What are time and causality?
Time feels self‑evident: second after second our clock ticks, and events follow one another in a fixed order. But what actually determines that sequential unfolding, in other words, why do time and cause‑and‑effect (causality) exist at all? In modern physics, time is woven together with space into spacetime, and the speed of light is the ultimate speed at which causality can propagate. Yet fundamental questions remain open. Why does the arrow of time always seem to point forward (from past to future) and not the other way around? Is there such a thing as a moment “before” time began, or a state without causality? These are deep questions that physicists and philosophers wrestle with.
A provocative new idea suggests that time and causality might not be fundamental at all, but phenomena that emerge from a deeper reality. In this idea, developed by the Dutch artist‑thinker Stephan Konings, there exists an underlying field that determines whether there is any passage of time and causality. This field is denoted by the Greek letter σ (sigma) and acts like a switch: only if σ is “on” (has a non‑zero value) can a “stage” arise on which events can unfold in cause and effect. In a hypothetical state where σ = 0, there would be no time and no cause‑and‑effect relations; there simply would be no light cones or other structures to connect events. Only when σ > 0 does the world as we know it come about: the light cones “open,” time begins to flow, and causality becomes possible.
Put differently, this σ‑field determines whether there is a stage on which physical processes can play out. One can view it as a kind of phase transition between no‑thing and something. As long as σ is off (0), a pre‑physical, perfectly symmetric silence reigns, nothing happens. Switch σ on, and something breaks that silence: spacetime and all the phenomena that go with it suddenly appear. We can compare this moment, in a conceptual sense, to a kind of Big Bang: reality switches on when σ flips. In this picture, the vacuum is not an absolute “nothing,” but a fragile state full of potential. Once a symmetry breaks and σ acquires a value, time and structure tumble out. Such an idea resonates with modern views in which the universe arises from an initial symmetric state, for example via spontaneous symmetry breaking as encountered in high‑energy physics or holographic theories.
This concept immediately offers a possible explanation for the arrow of time: before σ had a non‑zero value, “before” and “after” were meaningless, there was no time. From the moment the σ‑field appears, a one‑way development begins: order and structure increase, and with them entropy (information content). Time gains a direction (from lower to higher entropy), simply because it begins to exist and reality unfolds from that initial symmetric state toward ever more structure.
The σ‑field: a measure of “something” versus “no‑thing”
How exactly does this σ‑field work? You can regard σ as a field that has a certain value everywhere in space, indicating how much “something” there is at that location. In Konings’ theory, σ is also described as a measure of ontic density, a fancy term for “reality density” or “substance.” More σ means more “something,” a fuller, more “real” environment, whereas σ = 0 would stand for total emptiness, a state of “no‑thing.” This single field influences several fundamental properties of nature at that place: how fast local time passes, how light propagates, and even how heavy matter is. The beauty is that these are no longer three separate things, but facets of one underlying something, the sigma‑field, that coordinates these aspects.
Picture the σ‑field as a kind of “fabric density” of reality. In regions where σ is high, say, a richly filled area full of matter or energy, reality is denser. Where σ is low, you are closer to emptiness. This σ‑value has three important consequences at once:
● Passage of time: A high σ‑value makes local time tick more slowly compared to somewhere with lower σ. Compare it to how, near a heavy mass (like a planet or a black hole), clocks lag behind clocks far away in space, that’s gravitational time dilation. In this model, that happens because lots of “something” (high σ) lowers the lapse function N(σ); N is a factor in the metric indicating how fast proper time elapses. More σ = lower N = a slower local clock.
