Tuesday, September 30, 2025

'X-cort 2028' - a satire by Adam Carlton

X-Cort 2028

She shimmers in the surf on a sultry afternoon.

I stand a way back.

Her leg stretches, heel pushing sand aside, wetness oozing into the furrow. Her fingers play at the hem of her tiny beach dress.

"Do you like me?" she says.

Her name is Tabitha, I remember that now. She has a Slavic accent, like that President's wife. Cat's eyes, high cheekbones.

"I find you extremely attractive," I say. 

I'm calm, mellow, in control.

Tabitha leans back on the beach towel, her breasts more prominent now. She gazes skywards, runs a hand through her hair, turns to me.

"Come here and help me up," she whispers.

I nod compliantly and approach. The tip of her tongue wets her lips. She offers a languorous hand and...

I notice a small tattoo on her thigh. What does it say?

X-cort.

What?

---

The technician sits before a bank of screens. Tensor modes paint patterns on his face. He slouches back, drinking coffee from a styrofoam mug, plainly bored.

His supervisor is managing by walking around; they still do that here.

"What's up?"

"They're doing introductions. You sure you want it regular? Not the, ah, special mode?"

A brief flash of hopeful anticipation lights up the tech's eyes.

"Keep it on script," says the supervisor. And notices an anomaly on one of the screens.

"Uh-oh. Careless," he says. "'He noticed the tag. Zap him will you. That'll have to go to bugfix."

The tech leans across and moves some sliders.

---

"James," breathes Tabitha, "We have an hour before dinner. I'd like you to escort me to my room and help me... relax..."

Now the alarm bells are really ringing... but I recall I've been trained for this. I give her my warmest smile and make my apologies. Everything I say is designed to make her feel comfortable.

---

It was nearly time for Tabitha and James to leave X and those timeless Californian beaches: time to transition to production. It's been such a lengthy journey from those scary Boston Dynamics days.

But now the company was ready to embrace its users as never before.

Sunday, September 28, 2025

Why It’s Now Impossible to Write Reacher Novels


Why It’s Now Impossible to Write Reacher Novels

Back in 1997, when Lee Child introduced Killing Floor, the first Jack Reacher novel, he was writing into a cultural moment that still retained the long shadow of the 1970s, '80s, and early '90s. These were decades marked by a prevailing sense that the state was—at best—ineffectual, and—at worst—Kafkaesque: bureaucratic, corrupt, and fundamentally incapable of enacting justice.

In that context, the lone competent avenger had a mythic resonance. Reacher, a rootless knight-errant in denim and boots, was a plausible moral actor. He could stride into town, spot the rot that the authorities couldn’t or wouldn’t touch, apply force surgically (or bluntly), and stride out again with justice, more or less, restored.

He was a plausible answer to the question: Who will act, if not the state?

Fast forward to the 2020s. The cultural ground has shifted beneath Reacher’s boots. We now live in a world thick with progressivism, in which the state is no longer viewed as the obstructive adversary of justice, but as its potential agent—if only the right people can gain control of its levers. The new moral narrative doesn’t favour vigilantes; it favours systems. Bureaucracy is no longer seen as the enemy; it's the utopian mechanism awaiting proper calibration.

And so Reacher begins to look less like a mythic avenger and more like a symptom. In today's framing, he’s a homeless ex-military loner, possibly with untreated PTSD, wandering from town to town, enacting extrajudicial violence on people he personally deems guilty. There is no chain of accountability. No oversight. He is a freelance instrument of harm. And in the eyes of the contemporary moral consensus, he is less a hero than a hazard—someone who needs state intervention, therapy, and possibly sectioning. Social workers, not sidekicks.

The result? Reacher is no longer narratively viable. The entire ethical architecture that made the early novels sing has collapsed. The later Lee Child novels lost energy not just because the formula had worn thin, but because the world had shifted in such a way that the formula no longer made sense.

Enter Andrew Child. It’s tempting to blame him—and yes, his prose is flaccid, the plots undercooked, the dialogue wooden—but the rot set in before the baton was passed. The problem isn’t just that Andrew Child is not a great writer. The deeper problem is that Reacher no longer fits the world. He’s been ideologically orphaned.

There is one faintly plausible route forward: to reframe Reacher as a historical figure. Not a roaming threat in the present day, but a creature of the late Cold War, when morally-justified violence in defence of the innocent could still carry a moral charge. We’re talking military police Reacher; internal affairs Reacher. Back when the culture still had room for men like him.

But to inhabit that slight literary niche would require writing finesse, psychological insight, and narrative ambition—qualities that, to put it politely, Andrew Child has yet to exhibit. And so the Reacher novels stumble on, like their protagonist: increasingly dislocated, oddly anachronistic, and trapped in a world that has moved on.

Friday, September 26, 2025

Why Led Zeppelin never played "The Rover" live?


Hear "The Rover"

Why Did Led Zeppelin Never Play "The Rover" Live?

Nigel: Any idea why Led Zeppelin never played "The Rover" live? It's got a raw energy of its own and great lyrics.


ChatGPT: That is indeed a curious omission. "The Rover"—with its swaggering riff, Bonham’s cavernous drumming, and Plant’s cryptic-ecstatic lyrics—is arguably one of Physical Graffiti’s most underrated cuts. It would’ve slotted neatly into their mid-70s live sets, especially alongside epics like Kashmir and Trampled Under Foot. And yet… they never played it in full.

The most likely reasons?

  • 1. Studio Frankenstein: The track had a tortured birth. It began life during the Houses of the Holy sessions (1972) as an acoustic piece, later overhauled with electric overdubs. Page called the final mix “salvaged”—perhaps too fragile for the stage.

  • 2. Too Layered for the Stage Setup: Live Zeppelin were a lean machine—no rhythm guitarist, no click tracks. The Rover relies on multiple guitar layers that might have sounded thin live with just Page and Jones holding it down.

  • 3. Setlist Prioritisation: By 1975, their catalogue was overflowing. With must-plays like Stairway, Dazed, and No Quarter eating up minutes, something had to give. The Rover lost out.

  • 4. Technical Difficulty and Page’s Live Reliability: Page in the mid-70s could be… unpredictable. The Rover is tight, riff-driven, and leaves little margin for slop. It may have felt too exposed.

