L–Single-Spin–Single-Mode Dynamics

An Operator's Handbook – Common views

author: U. Warring affiliation: Institute of Physics, University of Freiburg version: 0.2.3 last_updated: 2025-12-19 license: CC BY 4.0 Disclaimer – This handbook is a conceptual and pedagogical synthesis intended to clarify minimal models, regimes, and design principles for single-spin–single-mode quantum dynamics; it does not claim completeness, universality, or direct applicability to all experimental platforms. Any emulation of bath-like or power-law behaviour discussed here is finite, state-dependent, and explicitly bounded by the assumptions and cutoffs stated in the text.

Foreword

How to Cite This Handbook

Stable citation format:

Warring, U. (2025). Single-Spin–Single-Mode Quantum Dynamics: An Operator's Handbook (Version 0.2).

Version-specific: Always cite the version number. Sections may evolve in subsequent versions; backward compatibility is maintained for numbered sections.

For specific results: When citing particular equations or claims, reference the section number (e.g., "Section 4.9, Eq. for IPR").


PART I — INVARIANT CORE

1. Scope, Philosophy, and Minimality

1.1 Scope

Single-spin–single-mode quantum dynamics refers to the physics of a single two-level quantum system (a spin-½ or qubit) interacting with a single quantised harmonic oscillator mode. This minimal composite system—essentially the quantum Rabi model—is a cornerstone of quantum optics and quantum information, underlying phenomena from cavity QED to trapped-ion quantum logic.

1.2 Thesis

Even this ostensibly simple system generates complex quantum dynamics. By tuning system parameters, preparing different initial states, or allowing environment coupling, one can explore diverse regimes—from exactly solvable models to chaotic-like behaviour—and gain insight into larger many-body or open quantum systems.

This handbook develops a unified framework for single-spin–single-mode dynamics, organised into six parts:

Part
Title
Content

I

Invariant Core

Hamiltonians, symmetries, regime boundaries

II

Dynamics from Simplicity

Manifold-resolved dynamics, initial-state complexity

III

From Closed to Open

Single mode as environment, memory effects, dissipation

IV

Analytic Tools

Solution recipes, approximation validity

V

The Trapped-Ion Platform

Hamiltonian engineering, experimental knobs

VI

Horizons

Open questions, what this handbook enables

1.3 The Minimality Theorem

What Makes This Minimal?

The spin–mode duo is minimal in the information-theoretic sense: it is the smallest Hilbert space that admits (i) non-commuting observables on both subsystems, (ii) entanglement, and (iii) a tunable symmetry-breaking term (the coupling).

Adding spatial degrees of freedom or more modes does not introduce new classes of phenomena—only more parameters. This handbook shows that collapse, revival, effective non-Markovianity, and time-dependent energy renormalisation are already present here; they are not emergent from thermodynamic limits.

1.4 The Exportability Principle

The single-spin–single-mode system serves as a reference implementation. Any phenomenon demonstrated here can be exported to larger systems with confidence that the essential physics is understood. Conversely, claims about multi-mode systems should be benchmarked against this minimal case: if the phenomenon requires more than one mode, that is significant; if it does not, the single-mode setting offers cleaner analysis.


2. Canonical Hamiltonians and Symmetries

2.1 The Quantum Rabi Hamiltonian

The generic Hamiltonian for a spin-½ (two-level atom or qubit) interacting with a single mode is:

H=ω02σz+ωaa+g(σ++σ)(a+a)H = \frac{\hbar \omega_0}{2}\sigma_z + \hbar \omega \, a^\dagger a + \hbar g \, (\sigma^+ + \sigma^-)( a + a^\dagger)

where:

  • ω0\omega_0 is the spin's transition frequency

  • ω\omega is the oscillator frequency

  • aa^\dagger, aa are the mode's creation/annihilation operators

  • gg is the coupling strength

  • σz\sigma_z, σ±\sigma^\pm are Pauli operators on the two-level system

This quantum Rabi Hamiltonian is invariant under a Z2\mathbb{Z}_2 parity transformation (flipping the sign of both spin-σx\sigma_x and oscillator quadrature)—a symmetry that yields a conserved parity operator:

Π=σz(1)aa\Pi = \sigma_z (-1)^{a^\dagger a}

2.2 The Jaynes–Cummings Model

In the rotating-wave approximation (valid when gω,ω0g \ll \omega, \omega_0 and ωω0\omega \approx \omega_0), one drops the non-energy-conserving terms σ+a\sigma^+ a^\dagger and σa\sigma^- a, recovering:

HJC=ω02σz+ωaa+g(σ+a+σa)H_{\text{JC}} = \frac{\hbar \omega_0}{2}\sigma_z + \hbar \omega \, a^\dagger a + \hbar g \, (\sigma^+ a + \sigma^- a^\dagger)

The Jaynes–Cummings model conserves the total excitation number:

N=aa+12(1+σz)N = a^\dagger a + \frac{1}{2}(1 + \sigma_z)

This exact invariant simplifies the solution: the Hilbert space decomposes into independent two-dimensional subspaces e,n,g,n+1{|e, n\rangle, |g, n+1\rangle} for each excitation number.

Exactness Disclaimer

The JC model is analytically diagonalisable but physically approximate. Its exactness is a mathematical property of the truncated Hamiltonian, not a claim about nature. The counter-rotating terms are always present; we are operating in a regime where their effect is emulable by a Lamb shift, not a dynamical process.

This is the first example of the handbook's core principle: exactness is a function of the question asked, not the Hamiltonian written down.

2.3 The Operator's Decision Tree (The Navigator)

2.4 Regimes of Interaction

Regime
Condition
Symmetry
Hamiltonian
Approximation

Dispersive

gΔg \ll |\Delta|, gω,ω0g \ll \omega, \omega_0

Eff. U(1)U(1)

HdispH_{\text{disp}}

exact within RWA (no population transfer)

Resonant JC

gωg \ll \omega, Δ0\Delta \approx 0

U(1)U(1)

HJCH_{\text{JC}}

RWA valid

Ultrastrong

0.1<g/ω<10.1 < g/\omega < 1

Z2\mathbb{Z}_2

HRabiH_{\text{Rabi}}

Braak exact

Deep Strong

g/ω>1g/\omega > 1

Z2\mathbb{Z}_2

HRabiH_{\text{Rabi}}

Polaron frame

where Δ=ωω0\Delta = \omega - \omega_0 is the spin-mode detuning.

Dispersive Regime (gΔg \ll |\Delta|): The spin and oscillator exchange virtual excitations, leading to energy shifts (Lamb/AC Stark) but not real energy exchange. The effective Hamiltonian is:

Hdispωaa+2(ω0+g2Δ(2aa+1))σzH_{\text{disp}} \approx \hbar \omega \, a^\dagger a + \frac{\hbar}{2}\left(\omega_0 + \frac{g^2}{\Delta}(2 a^\dagger a + 1)\right)\sigma_z

Validity Condition (Complete)

The dispersive approximation requires both gΔg \ll |\Delta| and gω,ω0g \ll \omega, \omega_0. The first ensures the perturbation expansion converges; the second ensures the RWA remains valid as a precursor step.

Resonant Strong Coupling (gg moderate, ωω0\omega \approx \omega_0): The dynamics feature Rabi oscillations. If the oscillator has nn quanta, the spin flips at the Rabi frequency Ωn=2gn+1\Omega_n = 2g\sqrt{n+1}. The energy spectrum splits into doublets separated by the vacuum Rabi splitting 2g2g.

Ultrastrong Coupling (0.1<g/ω<10.1 < g/\omega < 1): The counter-rotating terms σ+a\sigma^+ a^\dagger and σa\sigma^- a cannot be neglected. The ground state is no longer the spin-down vacuum g,0|g, 0\rangle but a squeezed entangled state containing virtual excitations. The Bloch–Siegert shift becomes appreciable.

