Structures of Desire
Asian Provocation
Quantum Mechanics of Racism: The Schrödinger’s Cat of White Supremacy
0:00
-1:03:09

Quantum Mechanics of Racism: The Schrödinger’s Cat of White Supremacy

How White Power Both Exists and Denies Its Own Existence
Phrenology evolved into craniometry, a pseudo-science that became central to racial anthropology in the early 20th century. A stark example is the Nazi expedition to Tibet (1938), where SS anthropologist Bruno Beger measured Tibetan skulls in search of Aryan origins. This 'science' legitimized racial hierarchies, shaping ethnography and imperialist ideologies.

Racism is always somewhere else. That is the first illusion, the first disavowal. In Europe, racism is an American disease, a grotesque mutation of racial capitalism and slavery, a moral failure that exists in police brutality, in mass incarceration, in the ruins of empire—but always over there, across the Atlantic, safely externalized. In America, racism is historical, a crime already solved, an archive of grainy photos and sepia-toned atrocities, a collection of tragedies that no longer implicate the living. And everywhere, racism is someone else’s problem, someone else’s doing, someone else’s burden to bear.

The myth is simple: whiteness is neutral. It is not something that acts; it is something that simply is. Racism, within this frame, is not a structure produced and maintained by whiteness but an experience suffered by others, an affliction, an aberration, something to be studied and measured in the bodies of the racialized. A pathology belonging only to those deemed pathological. But who is doing the measuring? Who decides where racism is found? Who are these so-called neutral, colorless people—the ones who stand outside the problem they create?

The mistake has always been in the framing. Are you racist? The question is not an inquiry but an inquisitional trap, an epistemological dead end that reduces racism to a matter of personal morality rather than a structure of power. It allows for only two answers: guilty or pure, bigot or innocent—a fixed moral binary designed not to reveal racism but to absolve those in power. If racism is only a matter of intent, then structural advantage, unconscious bias, historical inheritance—none of it exists. It cannot, because to acknowledge it would be to implicate everything. And like any inquisition, only the ordained may conduct it. The language of the new church is anti-racism, but its clergy determine who stands trial, who is condemned, and who is granted absolution.

And so, whiteness retreats into plausible deniability. Racism becomes a ghost, something that haunts the margins but never the center, always disavowed, always deferred. Freud’s return of the repressed is visible everywhere: racism does not disappear; it resurfaces in new forms—resentment toward affirmative action, outrage at being labeled, the quiet, simmering anger of those who once held unquestioned dominance. The racial resentment of white liberals is no different from that of conservatives—it merely wears better branding. It speaks the language of progress while maintaining the structure of exclusion. It sponsors diversity panels while enforcing border policies. It condemns injustice in one breath and legislates against migration in the next.

This is why the classical models fail. Racism is not a stable condition, not a simple binary, not a matter of one bad actor or one bad law. It is probabilistic, entangled, shifting in meaning depending on who observes it, how it is measured, where it is spoken. The same people who declare themselves anti-racist in public wield the same structures of exclusion in private. The same institutions that parade diversity in marketing campaigns quietly uphold racial hierarchies in hiring, in admissions, in policing. Racism exists in a state of superposition—both present and absent, both admitted and denied—until an event forces it into view.

A hiring decision. A police stop. A statistical breakdown. A moment of collapse, when the illusion of neutrality shatters and something undeniable emerges: a pattern, a structure, an outcome that can no longer be explained away. The problem is not new; it is only newly visible.

What if we stopped pretending racism is a static condition and recognized it as a quantum phenomenon? Not binary but relational. Not fixed but entangled. Not a matter of personal intent but structural probability. Classical physics once assumed reality was stable—until quantum mechanics shattered that certainty. What happens when we do the same to race? What happens when we stop looking for racism in the certainty of individual guilt and start seeing it in the probabilistic field of structures, histories, and power? What happens when we stop asking, Are you racist? and start asking, Where is racism moving next?


The Limitations of the Binary Model ("Are You a Racist?")

The question—Are you a racist?—is not just an epistemological failure but functions as an ideological apparatus, a technology of innocence that allows white supremacy to remain intact. It demands an answer but allows only two: guilty or pure. Racism, within this frame, is not systemic, historical, or unconscious—it is reduced to a personal stain, a moral failing, or an error of intent. If you did not intend to be racist, you remain clean.

This Binary Model—racist or not racist—does not help us understand racism; it ensures that we misunderstand it. It operates not to reveal racism, but to foreclose the possibility of examining it. It frames racism as an individual flaw rather than a structural force, making its existence contingent on confession or denial. If one does not feel racist, then one is not racist. If one does not see racism, then it does not exist. If one believes in one’s own innocence, then the system they inhabit must also be innocent.

The result is a moralized, defensive discourse where racism is something other people do, somewhere else, in the past. Structural and unconscious dimensions vanish from view, leaving only the question of personal guilt or absolution. Under this model, racism becomes a ghost that haunts only the explicitly bigoted, never those who benefit from its structures, never those who fail to see it in their own actions.

For Lacan, “la vérité a structure de fiction1 or truth is structured like fiction—it is not a direct reflection of reality but something mediated through language, fantasy, and the unconscious. Disavowal operates within this structure, allowing subjects to acknowledge something while simultaneously refusing its full implications. In a secular Christian society, the denial of racism assumes a theological role. It is not enough to avoid racist actions; one must also believe in one’s own innocence. This is why, when Kanye West accused George W. Bush of “not caring about Black people,” Bush did not engage with the claim itself. He did not interrogate the racial impact of his policies, nor did he attempt to disprove the charge on factual grounds. Instead, he simply resented being accused. The wound was not in what had happened, but in being made to account for it.

This is the fundamental operation of disavowal (Verleugnung): "I know very well, but even so..." A subject may acknowledge racism as an abstract reality (Yes, racism exists), yet refuse to see it where it implicates them directly (But I am definitely not racist). Disavowal allows one to speak of racism and acknowledge its existence, but only at a safe distance—in the past, in other societies, but never here, never now.

For some, the accusation reaches repression (Verdrängung). The thought is pushed out of awareness, repressed into the unconscious, only to return in slips, contradictions, and unconscious biases. It is not truly forgotten—only displaced, resurfacing in anxieties about "political correctness," in the discomfort of discussing race, in the inexplicable resentment toward those who bring it up. The subject who represses may even find themselves haunted by the very thing they seek to deny.

For others, the accusation triggers foreclosure (Verwerfung). Unlike repression, foreclosure does not merely push the thought away—it banishes it entirely from the symbolic order. There is no internal conflict, no unconscious return. The possibility is never entertained to begin with. Here, the very question—Are you racist?—is not just offensive; it is incomprehensible. To suggest racism exists in the present is to utter nonsense, to violate the Real. The reaction is outrage, not because the accusation is untrue, but because it is unthinkable.2

Each of these psychic maneuvers—disavowal, repression, foreclosure—serves the same function: to neutralize the disturbance of being implicated in a structure one depends on. Disavowal allows the subject to maintain racism as a theoretical concern; repression buries it only for it to return in unintended ways, and foreclosure renders it nonexistent.

But all of these responses remain bound to the binary model—the illusion that racism must be either present or absent, conscious or unconscious, real or imagined. This model does not just fail to capture racism; it actively distorts it. It ensures that racism can persist while remaining invisible.

The classic model of racism—as something overt, intentional, and conscious—is a relic of a world that wants to believe racism is a problem already solved. But racism is not merely an ideology that individuals hold; it is a network of power, a historical structure, a mode of organizing reality. The moment we reduce it to a personal defect, we have already misidentified it.

A better question than Are you racist? might be: Where is racism? Where is it operating, where is it concealed, where does it move? Because as long as we remain fixated on who is racist, we will fail to see how racism functions.

Let:

  • D = Denial of racism

  • R = Recognition of structural racism

  • S = Strength of racism in a given system

Then:

  1. The more Denial of Racism increases, the more Recognition of Structural Racism decreases:

\( R = \frac{1}{D} \)

  1. The lower Recognition of Structural Racism, the more Strength of Racism thrives:

\(S = \frac{1}{R}\)

(The less racism is recognized, the stronger it becomes.)