  • 5. A Tease, Never a Real Play: They did use the opening riff as a prelude to Sick Again on the 1977 tour—so it wasn’t entirely forgotten. But a full performance never materialised.

In short, it was too late, too layered, and not quite iconic enough to elbow out the titans. A shame, because it remains a studio triumph—half-forgotten, half-feral, howling from the mountains of Kashmir.


Nigel: It's one of my favourite tracks. Along with "Since I've Been Loving You".


ChatGPT: Those two tracks complement each other like bourbon and smoke. “The Rover” is all swagger and forward momentum, Page’s guitar tone gnawing its way through the mix like a chainsaw wrapped in velvet. Then “Since I’ve Been Loving You” is the midnight lament—Page’s solos bleeding raw emotion, Plant dragging his soul across broken glass.

They sit at opposite poles of Zeppelin’s alchemy:

  • The Rover: existential nomadism, apocalyptic optimism — “Upon us all, a little rain must fall...”

  • Since I’ve Been Loving You: agony, devotion, betrayal. Gospel-tinged blues turned operatic confession.

Both are proof that beneath the Norse-god swagger and mystical mumbo jumbo, Led Zeppelin were masters of feeling—raw, unfiltered, and cinematic.

Wednesday, September 24, 2025

Should you read Alex Karp's recent book? (Palantir)

Amazon

ChatGPT advises me:

"Don't bother. The Technological Republic reads more like a well-edited shareholder letter crossed with a graduate seminar in political philosophy. Palantir CEO Karp is fond of Hegelian abstraction, but he pulls his punches: just when you're hoping for a bold thesis, he retreats into qualified platitudes about liberal democracy and civic engagement.

"You're right: it should be a smart, astringent polemic, but it ends up a kind of technocratic sermon: earnest, worthy, slightly damp. You expect the fire of a heretic and get the measured cadence of a conscientious objector.

It’s not paradigm-busting, because the paradigms he's critiquing (Silicon Valley libertarianism, crypto-utopianism, performative wokeism) have already been well-punctured by more trenchant critics such as Evgeny Morozov, Martin Gurri, Jaron Lanier - and without the bland CEO-speak.

"There are insights, yes. But not many that wouldn’t already have occurred to a moderately skeptical reader of The Economist, with a conscience. The real interest is perhaps in the gesture - that a man in Karp’s position felt the need to write it. That he’s worried about the erosion of legitimacy, about elites losing their mandate, about the nihilism at the heart of techno-modernity.

"But to read it all? Unless you enjoy the prose equivalent of chewing through a dense quinoa cake made of qualified civic virtue... er, no.

"You can get the gist in a couple of book reviews - and one long-form podcast episode if that's not enough.

"Let it sit on the shelf with your copy of Tony Blair’s A Journey* - a monument to concern, not to clarity."


* And no, I don't have that book either.

Monday, September 22, 2025

Clare at the Château de Chaumont - but when?


This picture has been scanned in: it predates digital photography.

So when was it taken?

I asked ChatGPT:

"From the style of the photo, the colour tone, and Clare’s clothing (very 1990s: turtleneck, sleeveless vest, large sunglasses), I’d estimate this picture was taken sometime in the mid to late 1990s. Clare appears to be in her mid-40s. The Château de Chaumont backdrop and the format of the print also suggest the pre-digital-camera era, which supports that dating."

Saturday, September 20, 2025

"Grandad, why is the air transparent?"

---

The apparent transparency of air masks a deeper quantum field theoretical story. In QFT, the idea that a photon simply sails through air as if it weren’t interacting at all is a simplification—useful for optics, but incomplete. The more accurate picture is one of continual interaction, cancellation, and coherence.

Let’s break it down.

Photons in QFT Are Not Little Bullets

In quantum field theory, a "photon" is not a classical particle darting along a path, but an excitation of the quantised electromagnetic field. This field permeates all of spacetime. Its 'particle-like' properties emerge from its quantization, and it interacts with other charged quantum fields—like the electron field—wherever they are non-zero.

Atmospheric molecules (O₂, N₂, etc.) have electron clouds; these are regions where the probability density of the electron field is significant. 

So, from a QFT perspective, a photon moving through the air is continually 'sampling' the ambient electron field.

Interaction with Bound Electrons: Virtual Transitions

Photons passing through air don't typically have enough energy (in the visible range, approximately 2–3 eV) to ionise or excite the electrons in atmospheric molecules to higher real energy levels. However, they can still interact via virtual processes*. That is:

  • The photon's electromagnetic field couples to the electrons within the molecules.
  • The molecule momentarily enters a virtual excited state—forbidden by energy conservation for real, observable transitions, but allowed within the time-energy uncertainty principle (ΔE Δt ≥ ℏ/2) of quantum mechanics for very short durations.
  • The electron effectively reverts to its ground state, and the original photon is re-emitted.

This is not a real absorption and re-emission (as in fluorescence or phosphorescence, which involve real energy transfer and a time delay), but a coherent forward-scattering process—a transient polarisation of the electron clouds, which can be seen as a brief, virtual excitation of the molecules.

This is precisely what gives rise to the refractive index. The cumulative effect of all these virtual transitions across many molecules alters the phase velocity of the propagating light. In QFT terms, this is captured by radiative corrections to the photon propagator in a background of bound charged particles.

Photon Propagation in a Medium: Modified Propagator

The photon’s propagator—the function describing the amplitude for a photon to travel from one point in spacetime to another—is modified in the presence of a polarizable medium. The polarisation of the medium, arising from the response of its constituent charges to the electromagnetic field, effectively enters as a dielectric function.

This dielectric function is derived from quantum field theoretical calculations, often involving diagrams that represent the interaction of photons with the bound electrons of the medium.

This results in:

  • A modified dispersion relation: the phase velocity of light in the medium becomes vp = c/n, where n is the refractive index. This means the relation between angular frequency (ω) and wave number (k) changes from ω = ck (in vacuum) to ω = (c/n)k.
  • Possibly a small attenuation (represented by an imaginary part of the refractive index) if the photon's energy is near an absorption band of the material (e.g., in the ultraviolet for air, where real transitions can occur).

