Your new experience awaits. Try the new design now and help us make it even better

ORIGINAL RESEARCH article

Front. Quantum Sci. Technol., 21 July 2025

Sec. Basic Science for Quantum Technologies

Volume 4 - 2025 | https://doi.org/10.3389/frqst.2025.1603372

This article is part of the Research Topic100 Years of Quantum Science and TechnologyView all 6 articles

Dynamics of the ideal quantum measurement of a spin-1 with a Curie–Weiss magnet

  • Institute for Theoretical Physics, University of Amsterdam, Amsterdam, Netherlands

Quantum measurement is a dynamical process involving an apparatus coupled to a test system. The ideal measurement of the z-component of a spin-12 (sz=±12) has been modeled by the Curie–Weiss model for quantum measurement. Recently, the model was generalized to higher spins, and its thermodynamics were solved. Here, the dynamics are considered. To this end, the dynamics for the spin-12 case are cast in general notation. The dynamics of the measurement of the z-component of a spin-1 (sz=0,±1) are solved in detail and evaluated numerically. The energy costs of the measurement, which are macroscopic, are evaluated. The generalization to higher spin is straightforward.

1 Introduction

This year, we celebrate the centennial of the formulation of quantum theory; see Capellmann (2017) for the prehistory. After the “Zur Quantummechanik” by Born and Jordan (1925), the Dreimänner Arbeit by Born et al. (1926) on the matrix mechanics was soon followed by Schrödinger's (1926) formulation of wave mechanics, inspired by the insights of De Broglie (1924). The predictive power of the theory was expressed by Born's (1926) rule. For a compilation of historical contributions, see Wheeler and Zurek (2014).

The interpretation of quantum mechanics has been discussed throughout the century since then. The Copenhagen interpretation— with the Born rule and the collapse postulate—emerged as the most reasonable. Many attempts to deepen understanding begin with these postulates. However, they are merely shortcuts for what happens in a laboratory. With our collaborators Armen Allahverdyan and Roger Balian, we have taken the viewpoint of starting from the uninterpreted quantum formalism and applied it to the dynamics of an idealized measurement. The elements of this approach that have already been solved do not need to be interpreted; interpretation is needed to put the results in a proper, global context. As discussed below, this effort has led to a specified version of the statistical interpretation of quantum mechanics, popularized by Ballentine (1970).

The present study deals with the dynamics of an ideal quantum measurement. It is based on the Curie–Weiss model for measuring the z-component of a spin 12, introduced by Allahverdyan et al. (2003a). After reviewing various models for quantum measurement, it was considered in great detail by Allahverdyan et al. (2013). The apparatus consists of a mean-field type magnet having N1 spins 12 coupled to a harmonic oscillator bath. The magnet starts in a metastable, long-lived, paramagnetic state, which is separated by free energy barriers from the stable states with upward or downward magnetization. It is in a “ready” state for use in a measurement.

When employed as an apparatus, the magnetization acts as a pointer for the outcome. The coupling to the tested spin causes a quick transition to one of the stable states, thereby registering the measurement. For this to succeed, the coupling must be large enough to overcome the free energy barrier. While the final state of the magnet is described by thermodynamics, much detail is contained in the dynamical evolution toward this state.

In an ideal measurement, the Born rule appears due to the non-disturbance of the measured operator. It provides probabilities for the pointer, that is, for the final magnetization to be upward or downward. The state of the microscopic spin is correlated with it and inferred from the pointer indication.

Understanding the dynamics also provides a natural route toward the interpretation of quantum mechanics. Indeed, when assuming the quantum formalism, the task is to work out its predictions, and only then to interpret the results. This leads to viewing the wave function, or, more generally, the density matrix, as a state of best knowledge and the “collapse of the wave function” or “disappearance of cat states” as an update of knowledge after the selection of the runs with identical outcomes, compatible with the quantum formalism. Notably, quantum theory is not a theory of Nature based on an ontology; rather, it is an abstract construct to explain its probabilistic features.

The “measurement problem,” that is, describing the individual experiments that occur in a laboratory, is, in our view, still the most outstanding challenge of modern science. Many attempts have been made to solve it by making adaptations or small alterations to quantum mechanics or by interpreting it differently. We hold the opinion that this entire enterprise is in vain; one should start completely from scratch to “derive quantum mechanics,” that is to say, establish the origin of quantum behavior in Nature1.

Various formalisms of quantum mechanics were reviewed by David (2015). The insight that quantum mechanics is only meaningful in a laboratory context, stressed in particular by Bohr, is central to the approaches of Auffeves and Grangier (2016) and Auffeves and Grangier (2020), it leads to new insights regarding the Heisenberg cut between quantum and classical (Van Den Bossche and Grangier, 2023). One century of interpretation of the Born rule, including the modern one, was overviewed by Neumaier (2025).

1.1 The Curie–Weiss model for quantum measurement

A macroscopic material consists of atoms, which are quantum particles. The starting point for their dynamics lies in quantum statistical mechanics. For a measurement, the apparatus must be macroscopic and have a macroscopic pointer so that the outcome of the measurement can be read off or processed automatically. Hereto, an operator formalism is required, with dynamics set by the Liouville–von Neumann equation, the generalization of the Schrödinger equation to mixed states.

Progress on solvable models for quantum measurement has been made in recent decades when we, together with A. Allahverdyan and R. Balian introduced and solved the so-called Curie–Weiss model for quantum measurement (Allahverdyan A. E. et al., 2003) in our “ABN” collaboration. Here, the classical Curie–Weiss model of a magnet is taken in its quantum version and applied to the measurement of a quantum spin 12. Various further aspects were presented in Allahverdyan A. E. et al. (2003), Allahverdyan et al. (2005a), Allahverdyan et al. (2005b), Allahverdyan et al. (2007), and Allahverdyan et al. (2006). They were reviewed and greatly expanded in Allahverdyan et al. (2013). Lecture notes were presented by Nieuwenhuizen et al. (2014). A straightforward interpretation for a class of these measurement models was provided by Allahverdyan et al. (2017); it is a specified version of the statistical interpretation made popular by Ballentine (1970).

Simultaneous measurement of two noncommuting quantum variables was worked out (Perarnau-Llobet and Nieuwenhuizen, 2017a), as well as an application to Einstein-Podolsky-Rosen type of measurements (Perarnau-Llobet and Nieuwenhuizen, 2017b). A numerical test on a simplified version of the Curie–Weiss model reproduced nearly all of its properties (Donker et al., 2018).

Our ensuing insights, which are suitable for teachers of quantum theory (at the high school, bachelor’s, or master’s levels), are presented in Allahverdyan et al. (2024) and summarized in a feature article (Allahverdyan et al., 2025).

1.2 Higher-spin Curie–Weiss models

The mentioned Curie–Weiss model was recently generalized by us to measure a spin l>12 (Nieuwenhuizen, 2022). This study will be termed “Models” henceforth. For spin l, the state of the magnet is described by 2l order parameters. To assure an unbiased measurement, the Hamiltonian of the apparatus and the interaction Hamiltonian with the tested system have Z2l+1 symmetry. The statics were solved for spin-1, 32, 2, and 52.

Here, the dynamics are worked out for spin-1, laying the groundwork for higher-spin dynamics. In the spin 12 Curie–Weiss model, it was found that Schrödinger cat terms disappear through two mechanisms: dephasing of the magnet, possibly followed by decoherence due to the thermal bath. Similar behavior is now investigated for spin-1.

The setup of the article is as follows. In Section 2, we recall the formulation of the Curie–Weiss model for general spin-l and discuss aspects of its physical implementation for spin 12 and spin-1. In Section 3, we revisit the spin-12 case and cast its dynamics in a general form. In Section 4, we analyze the dynamics of the spin-1 situation. We close with a summary in Section 5.

2 Higher-spin Curie–Weiss Hamiltonian models

We start by recalling some properties of higher-spin models that we introduced in “Models” (Nieuwenhuizen, 2022). The statics were considered there; here, we define and study the dynamics, recalling parts of the spin 12 case. We often refer to the review by Allahverdyan et al. (2013) to be termed “Opus.”