● Light propagation: Where there is more “something,” light behaves as if it is traversing a denser medium. In optical terms: the refractive index n(σ) is higher at high σ. This means that light effectively travels more slowly through such a region and is also bent, as if space there were optically thicker. Conversely, in an almost empty region (low σ) light can approach the maximal speed c. In the extreme case that σ tends to 0 (the “pale” boundary of almost nothing), n(σ) would formally go to 0, meaning the phase of a light wave proceeds without delay. That sounds odd: infinite phase velocity? The upshot is that in a perfectly empty environment there is nothing to impede a light wave or to cause phase differences; you can view it as a perfectly transparent emptiness in which light no longer builds up wave crests. (Note, this is a theoretical limit: a complete σ = 0 also means there is no genuine physical interaction and thus no definable signal or event.) In normal situations we simply see: light is slower in glass than in air; likewise, in this model it is slower in a region with much σ compared to one with little σ.
It also bends toward where σ is higher, just as light bends toward optically denser media, or, from another perspective, toward masses as in Einstein’s gravity.
● Mass and inertia: The rest mass of particles can also depend on the σ‑field value. In an environment with more “something” (higher σ), particles acquire a slightly larger effective mass. As it were, they become heavier or richer in inertia due to the presence of that field. In one hypothesis Konings calls this a Higgs portal: the σ‑field would slightly influence the mass particles receive via the Higgs mechanism. The important advantage is that this must happen identically for all kinds of matter, otherwise one kind of matter would fall differently from another, contradicting the equivalence principle (the idea that in gravity all objects accelerate equally regardless of mass or composition). The model in its more universal form (called LIMEN‑U) preserves that equality, so two different objects still fall exactly the same even if σ plays a role. All matter “feels” σ in the same way. This prevents measurable violations of the equality of free fall, which have been tested extremely precisely (the MICROSCOPE satellite, for example, has shown that objects accelerate equally to within 10^{-15}). To date, variations in fundamental constants or mass due to environmental factors have likewise been minimal, so if the σ‑effect exists it must be very subtle under ordinary conditions.
In summary: the σ‑field acts as a universal conductor that orchestrates time, light, and mass in concert. If σ is high at a location, time runs more slowly, light moves more slowly, and mass is slightly larger. Where σ is low, time ticks faster, light goes at full speed, and particles are lighter. This is an elegant picture because it resembles what gravity does in Einstein’s relativity, yet this theory puts all those effects into a single scalar field instead of only the curvature of spacetime. You can compare it to how an optical medium works: Einstein says mass changes the geometry of spacetime, whereas Konings says the σ‑field creates a kind of optical metric within which matter and light move. For an observer, that makes no difference: they see identical effects (time dilation, light deflection, inertia) as if there were gravity. Such a σ‑field could therefore be the deeper cause from which the phenomena of spacetime, gravity, and inertia jointly flow.
Light: from waves to packets of information
Light is not only our source of sight, it is also the carrier of information throughout the universe. In classical physics light travels in straight lines at a constant speed c in vacuum, except when it passes through a medium (like water or glass), where it becomes slower and can change direction (refraction). In Einstein’s general relativity, vacuum is not truly empty: mass curves spacetime, and light follows that curvature (so its path bends) and loses energy climbing out of a gravitational field (redshift) or gains it when descending into one (blueshift).
The σ‑model brings these two viewpoints under one heading. It says: what we normally see as the gravitational effect on light can be understood as if space itself were a material with varying density, determined by σ. High σ makes space optically thick; low σ makes it optically thin. Thus:
● Light bends toward “more something”: Light rays experience a kind of refraction toward regions with higher σ, just as a ray in the atmosphere bends toward denser air layers or in water bends toward the normal. This directly explains why light deflects near a star or planet, those objects locally raise σ (they are a lot of “something”), causing light to bend slightly as if passing through a lensing medium.
● The speed of light changes locally: In a region with high σ, the phase velocity of light v_\phi = c/n(σ) is smaller (since n(σ) is large). The actual information transfer (group velocity v_g) always remains below c, so there is no conflict with the principle of causality (you still can’t transmit information faster than light). In regions with low σ (little something, almost emptiness) n(σ) approaches 1 and light proceeds almost unimpeded at speed c.
