G–Quantum Dynamics Reference: Frames, RWA, and Dressed States
Rigorous operational definitions distinguishing interaction/rotating frames (exact transformations), the rotating-wave approximation (controlled truncation), and dressed states (diagonalization).
Document Scope and Navigation
Purpose: This document clarifies the conceptual boundaries between interaction/rotating frames, the rotating-wave approximation (RWA), dressed-state bases, and their connections to symmetry via Noether's theorem. It addresses decades of terminology drift in the AMO literature.
Target audience and prerequisites: This document is an internal reference standard and advanced teaching note. It assumes familiarity with:
Basic two-level Rabi oscillations (Bloch sphere, rotating frame intuition)
Schrödinger and interaction pictures
Elementary quantum optics or AMO phenomenology
It is not a first introduction to driven quantum systems. Students encountering RWA for the first time should begin with [SCU97] Chapter 5 or [COH92] Chapter VI before using this document.
Relationship to standard textbooks: Standard two-level system derivations and RWA procedures are given in [SCU97] Chapters 5–6 and [COH92] Chapters VI–VII. This document does not reproduce those derivations but instead focuses on terminology disambiguation, symmetry structure (Noether perspective), and operational diagnostics. Readers unfamiliar with basic Rabi oscillations or the rotating frame should begin with those references before using this document.
Recommended usage:
For graduate courses: Assign after initial RWA lectures as consolidation/disambiguation
For manuscripts: Cite specific sections when defining terminology or justifying approximations
For lab standards: Use §§6–7 operationally; treat §§1–3 as conceptual reference
How to navigate this document:
Core material
[CORE](§§1–3): Foundational definitions and distinctions—required readingExtensions
[ADVANCED](§§4–5): Floquet theory and continuous-variable analogies—recommended for advanced studyPractical tools
[REFERENCE](§§6–7): Diagnostics, experimental signatures, and manuscript guidance—reference as neededAppendix A
[ADVANCED OPTIONAL]: Beyond-RWA perturbation theory—optional advanced material
Quick Reference: One-Page Summary [REFERENCE]
[REFERENCE]The Four Conceptual Steps
Interaction frame
Exact unitary
Basis for states/operators
Physics, probabilities
Always
Rotating frame
Exact unitary (at ω)
Removes fast ω oscillations
Physics, probabilities
Always
RWA
Approximation
Physics (terms dropped)
——
Ω/ω₀ ≲ 0.1
Dressed basis
Diagonalization
State representation
Eigenvalues, physics
After RWA
Visual Flow Diagram
Figure 0.1: Information flow showing the four conceptual steps. Only RWA introduces approximation (dropping counter-rotating terms); all other steps are exact mathematical operations that preserve the underlying physics.
Key Formulas (RWA regime)
Dressed states:
Dressed energies:
RWA validity:
Energy shifts: Error ~ (Ω/4Δ)²
General threshold: Ω/ω₀ < 0.05–0.1
Common Pitfalls
Don't say: "We go to the dressed frame" Do say: "We diagonalize to the dressed basis"
Don't say: "RWA is exact" Do say: "RWA introduces ~(Ω/ω₀)² errors"
Don't assume: Dressed states = quantized field Specify: Classical drive vs. cavity QED
When to Use What
Need slowly varying Hamiltonian? → Rotating frame
Want analytic solutions? → Apply RWA (check validity!)
Measuring spectroscopy? → Use dressed basis to predict peaks
Strong driving (Ω/ω₀ > 0.1)? → Use Floquet (§4)
Part I: Foundations
1. Interaction Frame and Rotating Frame [CORE]
[CORE]Consider a driven two-level system with
1.1 Formal definition: Interaction picture
The interaction picture with respect to uses
The interaction-picture Hamiltonian is
Mathematical status: Exact unitary transformation. No approximations, no physics changed.
1.2 Rotating frame as a special case
Choosing
defines the rotating frame at the drive frequency . This removes the fast precession and converts some drive terms into slowly varying or static contributions.
Physical interpretation: Move to a reference frame co-rotating with the drive field; in this frame, near-resonant terms appear stationary while far-detuned terms oscillate rapidly.