Common Misconception Alert

"Ultrastrong coupling means non-perturbative." False. The Rabi model is analytically solvable at any g/ωg/\omega (Braak 2011). The challenge is not solvability but interpretation: as gωg \to \omega, the notion of separate spin and mode excitations becomes a bad basis choice, not a physical mystery.

Deep Strong Coupling (g/ω>1g/\omega > 1): This extreme regime pushes the interaction beyond perturbative expansions. Even basic notions of separate "spin" and "oscillator" excitations become blurred.

2.5 Key Point: Symmetry as Signpost

The single-spin–single-mode system is governed by a simple Hamiltonian with symmetries that depend on coupling regime:

  • JC limit: Approximate U(1)U(1) symmetry (conservation of NN) makes the problem exactly solvable via independent 2×2 blocks.

  • Full Rabi: Only Z2\mathbb{Z}_2 parity remains, yet the model is still integrable (Braak 2011).

Integrability Clarification

The Rabi model is integrable in Braak's sense: the spectrum can be determined by transcendental equations involving confluent Heun functions. This is not Liouville integrability (which requires NN independent constants of motion for NN degrees of freedom). The discrete parity symmetry suffices for exact solvability but does not simplify dynamics to independent one-dimensional motions.


PART II — DYNAMICS FROM SIMPLICITY

3. Manifold-Resolved Dynamics

Within the Jaynes–Cummings approximation, the Hilbert space decomposes into independent manifolds:

H=M0M1M2\mathcal{H} = \mathcal{M}_0 \oplus \mathcal{M}_1 \oplus \mathcal{M}_2 \oplus \cdots

where Mn=spane,n,g,n+1\mathcal{M}n = \text{span}{|e, n\rangle, |g, n+1\rangle} for n0n \geq 0, and M1=spang,0\mathcal{M}{-1} = \text{span}{|g, 0\rangle} is the ground state (decoupled).

Within each manifold Mn\mathcal{M}_n, the Hamiltonian acts as a 2×2 matrix:

Hn=((n+12)ω+Δ2gn+1 gn+1(n+12)ωΔ2)H_n = \hbar \begin{pmatrix} (n+\tfrac{1}{2})\omega + \tfrac{\Delta}{2} & g\sqrt{n+1} \ g\sqrt{n+1} & (n+\tfrac{1}{2})\omega - \tfrac{\Delta}{2} \end{pmatrix}

The eigenstates (dressed states) are:

n,±=cosθne,n±sinθng,n+1|n, \pm\rangle = \cos\theta_n |e, n\rangle \pm \sin\theta_n |g, n+1\rangle

with tan(2θn)=2gn+1/Δ\tan(2\theta_n) = 2g\sqrt{n+1}/\Delta, and eigenfrequencies:

En,±=[(n+12)ω±12Δ2+4g2(n+1)]E_{n,\pm} = \hbar\left[(n+\tfrac{1}{2})\omega \pm \tfrac{1}{2}\sqrt{\Delta^2 + 4g^2(n+1)}\right]


4. Motional Initial States as Complexity Generators

The oscillator's initial state acts as a generator of complexity by controlling the distribution of interaction frequencies and quantum phases in the spin–mode evolution.

4.1 The Distribution Function

Define the Fock-state distribution:

pn=nψmode(0)2p_n = |\langle n | \psi_{\text{mode}}(0) \rangle|^2

This distribution pn{p_n} plays the role of a spectral weight in excitation space. The spin coherence evolves as:

σ+(t)n=0pn,eiΩnt(amplitude factors)\langle \sigma_+(t) \rangle \approx \sum_{n=0}^{\infty} p_n , e^{-i \Omega_n t} \cdot (\text{amplitude factors})

where Ωn=2gn+1\Omega_n = 2g\sqrt{n+1}. The statistics of pnp_n determine the interference pattern.

4.2 Fock States: The Single-Frequency Benchmark

If the mode starts in a Fock state n|n\rangle, the spin undergoes simple Rabi oscillations at a single frequency Ωn=2gn+1\Omega_n = 2g\sqrt{n+1}:

Pe(t)=cos2(gn+1,t)P_e(t) = \cos^2\left(g\sqrt{n+1} , t\right)

This is the reference case: predictable, oscillatory, and non-complex.

4.3 Coherent States: Collapse and Revival

A coherent state α|\alpha\rangle has Poisson-distributed Fock components:

pn=eα2α2nn!p_n = e^{-|\alpha|^2} \frac{|\alpha|^{2n}}{n!}

The spin excitation probability becomes:

Pe(t)=12[1+n=0pncos(2gn+1,t)]P_e(t) = \frac{1}{2}\left[1 + \sum_{n=0}^\infty p_n \cos(2g\sqrt{n+1} , t)\right]

Collapse: The spread of frequencies Ωn{\Omega_n} causes dephasing on timescale:

tcollapse(gn)1t_{\text{collapse}} \sim (g\sqrt{\langle n \rangle})^{-1}

Revival: The discrete quantisation allows rephasing at:

trevival2πngt_{\text{revival}} \sim \frac{2\pi\sqrt{\langle n \rangle}}{g}

The discovery of revivals was a seminal demonstration of field quantisation: revivals are a direct consequence of discrete photon number states, absent in classical or continuum fields.

4.4 Squeezed States: Variance Control

A squeezed state has reduced number uncertainty in certain quadratures. This modifies collapse/revival patterns:

  • Narrower pnp_n: Slower collapse (fewer frequencies interfere)

  • Even/odd parity structure: Frequency-comb dynamics with substructure in revivals

4.5 Displaced-Squeezed States: Tunable Asymmetry

Combining displacement and squeezing allows independent control of:

  • Mean excitation n\langle n \rangle (sets revival timescale)

  • Variance (Δn)2(\Delta n)^2 (sets collapse rate)

  • Parity structure (sets revival substructure)

4.6 Engineered Classical Mixtures: The Power-Law Emulator

Key Handbook Statement

In a single mode, power-law-like behaviour is realised in excitation space, not frequency space.

Motivation: Many-body physics studies power-law spectral densities J(ω)ωsJ(\omega) \sim \omega^s. In a single mode, we cannot replicate a continuum, but we can engineer pnp_n to emulate the dynamical signatures of non-Markovian baths.

Recipe: Prepare a mixed motional state:

ρmode=npnnn,pn(n+n0)β\rho_{\text{mode}} = \sum_n p_n |n\rangle\langle n|, \quad p_n \propto (n + n_0)^{-\beta}

using stochastically applied sideband pulses. The spin coherence then evolves as:

σ+(t)npn,ei2gn+1,t\langle \sigma_+(t) \rangle \approx \sum_n p_n , e^{-i 2g\sqrt{n+1} , t}

For large nn and β>1\beta > 1, the sum yields an envelope t(β1)\sim t^{-(\beta-1)} up to a cutoff time tcnmax/gt_c \sim n_{\max}/g.

Guardian Warning: Emulation Boundary

This is emulation. The underlying Hamiltonian is still JC; there is no bath-induced backflow, only frequency mixing. The power-law envelope is a signature, not a physical spectral density. The cutoff nmaxn_{\max} is not a bath bandwidth but a preparation limit.

4.7 Thermal States: The "Bath" Emulator

A thermal state has:

pn=nˉn(1+nˉ)n+1p_n = \frac{\bar{n}^n}{(1+\bar{n})^{n+1}}

This causes irreversible-looking damping of Rabi oscillations. The uncertain photon number acts like static disorder, washing out coherent oscillations without revival (since a thermal state has no fixed phase relationship to rephase).

This mimics decoherence even in a closed system, purely due to uncertainty in the initial mode state.