● Frequencies shift: Because σ also affects local time, it influences the frequency of light waves. In a deep σ‑well (for example near a heavy mass or in a region with lots of energy) local time runs slower, so local oscillations (e.g., of a light wave) occur more slowly in that local frame; to a distant observer they then appear to oscillate more slowly (redshift) when climbing out, or more quickly (blueshift) when descending. These are precisely the gravitational red/blue shifts that exist in Einstein’s theory, here explained intuitively by light encountering different σ‑values.
A striking prediction of the σ‑model is the idea of a “pale boundary” at extremely low σ. In such a hypothetical case, which perhaps occurs only in theory or edge cases, the environment is so empty (σ nearly zero) that the refractive index n(σ) becomes extremely small. The light wave would then have an infinitely large phase velocity and be in phase everywhere (there is no longer any difference between what the wave does “here” and a bit farther away, because there is nothing to make phases differ). Interference becomes trivial: everything moves in lockstep and no phase differences can be accumulated. In essence this means that in an almost entirely no‑thing environment, no complex wave phenomena occur, there is nothing to disturb or slow down. This is like an imaginary extreme in which a light wave loses its dynamics and appears “frozen”; as if the light is present but can do nothing. Interestingly, that resembles the idea of a measured quantum particle: when you measure a particle, you fix a value at one place and elsewhere nothing of that wave remains. An environment with n \to 0 is an analogue here: the wave is everywhere and, at the same time, effectively nowhere, until interaction occurs. In cosmological terms \sigma = 0 would literally mean no reality; on a lab scale, n \to 0 translates as “no signal, no wave,” the whole situation behaves as if there is nothing to wave.
In short, the σ‑field gives a new picture of light: not only as wave or particle, but as something whose properties derive from the environment through which it propagates. Light “feels” the σ‑structure of reality. Where that structure is homogeneous and open, light can behave as a perfect wave that interferes with itself. Where that structure varies or is abrupt, light adapts, and can partially or completely lose its wave character in favor of particle behavior.
From quantum wave to particle: σ and duality
One of the best‑known wonders of physics is wave‑particle duality. An electron or photon can, in one setup, show interference fringes like a wave, and in another act as a point‑particle leaving a single dot on a detector. How can something be both? In traditional quantum mechanics this remains somewhat mysterious: one speaks of a wavefunction that spreads over possibilities and “collapses” to a particle upon measurement. Konings’ σ‑field approach tries to explain this in a conceptually clear way. The idea is that one underlying reality (the σ‑field with its geometry) delivers both behavioral regimes depending on the circumstances.
Consider a classic double‑slit experiment with electrons. As long as you do not measure which slit the electron passes through, it behaves as a spread‑out wave and you see an interference pattern of many electrons together on the screen. In σ‑terms, the situation is as follows: the electron’s path via slit A and via slit B lie in practically the same environment, with comparable σ‑values along both paths. The σ‑structure is homogeneous, and the causal structure (light cones, etc.) for both paths is practically the same and open. The two possible routes of the electron can therefore remain coherent: their phase difference remains well‑defined, because there is no strongly different “something” on one route versus the other. The result is interference: bright and dark bands, the typical wave pattern. In the σ‑field model we designate this as the wave regime: the particle identity is vague and spread out, and what we observe are the relative phases of the wave traversing different paths.
But what happens when we try to measure which path the electron takes, for instance by placing a detector at one slit? By that measurement we introduce a local piece of apparatus, which by definition is concentrated “something” (think of an atom or photon in the detector: matter is “something”). That means a local increase in σ and often even an abrupt difference between path A (with detector, high σ) and path B (without, lower σ). Extreme σ‑gradients decohere the paths; nature as it were “measures” the difference itself. The joint phase picture collapses, because the phases of path A and B diverge too much to form a single interference pattern. In plain language: the interference pattern disappears as soon as you actively look where the particle is. In σ‑language: too large a difference in “something” between possible paths makes a single unified oscillation impossible. This is precisely the transition to particle behavior. The electron is now actually registered at one specific spot (a local deposition of energy). Konings describes this as the particle “appearing where σ is extreme or abrupt, a singular event in causality.” The behavior is now that of a classical particle: one location, one impact, no interference.
