Donald Hoffman turns perception on its head by arguing that what we see isn’t a window onto reality but a species-specific user interface shaped by evolution. His Interface Theory of Perception likens our senses to icons on a computer desktop: they’re simplified tools for survival, not accurate portrayals of the world. According to his Fitness-Beats-Truth theorem, organisms whose perceptions prioritize reproductive fitness always outcompete those tuned to objective reality. Building on this, Hoffman’s Multimodal User Interface framework shows how different senses combine to construct our seamless “reality,” even though each modality hides the true complexity beneath. He then takes a radical step with Conscious Realism: the claim that consciousness is fundamental and that what we call the physical world is a network of interacting conscious agents. In his view, brains don’t produce consciousness; consciousness gives rise to brains and the very objects we believe populate space-time. Together, these ideas form a form of philosophical idealism. Hoffman borrows from Berkeley and Leibniz but adds mathematical rigor and evolutionary theory to argue that our everyday world is an evolved illusion—powerful and consistent, yet hiding the deeper truth of a universe built on mind rather than matter.
The search for a “Theory of Everything” (ToE) has traditionally assumed that physical reality is fundamental – that space, time, matter, and energy are the basic constituents of the universe. Two unconventional approaches challenge this assumption from very different angles. Cognitive psychologist Donald D. Hoffman has advanced an ontological perspective often termed “Conscious Realism,” which posits that the objective world consists solely of conscious entities and their experiences [1]. In this view, the physical world that we perceive is merely a user interface generated by these conscious agents rather than an insight into objective reality. On the other hand, theoretical physicist John Onimisi Obidi’s “Theory of Entropicity” (ToE) reimagines the foundations of physics by elevating entropy from a statistical descriptor to the status of a fundamental field that actively drives all physical processes [2][3]. Obidi’s ToE contends that spacetime, gravity, and even quantum behavior emerge from the dynamics of an underlying entropic field, embedding irreversibility and the arrow of time into the fabric of reality [4][5].
Both Hoffman’s and Obidi’s frameworks are ambitious attempts to resolve deep problems—such as the nature of consciousness, the measurement problem in quantum mechanics, and the unification of physics—by altering our ontological premises. Yet, they do so in almost opposite ways: one by making mind primary and downgrading matter to a mere appearance, the other by reformulating physical law around entropy and relegating spacetime geometry and forces to emergent phenomena. This paper presents a detailed comparative analysis of these two paradigms. We will first outline each theory’s core principles: Hoffman’s Interface Theory of Perception and conscious-agent ontology, and Obidi’s entropic-field-based ToE along with its key concepts like the No-Rush Theorem, Entropic Observability, and Entropic Existentiality. Then, we explore the ontological differences and points of contention between them: What is fundamentally real (consciousness vs. entropy)? What is the status of space and time? How is causality implemented? What constitutes an “observer,” and how does observation work? We also examine how each theory approaches explanatory challenges such as the arrow of time, causality, and quantum measurement. Finally, we discuss the broader implications and outstanding challenges of each approach—conceptual, mathematical, and empirical—and consider whether these seemingly divergent views might be reconciled or are fundamentally incompatible.
Donald Hoffman’s ontological framework is rooted in two main ideas: the Multimodal User Interface (MUI) Theory of Perception and Conscious Realism [1]. According to Hoffman, evolution by natural selection did not shape organisms to perceive reality as it truly is; rather, it shaped us to perceive an interface that aids survival [6][7]. This is analogous to a computer’s desktop interface: icons (like a file icon) are simplified symbols that hide complex reality (the file’s binary code) because seeing the “truth” would be inefficient or even detrimental to the user. In Hoffman's words, “perceptual experiences do not match or approximate properties of the objective world, but instead provide a simplified, species-specific user interface”[8]. Our senses present a useful fiction tuned to fitness, not an veridical picture of an objective world [9]. The startling implication is that space, time, and physical objects as we perceive them are not fundamental reality; they are like icons on the desktop of our perception [10].
Underlying this interface, Hoffman posits that the objective world in itself consists of conscious agents and their experiences [11]. This is the thesis of Conscious Realism, which holds that consciousness is fundamental and that what we think of as the physical universe is essentially a network of interacting conscious entities [1]. Each conscious agent can be described formally (Hoffman provides a mathematical model for conscious agents as systems that have states of experience and interact via probabilistic rules), and what we call “physical objects” or “particles” are in fact convenient representations of interactions between these agents. Notably, space and time in this framework are constructed by consciousness rather than pre-existing containers: “In Hoffman’s theory, space and time are rendered by consciousness, rather than being fundamental”[10]. Physics as we know it – with its laws of spacetime and matter – thus emerges as a description of the regularities in the interface of conscious experiences. For example, the reason different observers agree on physical events is not that they access a mind-independent object, but that they share a similar interface and can exchange data (a notion Hoffman likens to different users on a networked multiplayer game seeing the same virtual environment).
A critical consequence of this view is that entropy and the arrow of time are not intrinsic properties of an objective universe but rather artifacts of the observers. Hoffman, in agreement with certain idealist interpretations, suggests that what we call the Second Law of Thermodynamics (entropy increase) reflects the fact that a conscious observer’s data about the world inevitably loses fidelity over time due to limited information capacity [12]. In other words, the “increase of entropy” might be a property of the perspective of a conscious agent rather than a fundamental physical law [12]. This aligns with the idea that our perception of time’s arrow (and of irreversible loss of information) is tied to the constraints of our interface. Hoffman’s model of conscious agents is in fact Markovian – it describes a stepwise state-update dynamic for experiences [13]. This Markovian dynamic of the conscious network can naturally lead to the appearance of time asymmetry and “entropy” in the interface, because each agent only has access to a subset of the total information and loses some information as interactions progress [12][14]. Thus, from Hoffman’s perspective, time and entropy emerge as perceptual phenomena of conscious agents exchanging information, not as absolute cosmic features.
One might ask: if our scientific theories are part of this interface, can we ever discover a “true” Theory of Everything? Hoffman’s answer is essentially no – at least not in the conventional sense. He argues that scientific theories by their nature offer models with certain assumptions and scopes, and each theory eventually gets superseded as we probe deeper. In his recent commentary, Hoffman suggests that “the very nature of scientific theories and their evolution ensures that there can never be a final Theory of Everything”[15]. This is both an epistemic point (all theories are provisional) and an ontological one rooted in his conscious realism: an infinite or ever-expanding network of conscious agents could produce an endless hierarchy of phenomena, so no finite theory within the spacetime interface can capture the whole truth [16][17]. Reality, in this view, is “infinite” or unbounded (he even invokes Gödel’s incompleteness theorem and Cantor’s hierarchy of infinities in arguments that consciousness entails an endless variety of experiences [18]). Therefore, any attempt to formulate a final theory purely in terms of physical objects in spacetime is, in Hoffman's view, fundamentally misguided – we must instead formulate theories in terms of conscious agents, and even those will likely evolve without end.