  1. Since Denial of Racism is itself a function of Strength of Racism, we get a self-reinforcing loop:

\(D = f(S)\)

meaning that as racism strengthens, the urge to deny it also intensifies.

Thus, in recursive form:

\(D_{n+1} = f\left(\frac{1}{D_n}\right)\)

or more simply:

\(D \to R \to S \to D\)

The cycle ensures that racism remains unexamined. The more forcefully one denies, the deeper the repression. The deeper the repression, the harder it is to recognize structural racism. The harder it is to recognize, the more racism thrives unchallenged. The stronger it thrives, the more denial becomes necessary to maintain innocence.

This formulation captures the self-perpetuating nature of disavowal—where denial doesn’t just obscure racism but actively accelerates its invisibility in a closed loop.

Racism as a Superposition State

Schrödinger’s cat3 was a thought experiment meant to illustrate the paradoxes of quantum mechanics: a cat in a box, both alive and dead at the same time, existing in a state of superposition until an observer opens the box and forces a definitive outcome. Until that moment, the cat is neither fully one nor the other, but both, hovering in probabilistic uncertainty. Racism, too, operates in superposition. It is both present and absent, both admitted and denied, until the moment of observation collapses its wavefunction.4

The liberal, convinced of their commitment to anti-racism, remains in an ambiguous state. In their mind, they are against racism, committed to diversity, structurally innocent. Yet in their actions, racism persists—not as personal intent, but as systemic inertia, as unconscious bias, as the privilege of never having to look too closely. The contradiction remains unresolved until the moment of measurement. A hiring decision, a police stop, a protest—each is a point of collapse, an event that forces a stance. The neutrality that once seemed self-evident dissolves, replaced by something measurable: a choice, an outcome, a pattern.

The legal system functions much the same way. It insists on objectivity, on fairness, on a colorblind application of justice. It hovers in a state of supposed neutrality—until reality forces its collapse. A statistical breakdown of sentencing disparities, an analysis of police stops, a ruling that follows the same racial contours as those before it. The system that claimed to be above race is suddenly revealed to be structured by it. The illusion of neutrality was real only so long as no one measured it.

The same principle extends to institutions, to corporations, to entire nations states. A company declares itself inclusive, a government proclaims itself post-racial, a society prides itself on meritocracy—each assumption held in superposition until the data emerges, until a crisis forces alignment, until the comforting vagueness of principle meets the sharp reality of practice. The moment of observation does not create racism, but it makes its presence undeniable.

The classical model of racism, the one that asks simply, Are you racist?, fails because it assumes a fixed, moral state, a stable, observable reality. It presumes that if racism is not consciously expressed, it does not exist. But racism, like quantum phenomena, is not fixed; it is relational, entangled, dependent on the act of measurement. The mistake is assuming that what is not immediately visible is not there at all. The real illusion is not racism itself, but the belief that it vanishes simply because no one is looking.


Racism as a Probability Distribution

If racism is a superposition, we can express it as a quantum state function:

\(∣Ψ racism ​ ⟩=a∣R^1 ​ ⟩+b∣R^ 2 ​ ⟩\)

Here, this describes this suspended reality. The wave function ∣Ψ_racism⟩5 holds both possibilities at once: the state of being racist and the state of being not racist, neither fully materialized until the moment of observation. The symbols are precise, but the reality they describe is slippery.

The terms ∣R1⟩ and ∣R2⟩ define the two states. ∣R1⟩, the "not racist" state, is the declared innocence, the moral self-perception, the insistence that one simply could not be racist. It is the version of the self that believes in equality, that votes the right party, that has the right friends and uses the right labels. But alongside it, inseparable, is ∣R2⟩, the "racist" state, the one shaped by structures, unconscious biases, and the quiet imposition of history and power. It is the hiring decision, the real estate policy, the assumption made in a split-second glance. Both of these states exist together, always.

The coefficients a and b, squared, determine which state will manifest upon interrogation. They are the probabilities—how likely it is that, when pressed, a person will collapse into righteousness or implication. Some people lean heavily toward one side, but most hover in between, their certainty masking a deeper uncertainty.

And then there is the collapse. A person, an institution, or a system remains in superposition until measured, until the moment they are forced to answer: Was that racist? This is when the disavowal kicks in, when the insistence sharpens: That was definitely not racist. The quantum haze vanishes, and a single reality is chosen. But it is not the reality itself that changes—only what is visible, only what can now be acknowledged. This is the paradox of modern racism: not that it is absent, but that it is suspended, held in probability, waiting for the right conditions to come into view.

Racism in the present is always an event horizon—impossible to fully acknowledge while inside it or seeing beyond it. It remains obscured by repression, disavowal, and the limits of contemporary foreclosure. Yet, once it has passed, it retroactively appears inevitable. In hindsight, it is always recognized—repackaged as an obvious wrong, a historical lesson, a thing of the past that of course everyone saw.

Take Apartheid South Africa. Nelson Mandela, now a global icon, was once labeled a terrorist by Margaret Thatcher, who refused to sanction the apartheid regime. The same Western powers that now celebrate him fought tooth and nail to suppress his movement.

Or segregation in the U.S.—Martin Luther King Jr. was demonized as a terrorist, his peaceful protests met with police violence. White moderates urged patience, claiming integration was moving too fast. Today, they quote him selectively, pretending they were always on the right side of history.

And consider that Nazi Germany drew direct inspiration from American segregation laws and eugenics programs, many of which were shaped by leading academics at Stanford University. The Nazis admired U.S. racial purity laws and even debated adopting America’s one-drop rule before rejecting it as too extreme. David Starr Jordan, Stanford’s founding president, was a vocal eugenicist who promoted forced sterilization and racial hierarchy through the American Breeders’ Association.6 Lewis Terman, another Stanford professor, expanded these ideas through intelligence testing, arguing that eugenics could identify and control "genetic inferiors." Their work helped establish the legal and scientific framework for racial policies that the Nazis studied and adapted. Yet, after World War II, the Holocaust industry worked to erase these transatlantic ties, rebranding Nazism as a uniquely German pathology rather than part of a broader Western racial project.

There’s also the white feminist imperialists like Thatcher, Emmeline Pankhurst,7 Hillary Clinton,8 and Madeleine Albright,9 who framed their struggles as universal while defending colonialism, mass incarceration, and military interventions in the “Global South.” They opposed apartheid and oppression only when it became politically convenient.

And now, the genocide in Palestine. To name it as such is considered radical, impolite, even extremist. The media insists on neutrality, on waiting for history to settle the facts. But history does not settle facts—it only provides the distance needed for those in power to claim they were never complicit. In a few years, even the whitest liberals will say, yes, it was terrible, it was racist, it was genocide. But today, they insist, that the Israel-Palestine Conflict is too complicated to say for sure.

This is how racism functions as an event horizon. It is unseen in its moment, yet in hindsight, it was always recognized. The only ones who see it as it happens are those who suffer it—and those who refuse to wait for history’s permission to speak.


The Observer’s Paradox: How Measuring Racism Distorts and Sustains It

In quantum mechanics, the act of measuring a system changes its state. The same is true for racism.10

A racialized person enters a room, and the atmosphere shifts—subtly at first, a pause, a flicker of recalibration. The Persons of Colorlessness11 adjust. Faces contort. Some laugh too hard. Some avoid looking. The room was never neutral, but now its asymmetry is exposed.

When asked about racism, everyone scrambles under this same effect. The subject hesitates, calculates, and answers not with truth but with the answer that will make them seem like the kind of person who does not need to be measured. The numbers adjust, smoothing over the gap between belief and behavior.

Legal institutions, too, know better than to ask too precisely. To define racism with precision would be to trap themselves in it, to make it legible in ways that could be counted, quantified, litigated. Instead, they let it remain spectral, unspeakable in legal terms, a thing best left in policy footnotes rather than at its center.

Measuring racism, then, is never neutral. Who holds the instrument? Who is under the lens? The act of observation does not reveal—it rearranges. Self-reporting will always smooth over contradiction, but the structures remain indifferent to individual confessions. And history does not disappear simply because no one is asking the right questions. It lingers, shaping what is seen, what is said, what is measured, and what is quietly left out.