In the following, we denote quantum operators by a hat, specifically ŝ and ŝz for the measured spin and σ̂(i) and σ̂z(i) for the spins of the apparatus. For simplicity of notation, we follow Models and denote the eigenvalues without a hat, notably those of ŝz by s and the ones of σ̂z(i) by σi. Sums over i lead to the operators m̂k and their scalar values mk for k=1,2,,2l. Switching between these operators and their eigenvalues is straightforward.

The strategy is to measure the z-component of a quantum spin-l with (l=12,1,32,). The eigenvalues s of the operator ŝz lie in the spectrum2

sspecl=l,l+1,,l1,l.(2.1)

The measurement will be performed by employing an apparatus with N1 vector spins-l having operators σ̂(i), i=1,,N. They have components σ̂a(i) (a=x,y,z), with eigenvalues σa(i)specl. These operators are mutually coupled in the Hamiltonian of M. For each i=1,N, and for each σ̂a(i), a=x,y,z, they are also coupled to a thermal harmonic oscillator bath; for the case l=12, this was worked out by Allahverdyan et al. (2003a), Allahverdyan et al. (2003b), and Allahverdyan et al. (2013). The generalization of such a bath for arbitrary spin-l is straightforward and will be applied to the spin-1 model.

2.1 Spin–spin Hamiltonian of the magnet

A quantum measurement is often assumed to be “instantaneous.” In our idealized modeling, it will take a finite time, but the tested spin will not evolve in the meantime. In other words, the spin itself is “sitting still” and waiting to be measured. Neither should it evolve during the “fast” measurement. This is realized when its Hamiltonian ĤS commutes with ŝz; we consider the simplest case: ĤS=0.

In order to have an unbiased apparatus, the Hamiltonian of the magnet should have degenerate minima and maximal symmetry. To construct such a functional, we consider, in the eigenvalue presentation, the form

C2=ν2i,j=1Ncos2πσiσj2l+1,ν1N,(2.2)

which is maximal in ferromagnetic states σi=σ1 (i=2,,N). In general, these interactions do not seem realistic, but here, the cosine rule allows expressing this as spin–spin interactions,

C2=col2+sil2,(2.3)

which is bilinear in the single-spin sums

col=1Ni=1Ncos2πσi2l+1, sil=1Ni=1Nsin2πσi2l+1.(2.4)

The discrete values of the spin projections allow expressing these terms in the 2l spin moments,

mk=1Ni=1Nσik,k=1,,2l,(2.5)

while m01. For l=12, the values s=±12 imply

cosπs=0,sinπs=2s.(2.6)

Applying this for sσi and summing over i yields

co12=0,si12=2m1,m1=1Ni=1Nσi.(2.7)

In the case l=1, one has s=0,±1. The rule

cos2πs3=132s2,sin2πs3=32s,(2.8)

leads to sσi and summing over i leads to

co1=132m2,si1=32m1,(2.9)

Here, m2 ranges from 0 to 1 with steps of ν1/N, while m1 ranges from m2 to m2 with steps of 2ν. At finite N, one can label the discrete m1,2 as

m1=2n1n2ν,m2=n2ν,0n2N,0n1n2.(2.10)

The results for s=32, 2, and 52 are given in Models.

Let out of the N spins σi, a number Nσ=iδσi,σ take the value σspecl and let xσ=Nσ/N be their fraction. The sum rule σNσ=N implies m0σxσ=1. The moments read

mk=σ=llxσσk,k=1,,2l,(2.11)

Inversion of these relations determines the xσ as linear combinations of the mk. For l=12, one has

m1=12x1212x12,x±12=12±m1.(2.12)

For spin-1 (l=1), one has

m1=x1+x1,m2=x1+x1.(2.13)

With x1+x0+x1=1, their inversion reads

x0=1m2,x±1=m2±m12.(2.14)

In a quantum approach, one goes to operators and sets sŝz, σiσ̂z(i), and mkm̂k. For the Hamiltonian ĤM=NĤ, we follow Allahverdyan et al. (2003a) and Allahverdyan et al. (2003b) and adopt the spin–spin and four–spin interactions:

ĤM=NĤ,Ĥ=12J2Ĉ214J4Ĉ22.(2.15)

Multispin interaction terms like 16J6Ĉ2318J8Ĉ24 can be added without changing the overall picture.

2.2 The interaction Hamiltonian

The coupling between the tested spin S and the magnet M is chosen similar to Equation 2.2,

ĤSA=NÎ, Î=gNi=1Ncos2πŝzσ̂z(i)2l+1,(2.16)

where g is the coupling constant. It takes the values

Isσi=gNi=1Ncos2πs2l+1cos2πσi2l+1+sin2πs2l+1sin2πσi2l+1,(2.17)

This can be expressed as a linear combination of the moments m1, , m2l. For l=12, one has

Ism1=4gsm1,(2.18)

and for l=1, denoting m=(m1,m2),

Ism=g132s2132m2+34sm1.(2.19)

The total spin Hamiltonian,

Ĥ=ĤM+ĤSA=ĤMNÎ,(2.20)

has Z2l+1 symmetry: on the diagonal basis, a shift ss+s̄ with s̄=1,2,,2l+1 can be accompanied by a shift σiσi+s̄ for all i. This is evident in the cosine expressions and implies a somewhat hidden invariance in the formulation in terms of the moments mk, as discussed in Models.

2.3 Coupling to a harmonic oscillator bath

For a general spin l, the magnet–bath coupling is taken as the spin–boson coupling of Opus Equation 3.10,

ĤMBγi=1Na=x,y,zσ̂a(i)B̂a(i),(2.21)

with γ1, where the bath operators read

B̂a(i)=kckb̂k,a(i)+b̂k,an,(2.22)

for each i,a, there is a large set of oscillators labeled by k, having a common coupling parameter ck. These bosons have the Hamiltonian

ĤB=i=1Na=x,y,zkωkb̂k,a(i)b̂k,a(i),(2.23)

with the ωk also identical for all n,a. The autocorrelation function of B defines a bath kernel K, which is identical for all i,a,

trBR̂B0B̂a(i)tB̂bjt=δi,jδa,bKtt,B̂a(i)teiĤBtB̂a(i)eiĤBt.(2.24)

Writing ck=c(ωk), this leads to

Kt=kcωkeiωkteβωk1+eiωkt1eβωk12π+dωeiωtK̃ω.(2.25)

The kernel K̃(ω) can be read off and expressed in the spectral density ρc(ω)=kc(ωk)δ(ωωk),

K̃ω=2π|ω|ρcωωeβω1.(2.26)

We adopt an Ohmic spectrum with a Debye cutoff,

K̃ω=eω/Γ4ωeω/T1,(2.27)

where T=1/β is the temperature of the phonon bath, and Γ the typical cutoff frequency. In Opus, we also consider a Lorentzian (power law) cutoff, for which the statics allows analytic results.

With the couplings in Equations 2.21, 2.22, and 2.23 independent of a, ĤMB is statistically invariant under Z2l+1. Combined with the invariance of ĤM and ĤSA, this ensures an unbiased measurement.

2.4 Evolution of the density matrix

The evolution of the density matrix of the total system is given by the Liouville–von Neumann equation. On the eigenbasis of ŝz, its elements R̂ss̄ evolve independently as given in Equation 4.8 of Opus; this involves the apparatus spins and the bath. The procedure of Opus for spin l=12 appears to hold for general spin-l operators.

Let us consider the time evolution of R̂ss̄ as given in Equation 4.8 of Opus (we now denote is, js̄), where the action of the harmonic oscillator bath has been expressed in the bath kernel K(t) and which involves commutators of R̂ss̄ with the spin operators σ̂a(i), a=x,y,z; i=1,,N.

Formally, the initial state (Equation 5.4) is a constant function of the σ̂z(i). In addition, Ṙss̄(ti) is a function of them, so it is consistent to assume that, at all t, R̂ss̄ only depends on the σ̂z(i). As a result, the a=z terms of Equation 4.8 in Opus have vanishing commutators for any spin l. Left with the x,y commutators, we define (using the index n rather than i to label the σ̂x,y)

σ̂±(n)=σ̂x(n)±iσ̂y(n).(2.28)

Because ax,yσ̂a(n)Ôσ̂a(n)=12α=±1σ̂α(n)Ôσ̂α(n) for any operator Ô, Equation 4.8 in Opus takes the form

dR̂ss̄tdt=iĤsR̂ss̄t+iR̂ss̄tĤs̄+γ2α,β=±1n=1N0tduKβuĈss̄,βα,nu,(2.29)

where

Ĉss̄,+α,nu=eiuĤsσ̂α(n)eiuĤsR̂ss̄t,σ̂α(n),Ĉss̄,α,nu=σ̂α(n),R̂ss̄teiuĤs̄σ̂α(n)eiuĤs̄,(2.30)

are commutators involving the Hamiltonian of M coupled to S in state s, without the bath, viz.