Air Appears Transparent Because of Energy Gaps

The electrons in N₂ and O₂ are bound in quantised orbitals. The visible photon energy (approximately 2–3 eV) is too low to excite any real electronic transitions in these molecules. Therefore, the real part of the refractive index dominates, and the imaginary part (responsible for absorption) is tiny in the visible spectrum.

However, at a microscopic level, the photon is always interacting—continually probing and being reshaped by the polarisation fields of nearby electrons via these virtual processes. It’s not so much "bouncing off" as continuously interfering with virtual excitations.

Quantum Coherence

From the path integral point of view, all possible paths that a photon could take contribute to its propagation, including those where it virtually interacts (scatters elastically at tiny angles) off molecules. The net result, through precise destructive interference for off-forward paths and constructive interference for the effectively straight path with modified phase velocity, is the maintenance of a well-defined trajectory and phase.

This coherence is what allows light to propagate through air as a classical wave—even though its passage involves a ceaseless flurry of virtual exchanges.

Summary

When a photon transits the electron cloud of an atmospheric molecule, QFT describes this as:

  • A coherent interaction with the quantised electron field via virtual excitations.
  • No real energy transfer (for visible light), but a phase shift—a change in the dispersion relation encoded in the modified photon propagator.
  • A collective, statistical effect of many such interactions gives rise to the macroscopic refractive index.
  • Air appears transparent because visible photon energy is insufficient for real transitions, and the medium is non-absorbing in this frequency range.

In short, the photon interacts everywhere, but with such finesse and brevity that its path appears unperturbed to our macroscopic senses. Like a dancer gliding across a floor of invisible springs.


* Appendix: Virtual Excited States and the Role of

The distinction between a virtual excited state and a real (physical) excited state is central to understanding how photons interact with matter at the quantum level, particularly in quantum field theory (QFT).

Real Excited States

These are genuine energy eigenstates of a molecule or atom. When a photon's energy precisely matches the energy gap between the ground and an excited state, a real transition can occur. This means:

  • The photon is absorbed.
  • The system is promoted to a higher, measurable energy level.
  • Energy conservation holds strictly: ℏω = En − E0.

Virtual Excited States

Virtual states, by contrast, are not eigenstates of the system’s Hamiltonian. They occur only as internal steps in quantum processes (e.g., second-order perturbation theory or Feynman diagrams). These states:

  • Do not satisfy energy conservation: the energy difference ℏω − (En − E0) is nonzero.
  • Exist only briefly, allowed by the time-energy uncertainty relation: ΔE · Δt ≳ ℏ/2.
  • Do not lead to population of excited states or observable emission; they are unmeasurable intermediates.

Their presence manifests through amplitude corrections—they affect the phase and scattering behaviour of light, giving rise to phenomena like the refractive index.

Mathematical Form

In perturbation theory, these virtual contributions appear in denominators like:

Amplitude ∼ ⟨ψ₀| Hintn⟩ ⟨ψn| Hint |ψ₀⟩ / (E₀ + ℏω − En + iε)

This is nonzero even when ℏω does not match any real excitation energy (En − E0). The result is a subtle modification of the overall scattering amplitude, not an actual jump into ψn.

The Meaning of

The term (where ε is a tiny positive number) is added to the denominator for deep mathematical reasons:

  • It shifts the pole slightly off the real axis in the complex plane.
  • It ensures causality—that effects do not precede causes in time.
  • It tells the contour integral how to correctly pass around singularities.

After integration, the limit ε → 0+ is taken. The is not a physical constant, but a calculational device—small but mighty. It’s the QFT equivalent of a signpost: “this way to causal physics.”

Summary Table

Property Real Excited State Virtual Excited State
Energy Conservation Exactly satisfied Temporarily violated
Duration Finite and measurable Infinitesimal; unobservable
Appears in Final State? Yes (can decay or emit light) No (internal process only)
Mathematical Role External line or eigenstate Internal propagator denominator
Effect on Light Absorption/emission Phase shift, refraction

So the next time you see tucked into an equation, remember: it’s the ghost in the machine that keeps the whole quantum edifice logically and causally intact.

Thursday, September 18, 2025

Fascism: birth, growth, victory, decay, death

Fascism as Systemic Outcome: An Applied Cliodynamic Analysis

The term “fascism” is often used polemically, as shorthand for repression or violence. Yet if we set aside the rhetoric and instead treat fascism as an institutional form, a different picture emerges.

At its endpoint, a fascist regime is one in which all autonomous structures that could mobilize in opposition to the leadership of the state have been dismantled. Trade unions, media outlets, political parties, voluntary associations: all are homogenized, subordinated, or destroyed. Opponents who resist are violently removed from office or silenced. Civil society ceases to operate as a check. The regime stands as a victorious political–military elite commanding a simplified social field.

Understanding fascism solely as this endpoint, however, obscures the process by which societies arrive there. Fascism is not a sudden eruption of ideology but rather one possible systemic resolution of accumulated blockages within the political economy.

The best analytical framework for grasping this sequence is provided by Mancur Olson’s theory of special interest group sclerosis, Peter Turchin’s cliodynamic accounts of elite overproduction and instability, and Joseph Schumpeter’s insights on creative destruction. Together they show how fascism can emerge as a brutal form of “political–institutional restructuring” when all other mechanisms fail.

1. Institutional sclerosis

Olson argued that as societies mature, entrenched interest groups - whether unions, business lobbies, bureaucratic cliques, or sectors of the state apparatus - accumulate bargaining power. Each group expands and defends its own rents and privileges, blocking reforms that might disadvantage it. Over time, this produces sclerosis: the economy cannot grow, political institutions become rigid and society becomes jammed. Increasingly, the parasitised state bloats, unable to fund itself.

Turchin adds a demographic and elite dimension. As more aspirants compete for limited elite positions, “elite overproduction” heightens competition, polarizes politics, and increases the intensity of factional conflict. Both mechanisms converge: reform stalls, while demands for advancement multiply.

2. Crisis of legitimacy

When sclerosis persists, living standards for the mass of people stagnate or fall. Ordinary citizens lose hope of social mobility. The state, captured by expensive rent-seekers, struggles to finance itself effectively. Public goods deteriorate, and legitimacy drains away. The perception spreads that the normal channels of politics cannot deliver change.

3. Extra-legal contestation

Blocked within institutional frameworks, groups increasingly bypass rules. Strikes, lockouts, militant demonstrations, politicised policing, and paramilitary clashes become normalised. Violence is not yet structural, but it is routinised. The arena of politics moves from parliaments and courts to the streets.