Summary of Hoffman’s Ontology: In Hoffman's framework, consciousness is ontologically primary. The world consists of conscious agents interacting. Space, time, matter, and entropy are secondary constructs – a kind of graphical user interface that each conscious agent “sees” as it interacts with others [1][10]. What we call an “electron” or a “brain” or any physical entity is akin to an icon or a data structure that symbolizes a deeper interaction between agents. Observers do not passively detect a pre-existing universe; they actively create the observed world (or the appearance of it) through their perceptions. There is no “hard problem of consciousness” in this scheme, because consciousness isn’t something to derive from matter – it’s assumed at the start [10][19]. Instead, the hard problem is reversed: how to derive the appearance of a stable physical world from interactions of conscious experiences. While Hoffman and colleagues have developed mathematical models of networks of conscious agents, this approach remains highly speculative. It faces challenges such as: How exactly do specific equations of physics (e.g., quantum mechanics or relativity) emerge from the conscious agent dynamics? Can this framework make new testable predictions? And, importantly when comparing to Obidi’s theory – what is the status of entropy and dynamics in a world without fundamental spacetime? Hoffman’s answer is that our measured “laws” (like increasing entropy or finite light speed) are properties of the interface constraints that our consciousness has evolved, not of the ultimate reality [14][20]. This sets the stage for a sharp contrast with Obidi’s Theory of Entropicity, which we turn to next.
The Theory of Entropicity (ToE) proposed by John Onimisi Obidi takes a diametrically opposed starting point: it remains within a physicalist paradigm but reshuffles what is considered “fundamental.” Obidi’s ToE asserts that entropy is an actual physical field pervading the universe, and that this entropic field is the foundational entity from which spacetime, forces, particles, and even quantum behavior emerge [2][3]. In other words, ToE replaces “matter-energy in spacetime” with “entropy in its own right” as the bedrock of reality. The entropic field is denoted $S(x)$, conceived as a dynamical scalar field defined over spacetime (or perhaps what becomes spacetime) [21]. This field has its own degrees of freedom and obeys its own field equation – dubbed the Master Entropic Equation – derived from a variational principle called the Obidi Action [22]. The Obidi Action is constructed with three conceptual components: a kinetic term $(\nabla S)^2$ giving the entropy field dynamics, a self-interaction potential $V(S)$ shaping its behavior, and a coupling term $\eta S T^\mu_{\ \mu}$ linking entropy to the stress-energy of matter. By applying the principle of least action to this entropy-based action, one obtains field equations governing how entropy $S(x)$ evolves and distributes in the presence of matter and energy.
What are the consequences of making entropy a fundamental field? First, gravity is reinterpreted: instead of being a fundamental force or curvature of spacetime, gravity emerges as a consequence of entropy gradients and flows. In fact, ToE falls into the broad category of entropic gravity theories, but it goes further by giving entropy its own dynamics. The claim is that when one derives the field equations from the Obidi Action, one can recover Einstein’s field equations of General Relativity under certain conditions, but now explained as an emergent effect of the entropic field constraining matter. For example, Obidi demonstrated that the anomalous precession of Mercury’s orbit (43 arc-seconds per century) – historically a key success of Einstein’s GR – can be derived by adding entropy-based corrections to Newtonian gravity, without invoking spacetime curvature. In that derivation, inputs like Unruh’s effect, Hawking’s black hole entropy, and the Holographic principle are used to inform an entropy-dependent potential, leading to precisely the same perihelion advance that GR predicts. This is presented as evidence that “entropy constraints, rather than curved spacetime, are the fundamental driver of gravitational interactions”. In sum, ToE describes gravity as an emergent entropic force/field, closely aligning with and extending Erik Verlinde’s entropic gravity ideas but embedding them in a new field theory.
Second, spacetime itself is secondary. ToE suggests that spacetime geometry (and possibly time itself) emerges from the distribution and flow of entropy. The entropic field $S(x)$ “permeates all of spacetime”and in fact defines the causal structure: changes in $S$ propagate and thereby enforce what events can influence what others. In traditional physics, light-cone structure of spacetime sets causal order; in ToE, the Entropic Field sets a similar limit via what is called the Entropic Time Limit (ETL). The ETL posits a minimum irreducible duration $\Delta t_{\min}$ for any interaction or information transfer. This is formalized in the No-Rush Theorem, which states that “no physical process can occur in zero time; nature cannot be rushed”. In practice, this means ToE forbids truly instantaneous changes or influences: even in quantum mechanics, wave function collapse or entanglement correlations cannot be absolutely instantaneous, but must involve at least a tiny time increment related to entropy transfer. The entropic field acts as a kind of “universal speed governor.” If in relativity the speed of light $c$ is the upper bound for signals, in ToE one might say $c$ itself emerges from the properties of the entropic field (indeed, Obidi’s work outlines how the invariance and value of $c$ can be derived by considering propagation of small disturbances in the entropic field, yielding an “entropic wave speed” identified with $c$). The No-Rush Theorem thus reinterprets causality and the arrow of time: the arrow of time becomes an intrinsic, fundamental aspect of reality, because the entropic field enforces irreversibility at a microscopic level. Unlike conventional thermodynamics where the arrow of time is a statistical emergent phenomenon, in ToE time’s arrow is built into the laws: entropy flow literally causes time to “move forward.”
Two novel philosophical concepts introduced in ToE are Entropic Observability and Entropic Existentiality, which capture the idea that observation and existence are not absolute yes/no concepts but depend on entropy. They arise from considering the role of entropy in quantum measurement and reality. In quantum mechanics, one puzzle is what constitutes a “measurement” or the realization of one outcome among many possibilities (the collapse of the wavefunction). Obidi’s ToE addresses this via an Entropic Seesaw Model and a critical entropy threshold for collapse. In essence, two entangled quantum systems are like two ends of a seesaw connected by an “entropic bar” (the entropic field linking them). As they evolve, entropy is generated in the system+environment. When the system’s entropic evolution crosses a certain threshold, the superposition can no longer be sustained and one branch becomes realized (collapse occurs). This threshold condition makes collapse an objective, law-like process: no external observer or conscious mind is needed to “cause” wavefunction collapse – it happens once an entropy limit is reached, analogous to a phase transition when a thermodynamic threshold is exceeded.