Ĥs=ĤM+ĤSAs=NHm̂1+NIsm̂1.(2.31)

The action of the bath is expressed in the kernel K(±u), with the smallness of γ allowing truncation at its first order. Equations 2.29, 2.30 are valid for general spin l=12,1,32,.

Most importantly, the R̂ss̃ are decoupled in the separate s,s̄ sectors, a property of ideal measurement but absent in general. Examples of these non-idealities are a spin S having nontrivial dynamics during the measurement and a biased measurement, in which the Hamiltonian of the magnet and/or the bath depends on the state of S.

2.5 Physical implementation of the model

The spin-12 Curie–Weiss model for quantum measurement (Allahverdyan A. et al., 2003) was initially conceived as a tool to understand the dominant physical aspects of idealized quantum measurements. It has served this purpose well. Let us look here at possible realizations of the model.

Curie–Weiss models are mean-field types of spin models. Their distance-independent couplings apply to a small magnetic grain. The grain need not be very large. From studies of spin glasses and cluster glasses, it is known that “fat spins,” clusters of hundreds or thousands of coherent spins, are easily detectable (Mydosh, 1993).

The Ising nature of the couplings refers to fairly anisotropic spin–spin interactions. For spin 12, Equation 2.15 expresses the pair and quartet couplings between the z-components of the spins. Multispin interactions are a natural result of the overlap of electronic orbits; here, they are approximated as not decaying with the distance between the spins in the grain. How reasonable this approximation is must be considered in each separate application. The main feature of our modeling, a first-order phase transition in the magnet, suggests that it represents a large class of short-range systems. This is underlined by the model’s support of the Copenhagen postulates of collapse and Born probabilities.

These features also hold for the spin-1 Curie–Weiss model. However, on top of this, Equation 2.8 produces the combination Σ̂iσ̂z(i)22/3, which takes the values 1/3 for the “out-of-plane” cases σi=±1 and 2/3 for the “in-plane” case σi=0. Separate-spin terms of the form iDσ̂z(i)2 are well known, stemming from crystal fields. For the apparatus, the co12 term of Equation 2.3 relates to the interaction ijΣ̂iΣ̂j between the Σ̂i, so it involves both the aforementioned D-term and also the terms σ̂z(i)2σ̂z(j)2. How to implement these crystal-field-type spin–spin interactions in practice is an open question.

Concerning numerical implementations, Donker et al. (2018)’s approximation of the Curie–Weiss model can be generalized to higher spin.

3 The spin 12 case revisited

3.1 Elements of the statics

We set the stage by considering the spin-12 situation, the original Curie–Weiss model for quantum measurement in slightly adapted notation3. The spin operators are σ̂x,y,z, with σ̂z=diag(12,12). It holds that [σ̂a,σ̂b]=iεabcσ̂c and σ̂x2+σ̂y2+σ̂z2=34σ̂0 with σ̂0=diag(1,1).

The magnet has N these spins σ̂x,y,z(i), i=1,2,,N. They have magnetization operator

M̂1=Nm̂1,m̂1=1Ni=1Nσ̂z(i),(3.1)

taking eigenvalues 12m112. In the paramagnetic state, m1=0. The Hamiltonian is taken as pair and quartet interactions,

ĤM=NĤ,Ĥ=2J2m̂124J4m̂14.(3.2)

With x̂σ=N̂σ/N, it holds that

m̂1=12x̂1/212x̂1/2,x̂±1/2=12αm̂1.(3.3)

The spins have eigenvalues σi=±12, so that m̂1 has eigenvalues m1=νiσ1 ranging from 12 to 12 with steps of ν.

3.2 The interaction Hamiltonian

To use the magnet coupled to its bath as an apparatus for a quantum measurement, a system–apparatus (SA) coupling is needed. According to Equation 2.18, it is chosen as a spin–spin coupling,

ĤSA=4gi=1Nŝσ̂z(i)=4gNŝm̂1,(3.4)

and takes the values HSAs(m1)=4gsNm1. The full Hamiltonian of S + A in the sector s thus reads

Ĥs=2J2Nm̂124J4Nm̂144gsNm̂1.(3.5)

The eigenvalues of ŝz are s=±12 and those of σ̂z(i) are σi=±12, so that m̂1 has the eigenvalues νi=1Nσi. The degeneracy of a state with magnetization m1 is

GN=N!N12!N12!=N!Nx12!Nx12!,(3.6)

and entropy SN=logGN. At large N, we get the standard result for the entropy SN=NS with

S=12m12log12m121+2m12log1+2m12.(3.7)

Combining Equation 3.5 and Equation 3.6, the free energy in the s-sector reads

Fsm1=2J2Nm124J4Nm144gsNm1TlogGNm1,(3.8)

which yields, for large N,

FsN=2J2m124J4m144gsm1TSm1.(3.9)

3.3 Dynamics of the spin 12 model

At the initial time ti of the measurement, the state of the tested system, S, here ŝ, a spin-12 operator, is described by its 2×2 density matrix r̂(ti) with elements rss̄(ti) for s,s̄=±12. The magnet M has N1 quantum spins-12 σ̂(i) (i=1,,N). In each s,s̄ sector, S + M lie in the state R̂ss̄(t)=R̂s̄s(t), which is an operator that can be represented by a 2N×2N matrix. At ti, M is assumed to lie in the paramagnetic state wherein the spins are fully disordered and uncorrelated. Multiplying by the respective element of r̂(ti) leads to the elements of the initial density matrix of S + M

R̂ss̄ti=rss̄tiσ̂0(1)2σ̂0(2)2σ̂0(N)2.(4.1)

3.4 Truncation for spin 12

The dynamics of the off-diagonal elements (cat terms) were worked out in Opus. In the relevant short-time domain, the spin–spin couplings are ineffective; therefore, it suffices to study independent spins coupled by the interaction Hamiltonian and the bath. These elements vanish dynamically, truncating the density matrix R̂ to a form diagonal on the eigenbasis of ŝz. There is no reason to repeat that here; for spin-1, this will be worked out in Section 4.1.

3.5 Registration for spin 12

Registration of the measurement is described by the evolution of the diagonal elements of the density matrix of the full system. For the situation of higher spin, it is instructive to reconsider and slightly reformulate the spin 12 situation.

For s̄=s, the Hamiltonian terms drop out of Equation 2.29; hence, the dynamics are a relaxation set by

dR̂sstdt=γ2α,β=±1n=1N0tduKβuĈss,βα,nu.(4.2)

For l=12, the spin operators σ̂x,y,z anticommute; hence, for any function f of the σ̂z(i), it holds that

σ̂α(n)fσ̂z(i)=f̂1δi,nσ̂z(i)σ̂α(n)f(n)σ̂z(i)σ̂α(n).(4.3)

This brings the σ̂α(n) and σ̂α(n) next to each other, which allows to eliminate them using the sum α=±1σ̂α(n)σ̂α(n)=σ̂0(n). With only functions of the σ̂z(i) (i=1,,N) remaining, we can go to their diagonal bases to work with scalar functions of their eigenvalues σi=±12 (see also Opus, Section 4.4). This expresses Equation 2.30 as

Css̄,+(n)uα=±1Css̄,+αnu=eiuHseiuHs(n)Rss̄(n)teiuHs(n)eiuHsRss̄t,Css̄,(n)uα=±1Css̄,αnu=Rss̄(n)teiuHs̄(n)eiuHs̄Rss̄teiuHs̄eiuHs̄(n),(4.4)

where for any function f({σi}), f(n) has the sign of σn reversed,

f(n)σi=f1δi,nσi.(4.5)

We employed the obvious rules (fg)(n)=f(n)g(n) and [f(g)](n)=f(g(n)). The terms in Equation 4.4, being scalars, yield the relation Css̄,(n)(u)=Css̄,+(n)(u), which allows combining the integrals of Equations 2.29 and 2.30 as a single one from u=t to t. Because γ1, the typical scale of t, the registration time 1/γT is much larger than the bath equilibration time 1/T. Hence, we may now take the integral over the entire real axis to arrive at the Fourier-transformed kernel K̄(ω) at specific frequencies.