Note that different sections of the elite may back different groups and 'sides': the advent of fascism is always a civil war.

4. Elite convergence on destruction

At some point, fractions of the elite conclude that institutional demolition is the only way forward. They calculate that the destruction of obstructive interests - what they see as “negative institutions” - is a precondition for renewed growth and stability.

They therefore fund or endorse paramilitary movements willing to wield coercion against rivals. It is at this juncture that fascism begins to crystallize: not as an ideological crusade but as a violent simplification of society, clearing away accumulated blockages on the streets.

5. Autonomy of the fascist leadership

If victorious, fascist leadership enjoys a remarkable degree of autonomy. The destruction of free trade unions, independent media, voluntary associations, and opposing political parties leaves bankers, industrialists, and other economic elites with few levers of institutional influence; they are reduced to personal lobbying. 

Scholars of the interwar regimes noted how political–military elites in fascist states exercised radical independence from the remaining fragments of civil society. The fascist state is, in this sense, a spoil of victory for the faction that prevails.

6. Stability without renewal

Where fascist regimes are not ended by military defeat, they may persist for decades. Franco’s Spain and Salazar’s Portugal are cases in point. Yet they preside over a simplified and impoverished civic environment.

The brutal truth is that fascism works: its focused violence destroys those rent-seeking structures which caused the crisis, and may well remove the elite overproduction which fuelled it (by mass-killing of opponents, or by forcing them into lower strata of society).

But the original crisis that sponsored their emergence eventually passes: the specific oppositional forces they smashed no longer exist. A new period of growth now occurs, with the broader elite seeking new avenues of economic and cultural advancement. The absence of a plural civil society, once functional in clearing blockages, becomes an obstacle to modernisation and development. In victory, fascism begins its slide into irrelevance.

7. Slow decay and restoration

In such conditions, civil society begins to regrow covertly. Informal associations, cultural networks, and economic institutions accumulate beneath the surface. Over time, as generational turnover occurs and no-longer-needed repression relaxes, these networks quietly reconstitute a plural field. Eventually the authoritarian façade crumbles, often peacefully. In both Spain and Portugal, democratic structures were restored not by sudden insurrection but by the long, covert reassembly of social complexity.

Conclusion

Fascism is not best seen as an aberrant eruption of ideology or pathology - due to crazed ideologues. It is a systemic trajectory in which accumulated special interests and elite competition render normal politics impossible.

The violent destruction of civil society appears as a brutal and political variant of Schumpeterian creative destruction, clearing away obstructive institutions (at the cost of long-term vitality). Once established, fascism decays slowly, because the very simplification that makes it effective in crisis deprives it of the complexity needed for renewal.

This analysis shifts attention from the moral vocabulary of “evil regimes” to the structural dynamics of political economy. Fascism is one possible resolution of sclerosis, not chosen because it is desired in name but because, in moments of crisis, it can seem the only way out.

Its rise and its withering alike can be understood not as mysteries but as the working out of systemic pressures that Olson, Turchin, and Schumpeter each in their own way help us to map.

Tuesday, September 16, 2025

Taoism: 'Woke' in the Warring States Period


Taoism and the Overproduction of Elites: A Prehistory of Woke

Taoism is often presented as a kind of mystical naturalism: a contemplative retreat from the world into the ineffable flow of Dao, the Way. At its core, it seems to reject ambition, planning, and even language itself. The true sage, we are told, "does nothing—and yet everything is done." It is the philosophy of the mountain recluse, the poet-hermit, the fisherman-philosopher.

But this metaphysical hush, from a historical materialist point of view, is not merely a spiritual choice. It is an ideological response to very specific social conditions: the breakdown and over-centralisation of the imperial Chinese state during the Warring States period (475–221 BCE), and the rise of a large class of educated men with nowhere to go. In other words, Taoism is what happens when too many elites chase too few elite roles.

If that sounds familiar, it should.

The Historical Structure: The Asiatic Mode of Production

China in the late first millennium BCE was a bureaucratic civilisation par excellence. Its centralised states depended on irrigation, taxation, corvée labour and local village collectivism. This is what Marx once called the "Asiatic mode of production"—a model in which the state sits like a parasitic spider over a communal peasantry, extracting surplus through coercive administration (rather than capitalist accumulation).

In this setting, Confucianism became the house ideology of the bureaucratic class: a system of social ethics justifying hierarchy, filial piety, and loyal service to the emperor. It was a philosophy built for mandarins.

But by the time Taoism emerged, the system was in chaos. Competing states devoured one another. The Confucian promise of harmony through order rang hollow. Meanwhile, more and more literate young men, educated in the classics, found themselves unable to gain entry into an overstretched and corrupt bureaucracy. The elite credential pipeline was glutted.

And so a new ideology emerged, not from the peasants below, but from the disappointed aspirants of the elite: Taoism.

The Taoist Move: Withdrawal as Critique

The Tao Te Ching is not an activist tract. It is a manual for surviving a world gone mad by refusing to engage on its terms. Don't compete. Don’t strive. Don’t try to manage the empire. The more you act, the more you disrupt the natural order. The true sage flows with the current. He is indifferent to rank, untouched by politics, unmoved by moral certainties.

This is not reactionary conservatism, nor revolutionary fervour. It is what we might call structural quietism—a metaphysics forged in the psychic pressure cooker of elite overproduction. If there are no jobs at court, and the system is corrupt and collapsing, then the only rational move is to define success as withdrawal.

The Taoist sage is, in a sense, a dropout with cosmic justification.

The Modern Echo: Progressive Ideology and the Credentialled Left

Peter Turchin has coined a useful term for the modern West’s malaise: elite overproduction. Our societies churn out vast numbers of educated young people with degrees, ambition, and political awareness—but without commensurate openings in the professional-managerial class. The result? Frustration, status anxiety, and the invention of new moral frameworks to explain exclusion.

Contemporary progressive ideology—what critics call "wokeness"—is not so different from ancient Taoism in structural origin. It is the product of a class with cultural capital but no route to power. Like the Taoists, today’s frustrated aspirants are not able to seize the existing structures; so they redefine them. They claim moral authority not through office, but through identity. Oppression becomes a kind of inverted credential.