4.8 State Class Summary

Initial State
Distribution p_n
Dynamics
Complexity Source

Fock n|n\rangle

δn,n0\delta_{n,n_0}

Single-frequency Rabi

Coherent α|\alpha\rangle

Poisson

Collapse & revival

Squeezed ξ|\xi\rangle

Sub-Poisson

Modified revival pattern

Thermal

Bose–Einstein

Irreversible-like decay

No phase correlation

Power-law mixture

(n+n0)β(n+n_0)^{-\beta}

Algebraic envelope

Engineered weight


4.9 An Ordering Parameter in Excitation Space

4.9.1 Definition

For a motional state expanded in the Fock basis with probabilities pn=nψ2p_n = |\langle n | \psi \rangle|^2, the inverse participation ratio (IPR) is:

IPR=n=0nmaxpn2\text{IPR} = \sum_{n=0}^{n_{\max}} p_n^2

The effective excitation number provides intuitive interpretation:

Neff=1IPRN_{\text{eff}} = \frac{1}{\text{IPR}}

This quantifies how many Fock states (equivalently, how many JC manifolds) actively participate in the dynamics.

4.9.2 Interpretation Rule

Scope Boundary

In this handbook, the IPR is not a measure of chaos, ergodicity, or thermalisation. It is an ordering parameter in excitation space that quantifies how many JC manifolds actively participate in the dynamics.

  • Large IPR (small NeffN_{\text{eff}}) → localisation in excitation space; few manifolds dominate; regular dynamics

  • Small IPR (large NeffN_{\text{eff}}) → delocalisation; many manifolds interfere; complex envelopes

This maps directly onto dynamical signatures:

IPR Regime
N_eff
Dynamical Character

IPR1\text{IPR} \approx 1

1\approx 1

Single-frequency Rabi oscillation

IPRnˉ1/2\text{IPR} \sim \bar{n}^{-1/2}

nˉ\sim \sqrt{\bar{n}}

Collapse with clean revivals

IPR1\text{IPR} \ll 1

1\gg 1

Extended collapse, complex envelope

4.9.3 IPR Across State Classes

Initial State
Typical IPR Scaling
N_eff
Dynamical Implication

Fock n|n\rangle

IPR=1\text{IPR} = 1

1

Single frequency, fully ordered

Coherent α|\alpha\rangle

(2πnˉ)1/2\sim (2\pi\bar{n})^{-1/2}

2πnˉ\sim \sqrt{2\pi\bar{n}}

Collapse + revival

Squeezed ξ|\xi\rangle

Smaller than coherent at same nˉ\bar{n}

Larger

Enhanced interference

Thermal

(1+2nˉ)1\sim (1 + 2\bar{n})^{-1}

1+2nˉ\sim 1 + 2\bar{n}

Strong dephasing-like decay

Power-law mixture

Depends on β\beta; may diverge

Tuneable

Algebraic envelopes

4.9.4 Connection to Prior Work

Similar participation-based measures have been successfully used to characterise localisation and information spreading in trapped-ion spin–boson and spin–phonon models, where they quantify how strongly dynamics explore the available mode space (Clos, Porras, Warring & Schaetz 2016).

The conceptual move here is consistent but deliberately simpler:

  • There: Participation across mode space in multi-ion chains

  • Here: Participation across excitation space of a single mode

This reinforces the minimality thesis: the essential physics of localisation/delocalisation is already present in the single-mode setting.

4.9.5 Relation to Non-Markovianity Measures

The IPR quantifies static delocalisation in excitation space. Non-Markovianity measures quantify dynamical memory effects in the reduced spin dynamics.

In particular, information-backflow measures based on trace-distance revivals (Wittemer, Clos, Breuer, Warring & Schaetz 2018) provide a complementary, operational characterisation of memory that is sensitive to time-dependent correlations rather than initial-state structure.

Interface Boundary

In this handbook, non-Markovianity measures are used only as diagnostic overlays: they do not define regimes, but help interpret when excitation-space delocalisation translates into observable memory effects.

A low IPR (high NeffN_{\text{eff}}) implies many frequencies interfere, which correlates with slower decay of distinguishability and longer memory times—but does not cause non-Markovianity in any formal sense.

4.9.6 Operational Use

The IPR can be:

  1. Computed from any prepared distribution pnp_n

  2. Measured via motional state tomography

  3. Engineered by choosing state preparation protocols

This makes it a practical design parameter: before running an experiment, compute the IPR of your target state to predict whether dynamics will be regular (high IPR) or complex (low IPR).


PART III — FROM CLOSED TO OPEN

5. The Single Mode as an Environment

5.1 Perspective Shift

The same spin–mode Hamiltonian admits two complementary perspectives:

  1. Closed system: Track the full state ψSM(t)|\psi_{SM}(t)\rangle; evolution is unitary, information is conserved.

  2. Open system: Trace out the mode degrees of freedom; the spin's reduced state ρS(t)=TrM[ψψ]\rho_S(t) = \text{Tr}_M[|\psi\rangle\langle\psi|] follows non-unitary dynamics.

The second perspective reveals phenomena invisible in the first: effective decoherence, relaxation, and—crucially—energy renormalisation.

5.2 The Master Equation in Minimal-Dissipation Form

For a thermal initial state of the mode with mean occupation nˉ\bar{n}, the exact time-convolutionless master equation in minimal-dissipation form reads:

ρ˙S=i[ω~(t)2σz,ρS]+γz(t)D[σz]ρS+γ+(t)D[σ+]ρS+γ(t)D[σ]ρS\dot{\rho}S = -i\left[\frac{\tilde{\omega}(t)}{2}\sigma_z, \rho_S\right] + \gamma_z(t)\mathcal{D}[\sigma_z]\rho_S + \gamma+(t)\mathcal{D}[\sigma_+]\rho_S + \gamma_-(t)\mathcal{D}[\sigma_-]\rho_S

where D[O]ρ=OρO12OOρ\mathcal{D}[O]\rho = O\rho O^\dagger - \frac{1}{2}{O^\dagger O\, \rho} is the Lindblad superoperator, and all coefficients ω~,γ±,z\tilde{\omega}, \gamma_{\pm,z} are time-dependent.

The emergent Hamiltonian is:

KS(t)=ω~(t)2σzK_S(t) = \frac{\tilde{\omega}(t)}{2}\sigma_z

with renormalised frequency ω~(t)=ω0+δω~(t)\tilde{\omega}(t) = \omega_0 + \delta\tilde{\omega}(t).

5.3 Time-Dependent Energy Renormalisation

This section incorporates results from Colla et al. (2025), Nature Communications 16:2502

According to the minimal-dissipation Ansatz, the splitting between coherent (Hamiltonian) and incoherent (dissipator) evolution is uniquely determined by minimising the dissipator's action.

For the mode initially in vacuum (nˉ=0\bar{n} = 0), the energy shift has an explicit analytic form:

δω~(t)=2g2Δ11+(1+4g2Δ2)cot2(πtT)\delta\tilde{\omega}(t) = -\frac{2g^2}{\Delta} \cdot \frac{1}{1 + \left(1 + \frac{4g^2}{\Delta^2}\right)\cot^2\left(\frac{\pi t}{T}\right)}

where T(Δ)=2π/Δ2+4g2T(\Delta) = 2\pi/\sqrt{\Delta^2 + 4g^2} is the JC oscillation period.

Key features:

  • The shift is periodic in coupling duration with period TT

  • Maximum shift: δω~(T/2)=2g2/Δ\delta\tilde{\omega}(T/2) = -2g^2/\Delta

  • Near resonance (Δ0\Delta \to 0), the effect is resonantly enhanced

Time-averaged shift:

δω~(t)T=2g2sgn(Δ)Δ+2π/T\langle \delta\tilde{\omega}(t) \rangle_T = \frac{-2g^2 \, \text{sgn}(\Delta)}{|\Delta| + 2\pi/T}

This equals the dressed-state energy shift—the conventional Lamb shift emerges as a time average of fundamentally time-dependent dynamics.

5.4 Experimental Observation (Freiburg 2025)

Recent trapped-ion experiments observed these time-dependent shifts directly:

Parameter
Value

Mode frequency ωm\omega_m

2π×1.302\pi \times 1.30 MHz

Coupling g/ωg/\omega

0.05\approx 0.05 (near USC threshold)

Detuning Δ/g\Delta/g

0.8\approx 0.8

Maximum shift δω~/ω\delta\tilde{\omega}/\omega

15%\approx 15\%

Mean shift

4%\approx 4\%

The observed modulation matches the minimal-dissipation prediction (Eq. above), providing direct evidence for time-dependent energy renormalisation in the JC model.