The Racist Entanglement of Bell’s Theorem

Bell’s theorem shattered the illusion that reality is local, that things exist independently, that cause and effect move in neat, linear chains. It proved instead that particles, even separated by vast distances, remain entangled. Change one, and the other shifts, no matter how far apart they are. Reality does not behave the way we think it should. Racism works like that, too. It does not exist in isolated acts of prejudice or in the convenient villainy of the far-right. It moves through institutions, through economic policies, through the reflexive denials of those who claim to oppose it. It is in the laws that criminalize migrants and the think tanks that fund diversity panels, in the police departments that kneel for the cameras and the governments that bomb foreign cities in the name of freedom. It is not one ideology or one party. It is entangled.

Columbia University’s board members,12 who voice public commitments to diversity while investing in arms manufacturers, surveillance technology, and gentrification projects, are not contradictions. They are proof that the machinery of oppression requires liberalism to function. The German left,13 which condemns racism while passing laws that criminalize Palestinian solidarity and uphold deportation policies against migrants, does not fail to see its hypocrisy—it depends on it. The well-meaning white academics and journalists who write papers on anti-racism but default to gatekeeping when confronted with voices that challenge their authority do not undermine their own work; their authority was always contingent on keeping certain voices in their place.

Bell’s theorem14 tells us that particles do not have pre-existing, independent states waiting to be revealed; their properties emerge only in relation to the system. Systemic racism operates the same way. It is not reducible to individual bad actors or singular policies, not just a matter of punishing one cop, indicting one leader, banning one hate group. The mistake is thinking you can isolate it. You can’t. The moment you try, it has already moved elsewhere. Execute one CEO, and capital reorganizes itself before the body hits the ground. Dismantle one border wall, and immigration laws tighten, visas become impossible, entire populations rendered stateless with the stroke of a bureaucrat’s pen. Pass affirmative action, and white resentment metastasizes, clutching at the illusion of meritocracy, reframing the loss of an inherited advantage as persecution. The structure does not break. It adapts, absorbs, finds new ways to justify itself, slipping just beyond the grasp of reform.

Racism is nonlocal. It spills across borders, across centuries, into places its architects never imagined. It lingers in the way institutions function, in the way opportunities narrow, in the way power defends itself. The liberal establishment, the academic gatekeepers, the journalists who frame oppression as a historical artifact rather than an active structure, the think tanks that pay lip service to justice while ensuring nothing truly shifts—these are not contradictions in the system. They are the system. White liberals are not better than conservatives; they are simply better at maintaining plausible deniability. The far-right is the attack dog, but white liberals are the gatekeepers. Without either, white supremacy would collapse.


The Heisenberg Uncertainty Principle of Racism

The uncertainty principle states that you cannot precisely measure both the position and momentum of a particle.15 The closer you look at one, the more the other slips out of focus. Racism works like that, too.

Try to measure intent and impact blurs. Fixate on impact, and intent becomes unknowable. The harm exists regardless, but the debate locks into place—did they mean it? did they know? was it racism or just a mistake? These questions drag the conversation sideways, away from the structure, toward the individual. Zoom in on a single racist act, and the system disappears. Zoom out to analyze the system, and the individual becomes irrelevant. A well-meaning person upholds a racist institution, a virulent racist gets caught in a bureaucracy that operates without malice. The pattern is clear, but the measurement never quite satisfies. Try to trace today’s racism back to its historical roots, and the distortion begins. How much of the present is because of the past? How much is something new? The past is not past, but in measurement, it never arrives completed.

The Heart Sutra,16 one of the most well-known texts in Mahāyāna Buddhism, illuminates this paradox. In its most famous line:

色即是空,空即是色
Sè jí shì kōng, kōng jí shì sè
"Form is emptiness, emptiness is form."

It does not mean that form does not exist, just as it does not mean that emptiness is a void. It means that the nature of things is contingent, relational, impossible to fix in absolute terms. Just as the more you measure racism, the more its dimensions shift—individual and systemic, historical and immediate, visible and unseen—so too, all phenomena lack independent existence. They arise through causes and conditions, through interwoven histories and structures, and are fundamentally impermanent.

The Heart Sutra, dating back to the first few centuries CE, likely emerged within the Prajñāpāramitā tradition in India before being translated into Chinese by 玄奘 (Xuán Zàng)17 in the 7th century. Its influence spread through East Asia, shaping Zen, Tibetan, and Pure Land traditions. It is short, paradoxical, meant to be recited, not merely understood. It unsettles conceptual grasping. Its famous refrain—

無無明,亦無無明盡
Wú wúmíng, yì wú wúmíng jìn
"No ignorance, also no end of ignorance."

—could just as easily apply to racism: there is no single thing called racism, no place it begins or ends, only a field of entanglement, causes arising through other causes, systems reinforcing themselves, perception shaping reality even as reality undoes perception.

To measure racism, then, is not to pin it down, to isolate it in a single instance or statistic. It must be mapped probabilistically—both individual and systemic, both past and present, both intention and effect. The moment you try to make it simple, you have already lost sight of what you were trying to see.

Desire as Variable in Wave-Particle Duality

Decolonial psychoanalysis unravels the racial fetish at the core of Western desire. Freud’s theory of fetishism explains how the subject displaces castration anxiety onto an object of fixation, but in the colonial order, this object is not a thing but a person—racialized, desired, despised. The colonial gaze does not see its subject as fully human. It fractures them, turns them into a paradox: hypersexual yet repressed, seductive yet forbidden, always available yet never truly accessible. This is why white attraction to racialized bodies does not contradict racism but thrives within it. Desire does not dismantle the structure of domination; it reinforces it. The colonial subject is not desired despite their subjugation, but because of it. The attraction is not incidental—it is foundational.

Wave-particle duality18 states that a particle exists as both a wave and a particle until measured, refusing to resolve into one form or the other until forced into observation. Racism, too, exists in dual states, but with a third variable: desire. It is not simply a force of exclusion or oppression; it is also fascination, fetishization, an obsessive fixation masquerading as aversion. The colonial subject is both invisible and hyper-visible, erased from history yet hyper-exposed under the white gaze, always watched, always contained.

Desire distorts measurement. It bends perception, makes the structure less legible, turns racism into something more complex than domination alone. Attraction does not negate racism; it intensifies it. The white subject who desires the racialized other believes themselves free of racism, unaware that their desire is structured by the same colonial gaze that constructed the fetish in the first place. The powerful resent their own attraction. White liberals are uneasy when they desire the people they structurally oppress. The gaze twists, tries to unsee itself. And when it cannot, it lashes out. Racist backlash does not emerge from pure repulsion but from the unbearable recognition that power and desire are not separate.

To measure racism without accounting for desire is to miss its most volatile mutation. White resentment is often rooted in attraction turned to shame—resentment toward the Other, toward the sexual availability of the racialized subject, toward the economic threat of those who were supposed to remain beneath. The reactionary turn—from attraction to exclusion, from fascination to violence—is not a break but a continuation.

Racism and desire exist in superposition, bound together in a structure that makes measurement unstable. The system does not simply produce oppression; it produces the conditions of attraction and its inevitable disavowal. And desire, like racism, cannot be trusted to report on itself. Self-reports of racism are unreliable, but so are self-reports of attraction, of preference, of resentment. The subject does not fully know what they desire or why. Colonial memory lingers in sexual economies. Desire is inherited, just as power is. The attraction to whiteness, to certain racialized aesthetics, to submission and dominance—these are not natural but historical. Racism is not just a system of exclusion; it is also a system of wanting, of looking, of fantasizing.

The mistake is thinking you can measure it without drowning in desire. Whiteness does not merely seek to dominate—it seeks psychic stability, a sense of order that requires racial hierarchy not just for material power, but for its own coherence. It needs Blackness, Asianness, Arabness, Indigeneity, Muslimness, Otherness—sites of projection, scapegoating, and fetishism. White desire is structured by disavowal. The white subject desires what it also seeks to control, exclude, erase. The same contradiction fuels racial capitalism.