According to this theory we thus need no separate quantum rule that turns waves into particles; it follows automatically from the σ‑field as background. The σ‑field is the “conductor that determines how waves or particles manifest.” As long as the σ‑field allows a process to extend over multiple paths (coherent phase across the whole), we see wave behavior. If the σ‑field forces everything into a single path (for example through a measurement, or in extreme cases such as near a black hole, more on that shortly), then we see particle behavior. Duality is no longer a puzzling two‑in‑one property, but a logical consequence of the circumstances: the presence or homogeneity of “something” in the environment. The nice thing is that this points classical gravitational effects and quantum effects to the same underlying parameter. Even fundamental constants would, in this view, vary in a coordinated way with σ, so that all sectors (EM, gravitational, inertial) keep in step and do not contradict one another.
In short, wave‑particle duality gets an elegant story: a particle is not intrinsically a wave or a particle; the σ‑field determines how it behaves. In a uniform, little‑perturbed reality (σ roughly equal everywhere) we see quantum waves that interfere. In a disturbing, rich reality (σ‑gradients or high peaks) we see particles and points. Both are extremes of the same continuum. Konings even names two extreme scenarios: the pale boundary (almost no “something” anywhere), in which everything is so uniform and empty that waves become infinitely spread out, and the black boundary (extremely much “something”), in which everything is so concentrated that waves are completely wrapped up and become point‑like. That black boundary occurs, for example, just outside a black hole or with a hypothetical infinitely heavy measuring device: σ is gigantic, N(σ)\to 0 (local time virtually stands still), frequencies shoot up for local observers, and light’s phase velocity drops to zero, as if an optical horizon arises. The wave then “shrinks” until it is effectively confined entirely to a point (wavelength → 0). You can see this as a wavelength collapse: the wave becomes so compact that it is practically a particle. In classical quantum terms we call that wavefunction collapse upon measurement, here caused by the extreme σ‑environment rather than something mysteriously random.
On the other hand there is the pale boundary (very low σ, almost empty everywhere): there the oscillations of waves are so slow and stretched out that effectively no visible interference or dynamics remain, it is as if the wave is not there, has the same phase everywhere, and thus, if you measure something somewhere, elsewhere it is as if there was never a wave. This corresponds to a situation in which reality itself becomes tenuous: you could say this borders on the disappearance of reality itself as σ → 0. Both boundaries, black and pale, are complementary: one maximizes presence (everything piled into one), the other minimizes it (everything so spread that nothing measurable remains). In intermediate, normal circumstances (like our everyday lab) σ lies between these extremes, and therefore we see both wave‑ and particle‑aspects, depending on which situation we create.
Gravity as an emergent phenomenon
One of the most revolutionary ideas in recent physics is that gravity may not be a genuinely fundamental force, but a kind of by‑product, an emergent property of something else. The Amsterdam physicist Erik Verlinde has, for example, argued that gravity arises from entropy and information: a form of “entropic elasticity” of spacetime that responds to changes in information or energy density. That sounds abstract, but concretely he proposes that when matter is present in the universe it slightly lowers the entropy (information content) of spacetime, and nature “wants” to maximize that entropy again; we perceive that response as an attractive force. In his model this even explains the extra “push” we usually attribute to dark matter: on large distances and at low accelerations a systematic extra gravity‑like effect appears, without needing invisible matter.