5.5 The Generalised Lamb Shift

Conceptual Reframing

The Lamb shift and AC Stark shift—traditionally viewed as static energy corrections—are time-averaged manifestations of fundamentally dynamic phenomena. They arise from correlation buildup between system and environment, appearing only constant when the environment is traced out and time-averaged.

In the dispersive limit (Δg|\Delta| \gg g):

δω~Tg2Δ\langle \delta\tilde{\omega} \rangle_T \to -\frac{g^2}{\Delta}

recovering the standard Lamb shift formula. But this is a limiting case of a more general time-dependent structure.


6. Mode Damping and Memory

6.1 Closed vs Open: The Fundamental Distinction

Property
Closed (single mode)
Open (continuum bath)

Spectrum

Discrete

Continuous

Collapse

Temporary (dephasing)

Permanent (decoherence)

Information

Preserved (entangled)

Lost (to environment)

Entropy

Constant

Increases

Revivals

Yes

No

In a closed system, the "collapse" of spin coherence is temporary—the full state remains pure, but the spin's reduced state loses coherence due to entanglement with the mode. Given sufficient time, coherence returns (revival).

In an open system coupled to many modes, collapse is permanent—information leaks into uncontrollable degrees of freedom.

6.2 Adding True Dissipation

When the mode leaks photons (rate κ\kappa) or the spin has spontaneous emission (rate γ\gamma):

ρ˙=i[H,ρ]+κD[a]ρ+γD[σ]ρ\dot{\rho} = -\frac{i}{\hbar}[H, \rho] + \kappa \mathcal{D}[a]\rho + \gamma \mathcal{D}[\sigma^-]\rho

The competition between coherent dynamics and dissipation yields:

  • Strong damping (κ,γg\kappa, \gamma \gg g): Quantum coherence constantly "monitored," no entanglement buildup

  • Moderate damping: Damped revivals—each revival smaller than the last

  • Weak damping (κ,γg\kappa, \gamma \ll g): Near-ideal JC dynamics with slow decay envelope

Overclaim Protection

The Lindblad master equation is Markovian by construction. It cannot capture non-Markovian backflow. However, a single damped mode with κg\kappa \sim g can emulate memory effects on timescales tκ1t \lesssim \kappa^{-1}. Do not confuse this with true system-bath memory.


7. Emulating Power-Law Baths: What Is and Is Not Possible

7.1 The Spin–Boson Benchmark

The full spin–boson model couples a two-level system to a continuum of oscillators characterised by spectral density J(ω)ωsJ(\omega) \sim \omega^s:

  • s<1s < 1: Sub-Ohmic (strong low-frequency coupling)

  • s=1s = 1: Ohmic

  • s>1s > 1: Super-Ohmic

7.2 Single-Mode Emulation

A single mode cannot reproduce a continuum. However, the dynamical signatures can be emulated:

True Bath Property
Single-Mode Emulation

Spectral density J(ω)J(\omega)

Distribution pnp_n in excitation space

Power-law decay envelope

Power-law pnp_n yields t(β1)t^{-(\beta-1)}

Bath bandwidth

Preparation cutoff nmaxn_{\max}

Non-Markovian backflow

Mode recurrences (different mechanism)

7.3 The Emulation Boundary

What cannot be emulated:

  • True information loss (single mode always recurs)

  • Bath-induced localisation (requires continuum)

  • Arbitrary correlation functions (limited by HJCH_{\text{JC}} structure)

Guardian Warning

Do not claim that a single-mode experiment demonstrates "non-Markovian dynamics" without qualifying: the recurrences arise from mode finiteness, not from structured bath memory. The physics is different even if signatures overlap.


PART IV — ANALYTIC TOOLS

8. Solution Recipes

8.1 Decision Framework

8.2 The Rotating-Wave Approximation

Procedure: Drop σ+a\sigma^+ a^\dagger and σa\sigma^- a terms.

When valid: gω,ω0g \ll \omega, \omega_0 and timescales 1/ω\gg 1/\omega.

What it yields: The JC Hamiltonian decomposes into 2×2 blocks with exact eigenstates:

n,±=cosθne,n±sinθng,n+1|n, \pm\rangle = \cos\theta_n |e,n\rangle \pm \sin\theta_n |g, n+1\rangle

What it misses: Bloch–Siegert shifts, virtual-photon physics, ground-state entanglement.

8.3 The Braak Solution

For the full Rabi Hamiltonian, Braak (2011) showed integrability via:

  1. Parity decomposition: H=H+HH = H_+ \oplus H_-

  2. Bargmann representation: Map to differential equations

  3. Transcendental eigenvalue equation: Zeros of confluent Heun functions

When to use: USC regime where RWA fails but numerical truncation is uncertain.

Practical note: The solution is implicit (eigenvalues via root-finding), not closed-form. For dynamics, numerical methods are often more practical.

8.4 Perturbative Expansions

Dispersive regime (g/Δg/\Delta expansion):

Heff=H0+g2Δσzaa+g22Δσz+O(g4/Δ3)H_{\text{eff}} = H_0 + \frac{g^2}{\Delta}\sigma_z a^\dagger a + \frac{g^2}{2\Delta}\sigma_z + O(g^4/\Delta^3)

Bloch–Siegert regime (g/(ω0+ω)g/(\omega_0 + \omega) expansion):

δBSg22(ω0+ω)\delta_{\text{BS}} \approx \frac{g^2}{2(\omega_0 + \omega)}

Strong-coupling regime (treat ω\omega as perturbation on gg-dominated interaction).

8.5 Numerical Methods

Truncated diagonalisation: Truncate oscillator space to nNmaxn \leq N_{\max}, diagonalise numerically.

  • Rule of thumb: Nmax>n+5nN_{\max} > \langle n \rangle + 5\sqrt{\langle n \rangle}

  • Check: Verify eigenvalues converge as NmaxN_{\max} increases

Time-domain simulation: Integrate Schrödinger/master equation directly.

  • For pure states: Fourth-order Runge-Kutta on state vector

  • For mixed states: Lindblad equation or quantum trajectories

8.6 Open-System Tools

Markov limit: Standard Lindblad equation with constant rates.

Non-Markov (single mode): Exact TCL master equation (see Section 5.2).

Structured bath: Nakajima–Zwanzig projection, path integrals (beyond this handbook's scope).


PART V — THE TRAPPED-ION PLATFORM

9. Hamiltonian Engineering with a Single Ion

9.1 The Physical Mapping

Theoretical Object
Ion Trap Realisation

Spin-½

Two hyperfine/Zeeman states (Laser dressed )

|{\downarrow}\rangle, |{\uparrow}\rangle

e.g., F=3,mF=3,F=2,mF=2|F=3, m_F=3\rangle, |F=2, m_F=2\rangle (Laser dressed)

Mode a,aa, a^\dagger

Quantised motion (axial COM mode)

ω\omega

Trap frequency (2π×1\sim 2\pi \times 1 MHz)

ω0\omega_0

Zeeman-like splitting (tunable via an effective BB-field – Laser detuning)

gg

ηΩ\eta \Omega (Lamb-Dicke × Laser Rabi rate)

9.2 Sideband Transitions as JC/AJC Engineering

Laser-ion interaction in the Lamb-Dicke regime:

Carrier (δ=0\delta = 0):

Hcar=Ω2(σ++σ)H_{\text{car}} = \frac{\hbar \Omega}{2}(\sigma^+ + \sigma^-) Flips spin without changing motion.

Red sideband (δ=ω\delta = -\omega):

HRSB=ηΩ(σ+a+σa)H_{\text{RSB}} = \hbar \eta \Omega (\sigma^+ a + \sigma^- a^\dagger) JC interaction: spin flip ↔ phonon annihilation.