The next step is to reduce the 2N×2N matrix problem to a problem of N+1 variables by considering Rss({σi})=Rss(m1) to be functions of the order parameter m1=νσi. This is formally true at ti and valid for Ṙss̄(ti); hence, it remains valid over time. Denoting Ps(m1) as the probability that Rss({σi})/rss(ti) involves m1=νiσi, it picks up the degeneracy number GN in Equation 3.6 of realizations {σi} with the same m1,

Psm1=GNm1Rssm1rssti.(4.6)

To obtain the evolution of Ṗs, we multiply Equation 4.2 by GN(m1)/rss(ti). At given m1, one has m1(n)=m12νσn, so we can split the terms with σn=12 (and 12) and perform the sum over n. The fraction of terms that flips an up spin σn=12 is x12(m1), which multiplies P(m1ν); flipping a down spin σn=12 happens with probability x12(m1), which multiplies P(m1+ν). Due to Equation 4.6, these Ps involve the ratios

GNm1GNm1ν=x12m1νx12m1,GNm1GNm1+ν=x12m1+νx12m1,(4.7)

which has the effect of eliminating the x±12(m1). Introducing the operators E± and Δ±=E±1 by

E±fm1=fm1±ν,Δ±fm1=fm1±νfm1,(4.8)

the evolution of Ps gets condensed as

Ṗsm1=γN2α=±1Δαx12αm1K̃Ωsαm1Psm1.(4.9)

where

Ωs±m1=ΔHs=Hsm1νHsm1.(4.10)

This is now a problem for N+1 functions P(m1;t) subject to the normalization m1P(m1;t)=1.

In Figure 1, the distribution of the magnetization m1 is depicted at various times. In Figure 2, this evolution is represented in a 3d plot.

Figure 1
www.frontiersin.org

Figure 1. Evolution of the magnetization distribution Ps(m1;t) for s=+12 at times 0,1,,8 in units of 1/γT. The paramagnetic state at t=0 is peaked around m1=0; the coupling between S and A moves the peak toward m1=+12. In doing so, it first broadens and later narrows significantly.

Figure 2
www.frontiersin.org

Figure 2. The case of Figure 1 plotted in 3d at intervals Δt=0.2/γT.

3.6 H-theorem and relaxation to equilibrium

The dynamical entropy of the distribution Ps(m1;t)=GN(m1)Rss({σi};t)/rss(ti) is defined as

Sst=TrR̂sstrsstilogR̂sstrssti=m1Psm1;tlogPsm1;tGNm1.(4.11)

As in Opus, we introduce a dynamical free energy:

Fdynst=UstTSst=m1Psm1;tHsm1+TlogPsm1;tGNm1,(4.12)

which adds the PslogPs term to the average of the free energy functional FN(m1)=Hs(m1)TSN(m1). With β=1/T, Equation 5.36 yields.

Ḟdyns=Tm1Ṗsm1logPsm1eβHsm1GNm1
=γNT2α=±1m1Δαx12αK̃ΩsαPslogPseβHsGN.

For general functions f1,2(m1) and α=±1, partial summation yields

m1Δαf1f2=m1f1Δαf2=m1Eαf1Δαf2=m1Eαf1Δαf2.(4.13)

provided that the boundary terms f1,2(m±ν) at m±ν=±(1+ν) vanish. As discussed, this holds for Ps but also for the logarithm in Equation 4.13 because we may insert a factor (1δm1,mνδm1,mν) that makes this explicit. For α=+1, we now use the last expression, and for α=1, we use the second one, which yields, also using Equation 4.10 and the property K̃(ω)=K̃(ω)eβω satisfied in (Equations 2.26, 2.27), the result

Ḟdyns=γNTm1K̃Δ+Hs×eΔ+βHsE+x12E+Psx12PsΔ+logPseβHsGN.(4.14)

The various x–factors are such that a term GN(m1)x12 can be factored out to yield

Ḟdyns=γNTm1β=±1GNx12K̃Δ+Hs×eΔ+βHsE+PsGNPsGNΔ+logPseβHsGN.(4.15)

With Δ+Hs=E+HsHs, GN=exp(SN) and Fs(m1)=Hs(m1)TSN(m1), this can finally be expressed as

Ḟdynst=γNTm1x12K̃Δ+HseβFs×Δ+PseβFsΔ+logPseβFs.(4.16)

The last factors have the form (xx)log(x/x), which is nonnegative, so that Fdyns is a decreasing function of time. Dynamic equilibrium occurs when these factors vanish, which happens when the magnet has reached the thermodynamic equilibrium set by the Gibbs state Ps=eβFs/Zs and R̂ss=eβĤs/Zs, with Zs=m1exp(βFs)=m1GN(m1)exp(βHs)=Trexp(βĤs), as usual. The dynamical free energy (Equation 4.12) indeed ends up at the thermodynamic one,

Fdyns=TlogZs=Fsg.(4.17)

This constitutes an example of the apparatus going dynamically to its lowest thermodynamic state and the pointer state indicating the measurement outcome s=±12. The temporal evolution from Fdyn(0) to Fs(g) is depicted in Figure 3.

Figure 3
www.frontiersin.org

Figure 3. Evolution of the dynamical free energy Fdyns(t), identical in both sectors s=±12, after coupling the apparatus to a spin-12 at time t=0. Its approach to the Gibbs state with Fs(g) (bottom line), exponential in t, expresses the registration of the measurement.

3.7 Decoupling the apparatus

Near the end of the measurement, at a suitable time tdc, the apparatus is decoupled from the system, by setting g=0; in doing so, an energy Udc=m1Ps(m1;tdc)HSA(m1) must be supplied to the magnet, which will then relax further its nearby minimum of the g=0 situation, to provide a stable pointer indication with a macroscopic order parameter M1=Nm1 that can be read off.

4 Dynamics of the spin-1 model

We now focus on the spin-1 case, in which the tested system, S, is ŝ, a spin-1 operator with ŝz having eigenvalues sz=1,0,1. Our magnet M has N1 quantum spins-1 σ̂(i) (i=1,,N). According to Equation 2.5, one now deals with two order parameters,

m̂1=1Ni=1Nσ̂i,m̂2=1Ni=1Nσ̂i2.(5.1)

While m̂1 is the usual magnetization in the z-direction, m̂2 is a spin-anisotropy order parameter that discriminates the sectors with eigenvalues σi=±1 from the sector with eigenvalues σi=0.

The quantity Ĉ2, the operator-form of Equation 2.2, is our starting point for a permutation-invariant Hamiltonian that ensures unbiased measurement. Expanding the cosine, employing Equation 2.8 for each spin σ̂i, and summing over i yields a polynomial in the moments m̂1,2,

Ĉ2=132m̂22+34m̂12.(5.2)

For the Hamiltonian, we take as in Equation 3.2

ĤN=NĤ,Ĥ=12J2Ĉ214J4Ĉ22.(5.3)

It can be understood as containing the single-spin term m̂2, the pair couplings m̂12=1/N2ijσ̂iσ̂j and m̂22=1/N2ijσ̂i2σ̂j2, the triplet couplings m̂12m̂2 and the quartet couplings m̂14, m̂12m̂22, and m̂24. However, note its different conception in Section 2.5.

At the initial time ti of the measurement, its state is described by its 3×3 density matrix r̂(ti) with elements rss̄(ti) for s,s̄=1,0,1.

In each s,s̄ sector, M lies in its state R̂ss̄(t)=R̂s̄s(t), which is an operator that can be represented by a 3N×3N matrix. This exponential problem gets transformed into a polynomial one, a step that is exact for the considered mean-field-type Hamiltonian.