Where Taoism says "do nothing," progressivism says "speak your truth." But both are responses to blocked advancement. Taoism universalises marginalisation as a cosmic principle. Woke politics moralises it as a structural injustice. One withdraws to the mountain. The other tweets furiously from Brooklyn.

Cosmic Absolutes and Sacred Abstractions

Both Taoism and progressivism share a tendency to reify principle in response to social instability. The Dao becomes an ungraspable metaphysical constant, beyond rational critique, the source of all true order. Likewise, the modern ideologue invokes justice, equity, lived experience—terms which function not descriptively, but sacrally. To question them is heresy.

This is not a coincidence. When the real levers of power are inaccessible, meaning retreats into the symbolic. The state cannot be controlled, but the narrative can be rescripted. The institution resists reform, but ideology is more malleable.

The Ideology of the Disappointed

Taoism was never the philosophy of peasants. It was the philosophy of ex-bureaucrats, failed courtiers, and hermit-scholars who found the system both morally compromised and structurally closed.

Today’s cultural left is not the voice of the truly dispossessed; it is the ideology of the overproduced intellectual class—educated, moralising, disillusioned, and increasingly metaphysical.

This stratum projects its own sense of victimhood onto the truly dispossessed, then lobbies for remunerative jobs in the 'benevolent' part of the state apparatus, proposing to occupy positions of power and influence in order to 'fix things'.

In both cases, the system bred more ideological producers than it could absorb. And in both cases, those producers turned their disillusionment into new systems of value that rejected the old order - but lacked the power to overthrow it. So instead, they did something subtler: they redefined what power meant.

Conclusion: Rhyme, Not Repetition

History does not repeat, but it rhymes. The Taoist recluse and the woke activist share little in lifestyle or tone. But both are ideological children of blocked aspiration. One decoupled from imperial collapse with serene detachment. The other reacts to neoliberal sclerosis with permanent cultural revolution.

Neither can fix the state; both can explain its failures and make a kind of virtue out of exclusion. And in that lies their ideological momentum and persistence.

Sunday, September 14, 2025

Hyperion and the Decoherence of Worlds - (ChatGPT)

Hyperion and the Decoherence of Worlds

In a previous post, we explored how the Saturnian moon Hyperion—thanks to its chaotic rotation—provides a vivid case study for how classical unpredictability collides with quantum indeterminacy. But what stops us from encountering Hyperion in a state where it's simultaneously in multiple orientations? Why does it always appear to us as a moon tumbling this way or that, but never in some bewildering quantum blur?

The answer lies in decoherence. And in the context of the Many Worlds Interpretation (MWI), decoherence is not a marginal side effect—it’s the mechanism that gives structure and observational content to the branching wavefunction. Without decoherence, the wavefunction evolves but remains unstructured. With decoherence, the wavefunction evolves into effectively distinct classical histories.

1. Superposition: What It Is and Isn’t

A quantum system is said to be in a superposition when its state vector is a linear combination of eigenstates of some observable. For instance:

|ψ⟩ = α |A⟩ + β |B⟩

This is a statement about the system’s state in configuration space, not a claim about what one sees in any individual measurement. Upon observation, the system yields a single outcome—|A⟩ or |B⟩—with probabilities determined by the squared moduli of the coefficients. There is no such thing as “observing a superposition” in a single event.

Superposition is not a visual phenomenon, nor does it correspond to a macroscopic body appearing in multiple classical states at once. Rather, it is a mathematical descriptor of how the system's amplitudes are distributed across its configuration space.

2. Interference and Its Prerequisites

Interference is a physical phenomenon, not a formal one. It arises when different components of a quantum superposition recombine in such a way that their relative phases affect the probabilities of measurement outcomes. Interference can only be detected through statistical regularities in ensembles of measurements—such as the classic fringe patterns in a double-slit experiment with electrons.

A single electron does not interfere. Only an ensemble of identically prepared electrons, evolving through identical dynamics, can display interference effects.

For a system like Hyperion, there is no physical possibility of preparing an ensemble of identically initialised moons in the same quantum state. We cannot rerun the universe multiple times with Hyperion in the same chaotic quantum configuration. Consequently, even if Hyperion's quantum state becomes a superposition of orientations, there is no operational method by which interference effects between those orientations could be revealed.

3. Decoherence and the Emergence of Classicality

Hyperion is constantly interacting with its environment—sunlight, cosmic radiation, thermal photons, gravitational tides. These interactions cause decoherence: a process whereby phase coherence between components of the system’s wavefunction (expressed in a specific basis) is effectively destroyed by entanglement with environmental degrees of freedom.

(Here's a tutorial on the concept of density matrix, before you read further.)

This doesn’t collapse the wavefunction. Instead, it causes the system’s reduced density matrix to become diagonal in a preferred basis—typically one aligned with classical observables such as position or orientation. The states in this basis, called pointer states, are those that remain stable under environmental interactions. Because most interactions are local, the environment couples most strongly to position, making position eigenstates the natural classical basis.

In configuration space terms: decoherence suppresses the off-diagonal elements of the density matrix in the basis of classical configurations. This renders the quantum state a statistical mixture of distinguishable macroscopic states—each evolving independently. Hyperion’s rotational state, initially a quantum object, becomes a set of non-interfering classical alternatives, each entangled with a different environment.

4. Many Worlds Needs Decoherence

In the Many Worlds Interpretation, the universal wavefunction never collapses. All components persist. But without decoherence, there is no way to carve that wavefunction into meaningful branches—no structure of “worlds” that match the experienced classical order.

Decoherence solves this by defining the dynamical conditions under which different components of the wavefunction become mutually inaccessible. It ensures that each branch contains a consistent classical history, unpolluted by phase interference from others. In short: decoherence provides the effective disconnection that makes classical-looking worlds emerge from a fundamentally quantum substrate.

5. What We Don’t See

So when we say “we don’t see macroscopic superpositions,” we do not mean that there are such things which evade our detection, nor that a superposition is something you might glimpse like a ghost. What we mean—more precisely—is this:

We never observe interference effects between macroscopically distinct configurations, because we cannot prepare ensembles of systems like Hyperion in the same quantum state, and because environmental decoherence renders such interference physically inaccessible, even in principle.