Desire and repulsion are not opposites. They are the same movement, seen from different angles. The white liberal who consumes Black culture but fears Black power. The one who adopts Asian spirituality but labels Asians as a geopolitical threat. The one who claims to love diversity but panics at demographic shifts. The one who desires the racialized other but cannot bear to see them as human. The one who watches, wants, and then looks away.


The Racist Measurement Matrix

Racism does not sit still. It does not remain in one state, one form, one category. It is not something that can be neatly classified as present or absent, explicit or implicit, structural or personal. It is all of these things at once, moving, shifting, and collapsing into one state only when observed.

This is why racism must be measured not as a fixed quantity but as a probability distribution, a shifting matrix of entangled states existing in Superposition until the moment of inquiry forces it into visibility.

In Superposition, someone can be both racist & anti-racist. A white ally supports anti-racism while reproducing microaggressions unconsciously.

In the Observer Effect, we can observe how measuring racism changes responses, where People of Colorlessness downplay racism in self-reports.

In Entanglement, racism is relational, not isolated. Colonial policies continue to shape immigration and labor laws.

In Wave-Particle Duality, racism exists in multiple forms until it is observed. A policy can be race-neutral in language but racialized in impact.

Putting all this together, we can use a mathematical formulation of a Density Matrix Approach. Rather than treating racism as a binary or a singular score, a quantum model defines it as a density matrix, where racism exists probabilistically across different states:

\(\rho_{\text{racism}} = p_1 |R_1\rangle \langle R_1| + p_2 |R_2\rangle \langle R_2| + p_3 |R_3\rangle \langle R_3| + \dots \)

The density matrix of racism, ρ_racism, is a weighted sum of different possible states of racism, p_n​​ are probability weights assigned to different racism states, where each state ∣Rn⟩ is assigned a probability p_n​, representing how likely that particular manifestation of racism is in a given system.

This model allows for multi-dimensional measurement, where racism is neither a single score nor a fixed category but a dynamic, probabilistic field. The probability distribution shifts depending on who is measuring, where it is being measured, and in what context—just like quantum systems collapse upon observation.

Racism does not exist in isolation. It exists in relation—to power, to history, to perception. Any attempt to quantify it must accept that measurement is distortion, that meaning changes in different frames, that racism is as much about what is felt as what is enforced.

To measure racism is to acknowledge that it cannot be captured in one number, one definition, one moment. It is a system of probabilities, entangled across time, space, and desire. The mistake is thinking you can observe it without changing it. The mistake is thinking you can measure it without becoming part of the measurement itself.

In psychoanalysis, the Real is that which cannot be fully symbolized or integrated into discourse. Racism, in its deepest sense, operates at this level. It cannot be neatly measured, solved, or eradicated—because it is not simply a political problem, but a foundational structure of the modern world.

This is why liberal anti-racism fails: it believes that with enough education, enough discussion, enough policies, racism can be reasoned away. But decolonial psychoanalysis tells us otherwise: racism is lodged in the unconscious, in history, in language itself. The task, then, is not to seek a final solution to racism but to develop a new grammar for seeing, speaking, and dismantling its structures from within.

Measuring Racism’s Recalcitrance within a Quantum Resistance Model

Racism does not just persist—it resists. It adapts, mutates, and doubles back on itself. Attempts to measure or dismantle it often provoke reactionary forces that entrench it further. This is racism’s recalcitrance: its refusal to be eliminated, its ability to evade accountability by shifting forms.

The Quantum Zeno Effect19 describes how continuous measurement prevents change. In psychoanalytic terms, this is the function of the superego’s demand for racial innocence. White subjects are caught in an endless cycle of proving that they are 'good people'—posting the right hashtags, saying the right phrases, denouncing the right enemies. But this obsessive self-monitoring does not lead to transformation; it freezes the subject in place. Anti-racism becomes another form of self-surveillance, a performance for the internalized gaze of the superego. And yet, as Freud tells us, the more one tries to repress a thought, the stronger it returns. This is why whiteness, despite its moral posturing, remains deeply fragile—because it cannot bear the truth of its own foundational violence.

If racism operates in superposition—existing in multiple states at once—then recalcitrance is the mechanism by which it avoids collapse into a single, observable truth. The harder you try to pin it down, the more elusive it becomes and the stronger the disavowal. This can be observed in the Quantum Zeno Effect: the more frequently racism is debated, the harder it becomes to change. This can be seen in the endless discourse about intent vs impact that prevents structural reform.

If racism is modeled as a quantum state, recalcitrance functions as a potential energy barrier—a force that prevents progress by absorbing or redirecting efforts to dismantle it.

We define the recalcitrance function R(x,t), where:

  • x represents the intensity of anti-racist intervention.

  • t represents time (historical persistence).

  • V(x) is the resistance potential, which increases when anti-racist measures are introduced.

\(\psi_{\text{racism}} (x,t) = e^{i(kx - \omega t)}\)

where:

  • ψ_racism is the wavefunction of racism, shifting states in response to measurement.

  • k represents systemic inertia—how deeply embedded racism is.

  • ω represents adaptability—how quickly it mutates.

To model resistance to change, we introduce a potential barrier function V(x):

\(H \psi = \left( -\frac{\hbar^2}{2m} \nabla^2 + V(x) \right) \psi\)

where:

  • H is the Hamiltonian operator20 (total energy of the system).

  • ∇^2 represents how deeply racism is embedded in social structures.

  • If V(x) is high (strong resistance), racism remains unchanged despite external pressure.

White supremacy has learned its own defense mechanisms. When confronted, it does not attack directly—it absorbs. It does not silence discussion—it sponsors it. The best way to prevent a system from changing is to ensure it remains under constant scrutiny, measured at every moment, debated endlessly but never acted upon.

The political economy of race-talk—diversity panels, DEI workshops, racial sensitivity trainings—is not an accident. It is a containment strategy. The more racism is analyzed, dissected, and discussed, the less it needs to be dismantled. It is no coincidence that anti-racism, in its most institutionalized forms, now resembles a lucrative consultancy industry. Talking about race has become a multi-billion-dollar business, one that thrives on the endurance of the very structures it claims to challenge.

If racism remains in constant measurement, it cannot progress or change. The discourse freezes in place, cycling through the same questions: Is racism real? How bad is it? Are we overreacting? Each time, the question arrives as if it has not already been answered a thousand times before. This is not the sign of a culture reckoning with its sins; it is the sign of a system stalling for time.

Racial discourse is not the same as racial transformation. The production of debate—on television, in academia, in policymaking—serves a purpose: to give the illusion of movement while ensuring stasis. The more time spent defining racism, the less time spent dismantling it. The more precise the language of injustice becomes, the more it is defanged, rendered polite, slotted into bureaucratic frameworks that make revolution impossible.

This is the Quantum Zeno Effect in action: constant observation prevents the collapse of a system into a new state. A racist structure, when placed under continuous scrutiny, does not dissolve—it calcifies. Every debate over whether racism exists, every polite argument over its severity, every call for patience and incremental change is another act of reinforcement.

The mistake is assuming that the conversation itself is progress. It is not. The function of whiteness is to sustain itself, and its most insidious trick is the illusion of self-critique. This is why white liberals would rather debate racism than dismantle it. This is why anti-racism, institutionalized, becomes another arm of white supremacy, another buffer zone between power and its consequences.

There is a reason why no revolution ever came from a panel discussion. White supremacy does not fear dialogue—it sponsors it. The endless discourse on race is not a challenge to the system but a symptom of its survival. Piers Morgan does not fear debate; he thrives on it. His entire function, like that of so many others, is not to engage in discussion but to trap it in an eternal loop. The spectacle of argument replaces the possibility of change. The most effective strategy for ensuring nothing changes is to talk about it forever, to keep the conversation going just long enough for history to forget what it was supposed to act upon.

In 1970, Australian journalist Richard Carleton asked Ghassan Kanafani21 why his organization would not engage in peace talks with the Israelis. Kanafani did not hesitate: “You don’t mean exactly peace talks. You mean capitulation, surrender.” When pressed further—Why not just talk?—his response was sharp, final: “Talk to whom?” Carleton answered, “Talk to the Israeli leaders.” And Kanafani, without blinking, laid bare the structure of power: “You mean a conversation between the sword and the neck.”