Konings’ σ‑theory dovetails with this surprisingly well. In fact you could say both are seeking the same underlying truth from different angles. In the σ‑model gravity is not fundamental, but a consequence of the field σ that is already managing time, light, and mass. There is no separate “graviton” or unique geometry just for gravity; rather, gravity follows from how σ is distributed in space. If σ is higher somewhere than in its surroundings (for example around a planet or star), then a σ‑gradient arises: a spatial slope in σ’s value. A particle situated in such a field then feels a kind of tug toward the region with higher σ, just as in Newtonian gravity it falls toward mass. Konings makes that insight concrete by defining a potential energy V(σ) that is lowest in regions with lots of “something” (high σ). Matter “wants” to move toward higher σ because that is energetically favorable, exactly like a ball rolling down in a gravitational field. In this description the force is proportional to the slope (gradient) in the σ‑field: F \propto -\nabla σ. This is essentially a reformulation of Newton: instead of “mass attracts mass,” we get “something attracts more something,” because more something (σ) on that side lowers the potential energy.
Importantly, this is not a new force; it is the same coherence we already saw with time dilation and light bending. For a local observer moving with the particle, the particle simply appears to accelerate toward the higher‑σ region (it speeds up; kinetic energy increases). But to a distant observer (who is not in the deep σ‑field but is looking from outside) that same particle appears to go more and more slowly as it goes deeper into the σ‑field, because time there runs more slowly. This is precisely the relativistic view of falling in a gravitational field: think of someone watching an object fall into a black hole; that object appears to slow down and never actually cross the horizon because of time dilation. The σ‑field yields the same phenomenon: what locally is an accelerating fall, appears from afar as a braking motion due to time‑ and light‑effects. Both perspectives are unified, and the σ‑field plays the role of the gravitational field.
This way of seeing gravity also offers new avenues for puzzles like dark matter and dark energy. In Verlinde’s entropic picture an extra force term emerges on large scales that resembles MOND (a modified gravity at low accelerations) and that can potentially explain the flat rotation curves of galaxies without invisible matter. Konings’ σ‑model can incorporate this by, for example, assuming there is a kind of cosmic background value of σ that yields an elastic response. If you then have matter on galactic scales (say a galaxy), that matter locally raises the σ‑field value, but the environment (the rest of the universe) provides an elastic counter‑reaction. The result is a long‑range effect: farther from the galaxy σ is slightly elevated compared to what you would expect without that matter, amounting to a weak extra attraction over large distances. In technical terms, Konings formulates this so that there is an almost massless mode of the σ‑field on large scales, which contributes with a characteristic scale (which turns out to relate to the Hubble constant H_0 or the acceleration a_0 in MOND). For the layperson: the model can be calibrated so that it automatically provides a dark component, an extra bit of gravity, precisely in the regimes where we observe it in astronomy (such as at the edges of galaxies). In doing so, the σ‑theory could explain both local tests (like in the solar system, where everything must agree with Einstein’s gravity) and the large‑scale phenomena (like galactic rotation curves) within a single framework.
As with Verlinde, information and entropy are not foreign elements here, but implicit components. If gravity is an emergent effect, that means that microscopic information changes underlie a falling acceleration. Verlinde speaks of bits on holographic screens; Konings has a more concrete field σ, but he concedes that σ is likely related to what Verlinde calls “information.” He shows in his work that his σ can play the same role: where Verlinde calculates with changing entropy (\Delta S) and an effective temperature, in the σ‑model one can set up a similar relation in which a σ‑gradient provides the force and N(σ) (the time dilation) functions as a kind of Unruh‑temperature term. In this way he translates F \Delta x = T \Delta S (Verlinde’s formula) into “matter falls toward higher σ because the potential is lower there, that is precisely entropy maximization.” Both stories come to the same thing: it is not mass itself that causes the force, but nature’s tendency to reach a particular state. In Verlinde it is maximal entropy; in Konings it is a minimal potential energy of σ (which, in essence, amounts to the same).