At ti, M is assumed to lie in a paramagnetic state, wherein the spins are fully disordered and uncorrelated. For each spin, its state is thus σ̂0(i)/3 where σ̂0(i)=diag(1,1,1). Multiplying by the respective element of r̂(ti) leads to the elements of the initial density matrix of S + M in the s,s̄=0,±1 sector,

R̂ss̄ti=rss̄tiσ̂0(1)3σ̂0(2)3σ̂0(N)3.(5.4)

For general angular momentum, the commutation relations [L̂a,L̂b]=iεabcL̂c and L̂x2+L̂y2+L̂z2=l(ł+1)Î carry over to general spin

σ̂a,σ̂b=iεabcσ̂c,σ̂x2+σ̂y2+σ̂z2=ll+1σ̂0,(5.5)

While we considered l=12 in Section 3, we now focus on l=1.

We proceed as for spin 12. The a=z commutator in Equation 2.29 does again not contribute. We introduce σ̂α=σ̂x+iασ̂y for α=±1. From Equation 5.5, it follows for general l that

σ̂ασ̂α=ll+1σ̂0+ασ̂zσ̂z2σ̂ασ̂ασσ=l+1ασl+ασδσσ.(5.6)

In the present case l=1, this has nontrivial values

σ̂ασ̂ασσ=2δσ,α+2δσ,0,σ=0,±1,(5.7)

with Equation 5.6 implying that the σ=α term indeed drops out. The SO(3) generators

σ̂x=12010101010,σ̂y=120i0i0i0i0,σ̂z=100000001,(5.8)

allow verifying these relations. Each of the σ̂(i) (i=1,,N) has such a presentation. In Equation 2.30, the interchange of the σ̂α(i) with the σ̂z(n) will be needed. For in, they commute, while for i=n,

σ̂αnσ̂znk=σ̂z(n)ασ̂0(n)kσ̂αn,σ̂z(n)kσ̂αn=σ̂αnσ̂z(n)+ασ̂0(n)k.(5.9)

Valid for k=1, induction yields this for higher k. For functions of the {σ̂z(i)}, (i=1,,N), that can be expanded in a power series, it follows that

σ̂αnfσ̂z(i)fn,ασ̂z(i)σ̂αn,fn,ασ̂z(i)=fσ̂z(i)δi,nασ̂0(n).(5.10)

Now the σ̂± can be eliminated using Equation 5.6, which leaves functions of only the σ̂z(i), with the shifts in their arguments arising as the cost for this. As before, we can assume that R̂ss̄(t)=Rss̄({σ̂z(i)},t), where Rss̄({σi},t) is a scalar function of the eigenvalues σi=0,±1 of the σ̂z(i). Valid at ti, this holds for (dR̂ss̃/dt)(ti), so it remains valid in time. Hence, it is possible to go from the matrix equations to scalar equations. With the equality in Equation 5.6 applied for spin n, we end up with the scalar expressions

Css̄,+(n,α)u=δσn,α+δσn,0eiuHseiuHs(n,α)Rss̄(n,α)tδσn,α+δσn,0eiuHs(n,α)eiuHsRss̄t,Css̄,(n,α)u=Css̄,+(n,α)u,(5.11)

where for any function Rss̄ expandable in powers of the σi=0,±1 (i=1,,N), it holds that

Rss̄(n,α)=Rss̄σiσi+αδi,n,(5.12)

Now that all terms are scalar functions of the σ̂z(i), it is seen that Css̄,(n,α)(u)=Css̄,+(n,α)(u;HsHs̄). We no longer need to track the operator structure and can work with scalar functions of the eigenvalues.

4.1 Off-diagonal sector: truncation of Schrödinger cat terms

In the spin 12 Curie–Weiss model, it was found that the Schrödinger cat terms disappear by two mechanisms: dephasing of the magnet, possibly followed by decoherence due to the thermal bath. Similar behavior is now investigated for spin-1.

4.1.1 Initial regime: dephasing

Truncation of the density matrix (disappearance of the cat states) is a collective effect that takes place within an initial time window, in which the magnet stays in the paramagnetic phase, so that the mutual spin couplings J2,4 and the coupling to the bath can be neglected. The spins of M act individually by their coupling to the tested spin S and do not get correlated yet. In the sector where the eigenvalue of the operator ŝz is s, the Hamiltonian of the magnet is

ĤSA=nĤSAsn,ĤSAsn=g132s2σ̂0(n)32σ̂z(n)2+34sσ̂z(n).(5.13)

At a given s, this is a trace-free diagonal matrix with elements 12g (twice) and g,

HSAsnσσ̃=g2δσ,σ̃13δσ,s,δσ,s=13+23s2132σ2+12sσ,(5.14)

for (s,σ,σ̃=0,±1). In this approximation, the 3N×3N density matrix of the magnet in each sector ss̄ maintains the product structure (Equation 5.4) of uncorrelated spins at t=ti,

R̂ss̄t=rss̄tiρ̂ss̄(1)tρ̂ss̄2tρ̂ss̄(N)t,(5.15)

where, setting ti=0, for each n,

ρ̂ss̄(n)t=eitĤSAs,nσ̂0(n)3eitĤSAs̄,n=ρ̂s̄s(n)t,ρss̄(n)tσσ̃=13δσ,σ̃exp32igtδσ,sδσ,s̄.(5.16)

Diagonal elements s=s̄ thus essentially do not evolve in this short-time window. The off-diagonal ones imply for ss̄

rss̄t=TrMR̂ss̄t=rss̄013+23cos32gtN.(5.17)

For small t, this decays as rss̄(0)exp(t2/τdph2) with the dephasing time τdph=2/g3N, very short for large N. The undesired recurrences at tn=4πn/3g, where the cosine equals 1 again, can be suppressed by assuming that the ggn=ḡ+δgn values in Equation 5.16 have a small spread δgn (see Opus, Section 6.1.1). If the thermal oscillator bath has proper parameters, it will cause decoherence, as seen next.

4.1.2 Second step: decoherence

To include the bath in Equation 5.16, we now make the generalized Ansatz:

ρ̂ss̄(n)tσσ̃=δσ,σ̃13expBσt×expitHSAs,nσ+itHSAs̄,nσ.(5.18)

In the commutators (Equation 5.11), Hs now reduces to the HSAs,n of Equation 5.14, and the terms are identical for all n. We can neglect Bγ in the exponents of Equation 2.29 and find, putting αα in the minus terms,

Ḃσ=γ2αKt>ΔασHs+Kt<ΔασHs̄Kt>ΔασHs+Kt<ΔασHs̄ess̄ασt,(5.19)

with

Kt>ω=0tduKueiωu,Kt<ω=t0duKueiωu.(5.20)

Here, Kt>(ω)=Kt>*(ω) because the kernel K̃(ω) is real valued; see the example in Equation 2.27, and

ΔασHs=Hsσ+αHsσ=3g2132s21+2ασ12sα,(5.21)

with a similar expression for ΔασHs̄, and finally

ess̄ασt=expitΔασHsΔασHs̄.(5.22)

For s̄=s, one has ess̄ασ(t)=1. For t1/2πT, one gets, using K̃(ω)=eβωK̃(ω),

Ḃσ=γ2αK̃ΔασHsK̃ΔασHs=γ2αΔασHse|ΔασHs|/ΓγN,(5.23)

because HsN, ΔασHsN0, and αΔασHs1/N. Therefore, for s=s̄, this confirms that hardly any dynamics take place in this time window. In the next subsection, we show that they occur on a longer time scale τreg=1/γT.

For off-diagonal elements s̄s, it is seen that ess̄ασ(t) has terms e±3igt/2 and e±3igt, so that

ess̄ασt=j=2,1,1,2cje3ijgt/2,0tduess̄ασ;u=j=2,1,1,2cje3ijgt/213ijg/2.(5.24)

The exponentials are equal to unity, making Ess̄α=1, at the times tn=4πn/3g, n=1,2,, encountered below Equation 5.17, when appearing in the dephasing process, and thus also as times where Ḃσ(t)=0. To suppress recurrences like in the dephasing, we again set in each n-term ggn=ḡ+δgn with small Gaussian distributed δgn. For times well exceeding the coherence time 1/2πT of the bath, the Kt> and Kt< reach their finite limits, so that we have

0tdtKt>ωEss̄ασ;t=0tdtKt>ωK>ωEss̄ασ;t+K>ω0tduEss̄ασu,(5.25)

The first part is small, and the second is given in Equation 5.24. After canceling out its exponents by the δgn, an imaginary part remains. Hence, for t1/2πT, the Ess̄ασ terms can be neglected in RB. We keep

RBσtRḂσ×t,RḂσγ2αK̃ΔασHs+K̃ΔασHs̄,(5.26)

which is positive, so that |exp(NBσ)|=exp(NRBσ) with NRBσγNgt leads for large enough values of N to a decoherence of the off-diagonal elements rss̄(t) of the density matrix at the characteristic decoherence time tdec=1/γgN and γNgτregNg/T.