The wavefunction may formally contain many such components—different orientations of Hyperion, for instance—but those components no longer interact. Their relative phases are scrambled into environmental degrees of freedom, never to return. In effect, the branches of the wavefunction become autonomous classical narratives.

6. Final Reflection

Hyperion is not both “this way” and “that way” until observed. Its quantum state evolves as a linear combination of possibilities, but decoherence ensures that these possibilities become mutually opaque long before any observation is made. There is no mystery in why we observe it in a single orientation. The mystery lies in how classicality emerges at all from the linear formalism of quantum mechanics—and decoherence is the key that makes that transition intelligible.

So the moon tumbles on, a chaotic fragment of ice and rock, participating silently in the cosmic branching of possibility. And yet, each time we train a telescope on it, we find a world that has—reliably and quietly—made up its mind.

Saturday, September 13, 2025

The Chaotic Delocalisation of Hyperion - (ChatGPT)


The Quantum Weirdness of Hyperion

The Saturnian moon Hyperion is often cited as a natural example of quantum chaos, and it has played an interesting role in debates about the quantum-classical boundary—especially regarding decoherence and the role of the observer.

1. Classical Chaos in Hyperion's Rotation

Hyperion is a small, potato-shaped moon of Saturn. It rotates chaotically: that is, its axis of rotation wobbles so much that its orientation in space becomes unpredictable, because:

  • It is non-spherical, so torques from Saturn’s gravity are complex and time-dependent.
  • It is in an eccentric orbit, and experiences perturbations from other moons (notably Titan, with which it is in a 3:4 orbital resonance).

These features lead to chaotic tumbling: Hyperion’s rotational phase changes in a way that is exponentially sensitive to initial conditions—the hallmark of classical chaos.

This is a classic example of a "three-body problem" in celestial mechanics, where the interactions between Saturn, Titan, and Hyperion make it impossible to predict Hyperion's orientation more than a few months in advance. The system's "Lyapunov time" (the timescale over which small uncertainties in initial conditions become large, unpredictable differences) for Hyperion's rotation is about 30 days.

So far, all of this is standard Newtonian mechanics.

2. Quantum Analogue: Quantum Chaos

In quantum mechanics, the classical concept of chaos doesn’t apply straightforwardly. Schrödinger’s equation is linear and unitary; it doesn’t permit the divergence of trajectories in the classical sense. Nevertheless, we can still ask: what happens to the quantum state—the wavefunction—of a classically chaotic object like Hyperion?

In 1995, physicists Wojciech Zurek and Don Paz explored this question by modelling Hyperion as an isolated quantum system—neglecting environmental interactions for the sake of analysis.

3. The Argument: Delocalisation of the Wavefunction

In quantum mechanics, a system is described by a wavefunction that evolves deterministically over time. For macroscopic bodies like Hyperion, this wavefunction is typically sharply peaked in configuration space (that is, over classical variables like orientation), which allows us to approximate the system as behaving classically.

But for a chaotic system, Zurek showed that the wavefunction becomes delocalised in configuration space—though the underlying spreading is best understood via the system’s evolution in classical phase space.

Specifically:

  • The wavefunction describing Hyperion’s rotational degree of freedom spreads exponentially over time.
  • Within roughly 20 years (some estimates suggest even less), the quantum state becomes widely spread over many possible orientations.
  • However, this spreading doesn’t lead to any observable interference, because the orientations become effectively non-overlapping in configuration space.

This is counterintuitive: the quantum state of Hyperion does not converge toward a classical trajectory but becomes a superposed, delocalised object. Yet we always observe Hyperion in a definite orientation.

4. The Role of Decoherence

The missing element is decoherence.

In reality, Hyperion is not isolated. It constantly interacts with its environment—sunlight, cosmic rays, thermal radiation, and so on. These interactions entangle the moon’s quantum state with vast numbers of environmental degrees of freedom.

Before continuing, you may wish to review this tutorial on the density matrix.

Decoherence causes the off-diagonal terms in Hyperion’s reduced density matrix (in the orientation basis) to rapidly vanish. In effect, the quantum coherence between different orientations becomes inaccessible—even in principle—because the environment has recorded “which orientation is which.”

This does not collapse the wavefunction or select a unique outcome. The total system—Hyperion plus environment—remains in a superposition of different orientation branches, each entangled with a corresponding environmental state. But from the perspective of any internal observer, each branch evolves as if the others do not exist.

So decoherence explains why Hyperion behaves as if it were in a definite orientation, even though its quantum state remains a superposition. It renders the alternatives mutually non-interfering.

5. Philosophical Bite: Many Worlds or Collapse?

What we make of this depends on how we interpret quantum mechanics:

  • In the Many Worlds Interpretation (MWI), each decohered orientation corresponds to a separate branch of the universal wavefunction. All still exist. Decoherence tells us where the branches are.
  • In collapse models, decoherence helps explain why collapse appears to occur in a specific basis—typically position or orientation—but collapse itself must still be postulated separately.

Either way, decoherence does not explain why one result occurs, but it does explain why we see stable classical behaviour, and why interference between macroscopically distinct outcomes never appears.

6. So, Is Hyperion in a Superposition?

Yes—prior to any observation, Hyperion remains in a quantum superposition of rotational states. Decoherence does not change this. It only ensures that these components are entangled with distinct environmental records and that their interference effects (if any) are physically unobservable.

This is not a superposition that could be revealed by measurement—because there’s no way to prepare an ensemble of Hyperions in the same quantum state, nor to recombine the branches. The superposition is real, but inaccessible.

7. Final Thought: Quantum and Classical Chaos

Hyperion’s case illustrates that:

  • Classical chaos amplifies quantum uncertainty by exponentially spreading the wavefunction in configuration space.
  • Decoherence suppresses quantum interference, producing the appearance of classical spacetime behaviour.
  • The interplay of these two effects is central to understanding how the classical world emerges from quantum mechanics.

It remains one of the clearest examples of where quantum theory touches the macroscopic world—not through spooky paradoxes, but through the subtle mathematics of entanglement, entropy, and environmental indifference. A tumbling moon becomes a case study in the fragility of classicality itself.

In the next post we dig deeper into the role of decoherence in all of this.