Within this context, “Entropic Existentiality” refers to the idea that an event or state truly “exists” (in the classical definite sense) only once the entropy associated with its occurrence passes a critical point. Before that, in the quantum realm, multiple potential outcomes coexist (superposition) in some sense. The ToE implies there is a measurable boundary between mere possibility and realized existence: a boundary defined by entropy. One paper describes this as “a measurable, entropic boundary between existence and access (existence and accessibility)”. In plainer terms, the existence of a phenomenon is conditional on entropic criteria – if not enough entropy has flowed or been produced to break unitary symmetry, the phenomenon hasn’t irreversibly happened. Conversely, “Entropic Observability” means that an event can only be observed (i.e. information about it can only be acquired) when the entropic field has propagated the necessary information, subject to the Entropic Time Limit and threshold. Observability is conditional: if an event’s entropy imprint has not reached you (or your measuring apparatus) due to the No-Rush limit or other constraints, that event is effectively unobservable and might as well not “exist” for you yet. The ToE literature explicitly speaks of “twin thresholds of entropic existentiality and observability” that together define a new ontology. Reality is thus described not in binary terms (existent vs nonexistent) but in terms of process: events gradually become definite and observable as entropy flows. This is a markedly different take on ontology, echoing some aspects of relational quantum mechanics or process philosophy, but grounding it in a physical entropy field.
Another significant aspect of ToE is its attempt to bridge to consciousness and information in a physical way. While Hoffman solves the mind-body problem by saying “mind is fundamental,” Obidi’s ToE tries to explain mind using physics. It introduces the concept of Self-Referential Entropy (SRE) to characterize consciousness. The idea is that a conscious system is one that possesses an internal entropic loop – it “observes itself,” in a manner of speaking, through entropy flows. An SRE Index is defined as the ratio of a system’s internal entropy generation to external entropy exchange. A high SRE index means the system is heavily self-referential (processing information internally relative to its outputs), which is conjectured to correlate with the degree of consciousness. Thus, ToE proposes that consciousness can be quantified by entropy flows: a fully rigorous formulation is not yet there, but qualitatively, a brain might be seen as an entropy-processing organ, and its level of awareness corresponds to how entropy/information is cycled internally versus dissipated externally. This approach is speculative, but it exemplifies how ToE extends beyond physics into AI and biology: for instance, it suggests one could design AI systems constrained by entropic principles (so-called “entropy-constrained networks”) to achieve better learning or even consciousness-like properties. In summary, Obidi’s theory does not treat consciousness as fundamental, but it seeks to embed consciousness into the entropic fabric of the universe as an emergent phenomenon obeying entropy-based laws.
The Theory of Entropicity asserts that entropy is ontologically primary, and everything else in physics – space, time, forces, particles, even measurement outcomes – derives from the dynamics of a fundamental entropic field [2][3]. It provides a unifying framework: gravity emerges from entropy gradients (entropic gravity); quantum indeterminacy and collapse are governed by entropy thresholds (resolving Einstein-Bohr debate via an objective collapse mechanism); the arrow of time and irreversibility are built-in via the No-Rush Theorem and entropic time limit; the invariance of the speed of light finds explanation in properties of the entropic field (the ratio of “entropic stiffness” to “entropic inertia” fixes $c$). In principle, ToE aspires to reproduce the known successful laws of physics in appropriate limits (it is claimed to reduce to Einstein’s equations in a classical regime, and to yield something like quantum uncertainty relations in another regime), while also providing new predictions (such as slight deviations from instantaneous quantum entanglement, or perhaps small corrections to cosmological dynamics from entropy fields). It is a bold hypothesis at an early stage. Many aspects remain under development, especially the explicit mathematical formalism: the literature acknowledges that while the conceptual pieces of the Obidi Action and entropic field equations are laid out, a fully rigorous equation set (with defined functional forms for $A(S), V(S)$ etc.) and solutions are still being worked out. This means ToE’s consistency and completeness are not yet proven – a point to remember when comparing with Hoffman’s theory, which also is currently more of a conceptual framework than a finished quantitative theory.
Having described each framework separately, we now turn to a direct comparative analysis. The task is to identify the “ontological problems” and differences between Hoffman’s and Obidi’s approaches: where do they fundamentally disagree about the nature of reality, and what issues does each face in addressing the other’s domain (physics vs consciousness, objectivity vs subjectivity, etc.)?
The clearest ontological divergence is the identity of the primary existent. Hoffman’s ontology is a form of monistic idealism: reality is made of conscious experiences and nothing else (except perhaps mathematical structures to describe the relations of those experiences). The physical world has no independent existence apart from being experienced; it’s an interface projected by consciousness. In stark contrast, Obidi’s ontology could be described as a kind of informational or entropic monism: reality is made of an entropy field and everything else (particles, fields, etc.) are manifestations or states of that field [2][21]. Consciousness, in this view, is not a fundamental substance but an emergent phenomenon, likely arising when the entropic field organizes into self-referential patterns (high SRE systems). Thus, one framework places mind before matter, the other places a new kind of matter (entropy as a physical entity) before mind.
This difference leads to opposite interpretations of what exists when no one is looking. In Hoffman’s universe, if no conscious agent is observing something, that “something” does not have a definite physical existence – it only potentially exists as a possible experience for some agent. For example, if no one observes the Moon, Hoffman would say the Moon as a physical object is just an icon that isn’t currently rendered on any conscious interface. Reality in itself would be the complex of conscious agents that would produce the Moon-icon if needed. Meanwhile, in Obidi’s ToE, if no one observes the Moon, the entropic field configuration corresponding to the Moon (and its environment) is still there objectively evolving. The Moon has an entropy distribution and gravitational entropy field around it which constrains how it moves, etc., even if no observer is present. The Moon’s existence might be conditional in the sense that its influence or information hasn’t reached a given observer (entropic observability), but ToE assumes an objective reality of the entropy field underlying the Moon irrespective of observation. In fact, ToE would treat the Earth-Moon system’s dynamics (like orbital motion) as governed by entropy flows and gradients that do not require a conscious mind to operate.
This raises a crucial philosophical point: Hoffman’s theory essentially does away with any objective physical reality – everything is relational and dependent on observers (not unlike certain interpretations of quantum mechanics taken to an extreme). Obidi’s theory, however, is staunchly realist about the existence of a physical world; it merely says the fundamental physical stuff is entropy rather than particles or spacetime points. If one were to ask “what exists in the universe at the deepest level?”, Hoffman answers: conscious experiences (with perhaps an infinite variety we can never exhaust) [18]. Obidi answers: an entropic field with a definite value and dynamics everywhere, obeying a fundamental equation (even if we haven’t fully determined that equation yet).