Blue sideband (δ=+ω\delta = +\omega):

HBSB=ηΩ(σ+a+σa)H_{\text{BSB}} = \hbar \eta \Omega (\sigma^+ a^\dagger + \sigma^- a) Anti-JC interaction: spin flip ↔ phonon creation.

Bichromatic drive (both sidebands): Hbigeff(σ++σ)(a+a)H_{\text{bi}} \approx \hbar g_{\text{eff}} (\sigma^+ + \sigma^-)(a + a^\dagger) Approximates the full Rabi Hamiltonian.

Operational Boundary

The Lamb-Dicke regime (η1\eta \ll 1) is not a fundamental limit—it is a convenience that linearises the coupling. Stronger coupling (large η\eta) introduces higher-order terms η2(a+a)2σz\sim \eta^2 (a + a^\dagger)^2 \sigma_z, which can be accounted for but invalidate the simple mapping.

This handbook assumes η0.2\eta \leq 0.2 unless otherwise stated.

9.3 Lamb-Dicke Validity

The Lamb-Dicke parameter:

η=k2mω=k,x0\eta = k \sqrt{\frac{\hbar}{2 m \omega}} = k , x_0

where kk is the laser wavevector component along the mode, and x0x_0 is the ground-state wavepacket size.

Lamb-Dicke regime: ηn+11\eta \sqrt{\langle n \rangle + 1} \ll 1

This ensures the motional wavefunction samples only the linear part of the laser's spatial phase.

9.4 — Targeted Hamiltonian Engineering: The 1+1D Dirac Example

(Worked example; benchmark for operator-level control)

The possibility of emulating relativistic quantum dynamics with trapped ions was articulated early in the development of quantum simulation, motivated by the observation that bichromatic sideband driving naturally generates linear couplings between spin and motional quadratures. In a seminal proposal, Lamata et al. showed that a single trapped ion, driven on the red and blue sidebands with controlled detuning, realises an effective Hamiltonian formally equivalent to the 1+1D Dirac equation, with tunable effective mass and speed of light parameters [Lamata2007].

This work established three ideas that are central to the present chapter:

  1. The Dirac Hamiltonian emerges within the same Hilbert space as the Jaynes–Cummings and quantum Rabi models; no additional degrees of freedom are required.

  2. Relativistic features such as Zitterbewegung arise as interference phenomena, not as fundamentally new dynamics.

  3. The mapping is parameter-exact within stated approximations, making it suitable as a benchmark for Hamiltonian engineering rather than a qualitative analogy.

Subsequent theoretical work generalised this approach to related relativistic models (e.g. the Dirac oscillator) and clarified the exact equivalence between Jaynes–Cummings–type interactions and relativistic operator structures under unitary transformations [Bermudez2007, Gerritsma2010].

In this handbook, the Dirac mapping is treated not as an excursion into relativistic quantum mechanics, but as a worked example of targeted Hamiltonian synthesis using the single-spin–single-mode toolbox developed in Part V. The emphasis is therefore on operational control, approximation bounds, and diagnostic observables, rather than on relativistic interpretation per se.


9.4.1 Objective

Demonstrate that the standard single-ion, single-mode toolbox (red/blue sidebands) suffices to engineer an effective Hamiltonian formally identical to the 1+1D Dirac equation, thereby benchmarking precise Hamiltonian control without expanding the system's Hilbert space.

This example illustrates:

  • Systematic use of the bichromatic drive (Section 9.2)

  • Translation between theoretical and experimental parameter spaces

  • Protocol validation through characteristic observable signatures


9.4.2 The Recipe (RSB + BSB Synthesis)

Apply a bichromatic laser field tuned symmetrically around the carrier, addressing the first red and blue motional sidebands with equal Rabi frequencies Ω\Omega and effective detuning δ\delta.

In the interaction picture (with respect to the bare qubit and motional Hamiltonians), the driven interaction reads:

Hint(t)=ηΩ[σ+(a,eiδt+ae+iδt)+σ(ae+iδt+a,eiδt)]H_{\mathrm{int}}(t) = \hbar \eta \Omega \left[\sigma^+ \left(a , e^{-i\delta t} + a^\dagger e^{+i\delta t}\right) + \sigma^- \left(a^\dagger e^{+i\delta t} + a , e^{-i\delta t}\right)\right]

Under the Lamb–Dicke condition and rotating-wave approximation (see Section 9.4.5), this reduces to a time-independent effective Hamiltonian of Dirac form:

HDirac=ceffp^σx+meffceff2σzH_{\mathrm{Dirac}} = \hbar c_{\mathrm{eff}} \, \hat{p} \, \sigma_x + \hbar m_{\mathrm{eff}} c_{\mathrm{eff}}^2 \, \sigma_z

This Hamiltonian acts on the same two-level–plus–one-mode Hilbert space as the Jaynes–Cummings model; only the basis and interpretation differ.


9.4.3 The Dictionary (Relativistic ↔ Experimental)

Relativistic Symbol
Meaning (Dirac Language)
Experimental Control

ceffc_{\mathrm{eff}}

Effective speed of light

ceff=2ηΩx0c_{\mathrm{eff}} = 2 \eta \Omega x_0

meffceff2m_{\mathrm{eff}} c_{\mathrm{eff}}^2

Rest energy

δ\hbar \delta (via symmetric detuning)

p^\hat{p}

Momentum operator

p^=i2x0(aa)\hat{p} = \dfrac{i\hbar}{2x_0}(a^\dagger - a)

x^\hat{x}

Position operator

x^=x0(a+a)\hat{x} = x_0 (a + a^\dagger)

σx,z\sigma_{x,z}

Spinor components

Qubit Pauli operators

Here x0=/(2mionω)x_0 = \sqrt{\hbar/(2 m_{\mathrm{ion}} \omega)} is the motional ground-state extent.

Tunable parameters:

  • "Speed of light" ceffc_{\mathrm{eff}}: Set via laser intensity (Ω\Omega) and Lamb–Dicke factor (η\eta)

  • "Mass" meffm_{\mathrm{eff}}: Set via laser detuning (δ\delta); sign-reversible

  • Trap frequency ω\omega: Sets length scale x0x_0 and momentum scale


9.4.4 Minimality Checkpoint

Guardian Protection

In scope: 1+1D Dirac Hamiltonian realised with two internal states and one motional mode. This is a change of interpretation, not an enlargement of the system.

Out of scope:

  • 3+1D Dirac (requires four spinor components and multiple modes)

  • Lorentz invariance tests or particle physics claims

  • Curved-space or field-theoretic extensions

If four spinor components or multiple spatial modes are required, the Single-Mode Harbour has been left.


9.4.5 Approximation Bounds (Navigator Checklist)

The mapping HintHDiracH_{\mathrm{int}} \to H_{\mathrm{Dirac}} is operationally exact when these conditions hold:

  1. Lamb–Dicke regime: ηn+10.2\eta\sqrt{\langle n\rangle + 1} \lesssim 0.2 (Section 9.3) Ensures linear coupling between motion and spin

  2. Rotating-wave approximation: δ5ηΩ|\delta| \gtrsim 5 \eta \Omega Suppresses counter-rotating terms (Section 2.2)

  3. Excitation cutoff: nmax(δ/2ηΩ)2n_{\max} \ll (\delta / 2\eta\Omega)^2 Ensures prepared motional state samples linear regime

Violation protocol: If any bound fails, revert to full quantum Rabi treatment (Section 8.3) or tighten experimental parameters before claiming Dirac mapping.


9.4.6 Observable Signature: Zitterbewegung as Interference

Zitterbewegung ("trembling motion") appears here as interference between the two eigen-manifolds of HDiracH_{\mathrm{Dirac}}, not as a relativistic mystery.

Predicted Dynamics

Prepare the spinor-motional state:

ψ(0)=12(+)α|\psi(0)\rangle = \tfrac{1}{\sqrt{2}}\left(|\uparrow\rangle + |\downarrow\rangle\right) \otimes |\alpha\rangle

with α|\alpha\rangle a coherent motional state (nˉ=α2\bar{n} = |\alpha|^2).