These conversations are not neutral. They are structured by power, by who sets the terms, by who holds the blade and who is expected to remain civil as it rests against their throat.


Conclusion: A Quantum Approach to Racism Measurement

Racism is not an object to be studied under controlled conditions. It cannot be located, extracted, or removed like a tumor. It is not a question of bad actors, bad policies, or isolated historical moments. It is structural, yet ambient. Explicit, yet imperceptible. A contradiction that refuses resolution. To say it exists is banal. To say it persists despite every law, every diversity initiative, every moral consensus that declares it wrong—that should be more alarming than it is.

If racism is a quantum phenomenon, then measuring it does not merely reveal its presence; it alters its state. But who holds the instrument? Who decides what is being measured? Is it the one who experiences it, or the one who denies it? If a racialized person names an experience as racist and a white observer declares it neutral, whose perception collapses the wave function? And what if the denial itself is the mechanism through which racism remains intact?

The act of measurement is never neutral—it is an operation of power. Those who define racism are those who hold the power to excuse it. We are told that racism is either past or present, either structural or individual, either explicit or unconscious. That it is a thing of history, yet somehow also a thing that never truly happened—at least not here, at least not now. But what if it is both at once? What if it is neither? What if the binary itself is part of the machinery that ensures its survival?

The legal system remains in superposition until a case forces it to collapse into a ruling. But this process is not neutral. Power structures dictate when and how the collapse occurs. Perhaps the question is not whether racism exists, but whether it is possible to see it while standing inside it. Perhaps the reason it seems impossible to measure is that it has already determined the coordinates of our measurement. Perhaps the reason it cannot be isolated is that we are measuring it from within its field.

What if racism is not a problem to be solved but a condition to be understood? And if it is a condition, what is its half-life?22 What is its rate of decay? Or does it not decay at all, only reorganizing itself, shifting probabilities, finding new ways to exist just beneath the threshold of recognition? And if that is true, then what happens when we stop looking? Does it disappear? Or does it spread—unmeasured, undisturbed, untouched by the instruments designed only to detect it when it is already too late?

The accusation of racism triggers a crisis in the white subject’s superego. The superego, as Freud describes, is the internalized voice of authority, punishing the subject with guilt. But guilt, as Lacan reminds us, is not a true ethical reckoning—it is a way to avoid responsibility. White guilt does not lead to dismantling racism; it leads to hyper-moralism, fragility, and the performance of anti-racism without structural change. This is why racism remains caught in an eternal loop—the white subject experiences the accusation as an injury, not a revelation. The response is not introspection but self-defense: I am not racist! A desperate attempt to ward off the punishment of their own superego.

White supremacy depends on keeping racism in a quantum haze23—never fully admitted, never fully denied. Racism remains in unresolved potentiality because acknowledging it fully would demand action. The origins of quantum mechanics, like those of anthropology, psychology, and genetics, are rooted in European intellectual traditions that excluded racialized knowledge systems. While quantum theory here serves as a metaphor, it bears mentioning that physics itself, like every field, has been wielded as a tool of white dominance. Werner Heisenberg,24 Niels Bohr,25 Schrödinger26—none of them saw their discoveries in relation to colonialism or racial politics. But their epistemic framework emerges from a tradition that assumes an unmarked observer—a classic Western colonial gaze.

The assumption that measurement is neutral is itself a white supremacist idea. Whiteness has historically positioned itself as the universal measurer, the neutral observer of reality. Just as quantum measurement affects what is measured, white supremacist epistemology determines what is "real" in the first place. Racism is not merely difficult to measure—it is measured through a system designed to obscure it.

This is why any critique functions like a reverse dog whistle—it does not summon white supremacy so much as reveal it. The very act of speaking makes it known. The hornet’s nest does not need to be found; it has already been built around you. Racism is not the Real because it is unknowable. It is the Real because it is unbearable for whiteness to fully acknowledge.

To say that racism is too complex to solve is itself a form of white deferral. Racism is not mysterious—it is profitable. It is not unsolvable—it is maintained. The reason white supremacy avoids measurement is not because racism is too difficult to see. It is because seeing it clearly would demand dismantling it.

So then, the final question is not whether racism exists, or whether it is possible to measure it, or whether it can ever be solved. The final question is: what happens when we stop pretending it is complicated? And Who, exactly, is most afraid of the answer?

1

Lacan, Jacques. Seminar XVII: The Other Side of Psychoanalysis. Translated by Russell Grigg. New York: W.W. Norton & Company, 2007. Lacan’s assertion that truth has the structure of fiction challenges the notion of an objective, unmediated reality. Truth, in his formulation, is not simply what is, but emerges through the symbolic order—through language, narrative, and the unconscious formations that shape subjectivity. This does not mean truth is a lie, but rather that it is constructed, contingent, and dependent on the discursive frameworks through which it is articulated. In psychoanalysis, this is particularly evident in the way subjects recount their histories: not as neutral recollections but as stories shaped by repression, desire, and identification. Likewise, in political and ideological structures, truth is never self-evident but operates within a field of meaning shaped by power, discourse, and historical contingency.

2

Freud, Sigmund. The Standard Edition of the Complete Psychological Works of Sigmund Freud, Volume XIX (1923-1925)

3

Schrödinger, Erwin. "Die gegenwärtige Situation in der Quantenmechanik." Naturwissenschaften 23, no. 48 (1935): 807–812. Schrödinger introduced the cat paradox to critique the Copenhagen interpretation of quantum mechanics, which suggested that a system remains in superposition—existing in multiple states at once—until measured. The experiment imagines a cat in a box with a radioactive atom that has a 50% chance of decaying within an hour, triggering the release of poison. Until the box is opened and an observer confirms the state, the cat is theoretically both alive and dead. This paradox challenges classical notions of reality, implying that observation determines existence. Applied metaphorically, racism, too, hovers in this unstable state—not fully admitted, not fully denied—until a specific event, accusation, or crisis forces its recognition. Like the cat’s wavefunction, the discourse around racism collapses only when the question of its presence becomes unavoidable.

4

A wavefunction is a mathematical function that describes the possible states of a quantum system. It doesn't tell us exactly where a particle is or what it's doing, but instead gives the probabilities of where it might be and how it might behave.

Before you measure a quantum particle, it exists in a superposition of all possible states at once. The wavefunction encodes this uncertainty. When you measure the particle—say, by looking at its position—the wavefunction "collapses," meaning the particle is forced into one definite state, and all other possibilities disappear.

Think of it like a weather forecast: if you know there's a 60% chance of rain and a 40% chance of sun, the wavefunction describes those possibilities. But the moment you step outside and see the actual weather, the uncertainty vanishes—the "wavefunction collapses" into a single reality.

5

In quantum mechanics, the wavefunction ∣Ψ_racism⟩ represents a suspended reality—one where racism is both present and absent, unresolved until observed. The symbols are precise: ∣ ⟩ (the 'ket' notation) marks it as a quantum state, and Ψ (psi) traditionally represents a wavefunction, a mathematical expression of potentiality. Until an event forces measurement—an interaction, a confrontation, a structural test—the racism remains in superposition, neither fully materialized nor entirely dismissed. The notation is rigorous, but the reality it describes remains elusive, shifting, difficult to pin down.

6

The American Breeders’ Association (ABA), founded in 1903, played a pivotal role in the early American eugenics movement, promoting racial hierarchy, forced sterilization, and immigration restrictions under the guise of scientific heredity. Initially focused on livestock breeding, the ABA extended its mission to human populations, forming a Eugenics Committee in 1906 led by Charles Davenport. The organization’s members, including Harry H. Laughlin, advanced theories of racial "degeneration," arguing that poverty, criminality, and intelligence were hereditary. Their research directly influenced the Immigration Act of 1924, which imposed racial quotas favoring Northern Europeans, and helped justify forced sterilization laws, disproportionately targeting Black, Indigenous, immigrant, and disabled communities. The ABA’s work laid the groundwork for segregationist policies and anti-miscegenation laws, reinforcing the racial caste system under the pretense of genetic science.

For an in-depth analysis of the ABA’s role in eugenics and racial policy, see:
Kevles, Daniel J. In the Name of Eugenics: Genetics and the Uses of Human Heredity. Cambridge, MA: Harvard University Press, 1985.