From Hoffman’s perspective, Obidi’s entropic field – no matter how exotic – is still part of the interface. Hoffman would likely categorize Obidi’s entire ToE as a clever new user interface metaphor for describing observations, not the true reality. If conscious agents indeed underlie everything, then an entropic field theory could be just a high-level description of patterns in their interactions (just as quantum fields and spacetime were earlier high-level descriptions). On the other hand, from Obidi’s perspective, Hoffman’s consciousness-only world might seem to lack explanatory power for concrete physical phenomena. ToE aims to compute actual numbers (e.g. Mercury’s perihelion shift, speed of light, quantum collapse probabilities) from entropic principles – things traditionally done by physics. Hoffman’s framework currently doesn’t derive such specifics from conscious dynamics (there’s no detailed derivation of Newton’s laws or quantum constants purely from conscious agent properties yet). Thus, Obidi might view conscious realism as too epistemic or even solipsistic to tackle the hard nuts of physics, whereas Hoffman might view ToE as missing the bigger picture that “physics itself is derivative”.
In summary, the ontological commitments are in conflict: idealistic monism versus entropic (informational) monism. They cannot both be fundamentally correct unless there is some way to subsume one into the other (e.g. perhaps one could speculate that the entropic field is actually a manifestation of relationships between conscious agents, or conversely that conscious agents emerge from complex entropic fields – we will return to such potential reconciliations later). For now, it’s clear that any Theory of Everything must decide what is truly fundamental. These two theories give opposite answers, which is a primary ontological problem between them.
Space and time are treated very differently in the two frameworks. For Hoffman, space-time is not fundamental reality – it is part of the “desktop interface” of human (and other creatures’) perceptions [10]. This means that features like distance, duration, and locality are properties of our interface, not necessarily of the underlying reality of conscious agents. In conscious agent theory, interactions between agents are described abstractly (e.g., as Markovian state transitions), without presupposing a background space or time. Physical spacetime, with its 3+1 dimensions and metric, is a data structure that certain conscious agents use to communicate. One implication is that the usual speed-of-light limit is an interface property; presumably, in the realm of conscious agents “behind” the interface, there might not even be a concept of speed or distance. The consistency of physics (why we don’t see violations of relativity) would be because all human observers share a similar interface shaped by evolutionary pressures, which enforces rules like relativity and $c$ as a constant. Hoffman’s framework does not yet explain why our interface has the particular constants it has (like $c=3\times10^8$ m/s, Planck’s constant, etc.), but it posits that such regularities are just part of the format of our perceptions.
In Obidi’s ToE, space and time are emergent but in a more concrete, law-bound way. Initially, ToE describes the entropic field $S(x)$ as a function over spacetime – implying spacetime as a continuum is assumed as a stage on which $S$ exists. However, the theory suggests that the geometry of spacetime (and perhaps the existence of a time direction itself) arises from the patterns of the entropy field. The entropic field enforces an Entropic Time Limit (ETL) which essentially mirrors the idea of a “chronon” or fundamental time quantum – a smallest time step for causal influence. This is reminiscent of some approaches in quantum gravity where time could be discrete or there is a minimal time interval. The No-Rush Theorem directly challenges the idea of absolute simultaneity or instantaneous action at a distance by stating these are impossible. Instead of saying “nothing can go faster than light” (which is relativity’s way of forbidding instantaneous influence), ToE says “nothing can happen in zero time; every cause needs a finite duration via entropy propagation”. It effectively provides a mechanism for the speed-of-light limit: in linearized perturbations of the entropy field, one can derive a wave equation whose characteristic speed is $c$, and this speed is fixed by fundamental constants related to entropy, quantum $\hbar$, and gravity $G$. Indeed, in the “Entropic Lorentz Group (ELG)” concept, Lorentz invariance (hence special relativity) is argued to emerge from symmetry in the entropic field equations. If this holds, ToE doesn’t just assume relativity’s structure, it explains it: $c$ is constant because it’s the propagation speed of entropic interactions, rooted in the ratio of “entropic stiffness” to “entropic inertia” in the fundamental action.
Causality in Hoffman’s model is tricky: if space-time is an illusion, what does it mean for one event to cause another? In his conscious agent network, causality would be replaced by the structure of agents influencing each other’s states. It might be more akin to information causality – one agent’s experience probabilistically influencing another’s state via some coupling. There isn’t a notion of a signal traveling through space, but rather of an “exchange” in the network. Without space, the idea of locality is replaced by network connectivity. Possibly, in the conscious agent formalism, there is still an analog of “light-cones” if, for example, interactions are only defined between certain agents or are limited by some parameter analogous to time steps. However, since Hoffman’s framework is not explicit about an alternative to space-time (beyond abstract networks), it leaves open the question of how exactly the appearance of a consistent causal spacetime emerges. So from a physics perspective, one might say Hoffman explains causality by saying it’s part of our convenient fiction and deeper reality might not have a simple notion of cause and effect in spacetime terms.
In Obidi’s ToE, causality is preserved in a more traditional sense but given an entropic twist. There is still a notion of events influencing later events, but the fundamental mediator is entropy. One could imagine that when something happens (say, a particle decays), a pulse of entropy is released into the field, and until that entropy has sufficiently flowed outward, other systems cannot be affected by the decay. This is analogous to a light signal, but it is an entropy signal. The ETL ensures a strict ordering: cause must precede effect by at least the minimum interval. In fact, ToE’s entropic causality might resolve some paradoxes: for instance, quantum entanglement correlations are not “causal” in the traditional sense (no information travels), but ToE would still impose that the establishment of correlation takes a small finite time due to entropy exchange, so there’s no violation of relativity – entanglement would have a tiny delay or “hysteresis” (some evidence for a finite entanglement propagation speed on the order of attoseconds has been cited as empirical support).
A contrasting point is the direction of time: Hoffman’s worldview can allow, in principle, that at the fundamental level (if one could step outside our interface), there might not be a single time dimension or a unidirectional time – those might just be part of our simplifying interface. It resonates with some interpretations of quantum theory (e.g., time-symmetric or block-universe views) where the arrow of time is not fundamental. But ToE insists the arrow of time is fundamental and universal: entropy flow gives a cosmic time orientation that cannot be reversed. In fact, in ToE, one could say time is literally the measure of entropy increase – time “emerges from the redistribution of entropy through the underlying entropic field”. This harks back to earlier proposals like Julian Barbour’s “time is change” or the idea that without entropy change time doesn’t progress. But ToE cements it as law: no entropy change, no passage of time; wherever entropy flows, that defines forward time.
Summarizing differences: Both Hoffman and Obidi reject the naive view of spacetime as fundamental, but they do it differently. Hoffman demotes spacetime to a mental construct; Obidi demotes spacetime to a secondary phenomenon of a deeper physical field. Hoffman’s approach might free us from conventional constraints (e.g., maybe consciousness could do things that violate spacetime limits – though consistent with interface, we wouldn’t notice), whereas Obidi’s retains the spirit of physical law enforcement (no free lunch: everything including consciousness must respect entropic causality and finite rates). For someone rooted in physics, Obidi’s might seem more concrete: it doesn’t ask you to throw away spacetime entirely, only to accept a new source for it. For someone concerned with the role of the observer, Hoffman’s is attractive: it addresses the fact that physics as practiced always involves observations by conscious beings, and perhaps we should start there.