The position expectation value evolves approximately as:

x^(t)xZBsin!(2meffceff2t)e(Ωcolt)2\langle \hat{x}(t) \rangle \simeq x_{\mathrm{ZB}} \sin!\left(\frac{2 m_{\mathrm{eff}} c_{\mathrm{eff}}^2}{\hbar} t\right) e^{-(\Omega_{\mathrm{col}} t)^2}

with oscillation amplitude:

xZB=2ηx0nˉ+1x_{\mathrm{ZB}} = 2 \eta x_0 \sqrt{\bar{n} + 1}

and collapse rate:

ΩcolηΩ/nˉ\Omega_{\mathrm{col}} \sim \eta \Omega / \sqrt{\bar{n}}

Connection to Part II

The envelope and collapse follow directly from the same pnp_n distribution governing JC collapse and revival (Section 4.3). "Positive/negative energy branches" in Dirac language correspond to the σz=±1\sigma_z = \pm 1 components in the spin–mode product basis.

Key Insight: Zitterbewegung is not new physics—it is the JC interference pattern viewed through a different coordinate representation. The frequency 2meffceff2/=2δ2 m_{\mathrm{eff}} c_{\mathrm{eff}}^2 / \hbar = 2\delta is simply twice the commanded detuning.


9.4.7 Operational Procedure (Benchmark Protocol)

Five-step validation sequence:

  1. Initialise: Doppler cool + resolved sideband cool to nˉ0.1\bar{n} \lesssim 0.1; apply displacement drive to prepare α|\alpha\rangle with target nˉ\bar{n}.

  2. Spinor preparation: Apply carrier π/2\pi/2 pulse to create 12(+)\tfrac{1}{\sqrt{2}}(|\uparrow\rangle + |\downarrow\rangle).

  3. Evolve: Apply bichromatic RSB+BSB drive (equal amplitudes, symmetric detuning ±δ\pm\delta) for time tt.

  4. Readout:

    • Measure σz(t)\langle\sigma_z(t)\rangle via fluorescence detection

    • Reconstruct x^(t)\langle \hat{x}(t)\rangle via sideband spectroscopy or direct motional tomography

  5. Validate:

    • Fit oscillation frequency → extract δ\delta (compare to commanded value)

    • Fit amplitude → extract ceffc_{\mathrm{eff}} and nˉ\bar{n}

    • Fit envelope → characterise decoherence rates

This single sequence simultaneously benchmarks:

  • Hamiltonian calibration (frequency/amplitude accuracy)

  • Motional state preparation (coherent displacement fidelity)

  • Coherence times (via envelope decay)

Typical Parameter Set

Sanity-Check Values (for experiment planning):

Parameter
Typical Value
Resulting Quantity

ω\omega

2π×1.52\pi \times 1.5 MHz

x010x_0 \approx 10 nm (for 40^{40}Ca+^+)

η\eta

0.05–0.15

(geometry-dependent)

Ω\Omega

2π×502\pi \times 50 kHz

ceff1c_{\mathrm{eff}} \approx 133 µm/ms

δ\delta

2π×1002\pi \times 100 kHz

2δ=2π×2002\delta = 2\pi \times 200 kHz (ZB freq)

nˉ\bar{n}

5–20

xZB20x_{\mathrm{ZB}} \approx 206060 nm

Evolution time

0–50 µs

~10 ZB cycles before collapse

Expected signal: Oscillating x^(t)\langle\hat{x}(t)\rangle with ~30 nm amplitude, 5 µs period, collapsing over ~20 µs (for nˉ10\bar{n} \approx 10).


9.4.8 Error Budget and Diagnostic Protocol

Operator Triage (First-Response Guide)

Use this checklist to prioritise debugging before consulting the full error budget:

  • ZB frequency wrong or drifting? → Check detuning drift (Row 1) and AC Stark systematics (Row 2) first.

  • Contrast lost early (clean decay)? → Check intensity noise (Row 3), bichromatic phase noise (Row 5), and qubit dephasing (Row 11).

  • Waveform distorted (non-sinusoidal)? → Check RSB/BSB imbalance (Row 4) and Lamb–Dicke breakdown (Row 8).

  • Slow parameter creep across the day? → Check ω(t)\omega(t) drift (Row 7) and your reference chain for δ(t)\delta(t) (Row 1).

Protocol: Identify symptom class → check listed rows → apply mitigation → rerun a short, high-SNR sequence (small nˉ\bar{n}, short tt) → iterate.


Table 9.1 — Dirac Protocol Error Budget (Zitterbewegung Benchmark)

#
Error Source
Hamiltonian Artifact
Signature in Zitterbewegung

1

δ\delta drift (laser/reference instability)

Time-dependent mass term: meffceff2δ(t)m_{\mathrm{eff}} c_{\mathrm{eff}}^2 \propto \delta(t)

Frequency jitter/shift of ZB oscillation; phase wander between repetitions. Mitigation: lock δ\delta to a stable reference; interleave Ramsey calibration runs.

2

Differential AC Stark shift (carrier/sideband imbalance; spectator levels)

Extra σz\sigma_z term ("fake mass"): +δStark2σz+\tfrac{\hbar\delta_{\mathrm{Stark}}}{2}\sigma_z

Systematic frequency offset (apparent mass) mismatched from commanded δ\delta. Mitigation: apply Stark-compensation tones; calibrate via carrier-only spectroscopy (Section 9.2).

3

Intensity / Rabi-rate noise Ω(t)\Omega(t)

Fluctuating ceffηΩ(t)c_{\mathrm{eff}} \propto \eta\Omega(t); weak amplitude-to-phase conversion

Contrast loss; run-to-run envelope variability; fitted ceffc_{\mathrm{eff}} broadens. Mitigation: AOM intensity servo; normalise via simultaneous Rabi monitor.

4

Red/blue amplitude imbalance (ΩRSBΩBSB\Omega_{\mathrm{RSB}} \neq \Omega_{\mathrm{BSB}})

Rotated coupling axis; residual JC/AJC asymmetry; σy\sigma_y admixture

Distorted waveform (non-sinusoidal); offset in x^(t)\langle \hat{x}(t)\rangle; reduced dictionary fidelity. Mitigation: calibrate sideband Rabi rates separately; active balancing via DDS amplitude trim (Section 11).

5

Bichromatic phase noise / jumps

Drifting coupling axis: σxcosϕ,σx+sinϕ,σy\sigma_x \to \cos\phi,\sigma_x + \sin\phi,\sigma_y

Apparent dephasing without heating; "rotating" quadrature readout; poor repeatability. Mitigation: phase-coherent DDS/AWG; fixed phase convention; periodic phase re-zero.

6

Motional heating (electric-field noise)

Stochastic drive + diffusion; effective Lindblad heating term κD[a]\kappa \mathcal{D}[a^\dagger]

Faster envelope decay; reconstructed nˉ\bar{n} grows; sideband asymmetry. Mitigation: filter/ground electrodes; shorten evolution; sympathetic cooling where available (Section 6.2).

7

Motional frequency drift ω(t)\omega(t)(trap potential drift)

Changes x0ω1/2x_0 \propto \omega^{-1/2}; perturbs interaction picture and scaling of x^,p^\hat{x}, \hat{p}

Slow systematic drift of inferred ceffc_{\mathrm{eff}} and xZBx_{\mathrm{ZB}}; possible beating. Mitigation: stabilise trap RF/DC; interleave ω\omega measurement (Section 11).

8

Lamb–Dicke breakdown (too large ηn+1\eta\sqrt{\langle n\rangle + 1})

Higher-order couplings beyond x^,p^\hat{x}, \hat{p}; dictionary becomes nonlinear

Non-Gaussian distortion; harmonics; parameter-dependent waveform. Mitigation: reduce nˉ\bar{n}; increase ω\omega; enforce ηn+10.2\eta\sqrt{\langle n\rangle + 1} \le 0.2 (Section 9.3).