Decoherence is a combined effect of the N apparatus spins; despite it, the individual elements of R̂ss̄ hardly decay in this time window, behaving as exp(γgt)=exp(t/Ntdec)1.

4.2 Registration dynamics for spin-1

In Section 3, a difference equation was derived for the distribution of the magnetization of the magnet for any number N of spins-12. Our aim here is to derive an analogous equation for the spin-1 case.

In the paramagnet, one has the form Rss({σi})=rss(ti)/3N. Let Ps(m) with m=(m1,m2) be the probability for a state of the magnet M characterized by the moments m1,2. It gathers the value Rss(m)/rss(ti) for all sequences {σi} compatible with m1,2, the number of which is the degeneracy factor GN=expSN,

Psm;t=GNmRssm;trssti,GNm=N!N1!N0!N1!,Nσ=xσN,(5.27)

with the x±1=12(m2±m1) and x0=1m2 from Equation 2.14. The normalizations are

σ(1)=11σ(N)=11Rssσ(i);t=rssti,m2=01m1=m2m2Psm1,m2;t=1.(5.28)

Due to the relations described by Equations 2.13 and 2.14 between the spin moments m0,±1 and the spin fractions x0,±1, the shifts in m1,2 induce the shifts Nσ=Nσ+δNσ and xσ=xσ+νδNσ, with

δN±1=1±α2+ασn, δN0=12ασn,(5.29)

which are integers, as they should be. The degeneracies for σn=α,0,α lead to the respective factors

GNGN=N1!N0!N1!N1!N0!N1!=x0+νxαδσn,α+xα+νx0δσn,0+x1+νx1+ν+xα+2νx0x0νx02νδσn,α,(5.30)

where Nσ=Nxσ is used. The complicated last term is fortunately not needed, while the denominators of the first two will factor out.

Going to the functions Ps of the moments m1,2, we proceed as for the spin 12 situation. The C± terms of Equation 5.11 can again be combined and performing the u-integrals in Equation 4.2 leads for t1/T to the kernel K̃(ω) at the frequencies

Ωαβm=Hsm1αν,m2βνHsm,(5.31)

for α,β=±1. Multiplying Equation 4.2 by GN and summing over α, there results an evolution equation for the distribution Ps at each discrete value of m1,2,

Ṗsm1,m2;t=γNα=±1x0+νK̃Ωs,α+Psαm1,m2;t+xα+νK̃Ωs,αPsα+m1,m2;txαK̃Ωsα++x0K̃ΩsαPsm1,m2;t.(5.32)

Let us condense notation and introduce the shift operators Eαβ and Δαβ=Eαβ1 by their action

Eαβfm=fm1+αν,m2+βν,Δαβfm=fm1+αν,m2+βνfm.(5.33)

on any f(m). They have the properties

EαβΔαβ=Δαβ,Ωsαβ=ΔαβHs,EαβΩsαβ=ΔαβHs=Ωs,αβ.Eαβxα=xα+1+β2ν,Eαβx0=x0βν.(5.34)

Hence, Equation 5.32 can be expressed as

Ṗsm1,m2;t=γNα=±1Δα+xαK̃Ωsα+Ps+Δαx0K̃ΩsαPs,(5.35)

which has a remarkable analogy to Equation 4.9 and Equation 4.16 of Opus for the spin-12 case. By denoting xα+=xα above and xα=x0, this is condensed further,

Ṗsm1,m2;t=γNα,β=±1ΔαβxαβK̃ΩsαβPs.(5.36)

4.3 H-theorem and relaxation to equilibrium

We now exhibit a H theorem that assures the relaxation of the magnet towards its Gibbs equilibrium state and, thus, a successful measurement. The dynamical entropy of the distribution Ps(m;t)=GN(m)Rss({σi})/rss(ti) is defined as

Sst=TrR̂sstrsstilogR̂sstrssti=mPsm;tlogPsm;tGNm.(5.37)

Following Opus and Equation 4.12 above, we consider the dynamical free energy

Fdynst=UstTSst=mPsm;tHsm+TlogPsm;tGNm.(5.38)

It appears to depend on s. The simultaneous change ss, m1m1 implies that Fdyn1(t)=Fdyn1(t) at all t, as happened for s=±12 in the spin 12 case, but the Fdyn±1(t) differ from Fdyn0(t), except in the thermal situations at t=0 and t.

With β=1/T, not to be confused with the index β=±1, Equation 5.36 yields

Ḟdyns=TmṖsmlogPsmeβHsmGNm=γNTα,β=±1mΔαβxαβK̃ΩsαβPslogPseβHsGN.(5.39)

For general functions f1,2(m) with vanishing boundary terms, partial summation yields

mΔαβf1f2=mf1Δαβf2=mEαβf1Δαβf2=mEαβf1Δαβf2.(5.40)

For α=+1, we use the last expression, and for α=1, we use the second one, while taking ββ, and also using Equation 5.34 and the property K̃(ω)=K̃(ω)eβω satisfied generally in Equation 2.26, which yields the result

Ḟdyns=γNTmβ=±1K̃Δ+βHs×eΔ+ββHsE+βx+βE+βPsx1βPsΔ+βlogPseβHsGN.(5.41)

The various parts are such that a term GN(m)x1β can be factored out, to express this as

Ḟdyns=γNTmβ=±1GNx1βK̃Δ+βHs×eΔ+ββHsE+βPsGNPsGNΔ+βlogPseβHsGN.(5.42)

With Δ+βHs=E+βHsHs, GN=exp(SN), and Fs(m)=Hs(m)TSN(m), this is equal to

Ḟdynst=γNTmβ=±1x1βK̃Δ+βHseβFs×Δ+βPseβFsΔ+βlogPseβFs.(5.43)

The last factors have the form (xx)log(x/x), which is nonnegative, implying that Fdyns is a decreasing function of time. Dynamic equilibrium occurs when these factors vanish, which happens when the magnet has reached thermodynamic equilibrium, that is, the Gibbs state Ps=eβFs/Zs and R̂ss=eβĤs/Zs, with Zs=mexp(βFs)=mGN(m)exp(βHs)=Trexp(βĤs), as usual. The dynamical free energy (Equation 5.38) then ends up at the thermodynamic free energy,

Fdyns=TlogZs,(5.44)

which actually does not depend on s due to the invariance map of the static state, reflecting that the measurement is unbiased. This constitutes an explicit example of the apparatus going dynamically to its lowest thermodynamic state, the pointer state registering the measurement outcome.

Although the statics are identical for s=0,±1, this does not hold for the dynamics. While it is similar for s=±1 (to change the sign of s=±1, also change the sign of m1), this deviates from the s=0 dynamics. For s=0, all Ωαβ(m) are finite, but for s=±1, there are cases where Ωαβ(m) vanishes, which leads to a slower dynamics; see Figure 5.

4.4 Numerical analysis

The initial spin-1 Hamiltonian leads to a 3N×3N matrix problem, which is numerically hard. For the considered mean-field-type model, the formulation in terms of the order parameters m1,2 is exact; it lowers the dimensionality considerably. The variable m2=(1/N)i=1Nσi2 can take N+1 values between 0 and 1. The value of M2=Nm2 indicates that NM2 of the σi take the value 0, while the other M2 of the σi are ±1. Given this number, m1=(1/N)i=1Nσi can take M2+1 values between m2 and m2. Accounting for conservation of total probability, this leads to N(N+3)/2 dynamical variables, a polynomial problem.

(Concerning higher spin: For spin 32, one separates terms with si=±32 from those with si=±12; for spin-2, one selects terms with si=0, ±1, or ±2, etc.)