Interestingly, both theories agree that what we perceive as spacetime is not the ultimate stage of reality – “spacetime is doomed”, as some physicists (like Nima Arkani-Hamed) also say in a different context. They also agree that irreversibility is key: Hoffman would say our interface inherently loses information (hence we perceive increasing entropy) [12], Obidi says entropy and irreversibility are basic laws (hence no process is truly reversible) [4]. Both therefore resolve to some extent the question “why does time have a direction?” – but one answers “because our minds impose one (due to how we handle information)” and the other “because entropy does (as an objective process)”.
The role of the observer or observation is central to both theories but conceptualized in almost opposite ways. In Hoffman’s framework, observers (conscious agents) are the only things that truly exist – they are not a subset of reality, they are reality. Thus the observer has a privileged ontological status. Measurement or observation in physics (like observing a particle’s position) is just an interaction between conscious agents (the scientist and the particle-as-agent) that results in a certain conscious experience (the data reading). The “wavefunction collapse” is simply a conscious agent’s update of beliefs/experiences when they interact; there is no physical collapse happening in an external world, because the only real change occurs in the observer’s consciousness.
By contrast, Obidi’s ToE depersonalizes the notion of observer. In fact, one could say ToE attempts to remove the observer from fundamental equations – a very objective stance. The theory introduces entropic observability as a condition for any information transfer, which applies equally whether or not a human is involved. An electron scattering off an atom “observes” the atom in the sense that entropy is exchanged; the same rules of needing a minimum time or entropy threshold apply. The observer in ToE is just another physical system (with perhaps some special complexity if it’s conscious, but fundamentally it’s physical). Therefore, measurement is treated as an entropy-driven interaction that reaches completion when the entropy threshold for irreversibility is crossed, not when “the mind looks at it.” This is an objective collapse viewpoint: wavefunction collapse (if one uses that language) happens irrespective of whether a human is watching, as long as the entropy criterion is met. In fact, even unobserved processes (in the human sense) will collapse due to entropy – which means there is always a definitive history at the fundamental level (no Schroedinger’s cat paradox, because the cat+environment entropy will exceed threshold long before a human opens the box, so the cat’s fate is objectively decided).
The terms Entropic Observability and Entropic Existentiality capture how ToE carefully defines what it means for something to be observed or to exist. “Observability is conditional” means you might have an event that physically happened (entropy was generated) but if that entropy hasn’t spread to you, you cannot confirm the event. This is analogous to relativity’s “outside our light cone, we cannot know events” but here couched in entropy: outside our “entropic cone” we have no access. “Existentiality is conditional” implies that a quantum possibility becomes an actual existent state only when the entropic process hits the point of no return. Before that, its existence is indeterminate or superposed. So ToE effectively introduces a new ontology with degrees of existence, something foreign to classical physics but resonant with quantum intuition. It is somewhat akin to say: Real (classical) existence is not a binary property but the end result of a continuous entropic transition.
From Hoffman’s viewpoint, all this talk of entropy thresholds for existence might miss the point that existence itself is observer-relative. Hoffman would argue that existence of an object is only meaningful with respect to an observer – nothing has standalone existence with inherent properties (like “the particle had a definite position”) independent of observation. Obidi’s approach squarely rejects that kind of quantum Copenhagen/observer-centric stance, replacing it with a realist, mechanism-based story: decoherence and collapse by entropy. In philosophical terms, Hoffman is closer to instrumentalism or even solipsism (the world is an instrument for conscious agents; only perceptions are certain), whereas Obidi is a scientific realist (there is a mind-independent world obeying entropy laws, and even if no one measures it, things happen in definite ways given enough entropy).
A notable area to compare is how each handles the quantum measurement problem. This problem asks: how do we reconcile the unitary (reversible, observer-independent) evolution of a quantum wavefunction with the apparently irreversible, definite outcomes we see when we measure (like a Geiger counter registering a discrete click)? Hoffman might say that the wavefunction is just a tool – it represents our interface’s probabilities for experiences. When a measurement “happens,” that is simply a change in the observer’s experience (the conscious agent transitions to a state corresponding to seeing a click vs no click). There is no literal collapse in the world; it’s a Bayesian update in the knowledge of the conscious agent. In fact, one could place Hoffman’s interpretation in the camp of quantum Bayesianism (QBism), which views the wavefunction as an expression of an agent’s personal degrees of belief about outcomes, not a physical object – and collapse as just belief updating. This is consistent with conscious realism: the only real events are conscious observations, and quantum mechanics is just a rule for anticipating those observations.
Obidi’s ToE instead provides a physicalist collapse model. The “Entropic Seesaw” analogy and threshold condition constitute a mechanism by which a superposition of states will reduce to one outcome once enough entropy is involved. It bears similarity to approaches like objective collapse theories (e.g., GRW or Penrose’s gravity-induced collapse), but here entropy (not gravity per se) is the trigger. Notably, ToE even identifies Landauer’s Principle (which links entropy and information erasure) as a clue that there’s a minimal entropy cost to measurement. In essence, ToE says wavefunction collapse = an increase of entropy beyond a critical value, making the process effectively irreversible and classical. This would reconcile Einstein’s desire for an objective physics (no spooky observer role) with Bohr’s insistence on irreversibility and loss of information in measurement. Indeed, one of Obidi’s article titles about reconciling Einstein and Bohr suggests exactly this: the entropic approach satisfies Einstein by providing a real mechanism, and Bohr by acknowledging fundamental irreversibility/information loss at collapse.
So the ontological issue here is: Is the observer central (Hoffman) or incidental (Obidi) to the unfolding of reality? They diverge strongly on this. Hoffman would likely view Obidi’s “entropy-driven collapse” as still just part of the interface – maybe a next-level refinement of quantum theory, but it still doesn’t explain who or what “chooses” the outcome. From Hoffman’s lens, ultimately a conscious agent sees an outcome; if no agent sees it, can we even talk about an outcome? (He might say no, we cannot, outcomes are for observers.) Obidi’s view is that outcomes happen regardless; observers just find out about them later by absorbing the entropy from the event. This is a classic split between epistemic vs ontic interpretations of quantum states/outcomes. ToE is ontic (wavefunction represents something real which evolves and then objectively collapses), whereas Hoffman is epistemic (wavefunction represents knowledge of an observer, which gets updated).