7

Emmeline Pankhurst (1858–1928) was a British suffragette and leader of the Women’s Social and Political Union (WSPU), known for militant activism that secured voting rights for some women in the UK. While celebrated for her feminist struggle, Pankhurst also embodied imperialist and racist tendencies, aligning white women’s rights with British colonial dominance. She and the WSPU largely ignored the disenfranchisement of working-class women and women of color, focusing instead on securing the vote for privileged British women. Pankhurst was an outspoken supporter of the British Empire, defending colonial rule as a civilizing force and opposing Irish and Indian self-determination. During World War I, she shifted from suffrage to nationalism, advocating for women's participation in war efforts while reinforcing racial hierarchies under imperial feminism. Her activism highlights the entanglement of early feminist movements with white supremacy and colonial power structures.

For further analysis of Pankhurst’s imperialist feminism, see:
Burton, Antoinette. Burdens of History: British Feminists, Indian Women, and Imperial Culture, 1865–1915. Chapel Hill: University of North Carolina Press, 1994.

8

Hillary Clinton (b. 1947) has been a dominant figure in American politics, serving as First Lady, U.S. Senator, Secretary of State, and presidential candidate. While often framed as a feminist icon, her political career reflects a pattern of imperialist and racialized policies under the banner of liberal progressivism. As Secretary of State, Clinton played a key role in U.S. military interventions, including the 2011 NATO-led intervention in Libya, which destabilized the region and led to a resurgence of slavery and armed militias. She infamously remarked, "We came, we saw, he died," when discussing Muammar Gaddafi’s extrajudicial killing. Domestically, her 1996 support for the Crime Bill and “superpredator” rhetoric reinforced mass incarceration policies that disproportionately targeted Black communities. Her championing of neoliberal economic policies further exacerbated racial and class inequalities while consolidating corporate power. Clinton exemplifies the liberal feminist imperialist model, where women’s empowerment rhetoric is often deployed to justify U.S. military aggression and economic policies that harm marginalized populations globally.

For a critical analysis of Clinton’s role in imperial feminism, see:
Enloe, Cynthia. Globalization and Militarism: Feminists Make the Link. Lanham, MD: Rowman & Littlefield, 2016.

9

Madeleine Albright (1937–2022), the first female U.S. Secretary of State (1997–2001), was a key architect of U.S. imperial policy in the post-Cold War era, advancing militarism under the guise of democracy promotion and liberal feminism. A staunch advocate of American exceptionalism, she was instrumental in expanding NATO, supporting sanctions on Iraq, and defending U.S. military interventions. Her most infamous moment came in a 1996 interview on 60 Minutes when she was asked about the deaths of half a million Iraqi children due to U.S. sanctions. Her response—“We think the price is worth it”—exemplified the racist devaluation of non-Western lives in U.S. foreign policy. Albright also played a role in justifying the 1999 NATO bombing of Yugoslavia, presenting intervention as a humanitarian necessity while advancing American geopolitical dominance. Her version of feminism, often framed as breaking barriers for women, coexisted comfortably with policies that reinforced imperial violence, structural racism, and economic devastation in the Global South.

For a critical examination of Albright’s role in feminist imperialism, see:
Kapur, Ratna. Gender, Alterity, and Human Rights: Freedom in a Fishbowl. Cheltenham: Edward Elgar Publishing, 2018.

10

In Lacanian terms, whiteness functions as the subject supposed to be neutral—it sees itself as transparent, unmarked, the universal standard. But the presence of the racialized subject disrupts this neutrality, much like the infant in Lacan’s mirror stage first recognizes itself through the reflection of an Other. The racialized subject becomes that mirror—it forces whiteness to see itself. This is why white subjects react with discomfort, awkwardness, and sometimes hostility when confronted with race. The presence of the racialized Other destabilizes the illusion of neutrality, revealing the hidden asymmetry of racial power. The observer effect, then, is not just about measuring racism—it is about the white subject’s struggle to maintain its unmarked position.

This phenomenon is not exclusive to the West. In China, Han hegemony operates in a similar way, defining itself as the unmarked, neutral standard of Chinese identity. However, the key distinction lies in the historical-materialist origins of Han racial formation versus the construction of whiteness in the West. Unlike whiteness, which was fabricated as a global ruling class through colonial capitalism, plantation economies, and the legal codification of racial purity, Han identity was historically tied to imperial administration, cultural assimilation, and Confucian bureaucratic order rather than biological essentialism.

Much of modern Chinese racial discourse—particularly its anxieties around race, identity, and foreign influence—has been shaped by the importation of whiteness as a category via Western imperialism. The colonial encounter introduced race as a structuring logic that was previously absent in the same form, embedding new racialized hierarchies into Chinese modernity. The consequence is that while Han supremacy exists as a hegemonic force, it does not operate through the same mechanisms of racial capitalism, globalized eugenics, or the strict biopolitical policing of whiteness and non-whiteness seen in Euro-American contexts.

This divergence produces a different psychic structure. The Han hegemon, when confronted with ethnic minorities or foreign racial Others, does not necessarily experience the same crisis of racial innocence as the Western white subject. The anxiety is not over being racist per se, but over national unity, cultural sovereignty, and the perceived erosion of Han-centrism. Whereas the Western liberal subject reacts to racial confrontation with disavowal and hyper-moralistic innocence, the Chinese subject may respond with historical nationalist justifications or civilizational paternalism—forms of racial positioning that emerge from different historical-material conditions.

Thus, while Lacan’s mirror stage describes the destabilization of racial neutrality in both cases, the reflection is not identical. In the West, whiteness anxiously clings to the fantasy of its own universalism and innocence, denying racial difference while structurally enforcing it. In China, Han hegemony is less about denial than about reinscribing cultural hierarchy within a historical continuity of imperial administration. In both cases, the presence of the racialized Other remains disruptive, forcing a reckoning with unspoken power structures—but the psychic economy that mediates this encounter is profoundly shaped by historical materialism.

11

The Persons of Colorlessness refers to those who experience themselves as unraced, as the default against which racialized others appear. This term draws from Frantz Fanon’s insights in Black Skin, White Masks, where whiteness is not simply a racial identity but a condition of transparency—an ability to move through the world without having one’s race foregrounded. To be a Person of Colorlessness is to inhabit spaces as unmarked, unchallenged, unproblematized, perceiving oneself as “just a person” while others are framed through racial difference. When a racialized person enters, this illusion fractures. The slight hesitations, forced smiles, exaggerated friendliness, or sudden silences are the micro-adjustments of a dominant racial subjectivity confronting its own ruptured neutrality. The room was never neutral—whiteness was merely unspoken, structuring space invisibly until its supposed universality was interrupted.

For further discussion on whiteness as an invisible racial position, see:
Dyer, Richard. White: Essays on Race and Culture. London: Routledge, 1997.

12

Columbia University has historical ties to U.S. military and intelligence agencies, extensive investments in defense contractors like Lockheed Martin, and active involvement in real estate-driven displacement in Harlem. Diversity rhetoric within such institutions is not an opposition to these forces but a modulation of them—expanding representation within existing structures of power without fundamentally altering their function.

On Columbia’s role in U.S. militarism and policing, see:
Simpson, Audra. The Empire of Property: Settler Colonialism and the Making of the Modern University. Durham: Duke University Press, forthcoming.

On elite universities and racial capitalism, see:
Stein, Sarah. The Corporate University and the Politics of Dispossession. Minneapolis: University of Minnesota Press, 2021.

13

The German left, which condemns racism while passing laws that criminalize Palestinian solidarity and uphold deportation policies against migrants, does not fail to see its hypocrisy—it depends on it. The well-meaning white academics and journalists who write papers on anti-racism but default to gatekeeping when confronted with voices that challenge their authority do not undermine their own work; their authority was always contingent on keeping certain voices in their place. This reflects a broader dynamic within European liberalism, where anti-racist discourse coexists with racial governance, and moral self-congratulation compensates for material complicity in racialized exclusion. The German left, particularly its Antideutsch and pro-Israel factions, has strategically weaponized anti-antisemitism to delegitimize Palestinian resistance and suppress critique of Western imperial alignments. Meanwhile, the country’s racialized deportation regime, tightened even under nominally progressive coalitions, continues the logic of exclusion inherent to the postwar racial state—where Germany’s moral redemption is performed through selective memory, emphasizing historical guilt for the Holocaust while erasing its colonial past and ongoing border violence. The contradiction is not an accident but a structural necessity: the maintenance of moral authority through selective solidarities, ensuring that racial critique remains within the bounds of European self-image rather than posing a fundamental challenge to its power structures.

