observed perihelion precession could be due to (1) measurement uncertainties, i.e., observational error margin (\(\pm 1.2''\)) encompasses the GR prediction; (2) additional perturbations, e.g., gravitational influences from other bodies or relativistic corrections not fully accounted for; or (3) kime-representation limitations, e.g., incorrect kime-phase prior distributions.
The theoretic kime-representation model extends standard time evolution by incorporating complex time and stochastic kime-phase differences. While it provides a novel approach, its application to precise astronomical phenomena like Earth’s perihelion precession must be consistent with observational data and established physical laws. In this case, the model does not appear to offer a better explanation than General Relativity when realistic parameters are considered.
While spacekime representation offers novel insights, it also introduces several mathematical challenges that need to be addressed for the framework to be fully developed and applied effectively. Below is a summary of these challenges, along with the pros and cons of using complex time representation in mathematical terms.
Challenge: Determining the conditions under which kimesurfaces (surfaces in spacekime formed by complex time trajectories) are analytic functions is crucial for ensuring that the mathematical descriptions are well-behaved and that techniques from complex analysis can be applied.
Challenge: Extending the concept of ergodicity to spacekime involves determining whether time averages over complex time trajectories are equivalent to ensemble averages across the spacekime manifold.
Challenge: The kime-phase \(\phi = \omega t\) (with \(\omega\) being a characteristic frequency) becomes complex in this framework. Determining how to measure and observe this complex phase in physical systems is non-trivial.
Challenge: Maintaining causality in systems where time is complex is essential, as causality is a cornerstone of physical laws. Complex time can introduce ambiguities in the sequence of events.
Challenge: Aligning empirical data with theoretical predictions from spacekime models requires that the models accurately reflect the underlying processes generating the data.
The complex time representation introduces a novel mathematical framework with the potential to provide deeper insights into systems exhibiting complex temporal behavior. However, several open mathematical challenges must be addressed:
Pros of the complex time representation include a richer mathematical framework, better modeling of oscillatory phenomena, and the ability to handle uncertainties. Cons involve increased mathematical complexity, ambiguities in physical interpretation, potential violations of fundamental principles, and difficulties in measurement and validation.
To address these challenges, future research in the following directions may be fruitful.