9

RWA violation (δ≫̸ηΩ\delta\not\gg \eta\Omega)

Counter-rotating terms; Bloch–Siegert-like shifts

Extra fast components; frequency pulling; apparent mass deviates from δ\delta. Mitigation: increase δ|\delta| or reduce Ω\Omega; scaling check vs Ω\Omega (Section 2.2).

10

Off-resonant carrier coupling (imperfect spectral selectivity)

Unwanted σx\sigma_x rotation during evolution; spurious spinor mixing

Baseline offsets; reduced ZB visibility; spurious oscillations at carrier frequency. Mitigation: tighten spectral shaping; verify with carrier-only control runs.

11

Qubit dephasing (magnetic noise, LO phase noise)

Dephasing channel on σz\sigma_z (or effective quantisation axis)

Pure contrast loss; minimal frequency shift; envelope decays even at low heating. Mitigation: magnetic shielding; clock states; dynamical decoupling if compatible (Section 6.2).

12

SPAM errors (state prep, readout)

No Hamiltonian change; biases inferred σz(t)\langle \sigma_z(t)\rangle and reconstructed x^(t)\langle \hat{x}(t)\rangle

Apparent reduced amplitude/offset; error floor on contrast. Mitigation: SPAM calibration matrix; repeat reference points at t=0t = 0 (Section 10.2).

13

x^(t)\langle \hat{x}(t)\ranglereconstruction bias (sideband spectroscopy model)

Measurement-model artefact; incorrect mapping from spectra to quadrature

Phase/amplitude bias in x^(t)\langle \hat{x}(t)\rangle while σz(t)\langle\sigma_z(t)\rangle appears consistent. Mitigation: cross-check with independent tomography; simulate reconstruction pipeline (Section 10.2).

Notation lock: δ\delta (commanded detuning), Ω\Omega (Rabi rate), η\eta (Lamb–Dicke factor), x0x_0 (ground-state size), σx,z\sigma_{x,z} (Pauli operators), ω\omega (motional frequency).


9.4.9 What This Example Demonstrates (and What It Does Not)

Demonstrates

  • Precise Hamiltonian synthesis using the Part V toolbox (RSB, BSB, bichromatic drives)

  • Equivalence of operator languages: JC/Rabi dynamics ↔ Dirac dynamics via basis transformation

  • Parameter tunability: "Speed of light" and "mass" are experimental knobs, not fundamental constants

  • Validation protocol: Zitterbewegung as a calibration-quality benchmark

  • Error attribution: Table 9.1 provides systematic debugging framework

Does Not Demonstrate

  • Fundamental relativistic effects: No Lorentz invariance, causality tests, or spacetime structure probed

  • Particle physics: No pair creation, no second quantisation, no field-theoretic phenomena

  • Multi-dimensional Dirac physics: 3+1D requires expanding beyond single-spin–single-mode (see Minimality Checkpoint)

Conceptual Boundary: This example shows that the trapped-ion platform can emulate the mathematical structure of the 1+1D Dirac equation with parameter-level control. It does not test whether nature's actual electrons obey this equation (they do, but for different reasons). The value lies in Hamiltonian engineering precision, not in simulating high-energy physics per se.


Cross-Reference to Appendix B

For readers interested in the formal equivalence, Appendix B provides the explicit unitary transformation mapping the Dirac oscillator to the Jaynes–Cummings Hamiltonian. This confirms that:

HDirac oscillatorU=eiπσy/4HJCH_{\mathrm{Dirac\ oscillator}} \xrightarrow{U = e^{-i\pi\sigma_y/4}} H_{\mathrm{JC}}

Operational takeaway: Every JC experiment is already a Dirac-oscillator experiment. The Hamiltonian is invariant; only the question asked of it changes.


Note on Experimental Status and Prior Demonstrations

The Dirac-emulation proposal of Lamata et al. was followed by a series of experimental demonstrations using single trapped ions, most prominently by the group of Gerritsma et al. These experiments directly observed key signatures predicted by the effective Dirac dynamics, including Zitterbewegung and Klein-paradox–like behaviour, using a single ^{40}Ca^+ ion driven by bichromatic sideband fields [Gerritsma2010Nature, Gerritsma2011PRL].

Further experiments explored variations of relativistic Hamiltonians, including tunable effective masses, sign changes of the mass term, and extensions to Dirac-oscillator–type dynamics [Casas2016, Lamata2011]. In all cases, the physical implementation relied on the same core ingredients documented in this chapter: resolved sideband control, coherent bichromatic driving, and precise calibration of detunings and Rabi frequencies.

From an operational perspective, these experiments demonstrated that:

  • The Dirac mapping is experimentally accessible with standard trapped-ion technology.

  • Observables such as Zitterbewegung are robust diagnostics of Hamiltonian synthesis quality, sensitive to detuning, phase stability, and motional coherence.

  • No fundamentally new hardware is required beyond that already used for Jaynes–Cummings and Rabi-model experiments.

The present handbook section does not aim to reproduce or extend these results. Instead, it abstracts their core lesson: the 1+1D Dirac Hamiltonian serves as a stringent, interpretable benchmark for single-ion Hamiltonian engineering, complementary to cat-state generation, squeezing, and collapse-and-revival protocols treated elsewhere in Part V.


10. Engineering Initial States

10.1 Ground-State Cooling

Resolved sideband cooling: Drive red sideband repeatedly. ,nRSB,n1decay,n1|{\downarrow}, n\rangle \xrightarrow{\text{RSB}} |{\uparrow}, n-1\rangle \xrightarrow{\text{decay}} |{\downarrow}, n-1\rangle

Achieves nˉ0.1\bar{n} \lesssim 0.1 routinely.

10.2 State Preparation Protocols

Target State
Preparation Method
Typical Fidelity

Fock n|n\rangle

Sequential BSB π\pi-pulses from 0|0\rangle

>95%>95\%

Coherent α|\alpha\rangle

Classical drive (displacement)

>99%>99\%

Squeezed ξ|\xi\rangle

Parametric modulation / pulse sequence

>90%>90\%

Cat state

Conditional displacement + measurement

>85%>85\%

Power-law mixture

Stochastic sideband pulses

<Did anyone do this?>

10.3 Cat State Generation

Protocol (Monroe et al. 1996):

  1. Prepare ,0|{\downarrow}, 0\rangle

  2. Apply carrier π/2\pi/2 to create (+)/2(|{\downarrow}\rangle + |{\uparrow}\rangle)/\sqrt{2}

  3. Apply state-dependent displacement: ,α,,α|{\downarrow}\rangle \to |{\downarrow}, \alpha\rangle, |{\uparrow}\rangle \to |{\uparrow}, -\alpha\rangle

  4. Result: (,α+,α)/2(|{\downarrow}, \alpha\rangle + |{\uparrow}, -\alpha\rangle)/\sqrt{2}

Normalisation Note

The cat state (α+α)/2(|\alpha\rangle + |-\alpha\rangle)/\sqrt{2} is properly normalised only for α1|\alpha| \gg 1 (where αα=e2α20\langle \alpha | -\alpha \rangle = e^{-2|\alpha|^2} \approx 0). For small α\alpha, include the overlap correction.