1.3 Symmetry implications (Noether perspective)
Switching frames is an exact unitary transformation; it does not create or destroy continuous symmetries. Any conserved quantity remains conserved, but expressed in the new representation:
Operational diagnostic: Populations transform under frame changes, but total probability (trace) is invariant. Physically measurable expectation values in the lab frame are unchanged.
Critical note: Frame transformations neither add nor remove symmetries—they are passive reparametrizations, not dynamical approximations.
2. The Rotating-Wave Approximation (RWA) [CORE]
[CORE]2.1 Physical motivation
In the rotating frame, the Hamiltonian contains:
Co-rotating terms: oscillate as (near-resonant, slow)
Counter-rotating terms: oscillate as (far-detuned, fast)
When (near resonance) and (weak drive), the counter-rotating terms average to near-zero over timescales , contributing negligibly to dynamics on slower timescales .
2.2 RWA as approximation
Definition: Discard the counter-rotating terms, yielding a time-independent effective Hamiltonian:
Mathematical status: This is an approximation, not a transformation. Terms are neglected, not unitarily removed.
Validity conditions (qualitative):
2.3 Quantitative error estimates
Important caveat: The error estimates below are order-of-magnitude guidelines derived from perturbative treatments assuming sinusoidal CW driving at or near resonance. Precise values depend on:
Pulse shape and duration (Gaussian, square, adiabatic ramps, etc.)
Detuning regime (on-resonance vs. dispersive vs. far-detuned)
Observable of interest (populations vs. coherences vs. geometric phases)
Best practice: Treat these as initial validity checks; verify numerically for your specific protocol.
Dressed-state energies
Bloch-Siegert shift (on-resonance)
Rabi fidelity (resonant drive)
General guideline
RWA breakdown begins
(*) Indicative scaling, not strict bounds. See text for protocol-dependent caveats.
Validation protocol: In all precision applications, treat these analytic criteria as pre-filters and validate by explicit numerical simulations for the concrete pulse sequence used. For trapped-ion gates, compare RWA predictions to full time-dependent Schrödinger equation (TDSE) or Floquet calculations before experimental implementation.
For pulsed driving, additional timescale constraints:
Adiabatic limit: (instantaneous eigenbasis follows pulse)
Sudden limit: (no phase accumulation)
Intermediate regime: Requires full time-dependent analysis or Floquet treatment (see §4)
References: Quantitative criteria from [HAE68, GOL14]; error scaling confirmed in [IRI07].
2.3.1 Failure Mode Hierarchy: Phase Errors Before Population Errors [CORE + PRACTICE]
[CORE + PRACTICE]RWA breakdown manifests in a predictable sequence as increases. Understanding this hierarchy is critical for quantum information applications, where phase fidelity requirements are typically more stringent than population fidelity requirements.
Stage 1: Phase Errors ()
Physical mechanism: The Bloch-Siegert shift introduces an additional dynamic phase:
This phase accumulates linearly with time and is not captured by the RWA Hamiltonian.
Observable consequences:
Geometric phase accumulation deviates from RWA prediction
-pulses become , causing systematic gate errors
Multi-pulse sequences accumulate phase errors coherently
Detection methods:
Ramsey interferometry with varying pulse separation
Spin-echo phase tomography
Gate fidelity measurements via randomized benchmarking
Critical impact: In quantum gate implementations, a 1% phase error translates directly to ~1% infidelity even if population transfer is perfect. For fault-tolerant quantum computation with typical error thresholds , this limits RWA validity to for multi-qubit gates.
Correction strategies:
Analytic correction: Include Bloch-Siegert shift explicitly:
Composite pulses: Design pulse sequences that cancel systematic phase errors [WIM94]
Calibration: Empirically determine phase shift and pre-compensate in gate design
Stage 2: Population Errors ()
Physical mechanism: Counter-rotating terms cause off-resonant excitation of non-target states, leading to incomplete population transfer.