Interestingly, ToE’s entropic observability could address one potential loophole in Hoffman’s theory: how do multiple conscious agents agree on what they see? In conscious realism, if reality is agent-specific, one has to explain the apparent intersubjective agreement in science (we all measure the electron mass to be the same, etc.). Usually, Hoffman’s answer is that we inhabit a shared network – we have interfaces that are tuned similarly by evolution, and our interactions (which are also through the interface) synchronize our experiences. But Obidi’s view simplifies intersubjective agreement: there’s one real world (the entropic field state), and different observers sample it. They agree because they’re looking at the same underlying thing, subject to the same entropic signals (no need for a mysterious alignment of experiences; classical common cause via the entropy field does it). So in terms of solving the communication/solipsism problem, Obidi’s is safer – it preserves an objective reality that guarantees consistency between observers if they follow correct procedures.
Beyond ontology, it’s important to assess how each theory deals with known physics and what new predictions or retrodictions they make. Here, Obidi’s ToE has a more direct engagement with empirical science: it is essentially a proposed new physics theory that aims to reproduce all that we know and then extend it. We have noted some successes: derivation of Mercury’s perihelion precession, suggestion of a mechanism for wavefunction collapse, explanation of why the speed of light is what it is, and it even claims to naturally account for quantum uncertainty relations in an entropic way. Furthermore, ToE provides concrete concepts like the Vuli-Ndlela Integral (an entropy-weighted path integral) which could be used to calculate quantum amplitudes with an entropy term included, and it references actual experiments such as an attosecond-scale delay in entanglement formation as supportive evidence for ETL. It also makes falsifiable predictions: for instance, if one could measure entanglement formation times or find evidence of deviations from perfect reversibility in closed systems, it could support or refute the entropic threshold idea. ToE’s entropic gravity might predict deviations from Newton/GR at some scale (perhaps where entropy gradients become significant or in certain nonequilibrium situations). All of these make Obidi’s theory at least scientifically engageable. It’s speculative, but one can imagine testing it or using it to compute something.
Hoffman’s theory, in contrast, has a less direct relationship with standard physics predictions. It is more of a metatheory about why physics works at all rather than a new physical theory with different numbers. Hoffman has offered some elements that could, in principle, become testable: he has done evolutionary game simulations that purport to show organisms evolving perceptions tuned to fitness (not truth) tend to outcompete those who see truth, supporting his premise that veridical perception is not favored by natural selection [9]. That’s an indirect empirical angle via evolutionary psychology. Another possible empirical aspect is in neuroscience or psychophysics: if our perceptions are just an interface, one might search for where the interface “breaks” or how it’s constructed (for example, studying illusions or brain tricks might reveal the icons). But those are not tests of conscious realism per se, just consistent with it. Testing conscious realism directly would be extremely challenging: it essentially claims that no matter what physical phenomena we probe, we are still just seeing interface, never the true reality. This is almost unfalsifiable because every time one finds a deeper layer (e.g., we found quantum fields under atoms), Hoffman can say “yes, that’s just a more detailed icon, not the end.” It’s reminiscent of the underdetermination in philosophy of science – multiple ontologies can produce the same observable phenomena, and Hoffman’s is so flexible (because it gives primacy to experiences which can presumably emulate any physics) that it’s not straightforward to refute by a physical experiment. The only way to “test” it might be something like: if consciousness is fundamental, perhaps one could detect effects of consciousness that cannot be reduced to known physical processes (like genuine psychophysical influences or violations of Born’s rule in quantum measurement correlated with observer’s mind states, etc.). But Hoffman’s theory doesn’t really focus on such parapsychological predictions; it mostly insists that doing physics as usual is fine, just don’t take it as the final reality.
ToE, being a physical theory, is more constrained and thus more at risk of being falsified. For example, if it turned out that quantum entanglement truly has no delay whatsoever up to an extremely high precision, that might challenge the idea of an entropic time limit (unless the limit is so incredibly small that it’s effectively unobservable). If gravitational phenomena are perfectly explained by spacetime curvature and no anomalous entropy-related deviations are ever found, one might doubt the necessity of an entropic field. Additionally, ToE faces the requirement of internal consistency – can the entropic field theory be made mathematically consistent with thermodynamics and relativity and quantum mechanics simultaneously? The need for a fully worked-out Lagrangian and field equations is an ongoing challenge. If no consistent formulation emerges, or if it contradicts some known theorem, that would be a serious issue.
Common ground and differences in implications: Interestingly, both theories emphasize information. Hoffman’s approach is essentially information-theoretic about perception (the interface transmits only partial information about reality); Obidi’s is explicitly about entropy (which is missing information or uncertainty measure). Both concur that current physics might be missing something important about information: Hoffman thinks we’re missing that information is subjective (tied to consciousness), Obidi thinks we’re missing that information has a physical field and dynamic role at the base of physics. They attack the infamous mind-matter divide from opposite sides: Hoffman eliminates matter, Obidi tries to extend matter to cover mind via information.
Another subtle convergence is that both theories in some sense limit human knowledge. Hoffman states that there are truths about reality we cannot access because of our interface. Obidi’s ToE says there is a “limit of human knowledge” dictated by entropy – for instance, due to entropic time limit, we can’t know certain things instantaneously or with arbitrary precision [5]. Also, if observability is conditional, there may be phenomena fundamentally beyond observational reach if they don’t generate enough entropy signals. Thus, both accept that we, as observers, have fundamental constraints (one cognitive, one physical) on what we can observe or know. In ToE, that limit is measurable (maybe something like no information faster than c or beyond horizon or below some entropy threshold); in Hoffman’s, that limit is conceptual (our perceptions are never the reality itself, akin to Kant’s noumena/phenomena distinction, which Obidi’s writing explicitly parallels by mentioning Kant).
In terms of addressing unification of physics: Obidi’s ToE is explicitly a unifying attempt (it wants one framework for gravity+quantum+thermodynamics). Hoffman’s work is not directly about unifying physical forces – it’s more about unifying physics and psychology (or solving mind-body rather than combining gravity and quantum). If one asked, “can Hoffman's theory unify quantum mechanics and general relativity?” the answer might be: it sidesteps that by saying both are just models in the interface. One could speculate that perhaps when we reformulate physics in terms of conscious agents, there will be a single framework that yields something analogous to QM and GR at our interface level, but this is far off. Meanwhile, Obidi’s theory has already taken steps to unify known physics (with some success on paper, but needing verification). So for someone interested in concrete progress toward a conventional ToE, Obidi’s approach is more tangible. For someone perplexed by consciousness and observation in quantum mechanics, Hoffman’s approach provides a radical conceptual solution (just redefine what’s real).
Finally, metaphysical implications: Hoffman’s conscious realism aligns with a long tradition of philosophical idealism (Berkeley, etc.), updated with evolutionary theory and cognitive science. It has implications for the nature of existence (maybe panpsychism or something akin to it – everything is conscious at some level in Hoffman’s model, since he allows conscious agents to be arbitrary, even an electron could be a trivial conscious agent in theory). Obidi’s entropicity aligns more with process metaphysics or perhaps informational structural realism – reality is information/entropy and processes, not static material substances. It resonates with approaches like John Wheeler’s “It from Bit” (primacy of information) but gives it a specific physical incarnation (the entropic field). Both challenge materialist reductionism: Hoffman by saying matter is illusory, Obidi by saying matter and energy are secondary to entropy and information flows.