11. Experimental Knob Map

Theoretical Parameter
Ion Trap Realisation
Typical Range
Calibration Method

gg

ηΩ\eta \Omega

2π×(1300)2\pi \times (1-300) kHz

Rabi flopping on sideband

ω\omega

Trap frequency

2π×(0.55)2\pi \times (0.5-5) MHz

Secular frequency measurement

ω0\omega_0

Zeeman shift

Tunable via effective BB-field

Sideband spectroscopy

Δ=ωω0\Delta = \omega - \omega_0

Laser detuning

2π×(01)2\pi \times (0-1) MHz

Ramsey spectroscopy

κ\kappa (damping)

Sympathetic cooling

10104s110-10^4\, \text{s}^{-1}

Sideband asymmetry

γϕ\gamma_\phi (dephasing)

Engineered noise

Tuneable

Ramsey decay

nˉ\bar{n}

Cooling quality

0.01100.01-10

Sideband ratio

pnp_n distribution

Squeeze + displace + mix

Engineered

Motional tomography

11.1 Accessing Different Regimes

Regime
How to Access

Dispersive

Large detuning Δg\Delta \gg g (detune laser far from sideband)

Resonant JC

Δ0\Delta \approx 0, moderate ηΩ\eta \Omega

Near-USC

Small Δ\Delta, large ηΩ\eta \Omega (typical: g/ω0.050.1g/\omega \sim 0.05-0.1)

Engineered dissipation

Add sympathetic cooling or electrode noise


PART VI — HORIZONS

12. What Is Already Known

The single-spin–single-mode system is among the most thoroughly characterised quantum systems:

  • JC spectrum and dynamics: Fully solved analytically (Jaynes & Cummings 1963)

  • Collapse and revival: Observed in cavity QED and ion traps (1980s-present)

  • Braak integrability: Full Rabi model solved (2011)

  • USC phenomena: Observed in superconducting circuits (2010s)

  • Time-dependent Lamb shift: Observed in trapped ions (Colla et al. 2025)

  • State engineering: Fock, coherent, squeezed, cat states all demonstrated

13. What This Handbook Enables

13.1 For Graduate Students

  • Conceptual bridge: Connect textbook Hamiltonians to laboratory reality

  • Decision protocols: Know which approximation applies when

  • Operational fluency: Translate theoretical parameters to experimental knobs

13.2 For Experimentalists

  • Design guidance: Engineer specific dynamical signatures via state preparation

  • Regime navigation: Identify which physics dominates in your parameter range

  • Benchmark protocols: Test system calibration against exact predictions

13.3 For Theorists

  • Minimal testbed: Verify new ideas in the simplest non-trivial setting

  • Emulation boundaries: Know what can and cannot be simulated with one mode

  • Open-system foundations: Use the minimal-dissipation framework as reference

14. Open Questions and Outlook

14.1 Within the Single-Mode Framework

  1. Non-thermal initial states: How does KS(t)K_S(t) depend on squeezed, Fock, or cat initial states of the mode?

  2. Beyond JC: Full characterisation of time-dependent renormalisation in the Rabi model.

  3. Quantum thermodynamics: Connecting KS(t)K_S(t) to work and heat definitions at strong coupling.

14.2 Extensions

  1. Multi-mode environments: How do signatures change with 2, 3, ..., NN modes?

  2. Structured baths: Engineering effective spectral densities with multiple modes.

  3. Autonomous quantum machines: Using emergent driving for heat engines.

14.3 The Handbook's Continuing Role

This handbook is a living document. As the field develops, new sections will address:

  • Additional state classes and their dynamics

  • Refined experimental protocols

  • Connections to quantum error correction and fault tolerance


References

  • Bermudez, A., Martin-Delgado, M. A. & Solano, E. (2007). Exact mapping of the 1+1 Dirac oscillator onto the Jaynes–Cummings model. Physical Review A 76, 041801(R).

  • Braak, D. (2011). Integrability of the Rabi Model. Physical Review Letters 107, 100401.

  • Casas, M. et al. (2016). Quantum simulation of the Majorana equation and unphysical operations. Physical Review X 6, 041018.

  • Clos, G., Porras, D., Warring, U. & Schaetz, T. (2016). Time-Resolved Observation of Thermalization in an Isolated Quantum System. Physical Review Letters 117, 170401.

  • Colla, A. & Breuer, H.-P. (2022). Open-system approach to nonequilibrium quantum thermodynamics at arbitrary coupling. Physical Review A 105, 052216.

  • Colla, A., Hasse, F., Palani, D., Schaetz, T., Breuer, H.-P. & Warring, U. (2025). Observing time-dependent energy level renormalisation in an ultrastrongly coupled open system. Nature Communications 16, 2502.

  • Forn-Díaz, P., Lamata, L., Rico, E., Kono, J. & Solano, E. (2019). Ultrastrong coupling regimes of light-matter interaction. Reviews of Modern Physics 91, 025005.

  • Gerritsma, R. et al. (2010). Quantum simulation of the Dirac equation. Nature 463, 68–71.

  • Gerritsma, R. et al. (2011). Quantum simulation of Klein tunneling and Zitterbewegung. Physical Review Letters 106, 060503.

  • Jaynes, E. T. & Cummings, F. W. (1963). Comparison of quantum and semiclassical radiation theories with application to the beam maser. Proceedings of the IEEE 51, 89–109.

  • Lamata, L., León, J., Schätz, T. & Solano, E. (2007). Dirac Equation and Quantum Relativistic Effects in a Single Trapped Ion. Physical Review Letters 98, 253005.

  • Lamata, L. et al. (2011). Relativistic quantum mechanics with trapped ions. Physical Review A 84, 012335.

  • Leggett, A. J. et al. (1987). Dynamics of the dissipative two-state system. Reviews of Modern Physics 59, 1–85.

  • Leibfried, D., Blatt, R., Monroe, C. & Wineland, D. (2003). Quantum dynamics of single trapped ions. Reviews of Modern Physics 75, 281–324.

  • Meekhof, D. M., Monroe, C., King, B. E., Itano, W. M. & Wineland, D. J. (1996). Generation of Nonclassical Motional States of a Trapped Atom. Physical Review Letters 76, 1796.

  • Monroe, C., Meekhof, D. M., King, B. E. & Wineland, D. J. (1996). A "Schrödinger Cat" Superposition State of an Atom. Science 272, 1131–1136.

  • Scully, M. O. & Zubairy, M. S. (1997). Quantum Optics. Cambridge University Press.

  • Shore, B. W. & Knight, P. L. (1993). The Jaynes–Cummings model. Journal of Modern Optics 40, 1195–1238.

  • Smirne, A. & Vacchini, B. (2010). Nakajima-Zwanzig versus time-convolutionless master equation for the non-Markovian dynamics of a two-level system. Physical Review A 82, 022110.

  • Wittemer, M., Clos, G., Breuer, H.-P., Warring, U. & Schaetz, T. (2018). Measurement of quantum memory effects and its fundamental limitations. Physical Review A 97, 020102.


Appendix A: Notation Summary

Symbol
Meaning

ω0\omega_0

Bare spin transition frequency

ω\omega

Oscillator/mode frequency

gg

Spin-mode coupling strength

Δ=ωω0\Delta = \omega - \omega_0

Detuning

η\eta

Lamb-Dicke parameter

Ω\Omega

Laser Rabi frequency

nˉ\bar{n}

Mean thermal occupation

pnp_n

Fock-state probability distribution

Ωn=2gn+1\Omega_n = 2g\sqrt{n+1}

nn-photon Rabi frequency

T=2π/Δ2+4g2T = 2\pi/\sqrt{\Delta^2 + 4g^2}

JC oscillation period

KS(t)K_S(t)

Emergent (renormalised) system Hamiltonian

ω~(t)\tilde{\omega}(t)

Renormalised spin frequency

IPR=npn2\text{IPR} = \sum_n p_n^2

Inverse participation ratio

Neff=1/IPRN_{\text{eff}} = 1/\text{IPR}

Effective excitation number


Appendix B:

Version History

Version
Date
Changes

0.1

2025–12-17

Initial draft (review format)

0.2

2025–12-17

Modular restructure; Guardian protections; Navigator; minimal-dissipation Ansatz; power-law emulator; trapped-ion axis elevated

0.2.1

2025–12-17

Added Section 4.9 (IPR ordering parameter); citation apparatus; references for Clos et al. 2016, Wittemer et al. 2018

0.2.2

2025–12-18

Added Section 9.4 (1+1D Dirac worked example); bichromatic drive synthesis; Zitterbewegung as JC interference; Table 9.1 error budget with operator triage; Appendix C cross-reference

0.2.3

2025-12-19

Added clarifying Foreword

Last updated