Observable consequences:
Rabi oscillation amplitude deviates from ideal
Residual population in unwanted states after "complete" transfer
Oscillation frequency shifts:
Detection methods:
State tomography after nominally complete population transfer
Fluorescence measurements showing non-zero residual excited-state population
Fourier analysis of Rabi oscillation data revealing frequency shifts
Critical impact: Population errors are typically less severe than phase errors for the same ratio, because they scale as whereas phase errors scale linearly with time.
Stage 3: Multiphoton Processes ()
Physical mechanism: Higher Fourier blocks in the Floquet spectrum become significantly populated, enabling transitions at for .
Observable consequences:
New spectral features appear in spectroscopy (multiphoton resonances)
Qualitatively different dynamics (e.g., photon-assisted tunneling, dynamical localization)
RWA conceptual framework breaks down entirely
Detection methods:
Broadband spectroscopy revealing sidebands at
Time-resolved measurements showing non-sinusoidal oscillations
Parameter sweeps revealing unexpected resonances
Critical impact: At this regime, RWA is no longer a controlled approximation but a qualitatively incorrect model. Full Floquet analysis is mandatory (see §4).
Validation Protocol for Quantum Gates
For precision applications (quantum information processing, high-fidelity state transfer), use the following staged validation:
Step 1 – Analytic pre-check:
Spectroscopy: rad (10%)
Single-qubit gates: rad (1%)
Multi-qubit gates: rad (0.1%)
Step 2 – Numerical validation (if Stage 1 marginal):
If infidelity : Apply Bloch-Siegert correction and re-validate
If still : Escalate to Floquet analysis (§4) or reduce
Step 3 – Experimental characterization:
Typical Tolerance Budgets
Spectroscopy
10%
10% (0.1 rad)
State preparation
1%
5% (0.05 rad)
Single-qubit gates
0.1%
1% (0.01 rad)
Two-qubit gates
0.1%
0.1% (0.001 rad)
Fault-tolerant QC
0.01%
0.01% (0.0001 rad)
Key insight: For quantum information processors, phase fidelity is the limiting constraint, not population fidelity. Gate designers must verify phase corrections explicitly, not merely population transfer.
Reference: Fault-tolerant thresholds and phase sensitivity analysis from [GOT09, BAL16].
2.4 Symmetry restoration (Noether perspective)
Before RWA: The full Hamiltonian is periodic but not time-independent. Continuous time-translation symmetry is broken; energy is not conserved (though Floquet quasienergy is—see §4).
After RWA: The effective Hamiltonian is time-independent. Within the RWA model, we gain an exact conservation law: the rotating-frame energy becomes a constant of motion. This is a consequence of the approximate time-translation symmetry we have imposed by neglecting counter-rotating terms—it holds exactly within the model, but only approximates the true dynamics when .
Noether consequence: The rotating-frame energy (equivalently, the Floquet quasienergy at lowest order) becomes conserved:
Critical clarification: This is an approximate model symmetry, valid only within the RWA regime. The exact Hamiltonian does not possess continuous time-translation symmetry; the symmetry is introduced by the approximation, not by a frame transformation.
2.5 Additional symmetry: Excitation-number conservation
For quantized fields (Jaynes–Cummings model), RWA additionally removes terms that violate photon-number conservation, generating an approximate U(1) symmetry [IRI07, FEL14]:
This U(1) symmetry is absent in the full Rabi model (which includes and terms) and is a direct consequence of the RWA truncation.
Operational diagnostic: Counter-rotating contributions are ; measured Bloch-Siegert shift matches within error bars; spectroscopic lineshapes show symmetric splitting without higher-order sidebands.
3. Dressed States [CORE]
[CORE]3.1 Definition and mathematical structure
Dressed states are the eigenstates of the effective (RWA) Hamiltonian:
Explicitly, in the bare-state basis :
Physical interpretation: The dressed states encode the hybridization of bare atomic levels induced by the near-resonant drive. At resonance (), maximum mixing occurs (); far from resonance (), dressed states approach bare states.
Mathematical status: This is a basis change (diagonalization), not a frame transformation or approximation. It is exact within the RWA Hamiltonian.