Neither Hoffman’s nor Obidi’s theory is a finished product, and each faces non-trivial challenges – both in convincing the scientific community and in internal consistency.
For Hoffman’s conscious realism, a major challenge is mathematical and empirical development. Thus far, Hoffman and colleagues have provided a formal model of conscious agents (with states, perceptions, actions, etc.) and some theorems about evolutionary games, but the connection from that model to the actual laws of physics remains speculative. For example, one might ask: How do multiple conscious agents produce what looks exactly like quantum mechanics with complex amplitudes and the specific numerical constants we measure? There are attempts in Hoffman's circle to connect conscious agent networks to quantum theory (some have suggested that a combination (or “fusion”) of two agents might be represented by a tensor product of Hilbert spaces, etc. [17][20]). There is also the suggestion that because conscious agent dynamics are Markovian, they might map to quantum dynamics (which can be seen as Markovian in higher-dimensional space via decoherence theory) [13]. However, these ideas are far from a full derivation of physics. The risk for Hoffman’s theory is that it could remain a philosophical interpretation that is always compatible with whatever physics finds (since one can claim any new discovery is just another interface icon), but never making distinct predictions.
Another challenge is falsifiability and interaction with neuroscience. If consciousness is fundamental and not produced by the brain, one would want some phenomena where brain-based explanations fail but conscious-agent theory succeeds. For instance, can conscious realism solve the hard problem by describing how certain configurations of conscious agents appear as brain dynamics to us? Or can it address why, say, altering brain chemistry alters conscious experience (which in Hoffman's view would be like altering the user interface icon affects the conscious agent behind it)? There is a gap between saying “the brain doesn’t create consciousness” and explaining why brain damage stops the interface feed (i.e., causes unconsciousness). Hoffman might say the brain is just a representation of certain interactions of the conscious agent with itself or others, but this is not yet a fleshed-out explanatory bridge that would satisfy neuroscientists.
For Obidi’s Theory of Entropicity, the immediate challenge is mathematical rigor and integration with established physics. The theory is ambitious, but with ambition comes a high bar: to be taken seriously, it must show it can recover known laws (at least as approximations) and make new quantitative predictions that can be tested. The HandWiki analysis of ToE notes that explicit equations for the Obidi Action’s terms and the Master Entropic Equation are still under development. Without those, some may dismiss ToE as a hand-waving framework. However, progress is being made, as indicated by derivations like the entropic perihelion shift and entropic speed of light papers. One specific mathematical challenge is merging the inherently dissipative, time-arrowed physics of entropy with the time-symmetric formalisms of quantum field theory and general relativity. ToE suggests that perhaps the time-symmetry of those theories is an approximation (valid when entropy flows are negligible or balanced), but making that rigorous is hard. There is also the question of Lorentz invariance: if entropy field has its own dynamics, does it respect relativity or does it introduce a preferred frame? The claim is that an “Entropic Lorentz Group” ensures no violation of observer-independence, but that needs concrete demonstration.
Empirically, ToE will face scrutiny on whether its novel elements actually appear. For example, if one claims a finite minimum time for quantum interactions, experiments in quantum optics or nuclear physics could potentially detect delays or deviations. If none are found at ever-tighter bounds, the theory might have to push the scale so low (e.g., $10^{-50}$s) that it loses any hope of near-term testability, or else be revised. Similarly, ToE’s explanation of gravity as entropic might lead to tiny deviations in gravitational lensing or black hole thermodynamics – which could either be found or constrained by precision astrophysics.
Another aspect is philosophical robustness: does ToE inadvertently smuggle in an “observer” through the back door? It tries not to, but one might critique: entropy is defined in terms of information, which traditionally involves an observer’s knowledge. ToE defines entropy as an objective field $S(x)$; this is fine, but then one must clarify what microstates and macrostates underpin that $S$. Usually entropy is relative to a coarse-graining (an observer’s choice of description). ToE asserts an entropy field without explicitly saying “with respect to what coarse-graining.” If it implicitly uses a natural coarse-graining (like perhaps one defined by the entropic field’s own dynamics), that needs clarification. This is a conceptual issue: can entropy be truly fundamental and not relative? If not handled, critics might say ToE is using a fundamentally epistemic concept (entropy) as an ontic entity, which is conceptually delicate (though not unprecedented – some interpretations of statistical mechanics try to do that as well).
When it comes to bridging to consciousness, ironically Hoffman’s theory is strong by definition (it’s about consciousness), but weak in physics, while Obidi’s is strong in physics aspirations, but still very tentative in addressing consciousness. The introduction of SRE and entropic measures of consciousness is intriguing but speculative. Will an entropy-based measure of consciousness align with what we consider conscious in practice? This is testable in principle: e.g., measure internal entropy flows in a brain vs a computer and see if it correlates with consciousness signs. That might be a future direction. But until then, Obidi’s attempt to solve the hard problem is incomplete – it provides a possible correlate (entropy ratio) but not an explanation of why those entropy patterns yield subjective experience. In fairness, ToE doesn’t necessarily claim to solve that philosophical hard problem; it might just be giving a practical diagnostic.
After highlighting differences, one might wonder if these two frameworks could ever be reconciled or if one must discard the other. Is it possible that both consciousness and entropy are fundamental in different senses? Perhaps one could envision a scenario where the fundamental reality has two aspects: a “mind aspect” and an “entropy/information aspect.” This starts to sound like a dual-aspect monism (like some interpretations of quantum information theory or some panpsychist models). For example, maybe the entropic field is actually the extrinsic, quantitative face of reality, while conscious agents are the intrinsic, qualitative face of the same thing. In such a scenario, Obidi’s equations would describe how the extrinsic aspect behaves (giving rise to physics), and Hoffman’s conscious network describes the intrinsic aspect (the experiences underlying those dynamics). Indeed, some philosophers (like Bertrand Russell’s neutral monism or certain interpretations of quantum mechanics by W. K. Clifford or modern panpsychists) have suggested that what we call “energy” or “information” might just be the outward appearance of what inwardly is consciousness. If one takes that view, then Hoffman and Obidi might be describing two sides of the same coin. Obidi’s entropic field could be the physical mathematical structure that corresponds to the interconnections of Hoffman’s conscious agents. In fact, information theory often bridges between objective and subjective: Shannon information is objective, but what it means depends on an observer. A unification might say: each “unit” of the entropic field carries a bit of awareness (some proto-conscious aspect), and as entropy flows, that is literally the flow of interactions among units of consciousness.