Equation 5.32 can be solved numerically as a set of linear differential equations. Programming it is straightforward; the vanishing of boundary terms and conservation of the total probability must be verified as a check on the code.

The magnet starts in the paramagnetic initial state

Psm;0=13NGNm33/22πNexpN34m12+32m212.(5.45)

The sum of Ps over m1,2 equals unity and, with the mesh Δm1Δm2=2ν2, so does its integral.

The dynamics (Equation 5.32) can be solved numerically, and the results are presented in upcoming figures. We consider the parameters, with g large enough,

N=100,J2=0,J4=1,g=0.15,T=0.2,Γ=10.(5.46)

We plot in Figures 4A,B snapshots of Ps/(2ν2) at four times, for s=0 and s=1. The case s=1 follows from the case s=1 by setting m1m1.

Figure 4
www.frontiersin.org

Figure 4. Snapshots of the distribution Ps of the magnetization moments m1,2 for registration of the spin-1 measurement. Upper: Ps103 data in the s=0 sector at times t=(0,1,2,3)×2/γT from right to left. Lower: the s=1 sector at t=(0,1,2,3)×5/γT from left to right; it evolves more slowly. The parameters are listed in Equation 5.46.

Figure 5 shows the evolution of the dynamical free energy Fdyns(t).

Figure 5
www.frontiersin.org

Figure 5. The spin-1 dynamical free energy Fdyns of Equation 5.38 relaxes from its t=0 value to its thermodynamic value. Fs(g) of Equation 5.44, thereby registering the measurement. Parameters are as in Figures 4A,B, and time is expressed in units of 1/γT. The relaxation for s=±1 is slower than for s=0 due to the occurrence of zero frequencies. The initial “shoulders” describe the initial broadenings in Figures 4A,B.

4.5 Decoupling of the apparatus

Near the end of the measurement, the interaction between the system and the apparatus is cut off by setting g=0; in doing so, at decoupling time tdc, Equation 5.13 expresses that an amount of energy

Udc=mPsm;tdcHSAm=+gN×mPsm;tdc132s2132m2+34sm1,(5.47)

must be supplied to the magnet, leaving it with the post-decoupling free energy

Fdc=mPsm;tdcHMTlogGN.(5.48)

This post-decoupling state is not an equilibrium state; the magnet will now relax to the nearby minimum of the g=0 case. There follows a relaxation driven by bath, with the magnet evolving under the g=0 Hamiltonian HM(m) to its Gibbs state PG(m)=GNexp[HM(m)/T]/ZG, with free energy FG=mPG(m)[HM(m)TSN(m)].

When the decoupling time tdc is large enough, the magnet M lies in its Gibbs state at coupling g, Ps(tdc)exp[βHs(m)]. Due to the invariance of the g=0 situation, the approach to it is identical for starting in any of the sectors s=0,±1.

To compare with the dynamics that end up in one of the minima, one must restrict the Gibbs state, which has three degenerate minima, to the nearby minimum. This is achieved numerically even at moderate N by discarding exp(βHs) well away from the peak of Ps(tdc), also in ZG. For s=0, it suffices to keep exp(βHs) for m2<13; for s=±1 by doing that for m1s>13.

The change of the state is also seen in m2(t)=mm2Ps(m;t). Let us consider the sector s=0, where m1=0 at all t. Here, the coupling HSA=gN(32m21) has the tendency to suppress m2, so after decoupling, m2 will relax to a larger value. For N, we get from the Gibbs states at g and at g=0, respectively,

m20=3.63104,m2=11.5104.(5.49)

The full-time behavior for N=100 and couplings as in Equation 5.46 is presented in Figure 6, with the finite-N values increasing from m2(0)=9.975104 to m2()=12.69104.

Figure 6
www.frontiersin.org

Figure 6. After decoupling the apparatus from the system, the magnet relaxes to its nearby g=0 equilibrium. If this happens at a time tdc where finite-g equilibrium has been reached, this goes identical in the sectors s=0,±1. The dynamical free energy is plotted with parameters as in Figures 4, 5, relaxing from its decoupled value (indicated by the dot) to its g=0 thermodynamic limit F (lower line). Compared to Figure 5, this macroscopic energy cost is a permille effect.

The relaxation in the sectors s=±1 follows immediately from this. The map (Equation 5.50) yields. The maps (4.11) and (4.13) of Models lead to

m1s=±1=±132m2s=0,m2s=±1=112m2s=0.(5.50)

4.6 Energy cost of quantum measurement

The Copenhagen postulates obscure one of the facts of life in a laboratory: a firm cost for the energy needed to keep the setup running. In this work, we consider two intrinsic costs. In the previous subsection, we established the cost of decoupling the apparatus from the system. Here, we consider resetting the magnet for another run. It must be set from its stable state back to its metastable state. Being related to the magnet, both costs are macroscopic.

Our initial state, the paramagnet (pm), has zero magnetic energy and maximal entropy

Fpm=NTlog3,(5.51)

The energy needed to reset the Gibbs state of the magnet to the paramagnetic one is

Ureset=FpmFG=mPGmHMTlogGN/3N.(5.52)

It is evidently macroscopic. The condition that Ureset is positive was identified in Opus and in Models as the condition that the initial paramagnetic state is metastable but not stable.

5 Conclusion

This article dealt with the dynamics of an ideal quantum measurement of the z-component of a spin-1. The statics for this task were worked out recently in our “Models” article (Nieuwenhuizen, 2022); it generalized to any spin l>12 the Curie–Weiss model to measure a spin 12; the latter was considered in great detail in “Opus” (Allahverdyan et al., 2013). Here, we first reformulated the dynamics of the known case for spin 12 and worked out some further properties. The resulting formalism is suitable as a basis for models to measure any higher spin.

The dynamics of measurement in the spin-1 case were analyzed in detail. Off-diagonal elements of the density matrix (“cat states”) were shown to decay very fast (“truncation of the density matrix”) due to dephasing, possibly followed by decoherence.

The evolution of the diagonal elements of the density matrix was expressed as coupled first-order differential equations for the distribution of two magnetization-type-order parameters, m1,2. The approach to a Gibbs equilibrium was certified by demonstrating a H-theorem. The resulting scheme was found to be numerically a polynomial problem. These are easily solved with the present power of laptops for an apparatus consisting of a few hundred spins. The evolution of the probability density was evaluated, and the H-theorem was verified. The macroscopic energy costs for decoupling the apparatus from the spin and for resetting it from its stable state to its metastable state for use in the next run of the measurement were quantified.

For general spin l, this method simplified the numerically hard problem of dimension (2l+1)2N1 by a polynomial problem of order N2l for its 2l-order parameters. For more complicated models of the apparatus, it will likewise pay off to focus on the order parameter of the dynamical phase transition of the pointer that achieves the registration of the measurement. The fact that the phase transition in the magnet is of first order underlines that our mean-field-type models, although of mathematical convenience, are not essential for the fundamental description of quantum measurements.

Data availability statement

The original contributions presented in the study are included in the article/supplementary material; further inquiries can be directed to the corresponding author.

Author contributions

TN: writing – original draft and writing – review and editing.

Funding

The author(s) declare that no financial support was received for the research and/or publication of this article.

Conflict of interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Generative AI statement

The author(s) declare that no Generative AI was used in the creation of this manuscript.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Footnotes

1An analogy is offered by the dark matter problem in cosmology. Abandoning particle dark matter, we view dark “matter” as a form of energy and assume new properties of vacuum energy. This provides a description of black holes with a core rather than a singularity (Nieuwenhuizen, 2023), aspects of dark matter throughout the history and future of the Universe (Nieuwenhuizen, 2024a), and the giant dark matter clouds around isolated galaxies (Nieuwenhuizen, 2024b), explaining the “indefinite flattening” of their rotation curves (Mistele et al., 2024). Remarkably, this approach is a generalization of the classical Lorentz–Poincaré electron—a charged, non-spinning spherical shell filled with vacuum energy (Nieuwenhuizen, 2025).

2To simplify the notation, we replace the standard notation for spins with sl and szs. For an angular momentum L2=l(l+1), the model also applies to the measurement of L̂z with eigenvalues ms. We employ units =k=1.