Visual representation: Avoided crossing structure
Figure 3.1: Dressed-state energies vs. normalized detuning . The avoided crossing at has minimum gap , representing maximum hybridization of bare states. Far from resonance (), energies asymptotically approach bare-state values , and dressed states become essentially unmixed.
3.2 Noether perspective: Dressed energies as conserved charges
Since is time-independent (after RWA), the dressed-state populations
are strictly conserved under ideal unitary evolution.
The dressed energies are the eigenvalues associated with this (approximate) symmetry. In Noether language, they serve as the conserved charges corresponding to time-translation invariance in the RWA model. Any deviation from population conservation signals breakdown of RWA, decoherence, or additional perturbations.
Operational diagnostic: In the dressed basis, populations remain constant—no oscillations between and . Transitions require additional perturbations (spontaneous emission, dephasing, probe fields, parameter sweeps).
3.3 Terminology evolution and disambiguation
Historical usage (1960s–1980s, [COH92, SCU97]): "Dressed states" originally referred to eigenstates of an atom coupled to a quantized electromagnetic field—solutions to the Jaynes–Cummings Hamiltonian in cavity QED.
Modern usage (1990s–present, [SHO11, WU19]): "Dressed states" now commonly denotes eigenstates of any effective time-independent Hamiltonian obtained after RWA, including semi-classical treatments with classical driving fields (Floquet-Rabi systems).
Conceptual distinction:
Vacuum effects
Vacuum Rabi splitting at
No zero-point splitting
Spontaneous emission
Modified Purcell factor even at
Free-space decay restored at
Photon statistics
Sub-Poissonian, entanglement
Coherent state (Poissonian)
Field quantization
Essential ( operators)
Not required (classical amplitude)
Operational test: Measure decay rates with drive off (). If rates differ from free space → quantized-field dressing (cavity coupling). If rates match free space → classical dressing (AC Stark shift only).
Both usages are acceptable in modern literature, provided the context (quantized vs. classical field) is made explicit.
3.4 Experimental Signatures of Dressed States [CORE + PRACTICE]
[CORE + PRACTICE]Dressed states are not merely mathematical constructs—they produce measurable physical consequences:
3.4.1 Autler–Townes splitting [AUT55]
In probe transmission spectroscopy, a weak probe field reveals two absorption peaks separated by the dressed-state splitting:
At resonance (), symmetric splitting by .
3.4.2 Mollow triplet [MOL69]
In resonance fluorescence, the emitted spectrum shows three peaks:
Central peak at the laser frequency (elastic scattering, Rayleigh component)
Sidebands at (inelastic, dressed-state transitions)
The Rabi frequency is frame-independent. Intensity ratios reveal drive strength and detuning .
Notation note: Here we reference frequencies to the laser; some texts use the bare transition , yielding central peak at and sidebands at in the lab frame. The splitting is the same in both conventions.
Visual representation: Mollow triplet structure
Figure 3.2: Mollow triplet in resonance fluorescence spectrum. The central Rayleigh peak appears at the laser frequency , with sidebands symmetrically displaced by . The splitting is directly observable and provides independent verification of the dressed-state structure. Relative peak intensities depend on detuning and drive strength, enabling spectroscopic determination of system parameters.
3.4.3 Avoided crossings
Sweeping parameters (e.g., varying while monitoring transmission) reveals anticrossings with minimum gap at —direct evidence of level hybridization (see Figure 3.1).
3.4.4 Modified spontaneous emission
In the dressed basis, decay rates differ from bare-state values:
where accounts for different dipole matrix elements in the dressed basis. Observable via fluorescence decay curves [COH92].
Operational use: These signatures provide independent verification of RWA validity and dressed-state structure without requiring full quantum state tomography.
Part II: Extensions
4. Floquet-Theoretic Framework [ADVANCED]
[ADVANCED]For periodic drives with period , Floquet's theorem [SHI65] guarantees solutions of the form:
where are quasienergies (defined modulo ) and are Floquet modes (time-periodic states).