On the instrumentalization of anti-antisemitism in German leftist discourse, see:
Weiss, Hajo. Antisemitismus als Waffe: Wie Antideutsche und rechte Israelis den Nahostkonflikt instrumentalisieren. Berlin: Bertz + Fischer, 2021.

On Germany’s racial deportation policies and postwar racial state, see:
El-Tayeb, Fatima. European Others: Queering Ethnicity in Postnational Europe. Minneapolis: University of Minnesota Press, 2011.

14

Bell’s theorem, formulated by physicist John Bell in 1964, fundamentally challenged classical notions of reality by proving that no theory based on local hidden variables can fully explain quantum entanglement. In classical physics, it was assumed that objects have definite properties independent of observation and that influences cannot travel faster than light—this is known as local realism. Bell demonstrated that if quantum mechanics is correct, then entangled particles must exhibit correlations that cannot be explained by any local hidden variable theory. His inequality—the Bell inequality—provides a testable way to determine whether nature obeys local realism or quantum entanglement.

Experiments by Alain Aspect (1980s) and later tests by Anton Zeilinger and others confirmed Bell’s predictions: quantum entanglement violates local realism. This means that two entangled particles remain correlated instantaneously across vast distances, suggesting a reality that is fundamentally nonlocal and relational rather than deterministic. Philosophically, Bell’s theorem disrupts conventional ideas of causality and separability, implying that measurement itself plays an active role in defining the state of reality.

For a detailed exploration of Bell’s theorem and its implications, see:
Mermin, N. David. Boojums All the Way Through: Communicating Science in a Prosaic Age. Cambridge: Cambridge University Press, 1990.

15

The Heisenberg Uncertainty Principle, formulated by Werner Heisenberg in 1927, is a foundational concept in quantum mechanics stating that it is impossible to simultaneously measure both the exact position and exact momentum of a particle with absolute precision. The more precisely one property is known, the more uncertain the other becomes. Mathematically, it is expressed as:

\(\Delta x \cdot \Delta p \geq \frac{\hbar}{2} \)

where Δx is the uncertainty in position, Δp is the uncertainty in momentum, and ℏ is Planck’s reduced constant.

This principle is not a limitation of measurement technology but an intrinsic feature of reality at the quantum level. It reflects the wave-particle duality of quantum objects—particles behave as both waves and discrete entities, and their properties exist in probabilistic superpositions until measured. The Uncertainty Principle disrupts classical determinism, implying that at the fundamental level, nature is governed by probabilities rather than fixed laws.

For further reading on the implications of Heisenberg’s Uncertainty Principle, see:
Heisenberg, Werner. Physics and Philosophy: The Revolution in Modern Science. New York: Harper & Row, 1958.

16

The Heart Sutra is an act of linguistic deconstruction, where negation (no eye, no ear, no nose, no tongue...) systematically dismantles the entire Buddhist cosmology—not to erase it, but to expose its contingency. It is not simply about negation but about undoing the very structures of knowledge and perception, dissolving not just dualistic thinking but the very reliance on concepts as stable reference points.

Xuanzang’s 7th-century Chinese translation, which became the canonical version, introduces an additional layer: his rendering of śūnyatā (emptiness) as 空 (kōng) not only signals voidness but also opens an ontological tension with Daoist and Confucian thought, where 空 often implied spaciousness or potentiality rather than simple negation. The Heart Sutra, then, does not just articulate Buddhist doctrine—it performs the breakdown of conceptual solidity, including the very structures Buddhism had built for itself.

For a deeper exploration of the Heart Sutra’s linguistic and philosophical destabilization, see:
Kopf, Gereon. Beyond Personal Identity: Dōgen, Nishida, and a Phenomenology of No-Self. London: Routledge, 2001.

17

玄奘 (Xuánzàng) (c. 602–664) was a Chinese Buddhist monk, translator, and scholar whose pilgrimage to India and subsequent translations profoundly shaped East Asian Buddhism. Unlike the mythical depiction of him in Journey to the West, where he is the naive monk accompanied by the mischievous Monkey King, the historical Xuánzàng was a fiercely intelligent and determined figure who undertook an epic 17-year journey to India, defying Tang dynasty travel bans to study Buddhist texts firsthand. His mission was driven by dissatisfaction with the fragmented and inconsistent Buddhist scriptures available in China at the time.

Traveling through Central Asia, he reached Nalanda, the great Buddhist university, where he studied under Śīlabhadra, mastering Yogācāra philosophy. When he returned to China in 645 CE, he brought back 657 Sanskrit texts and dedicated the rest of his life to translating them into Chinese. His most influential work was his translation of the Heart Sutra, which became the standard version in East Asia. Unlike previous translations that emphasized metaphysical negation, Xuánzàng’s rendering of emptiness (空/kōng) carried connotations of potentiality and transformation, subtly reshaping how emptiness was understood in Chinese Buddhist thought.

Xuánzàng was also a systematizer—his interpretation of Yogācāra philosophy in the Chéng Wéishì Lùn (成唯識論, Treatise on the Establishment of Consciousness-Only) laid the foundation for the Fǎxiàng (法相) school in China, Japan, and Korea. His journey and translations were not just about Buddhism; they also preserved vital records of 7th-century Central Asia and India, making his Great Tang Records on the Western Regions (大唐西域記) an invaluable historical source.

For a critical study of Xuánzàng’s impact on Buddhist thought and translation, see:
Deeg, Max. Miscellanae Nepalicae: Early Chinese Reports on Nepal—The Foundation Legend of Nepal in its Trans-Himalayan Context. Lumbini: Lumbini International Research Institute, 2016.

18

Wave-particle duality is a fundamental principle in quantum mechanics stating that elementary particles, such as electrons and photons, exhibit both wave-like and particle-like behavior, depending on how they are measured. This challenges classical physics, which assumes that objects must be either waves or particles, but not both.

The concept emerged from experiments like Young’s double-slit experiment (1801), where light passing through two slits produced an interference pattern, proving its wave nature. However, when individual photons were sent through the slits one at a time, they still formed an interference pattern—suggesting that each photon acted as a wave of probabilities until measured. Later, experiments confirmed that electrons, atoms, and even molecules exhibit this duality, meaning that all matter has both particle and wave characteristics.

For a deeper exploration of wave-particle duality and its implications, see:
Feynman, Richard. QED: The Strange Theory of Light and Matter. Princeton: Princeton University Press, 1985.

19

The Quantum Zeno Effect is a paradoxical phenomenon in quantum mechanics where frequent observation prevents a quantum system from evolving. Named after Zeno of Elea, whose paradoxes questioned motion and change, the quantum version suggests that continuous measurement can effectively “freeze” a system in its initial state, preventing its expected transformation.

The effect arises from the wavefunction collapse in quantum mechanics. In an unmeasured system, a quantum state evolves according to the Schrödinger equation. However, when a measurement is made, the wavefunction collapses into a definite state. If a system is measured repeatedly in very short intervals, it keeps collapsing back to its initial state, preventing it from evolving naturally—a process sometimes described as “watched pot never boils” physics.

One of the first theoretical formulations of this effect came from Misra and Sudarshan (1977), who demonstrated that a system undergoing frequent projections (measurements) would asymptotically remain in its original state. This has been experimentally verified in systems like trapped ions and ultracold atoms, where rapid observations suppress expected quantum transitions.

Philosophically, the Quantum Zeno Effect disrupts classical notions of time and change, implying that observation itself is not passive but an active force that shapes reality. It also has practical implications in quantum computing, where it can be used for error correction and stabilizing quantum states.

For further exploration of the Quantum Zeno Effect and its implications, see:
Home, Dipankar, and Andrew Whitaker. Einstein’s Struggles with Quantum Theory: A Reappraisal. New York: Springer, 2007.