A key point is that all these adventurous ideas must not conflict with what we have already measured precisely. The σ‑field must pass the tests that Einstein’s theory passes as well. For that reason Konings emphasizes, for example, that the equivalence principle (WEP) must be strictly preserved in the realistic version of his model. Things like the speed of light must also remain constant for all observers (the model does not vary c itself, only the structure through which light propagates). So far all measurements are consistent with no change: for example, the fine‑structure constant \alpha (which sets the strength of electromagnetism) appears not to differ measurably at different heights or over the years, with an upper limit of |\Delta \alpha/\alpha| < 10^{-18} per year. If a theory like this is correct, it therefore predicts that even future, more precise clocks will not find deviations in free‑fall acceleration or constant values, except perhaps via very subtle effects.
Interestingly, experiments already are being done or proposed that look for exactly such minute differences: ultra‑stable optical clocks compared on the ground and in space to see whether ratios of fundamental constants diverge even a hair in different gravitational fields; or fast gamma‑ray bursts (GRBs) analyzed to see whether photons of different energy arrive at exactly the same time (which would suggest that the vacuum introduces no dispersion up to the Planck scale, something the σ‑model also requires). To date those tests are all in line with conventional expectations, which means that if σ exists, it hides itself very well or manifests only on extreme scales.
Black holes and the edges of reality
No discussion of spacetime phenomena is complete without the ultimate stress test: black holes. Black holes are regions where gravity (or in this account: the σ‑field value) climbs so high that nothing, not even light, can escape once it has passed the so‑called horizon. In the classical picture the escape velocity there approaches the speed of light, and at the horizon itself you could say that “you need c to get away, but c is not enough,” hence nothing can leave. It is sometimes said in popular terms that “time stands still at the horizon for an external observer” and “light seems frozen” there. How does the σ‑field model translate this extreme situation?
Konings provides a sharp insight: people sometimes say “c \to 0 in a black hole,” but in fact the (apparent) speed of light is of course locally c everywhere (the speed of light as a fundamental constant does not change). What happens is that the causal structure becomes so distorted that moving forward (away from the black hole) becomes practically impossible. In his model this is made explicit: at a black hole σ takes on extremely high values toward the horizon (maximum ontic density), this is the “black boundary” we mentioned earlier. As a result the lapse N(σ) \to 0: local time stands still in the limit. The light cones tip inward: every possible future lies within the black hole; no path leads outward any longer. But this is not because the (apparent) speed of light suddenly became smaller; it is because the metric factor N(σ) goes to zero. The horizon thus becomes a kind of optical boundary: in the coordinates of an outsider, light appears to stand still at the horizon. Interpret it like this: a photon trying to escape a black hole undergoes ever stronger time dilation and redshift (diverging in the limit) and never actually makes it out. In the σ‑model we saw: high σ → large refractive index → phase velocity tends to zero. For an external observer, light indeed “freezes” at the horizon.
This also has consequences for what happens to information and the quantum aspects. In the σ‑description, a black‑hole horizon is an extreme case in which waves shrink to points, a wavelength collapse. That means that within or right at the edge of the black hole, processes are confined to a single track; all coherence with the outside world disappears. This resembles quantum measurement: information that disappears into a black hole does not classically come out again. Verlinde’s view was that black holes carry maximal entropy; the σ‑model adds that maximal entropy goes together with maximal σ. In fact you might say: a black hole is a σ‑field concentration so strong that reality around it changes character (comparable to a phase transition). Time stands still, space contracts, and our familiar laws head into edge cases.