4.1 Connection to rotating frame and RWA
Rotating frame shifting to a basis where the Floquet Hamiltonian becomes simpler (removing the dominant oscillation).
RWA keeping only the zeroth Fourier block in the Floquet expansion—equivalent to the lowest-order Magnus or Van Vleck expansion [GOL14, BUK15].
Dressed energies quasienergies in the RWA limit:
Beyond-RWA corrections (Bloch-Siegert shift, multiphoton resonances) appear systematically as higher Fourier blocks in the Floquet spectrum. See Appendix A for perturbative expansions.
Connection to §2: In this sense, the RWA procedure described in §2 can be understood as keeping only the lowest (zeroth) Fourier block of the full Floquet Hamiltonian—a controlled truncation that becomes exact in the limit $\Omega/\omega_0 \to 0$. Higher Fourier blocks encode the beyond-RWA corrections (Bloch-Siegert shifts, multiphoton processes) discussed in Appendix A.
4.2 Experimental signatures of Floquet structure
Multiphoton resonances: Absorption/emission at (), revealing higher Floquet blocks [CHI17]
Sideband spectroscopy: Probe transmission shows peaks at , mapping the full Floquet ladder [UND16]
Photon-assisted tunneling: In coupled systems, drive-induced transitions at energy mismatch [PLU04]
Stroboscopic Ramsey interferometry: Accumulated phase measures quasienergy splitting [KAS91]
Operational diagnostic: Reconstruct full Floquet spectrum via multiphoton spectroscopy; verify quasienergy shifts match Magnus expansion to stated order; confirm higher-order corrections signal RWA breakdown.
Computational scaling: Full Floquet analysis requires diagonalizing a dimensional matrix, where is the Fourier truncation order and is the bare Hilbert space dimension. For a two-level system with , this yields an matrix (cost ). For coupled multi-qubit systems, the cost scales as where is the number of qubits—making RWA attractive for unless beyond-RWA effects are critical.
Reference: For comprehensive Floquet engineering review, see [BUK15].
5. Continuous-Variable (CV) Systems [ADVANCED]
[ADVANCED]The frame/RWA/dressed-state structure extends naturally to harmonic oscillators and cavity QED systems.
5.1 Driven harmonic oscillator (classical field)
Hamiltonian:
Frame transformation: removes fast optical oscillations.
RWA: In the rotating frame, drop counter-rotating terms and that oscillate at the sum frequency when far from resonance. For linear driving (as above), there are no explicit or terms to drop.
Note for parametric driving: In systems with parametric driving or quadratic couplings (e.g., squeezing Hamiltonians of the form ), one additionally drops rapidly oscillating two-photon terms under RWA. The linear-drive case above does not contain such terms.
Dressed states: Diagonalize displaced oscillator → coherent states with displacement (where ). These are the CV analogue of two-level dressed states [WAL08].
5.2 Jaynes–Cummings model (quantized field)
Hamiltonian:
RWA: Already in RWA form (counter-rotating terms , absent).
Dressed states: Fock-state doublets (the "Jaynes–Cummings ladder")
with -dependent splitting:
Key distinction from classical case:
Vacuum Rabi splitting observable even at (zero photons)
Modified spontaneous emission (Purcell effect) persists at even without external drive
Atom-cavity entanglement structure [HAR06, RAI01]
5.3 Summary table: Three dressing mechanisms
Coherent-state dressing
Driven cavity (classical)
Not required
AC Stark shift; displacement
Fock-state dressing
Jaynes–Cummings
Essential
Vacuum Rabi splitting; Purcell effect
Floquet dressing
Driven two-level (classical)
Not required
Autler–Townes splitting; Mollow triplet
Pedagogical note: Graduate students should distinguish these three regimes—they involve different physics despite superficial formal analogies.
Part III: Practical Tools
6. Unified Summary Table [REFERENCE]
[REFERENCE]Interaction frame
Exact transform
Unitary
None (reparametrization)
Same as lab
None
per step
Populations transform; trace invariant
Rotating frame
Exact transform
Unitary at
None (reveals slow)
None added
None
per step
Fast oscillations removed
RWA
Approximation
Neglect
Restores time-trans. + U(1)
Quasienergy
(analytic)
Counter-rot. ; B-S shift OK
Dressed basis
Diagonalization
Eigen-decomp. of
Makes symm. explicit
Eigenvalues
None (exact in RWA)
one-time
No pop. oscillations in basis
Floquet (full)
Numerical diag.