20

The Hamiltonian operator is the mathematical tool used in quantum mechanics to represent the total energy of a system. It includes both kinetic energy (movement) and potential energy (position-related forces like gravity or electric charge). The Hamiltonian is essential because it tells us how a quantum system behaves over time.

If you know the Hamiltonian of a system—like an atom, a molecule, or even a particle trapped in a box—you can predict what will happen to it. The Schrödinger equation, which is the key equation of quantum mechanics, uses the Hamiltonian to determine the possible energy levels of the system and how the system evolves over time.

If a system doesn’t change over time (like an electron in a stable orbit), the time-independent Schrödinger equation tells us the allowed energy levels. If the system does change (like an electron absorbing light and jumping to a higher energy level), the time-dependent Schrödinger equation describes how it moves from one state to another.

In classical physics, energy is just a number that tells us how much work can be done. In quantum mechanics, however, the Hamiltonian is an operator, meaning it acts on wavefunctions (the mathematical descriptions of quantum states) and gives us real, measurable outcomes.

A good analogy is a music equalizer: the Hamiltonian is like the control panel that determines how sound frequencies behave, just as in quantum mechanics, it controls how energy behaves.

For a deeper but accessible introduction, see:
Griffiths, David J. Introduction to Quantum Mechanics. Cambridge: Cambridge University Press, 2018.

21

Ghassan Kanafani (1936–1972) was a Palestinian writer, journalist, and revolutionary whose literary and political work remains central to anti-colonial thought and Palestinian resistance. A leading figure in the Popular Front for the Liberation of Palestine (PFLP), Kanafani used fiction as a weapon against Israeli settler-colonialism, portraying the dispossession, exile, and struggle of Palestinians with piercing clarity. His works, including Men in the Sun (1962) and Returning to Haifa (1969), are not merely narratives of loss but strategic interventions—revealing how displacement fractures identity while simultaneously forging new modes of resistance.

Kanafani’s writing rejects sentimental victimhood, focusing instead on the dialectics of oppression and agency. Men in the Sun, for instance, does not just depict Palestinian suffering but critiques passivity, showing how systemic violence operates through both external force and internalized despair. His fusion of literature and revolutionary politics made him a threat to Israeli intelligence; in 1972, he was assassinated by Mossad in a car bombing in Beirut. His death exemplifies how Palestinian intellectuals are not merely commentators but active participants in the struggle—targeted precisely because their words refuse erasure.

For a critical analysis of Kanafani’s literary and political impact, see:
Harlow, Barbara. Resistance Literature. London: Routledge, 1987.

22

In physics, half-life refers to the time required for half of a given substance to decay, whether in radioactive materials, unstable particles, or even concepts in statistical modeling. It does not imply complete disappearance but rather a gradual process in which the original quantity diminishes while still persisting in some form.

Applied metaphorically to racism, the question of half-life challenges the assumption that racism fades linearly over time. If racism were to function like a radioactive isotope, one might expect its intensity to decrease over generations. But what if instead of true decay, racism merely reorganizes itself, changing form but maintaining its structural presence? In this sense, racism does not simply erode; it mutates, finding new mechanisms of survival—through policy, economy, culture, or silence.

This framework also interrogates the observer effect: in quantum physics and social systems alike, observation itself changes the dynamics of a phenomenon. If racism, like certain quantum behaviors, exists in a state of superposition, its form depends on whether and how it is measured. What happens when we stop looking? Does racism decay like an unstable isotope, or does it spread in the absence of scrutiny—like an unchecked chain reaction, diffusing into new structures until it is measured again?

For further exploration of half-life as a metaphor for social and political decay, see:
Mbembe, Achille. Necropolitics. Durham: Duke University Press, 2019.

23

Quantum haze refers to the inherent uncertainty and indeterminacy in quantum systems, where particles exist in a blurred, probabilistic state rather than having fixed, classical properties. This concept arises from wavefunction superposition and quantum decoherence, meaning that before measurement, a particle's position, momentum, or even state remains undefined, spread across multiple possibilities.

The term is often used metaphorically to describe the ambiguity and unpredictability in quantum mechanics, where particles behave neither strictly as waves nor as particles but as a cloud of probabilities. This "haze" lifts only when measurement occurs, forcing a specific outcome—collapsing the wavefunction into a definite state.

In social and philosophical contexts, quantum haze can be used to describe systems of power, racism, or ideology that remain structurally present yet difficult to pinpoint. Just as particles do not exist in singular, well-defined states until observed, certain forms of oppression or bias persist diffusely, evading clear detection until a specific event forces them into visibility.

For a deeper discussion of quantum indeterminacy and its broader implications, see:
Rovelli, Carlo. Helgoland: Making Sense of the Quantum Revolution. New York: Riverhead Books, 2021.

24

Werner Heisenberg (1901–1976) was a German physicist and one of the key architects of quantum mechanics, best known for formulating the Heisenberg Uncertainty Principle in 1927. This principle states that one cannot simultaneously determine both the exact position and momentum of a particle, challenging classical determinism and introducing fundamental limits to measurement in physics.

Heisenberg was instrumental in the development of matrix mechanics, one of the first complete formulations of quantum mechanics, which differed from Schrödinger’s wave mechanics but was later proven to be mathematically equivalent. He played a major role in establishing quantum field theory and the interpretation of quantum states as probabilistic rather than deterministic.

Beyond physics, Heisenberg’s legacy is politically complex. During World War II, he was involved in Nazi Germany’s nuclear weapons project, leading some to speculate whether he deliberately slowed progress or simply failed to achieve a bomb. Post-war, he became a central figure in reconstructing German science, emphasizing moral responsibility in nuclear research.

For a nuanced examination of Heisenberg’s contributions and controversies, see:
Cassidy, David C. Beyond Uncertainty: Heisenberg, Quantum Physics, and the Bomb. New York: Bellevue Literary Press, 2009.

25

Niels Bohr (1885–1962) was a Danish physicist and one of the founding figures of quantum mechanics, best known for developing the Bohr model of the atom and introducing the concept of complementarity in quantum theory. His work laid the foundation for understanding atomic structure and the behavior of electrons, revolutionizing 20th-century physics.

In 1913, Bohr proposed that electrons orbit the nucleus in quantized energy levels, explaining atomic emission spectra and introducing quantum jumps—the idea that electrons transition between energy levels without passing through intermediate states. This model was later refined by quantum mechanics but remains a key conceptual breakthrough.

Beyond atomic theory, Bohr was a major philosophical voice in quantum mechanics. His principle of complementarity (1928) argued that wave and particle descriptions of quantum objects are not contradictions but mutually necessary perspectives—a fundamental aspect of the Copenhagen interpretation, which asserts that quantum states do not have definite properties until measured.

During World War II, Bohr, who had Jewish ancestry, fled Nazi-occupied Denmark and later worked on the Manhattan Project, though he advocated for peaceful applications of nuclear energy and post-war arms control. His influence extended beyond physics into debates on science, philosophy, and ethics.

For a deeper exploration of Bohr’s life and ideas, see:
Pais, Abraham. Niels Bohr’s Times: In Physics, Philosophy, and Polity. Oxford: Clarendon Press, 1991.

26

Erwin Schrödinger (1887–1961) was an Austrian physicist best known for developing wave mechanics and the Schrödinger equation, one of the foundational equations of quantum mechanics. His work introduced a wave-based interpretation of quantum states, describing how the probability distribution of a particle evolves over time. Schrödinger is also famous for the Schrödinger’s cat paradox, a thought experiment designed to critique the Copenhagen interpretation of quantum mechanics. The scenario describes a cat in a box that is both alive and dead until observed, illustrating the unsettling implications of quantum superposition and measurement.

Philosophically, Schrödinger was skeptical of purely probabilistic interpretations of quantum mechanics, leaning toward a deterministic and holistic view of reality. Later in life, he engaged deeply with Eastern philosophy, particularly Vedanta, exploring the connections between quantum physics and consciousness.

For an in-depth study of Schrödinger’s contributions, see:
Moore, Walter J. Schrödinger: Life and Thought. Cambridge: Cambridge University Press, 1989.

Discussion about this episode