Can we test anything like this? Interestingly, in recent years black holes have actually been imaged (the Event Horizon Telescope captures images of the shadows of supermassive black holes), and their properties are studied via starlight and gravitational waves. Konings points out that subtle deviations, for example in the precise thickness of the photon ring around a black hole, or in spectral lines of material falling into it, could provide hints of his σ‑effects. If σ is connected with entropy and information and becomes extremely large at a black hole, then small deviations might show up in how light behaves around the horizon (a slightly different lensing or spectrum). Such effects have not yet been clearly observed, but the resolution of our observations is increasing. Perhaps future telescopes that look more sharply at the horizon, or measurements of gravitational waves that reveal the innermost dynamics of black‑hole mergers, will allow something to peek through. If the σ‑field really plays a role, it would be hard for it to hide completely here, these are the ultimate laboratories of extreme concentrations of “something.”
Another practical domain is the search for very light particles or fields that could play the role of σ. Some theorists have proposed that if fundamental constants turn out not to be constant, that would point to a new field (a kind of “dilaton”). People are trying to detect such signals with sensitive experiments, for example, atomic clocks measuring variations throughout the year as Earth travels through different places in the Milky Way. Nothing has been found so far, which aligns with the idea that the σ‑field, if it exists, is very weakly coupled or almost constant on terrestrial scales. It is therefore quite possible that we see the effects only indirectly in cosmological phenomena and not as a “new particle” in a collider (Konings calls σ a kind of cosmic condensate, not a field from which you would pick up a particle in a detector).
From Big Bang to cosmos: something out of no‑thing
We return to the biggest question of all: where does it all come from? In the standard cosmological picture the universe arose from a Big Bang, a hot, dense beginning, and has been expanding since. But that theory starts at the Big Bang; it says nothing about why there was a Big Bang at all or what preceded it (if that is even a sensible question). Here the σ‑field offers a fascinating perspective. As we saw, you can interpret the transition from “no reality” (σ = 0) to “a reality” (σ > 0) as a phase transition. Just as water suddenly crystallizes into ice at the freezing point, you can imagine the “empty,” pre‑physical state becoming critical and diving into a new phase: our rich, matter‑filled spacetime. That first spark was the switching on of σ, a spontaneous symmetry breaking of total emptiness. From that moment, time and space exist, and a historical “story” can begin.
In this light the Big Bang is not just a mysterious singularity, but the logical consequence of a system that was metastable at σ = 0 and at some point dropped to a lower energy state by giving σ a value. The universe, as it were, started making “something” out of no‑thing, Konings literally calls it escaping from no‑thing, via that phase transition. The vacuum was not an absolute nothing; it was brimming with potential in the form of quantum fluctuations and correlations. When the symmetry broke, that noise organized into spacetime and matter. This idea resembles some speculations in physics that our universe could have emerged from a quantum fluctuation or from higher‑dimensional physics in which spacetime is not fundamental but appears via emergent processes (think of ideas from string theory/holography where spacetime is “woven” out of entanglement). Here we see a concrete worked‑out example: a single field σ that bridges nothing and something, pre‑geometry and geometry.
After that birth of σ, the universe would have undergone its expansion and evolution as we know it from cosmology. Interestingly, if σ were to change very slowly over cosmic time, this could leave subtle traces, for example in the effective gravitational “constant” or in correlations between constants of nature. Konings’ OOO‑hypothesis (One‑dimensionally Orchestrated Order) says that if constants of nature change over time, they do so in a coordinated way, as if they are linked along a single underlying thread. That means, for example, that a change in the fine‑structure constant α would be accompanied by a particular change in the proton‑to‑electron mass ratio, and so on, all in a single pattern. If σ is responsible for that coordination, the universe has a kind of deeper unity in its evolution: microphysical parameters and macro‑dynamics (such as the expansion rate of the universe, H(t)) are not independent of one another but march together. This leads to testable ideas: perhaps gravity shows just a slightly different behavior in distant galaxies (earlier cosmic times) than here and now, in exactly the way that fits a single line of varying constants. Such tests are difficult, many factors come into play when comparing different epochs, but not impossible. One could, for example, look at gravitational lensing by galaxy clusters at different distances (and thus different eras) to see whether there is a trend. So far no convincing difference has been found, but the uncertainties are large.