Diag. matrix
Reveals all harmonics
Quasienergies
Numerical only
Multiphoton res.; B-S shift exact
Notation:
= bare Hilbert space dimension (2 for qubit, unbounded for cavity)
= Fourier truncation order in Floquet (typically 5–10)
"per step" = cost per time-evolution step in numerical integration
"one-time" = cost is amortized over entire calculation
Key insights:
Only RWA changes physics by discarding terms; others are exact operations (frames) or representational choices (basis)
RWA is computationally cheapest (analytic solutions often available)
Floquet is most expensive but captures all beyond-RWA physics systematically
For multi-qubit systems (), Floquet cost can become prohibitive, making RWA attractive even with ~1% systematic error
Practical guidance:
For gate design: Start with RWA (fast iteration), validate with Floquet (1–2 representative cases)
For spectroscopy: Use RWA for peak positions, Floquet if lineshapes show asymmetry or sidebands
For many-body systems: RWA is often the only tractable approach; validate in few-body limit
7. Guidance for Manuscript Preparation [FOR AUTHORS]
[FOR AUTHORS]7.1 Standard citation templates
Minimal (for routine RWA usage):
"We employ the rotating-wave approximation [SCU97], valid in the regime ."
Explicit (when clarity is critical):
"We transform to a rotating frame at the drive frequency (an exact unitary transformation), then apply the rotating-wave approximation by neglecting counter-rotating terms oscillating at . This yields a time-independent effective Hamiltonian with dressed states as eigenbases [SCU97, COH92]. The approximation introduces errors [BUK15], verified to be in our parameter regime ()."
For beyond-RWA work:
"We retain counter-rotating terms and employ Floquet theory [SHI65, BUK15] to compute quasienergies, capturing Bloch-Siegert shifts [GOL14]."
For quantum gate fidelity:
"We account for Bloch-Siegert phase shifts in our gate calibration [see §2.3.1]. For our parameters (, s), this contributes rad, below our phase error budget of 0.01 rad."
7.2 Common reviewer objections and responses
Objection 1: "You use 'dressed frame' but frames and dressed states are different."
Response:
"We thank the reviewer for this important clarification. We have revised the manuscript to distinguish: (i) transformation to the rotating frame (exact unitary, §2.1), (ii) application of RWA (approximation, §2.2), and (iii) diagonalization to obtain dressed eigenstates (basis change, §3.1). We now use 'dressed basis' consistently rather than 'dressed frame.'"
Objection 2: "RWA validity is not justified quantitatively."
Response:
"We have added explicit validity criteria (new §2.3): in our regime , the relative error is . We confirm this by comparing RWA predictions to full numerical Floquet calculations (new Fig. X), showing deviation in quasienergies."
Objection 3: "You invoke dressed states but don't clarify whether the field is quantized."
Response:
"We have clarified in §3.3 that our treatment uses a classical driving field (Floquet dressed states). We note that quantized-field effects (vacuum Rabi splitting, Purcell enhancement) are negligible in our bad-cavity regime [HAR06]."
Objection 4: "Phase errors from Bloch-Siegert shifts are not addressed."
Response:
"We thank the reviewer for highlighting this. We have added §2.3.1 analyzing the failure mode hierarchy. For our parameters, the Bloch-Siegert phase shift is rad, which is below our target phase fidelity of 0.01 rad. We have verified this numerically (new Fig. Y) showing agreement between analytic correction and full TDSE to within our measurement precision."
7.3 Cross-referencing this document
In your manuscript (if this becomes a published resource):
"For a detailed disambiguation of terminology, see [Reference to this document]."
In internal lab documentation:
"Standard definitions and operational diagnostics are catalogued in the Frames/RWA/Dressed States reference document (v0.4)."