3For the connection with the parameters in Opus, see ref 1.

References

Allahverdyan, A., Balian, R., and Nieuwenhuizen, T. M. (2003a). The quantum measurement process: an exactly solvable model. Atti della Fondazione Giorgio Ronchi, 719. Available online at: https://arxiv.org/abs/cond-mat/0309188.

Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2003b). Curie-Weiss model of the quantum measurement process. EPL Europhys. Lett. 61, 452–458. doi:10.1209/epl/i2003-00150-y

CrossRef Full Text | Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2005a). The quantum measurement process in an exactly solvable model. AIP Conf. Proc. 750, 26–34. doi:10.1063/1.1874554

CrossRef Full Text | Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2005b). Dynamics of a quantum measurement. Phys. E Low-dimensional Syst. Nanostructures 29, 261–271. doi:10.1016/j.physe.2005.05.023

CrossRef Full Text | Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2006). Phase transitions and quantum measurements. AIP Conf. Proc. 810, 47–58. doi:10.1063/1.2158710

CrossRef Full Text | Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2007). “The quantum measurement process: lessons from an exactly solvable model,” in Beyond the quantum (World Scientific), 53–65.

Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2013). Understanding quantum measurement from the solution of dynamical models. Phys. Rep. 525, 1–166. doi:10.1016/j.physrep.2012.11.001

CrossRef Full Text | Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2017). A sub-ensemble theory of ideal quantum measurement processes. Ann. Phys. 376, 324–352. doi:10.1016/j.aop.2016.11.001

CrossRef Full Text | Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2024). Teaching ideal quantum measurement, from dynamics to interpretation. Comptes Rendus. Phys. 25, 251–287. doi:10.5802/crphys.180

CrossRef Full Text | Google Scholar

Allahverdyan, A. E., Balian, R., and Nieuwenhuizen, T. M. (2025). Interpretation of quantum mechanics from the dynamics of ideal measurement. Europhys. News, appear 56, 23–25. doi:10.1051/epn/2025210

CrossRef Full Text | Google Scholar

Auffèves, A., and Grangier, P. (2016). Contexts, systems and modalities: a new ontology for quantum mechanics. Found. Phys. 46, 121–137. doi:10.1007/s10701-015-9952-z

CrossRef Full Text | Google Scholar

Auffèves, A., and Grangier, P. (2020). Deriving born’s rule from an inference to the best explanation. Found. Phys. 50, 1781–1793. doi:10.1007/s10701-020-00326-8

CrossRef Full Text | Google Scholar

Ballentine, L. E. (1970). The statistical interpretation of quantum mechanics. Rev. Mod. Phys. 42, 358–381. doi:10.1103/revmodphys.42.358

CrossRef Full Text | Google Scholar

Born, M. (1926). Zur Quantenmechanik der Stoßvorgänge. Z. für Phys. 38, 803–827. doi:10.1007/bf01397184

CrossRef Full Text | Google Scholar

Born, M., Heisenberg, W., and Jordan, P. (1926). Zur quantenmechanik. II. Z. für Phys. 35, 557–615. doi:10.1007/bf01379806

CrossRef Full Text | Google Scholar

Born, M., and Jordan, P. (1925). Zur quantenmechanik. Z. für Phys. 34, 858–888. doi:10.1007/bf01328531

CrossRef Full Text | Google Scholar

Capellmann, H. (2017). “The quantum theory of Born, Heisenberg, and Jordan,” in The development of elementary quantum theory (Cham: Springer International Publishing), 23–39. Available online at: https://link.springer.com/content/pdf/10.1007/978-3-319-10539-0.pdf.

CrossRef Full Text | Google Scholar

David, F. (2015). The formalisms of quantum mechanics. Lect. notes Phys. 893, 42–43. doi:10.1007/978-3-319-10539-0

CrossRef Full Text | Google Scholar

De Broglie, L. (1924). Recherches sur la théorie des quanta. Paris: Migration-université en cours d’affectation. PhD thesis.

Google Scholar

Donker, H., De Raedt, H., and Katsnelson, M. (2018). Quantum dynamics of a small symmetry breaking measurement device. Ann. Phys. 396, 137–146. doi:10.1016/j.aop.2018.07.010

CrossRef Full Text | Google Scholar

Mistele, T., McGaugh, S., Lelli, F., Schombert, J., and Li, P. (2024). Indefinitely flat circular velocities and the baryonic tully–Fisher relation from weak lensing. Astrophysical J. Lett. 969, L3. doi:10.3847/2041-8213/ad54b0

CrossRef Full Text | Google Scholar

Mydosh, J. A. (1993). Spin glasses: an experimental introduction. London: CRC Press. doi:10.1201/9781482295191

CrossRef Full Text | Google Scholar

Neumaier, A. (2025). The Born rule–100 years ago and today. arXiv Prepr. arXiv:2502.08545 27, 415. doi:10.3390/e27040415

PubMed Abstract | CrossRef Full Text | Google Scholar

Nieuwenhuizen, T. (2022). Models for quantum measurement of particles with higher spin. Entropy Basel, Switz. 24, 1746. doi:10.3390/e24121746

PubMed Abstract | CrossRef Full Text | Google Scholar

Nieuwenhuizen, T. M. (2023). Exact solutions for black holes with a smooth quantum core. arXiv Prepr. arXiv:2302.14653.

Google Scholar

Nieuwenhuizen, T. M. (2024a). Solution of the dark matter riddle within standard model physics: from black holes, galaxies and clusters to cosmology. Front. Astronomy Space Sci. 11, 1413816. doi:10.3389/fspas.2024.1413816

CrossRef Full Text | Google Scholar

Nieuwenhuizen, T. M. (2024b). Indefinitely flat rotation curves from Mpc sized charged cocoons. Europhys. Lett. 148, 49002. doi:10.1209/0295-5075/ad895f

CrossRef Full Text | Google Scholar

Nieuwenhuizen, T. M. (2025). How the vacuum rescues the Lorentz electron and imposes its Newton and geodesic motion, and the equivalence principle, to appear Eur. J. Phys.

Google Scholar

Nieuwenhuizen, T. M., Perarnau-Llobet, M., and Balian, R. (2014). Lectures on dynamical models for quantum measurements. Int. J. Mod. Phys. B 28, 1430014. doi:10.1142/s021797921430014x

CrossRef Full Text | Google Scholar

Perarnau-Llobet, M., and Nieuwenhuizen, T. M. (2017a). Simultaneous measurement of two noncommuting quantum variables: solution of a dynamical model, Physical Review A 95 (5), 052129.

CrossRef Full Text | Google Scholar

Perarnau-Llobet, M., and Nieuwenhuizen, T. M. (2017b). Dynamics of quantum measurements employing two Curie--Weiss apparatuses, Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 375 (2106), 20160386.

PubMed Abstract | CrossRef Full Text | Google Scholar

Schrödinger, E. (1926). An undulatory theory of the mechanics of atoms and molecules. Phys. Rev. 28, 1049–1070. doi:10.1103/physrev.28.1049

CrossRef Full Text | Google Scholar

Van Den Bossche, M., and Grangier, P. (2023). Contextual unification of classical and quantum physics. Found. Phys. 53, 45. doi:10.1007/s10701-023-00678-x

CrossRef Full Text | Google Scholar

Wheeler, J. A., and Zurek, W. H. (2014). Quantum theory and measurement, 15. Princeton University Press.

Google Scholar

Keywords: ideal quantum measurement, dynamics, Curie–Weiss model, higher spin, exact solution

Citation: Nieuwenhuizen TM (2025) Dynamics of the ideal quantum measurement of a spin-1 with a Curie–Weiss magnet. Front. Quantum Sci. Technol. 4:1603372. doi: 10.3389/frqst.2025.1603372

Received: 31 March 2025; Accepted: 16 May 2025;
Published: 21 July 2025.

Edited by:

Karl Hess, University of Illinois at Urbana-Champaign, United States

Reviewed by:

Marcin Wieśniak, University of Gdansk, Poland
Bryan Sanctuary, McGill University, Canada

Copyright © 2025 Nieuwenhuizen. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Theodorus Maria Nieuwenhuizen, dC5tLm5pZXV3ZW5odWl6ZW5AdXZhLm5s

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.