However, such a reconciliation is highly speculative and would require both sides to compromise. Hoffman would have to accept that maybe consciousness has states that correspond to something like entropy states, and Obidi would have to accept that entropy field on its own isn’t the full story until you also say what it “feels like” to be that field (i.e., consciousness). At present, neither theory addresses this duality explicitly: Hoffman is silent on why conscious agents when they interact produce a stable-looking world (beyond saying “that’s just how the interface is shaped”), and Obidi is silent on why certain entropy configurations feel like something from inside. Bridging them might need a new idea, but if successful, it could yield an even more powerful framework addressing both the external and internal aspects of reality.
In conclusion, the ontological gap between Hoffman’s and Obidi’s theories is large: one makes mind fundamental and treats physics as a derivative illusion, the other makes a physical principle (entropy) fundamental and aims to derive mind as a derivative phenomenon. They tackle different “big problems” – Hoffman primarily the mind-body problem, Obidi primarily the unification of physical laws and time’s arrow. Each encounters difficulties in the domain the other prioritizes. Hoffman’s takes consciousness seriously but risks dismissing the hardness of physical objectivity; Obidi’s takes physical objectivity to a new level but must still account for subjectivity.
Hoffman’s Conscious Realism and Obidi’s Theory of Entropicity present us with two visionary, yet contrasting, paths toward a deeper understanding of reality. In evaluating them side by side, we find that they invert the classical hierarchy of mind and matter in opposite ways. Hoffman invites us to consider that the tangible world is a perceptual fiction orchestrated by conscious agents [1][10], thereby placing our scientific theories in a humbling light – as immensely useful, but ultimately provisional stories about an interface that hides the true complexity of existence [15][18]. Obidi challenges us to rethink physics by asserting that the ubiquitous increase of entropy is not a passive byproduct but the prime mover of the cosmos [3][4], embedding irreversibility and information at the core of every transaction in nature.
Ontologically, the two theories clash on what is fundamental – consciousness or entropy – but they converge in dethroning “ordinary matter” and “spacetime” from the seat of ultimate reality. They also share an appreciation for the role of information/entropy in shaping what we observe, and each, in its own way, limits human access to reality (be it through interface constraints or entropy constraints) [5][12]. This reflects a broader shift in foundational thinking: away from naive realism and toward frameworks where what is empirically real emerges from deeper principles or perspectives.
The ontological problems between the two can be summarized as follows: - Nature of Reality: Hoffman posits a reality of experiencers without an independent physical world, whereas Obidi posits an independent physical (entropic) world that would churn on even without experiencers. Is reality constituted by observers, or do observers arise within a reality ruled by entropy? This remains a philosophical divide. - Role of Time and Causality: Hoffman sees time and causality as part of the interface (potentially not fundamental), while Obidi builds them into the ground floor via entropy’s irreversibility. Thus, one might ask: is the arrow of time an illusion of perspective [12], or a fundamental arrow embedded in physics [4]? The two theories answer in opposite ways. - Bridge to Existing Science: Hoffman’s theory, at least so far, provides a radical reinterpretation of known science rather than a new empirical framework – it doesn’t tell a physicist how to calculate a new result, it tells them how to interpret the meaning of results (as interface outputs). Obidi’s theory provides a new set of equations and principles that, while still speculative, aim to reproduce and extend known results. This means if one is looking for a theory to do practical unification of gravity and quantum mechanics, ToE is directly addressing that, whereas conscious realism might imply that such unification is just about deepening the interface model and can never be final [15]. - Testability and Falsification: There is a tension between the almost unfalsifiable nature of an all-encompassing idealist ontology and the more traditional falsifiability of a physical theory. If one demands near-term empirical tests, Obidi’s approach offers concrete avenues (laboratory or astrophysical tests for entropic effects). Hoffman’s approach might instead be evaluated on explanatory power and coherence with subjective experience, since it’s not easily pinned down by an experiment (short of discovering something truly anomalous like consciousness affecting quantum outcomes, which would shake physics).
In moving forward, both theories will likely evolve. Hoffman’s work might inspire more rigorous models connecting conscious agent networks to quantum computation or cosmology, potentially making the theory more predictive. Obidi’s ToE will undergo the scrutiny of developing a full mathematical model and confronting experimental data – it will either find observational support (even if just circumstantial, like the attosecond delay hint or future observations of non-instantaneous state reductions) or will need refinement if some effects fail to materialize. It is conceivable that future researchers could attempt a synthesis: perhaps an entropy-consciousness duality, where entropy production is the outside view of what, from the inside, is experienced as the flow of time or quality of experience. If such a dual-aspect view held, the ontological gap might be bridged by a deeper theory that includes both as facets.
For now, however, Hoffman and Obidi’s approaches serve as profound reminders that solving the deepest puzzles (the nature of reality, the unification of physical law, the emergence of mind) may require us to venture far beyond conventional paradigms. They exemplify two strategies: one subtracts matter from the equation, the other adds a new kind of matter (entropy) into the equation. Each strategy has its merits and pitfalls. The ultimate measure of these theories will be their ability to explain and predict our world (and our experience of the world) better than existing frameworks. Obidi’s Theory of Entropicity will rise or fall on whether entropy-as-field can quantitatively account for phenomena across scales – from cosmological constants to quantum collapse – and whether it can be integrated without contradiction into the edifice of physics. Hoffman’s Conscious Realism will stand or falter on whether it can produce a fruitful scientific research program – can it, for example, lead to new insights in cognitive science or new interpretations of quantum phenomena that can be validated, or will it remain a clever reinterpretation of what we already know?
In conclusion, the dialogue between these two frameworks – one centered on mind and the other on entropy – is emblematic of the current ferment at the foundations of knowledge. Are we witnessing the death throes of materialism, to be replaced by an era where information and experience are the primary currency of reality? Both Hoffman and Obidi would say yes, albeit in different ways. The comparison highlights that the quest for a “Theory of Everything” might force us to answer “What is everything fundamentally made of?” with something other than particles and fields in spacetime. Whether that answer is “consciousness”[1] or “entropy”[2]– or perhaps a unification of the two – is a profound ontological choice that will shape the future of physics and philosophy. This paper has mapped the landscape of that choice, elucidating the bold ideas and ontological challenges at play. The hope is that by studying such unconventional theories, we expand our imagination of what a final explanatory framework could be, and inch closer to resolving the puzzles that standard models thus far leave unanswered.
This entry is adapted from: https://doi.org/10.33774/coe-2025-hmk6nI, https://doi.org/10.33774/coe-2025-hmk6n, https://doi.org/10.33774/coe-2025-n4n45, https://doi.org/10.33774/coe-2025-vrfrx, https://doi.org/10.33774/coe-2025-7lvwh and https://wwnorton.com/books/9780393607002