The important thing about this cosmological angle is philosophical: it unites the very small (fundamental constants, quantum behavior) with the very large (the expansion and contents of the universe). In a σ‑based unification theory those two seemingly separate domains are in reality two sides of the same coin. The arrow of time, black‑hole entropy, atomic‑clock variations, all these turn out to be facets of one underlying story. The σ‑field is the pen with which that story is written, from the first page (the appearance of time) up to the present chapters (the formation of structures, the role of information, the fate of black holes).
Conclusion: a single story for all phenomena
We have traveled from the deepest emptiness to the most extreme objects in the universe, guided by one concept: the σ‑field. This theory attempts, in a conceptually elegant way, to connect disparate puzzles. Time and causality are not inexplicable starting points of the universe, but flow from a phase transition of the σ‑field; before σ was on, there was no time and thus no “before” or “after.” Light and electromagnetic waves do not propagate in an abstract nothing, but in a medium whose density is determined by σ, so that gravitational effects like time dilation and light bending share a natural, common cause with optical phenomena. Matter and mass gain context: inertia is not an absolute given but something that comes along with the “reality density” of the environment. The age‑old duality between wave and particle turns out not to be magical quantum folklore, but a logical consequence of whether processes can stretch over multiple causal paths in a homogeneous reality or become confined to a single path by disturbance, governed by σ. Gravity itself is demoted from primary mystery to secondary phenomenon: the manner in which σ varies spatially is what we feel as gravity, and thereby gravity neatly joins the list of things (time, light, mass) steered by a single field.
Of course, this theoretical vision is still in its infancy. It must be consistent with all the precision tests to which relativity and quantum mechanics have been subjected for decades. For now that seems possible: by making the couplings sufficiently weak or sufficiently universal that no existing experiments are violated, Konings has brought the σ‑model into accord with measurements. At the same time it offers predictions for the future. If this picture is correct, we will find no deviations in free‑fall accelerations or the speed of light (those principles remain intact), but perhaps we will find subtle coordinated changes: ultra‑precise atomic clocks could provide a hint of a background σ‑field, or astronomical observations of distant gamma‑ray bursts and gravitational lenses could point toward an emergent effect that does not fit a purely classical model. The coming generations of telescopes and instruments bring us closer to the boundary regimes, closer to black holes, or to immeasurably empty regions, where the σ‑theory distinguishes itself.
The σ‑field story is in fact a modern search for a unified description of nature. Just as electricity and magnetism once turned out to be a single electromagnetic whole, or space and time a single spacetime, so this suggests that the dichotomy between quantum (waves/particles) and gravity (spacetime) might also be resolved by a new insight. Everything could be part of a single fabric whose density is given by σ. The universe then becomes a kind of dynamic carpet that transforms from almost‑nothing to full‑blown structures, and that carpet itself determines how the dance proceeds, whether it is light speeding across it, time ticking, or planets rolling along. It is an ambitious vision, but precisely because of its broad reach it is incredibly compelling: it links the question “Why is there something rather than no‑thing?” with the behavior of a photon in a lab and the fall of an apple from a tree. And perhaps, just perhaps, it points the way to the next revolution in our understanding of reality, one in which we discover that reality is simpler and more coherent than it first appeared.

Sources
The concepts and quotations in this explanation are based on the work of Stephan Konings (2025) on the LIMEN model and the σ‑field, including integrated notes on causality, emergent gravity (Verlinde), and questions of duality. These insights are contained in internal reports such as “Sigma Geometry and Wave–Particle Duality,”, “Gravity as σ‑Response,” and Konings’ synthesizing notes, in which the σ‑field is proposed as the central ordering principle. In addition, current studies and experiments are cited, such as atomic‑clock tests and astrophysical observations, which are relevant to emphasize the falsifiable aspects of the model. Together these sources sketch a rich yet consistent picture in which time, space, matter, and information converge in a single theoretical framework.