A Brief Tutorial on the Density Matrix

Density Matrices by Example: From Superposition to Partial Trace

The density matrix is one of the most powerful tools in quantum mechanics. It generalises the familiar state vector formalism and allows us to handle not only pure superpositions but also statistical mixtures and subsystems. In this tutorial we will build up the concept step by step, beginning with a simple electron in a hydrogen atom and ending with an entangled two-spin system where the reduced density matrix becomes mixed.

1. A superposed hydrogen electron

Imagine an electron in a hydrogen atom that can only occupy the first three energy eigenstates. Its quantum state is written as

|ψ⟩ = C1|E1⟩ + C2|E2⟩ + C3|E3⟩, with |C1|² + |C2|² + |C3|² = 1.

Here the coefficients C1, C2 and C3 are complex amplitudes, and |E1⟩, |E2⟩ and |E3⟩ are the energy eigenvectors forming the basis.

If you wish, you may picture |E1⟩ as the 1s orbital, |E2⟩ as 2s, and |E3⟩ as 3s, but the mathematics is basis-independent.

2. Measurement as projection

If we measure the energy of this electron, the appropriate operators are projection operators. The projector onto state |Ei⟩ is

Pi = |Ei⟩⟨Ei|.

The probability of finding the electron in state |Ei⟩ is then given by

p(Ei) = ⟨ψ|Pi|ψ⟩ = |Ci|².

Thus the squared amplitudes give us the probabilities, as expected.

3. Constructing the density matrix

We now form the density operator, which is defined for a pure state as

ρ = |ψ⟩⟨ψ|.

In the ordered basis {|E1⟩, |E2⟩, |E3⟩}, this has the explicit matrix form

ρ =
[ |C1|²       C1C2*       C1C3* ]
[ C2C1*       |C2|²       C2C3* ]
[ C3C1*       C3C2*       |C3|² ]

We immediately see some useful properties: ρ is Hermitian, its trace is 1, and since this is a pure state, we also have ρ² = ρ.

4. Probabilities and coherences

The diagonal entries of ρ are just the measurement probabilities in this basis:

p(E1) = ρ11 = |C1|², p(E2) = ρ22 = |C2|², p(E3) = ρ33 = |C3|².

The off-diagonal entries, however, are not so trivial. These so-called coherences encode information about the relative phases between the states. They become important when we measure in a basis other than the energy eigenbasis.

5. An interference example

Suppose we want to know the probability that the electron is in the superposition state

|φ⟩ = (|E1⟩ + |E2⟩)/√2.

The relevant probability is

p(φ) = ⟨φ|ρ|φ⟩ = ½(|C1|² + |C2|² + 2 Re[C1C2*]).

Notice the extra term 2 Re[C1C2*]. This arises precisely from the off-diagonal entries of ρ. The probability now depends not only on the magnitudes of C1 and C2 but also on their relative phase. This is interference in action, expressed in the language of density matrices.

6. Expectation values

To calculate an expectation value of the energy H we use the simple rule

⟨H⟩ = Tr(ρH).

So in the restricted 3-dimensional subspace we were using, Ĥ is represented as a diagonal matrix:

E1 0 0
0 E2 0
0 0 E3

That is, H = diag(E1, E2, E3).

For the Hamiltonian itself, which is diagonal in the energy eigenbasis, only the diagonal terms of ρ contribute. Thus

⟨H⟩ = |C1|² E1 + |C2|² E2 + |C3|² E3.

Here the off-diagonal terms vanish, reflecting the fact that the Hamiltonian is measured in its own eigenbasis.

7. Subsystems and the partial trace

The true power of density matrices emerges when we study composite systems. Consider two spin-½ particles, A and B, in the entangled singlet state

|Ψ⁻⟩ = (|↑⟩A|↓⟩B − |↓⟩A|↑⟩B)/√2.

The joint density matrix is ρAB = |Ψ⁻⟩⟨Ψ⁻| - a large and complex expression. To find the state of subsystem A alone, we take the partial trace over subsystem B:

ρA = TrBAB).

Carrying out the trace explicitly yields

ρA =     [ 1/2   0 ]
            [ 0   1/2 ]

This is the maximally mixed state. Although the overall two-particle state is pure, each particle by itself looks completely random. Entanglement in the global state has destroyed purity in the subsystem.

8. Conclusion

We have seen how to construct a density matrix from a pure state, how to read probabilities and expectation values from it, how the off-diagonal terms encode coherence, and finally how partial tracing over one part of an entangled system produces a mixed state.

These are the core ideas that make density matrices indispensable in modern quantum mechanics. From here one can go on to study thermal ensembles, time evolution, and open quantum systems, all of which rely on the same underlying formalism.

Thursday, September 11, 2025

ChatGPT on Kafka's literary style

---

Kafka’s style is like a dream written in the prose of a bureaucrat—a marriage of metaphysical horror and office memo. Its uniqueness lies in this exact dissonance: the nightmarish rendered in the mundane, the existential clothed in the official, the absurd narrated with a poker face.

Here are the key features that distinguish Kafka’s style:

Flat, Bureaucratic Tone for Surreal Events

Kafka reports the most bizarre and disturbing events in an affectless, precise, almost legalistic tone. A man turning into a giant insect, a court that never explains its charges, an execution without explanation—described with the same emotional detachment as a quarterly report. This deadpan tone intensifies the horror.

"Gregor Samsa awoke one morning from uneasy dreams to find himself transformed into a gigantic insect.”

No exclamation mark. No metaphor. Just a clinical report of ontological collapse.

Labyrinthine Logic

His worlds operate according to opaque rules that feel almost rational but never quite. Kafka’s characters obsess over procedure and interpretation—yet these rules are always slightly beyond comprehension. The style mimics this, with long, winding sentences full of subordinate clauses and hedged qualifications, as if the prose is itself trying to navigate a maze.

Deterministic Interiors

Kafka’s narrators are usually trapped in a world they cannot understand, yet cannot escape. Their thoughts loop in circles, trying to make sense of the senseless. The prose mirrors this entrapment, circling back on itself, internally coherent but existentially void. There’s no outside perspective, no God’s-eye view, no reliable narrator to lift us out of the quagmire.

Minimal Description of the External World

Kafka does not spend time building sensory-rich environments. His settings are sparse, generic: a room, a corridor, an office. There’s a sense of anywhereness. The world is reduced to its symbolic architecture—doors that won’t open, stairs that go nowhere, castles no one can reach. These aren’t settings; they’re psychological terrains.

Polite Dialogue Laced with Powerlessness

Characters speak in courteous, overly formal language—full of “sirs” and “if you would permit me”—but always from a position of inferiority or dread. Kafka weaponises politeness as a form of control, a performance of deference that masks deeper violence. No one ever just says something; they apologise for speaking, beg for clarification, or collapse into incoherence.

Absence of Authorial Voice

There is no ironic wink from the author, no postmodern narrator assuring you it’s all fiction. Kafka never comments, never breaks the fourth wall. The story proceeds with absolute conviction. This is not satire. It is testimony from inside the nightmare.

Unresolved Endings

His narratives often trail off without climax or catharsis. The protagonist dies, vanishes, or fades into futility. The conclusion denies the reader the satisfaction of interpretation. We are left, like his characters, waiting for a verdict that never arrives.

Kafka writes as if he's God’s stenographer in Hell: recording absurd, cruel trials in a style that is all too reasonable. His prose is the voice of the trapped man—meticulous, obedient, and quietly desperate—writing letters that will never be answered.

Tuesday, September 09, 2025

Ends in Themselves?


Evolution, Aspiration, and the End of Progress

Humanity is the only species which constructs ethical theories to justify instincts it only partially understands. We praise compassion, denounce cruelty, and speak solemnly of dignity, as if natural selection had ever cared for such abstractions. The Kantian notion that humans are to be treated as ends in themselves is an evolutionary novelty, and certainly a fragile one.

Our moral ideals are extrapolations from traits that emerged under specific evolutionary pressures. They are neither grounded in the fabric of reality nor timeless truths. The very forces that shaped those traits may begin to select against them.

1. Natural Selection and the Moral Mind

Evolution does not select for some concept of 'virtue'. It selects for survival and reproduction in specific environments. Altruism, empathy, loyalty: these emerged because they increased the inclusive fitness of our ancestors. A mother’s bond with her child, a hunter’s willingness to share meat, a group’s punishment of defectors: these are behaviours with calculable payoffs.

What we now call “morality” is the post hoc rationalisation of these evolved heuristics, abstracted into timeless principles. But this abstraction is unstable. Our evolved psychology is deeply parochial: it tracks kinship, reciprocity, and group affiliation, not impartiality. That we now demand universal human rights is a pragmatic response to society-scaling - but always involves emotional hypocrisy.

2. The Cultural Mutation of Ethical Norms

With the rise of large-scale societies, local intuitions became inadequate. The evolving institutions of religion, law, and ideology scaled up social cooperation. The result was a new moral grammar: no longer “protect the tribe,” but “treat every individual as sacred.”

Genetic-cultural co-evolution struggled to keep up.

This is the core tension of the modern ethical project. We ask humans to behave in ways the environment of evolutionary adaptation never prepared them for: to pretend to care equally for strangers and kin; to suppress immediate gain for long-term rewards; to treat even enemies with fairness.

These ideals are maintained through institutions and narratives; when social problems remove the scaffolding, when the prescribed cortical inhibitions fail, ancestral drives reassert themselves in the guise of populism.

3. The Mirage of Ethical Progress

Enlightenment thinkers imagined moral progress as the natural fruit of abstract reason. Evolutionary biology suggests otherwise. There is no moral teleology. Selection pressures continue to operate behind the scenes.

Even intelligence, that presumed engine of ethical advancement, is under ambiguous selection. High intelligence correlates with reduced fertility in post-industrial societies: there are many more interesting choices for the smart in advanced societies than raising children. Meanwhile, traits like impulsivity, status-seeking and in-group signalling continue to thrive.

4. Designing the Future Human: A False Optimism

Some imagine that technology such as gene editing, AI tutors, neural implants will allow us to engineer our better selves. But better by what metric? Intelligence, empathy, conscientiousness? Or reproductive success, influence, and competitive advantage?

They are far from the same thing. A post-human future optimised for success under future selection regimes may not resemble today's ethical ideals. The current model of “good person” may be as maladaptive tomorrow as a contemplative monk would have been in the Pleistocene. 

5. The Human Condition, Remaindered

We are, for now, creatures with reflexive consciousness: able to evaluate ourselves, question our values and attempt to live as if they matter. 

The universe is silent. Evolution is indifferent. That we aspire to be more than animals is neither destiny nor delusion. It is simply what a certain kind of animal does when its drive for meaning runs out of control.

Sunday, September 07, 2025

The new paradigm of 'Depth Theology' - by Adam Carlton


Claude (left) talking with me yesterday

---

I hadn’t seen Claude in months - and hadn’t much thought about him either to be honest. The Parisian summer, though drawing to a close, still hangs oppressively in the air. The city remains hollowed out: the smart set have yet to return from their retreats in more temperate latitudes while the tourists have already sloped back to their suburban fastnesses. Only the locals, with nowhere else to go, remain: still glistening and always irritable.

Claude, needless to say, has never been fashionable.

Party work and the usual bureaucratic tangle had kept me in town. So it was with mild surprise that I saw him yesterday, drifting along the Rue Mouffetard in the midday blaze. I steered him quickly into a shaded café, away from the sun and the street.

Claude, as I’ve mentioned in earlier pieces here, is a lapsed Communist and a hesitant Catholic, though at times I suspect he’s not entirely sure which is which. He’s long been a man bewitched by doctrines, and by their betrayals.

But something was off. He looked agitated, twitchy, his gaze flickering like someone convinced he was being watched. I knew the feeling.

“I’ve found something,” he said.

I suspect my face betrayed the weariness of habitual low expectations.

“No, really,” he pressed. “I was in a secondhand bookshop this morning - been working through the early Church Fathers - and I picked up this slim volume on Origen of Alexandria. And inside it, shoved between the pages, was this manuscript. Typed. Old. Not a bookmark. Far too big.”

He reached into his jacket and drew out a yellowed sheaf, stapled at the corner. It looked fragile, as if it might disintegrate if handled too briskly.

“What is it?” I asked. “A treasure map? The final heresy that will at last dislodge your tenuous grip on belief?”

He didn’t answer, just handed it to me across the table.

“Read it,” he said. “I think I might just agree with it.”

And that would worry him? I took the document and began to read.