References
[AUT55] Autler, S. H., & Townes, C. H. "Stark Effect in Rapidly Varying Fields." Phys. Rev. 100, 703 (1955). doi:10.1103/PhysRev.100.703
[BAL16] Ballance, C. J., et al. "High-Fidelity Quantum Logic Gates Using Trapped-Ion Hyperfine Qubits." Phys. Rev. Lett. 117, 060504 (2016). doi:10.1103/PhysRevLett.117.060504
[BUK15] Bukov, M., D'Alessio, L., & Polkovnikov, A. "Universal High-Frequency Behavior of Periodically Driven Systems." Adv. Phys. 64, 139 (2015). doi:10.1080/00018732.2015.1055918
[CHI17] Choi, S., et al. "Observation of Discrete Time-Crystalline Order." Nature 543, 221 (2017). doi:10.1038/nature21426
[COH92] Cohen-Tannoudji, C., Dupont-Roc, J., & Grynberg, G. Atom–Photon Interactions. Wiley-VCH (1992). ISBN: 978-0471625564
[FEL14] Felicetti, S., et al. "Dynamical Casimir Effect Entangles Artificial Atoms." Phys. Rev. Lett. 113, 093602 (2014). doi:10.1103/PhysRevLett.113.093602
[GOL14] Goldman, N., & Dalibard, J. "Periodically Driven Quantum Systems." Phys. Rev. X 4, 031027 (2014). doi:10.1103/PhysRevX.4.031027
[GOT09] Gottesman, D. "An Introduction to Quantum Error Correction and Fault-Tolerant Quantum Computation." Proc. Symp. Appl. Math. 68, 13 (2009). arXiv:0904.2557
[HAE68] Haeberlen, U., & Waugh, J. S. "Coherent Averaging Effects in Magnetic Resonance." Phys. Rev. 175, 453 (1968). doi:10.1103/PhysRev.175.453
[HAR06] Haroche, S., & Raimond, J.-M. Exploring the Quantum. Oxford (2006). ISBN: 978-0198509141
[IRI07] Irish, E. K. "Generalized Rotating-Wave Approximation." Phys. Rev. Lett. 99, 173601 (2007). doi:10.1103/PhysRevLett.99.173601
[KAS91] Kasevich, M., & Chu, S. "Atomic Interferometry Using Stimulated Raman Transitions." Phys. Rev. Lett. 67, 181 (1991). doi:10.1103/PhysRevLett.67.181
[MOL69] Mollow, B. R. "Power Spectrum of Light Scattered by Two-Level Systems." Phys. Rev. 188, 1969 (1969). doi:10.1103/PhysRev.188.1969
[PLU04] Platero, G., & Aguado, R. "Photon-Assisted Transport in Semiconductor Nanostructures." Phys. Rep. 395, 1 (2004). doi:10.1016/j.physrep.2004.01.004
[RAI01] Raimond, J. M., Brune, M., & Haroche, S. "Manipulating Quantum Entanglement with Atoms and Photons." Rev. Mod. Phys. 73, 565 (2001). doi:10.1103/RevModPhys.73.565
[SCU97] Scully, M. O., & Zubairy, M. S. Quantum Optics. Cambridge (1997). ISBN: 978-0521435956
[SHI65] Shirley, J. H. "Solution of the Schrödinger Equation with a Hamiltonian Periodic in Time." Phys. Rev. 138, B979 (1965). doi:10.1103/PhysRev.138.B979
[SHO11] Shore, B. W. Manipulating Quantum Structures Using Laser Pulses. Cambridge (2011). ISBN: 978-0521763578
[UND16] Underwood, D. L., et al. "Imaging Photon Lattice States." Phys. Rev. X 6, 021044 (2016). doi:10.1103/PhysRevX.6.021044
[WAL08] Walls, D. F., & Milburn, G. J. Quantum Optics, 2nd ed. Springer (2008). ISBN: 978-3540285731
[WIM94] Wimperis, S. "Broadband, Narrowband, and Passband Composite Pulses for Use in Advanced NMR Experiments." J. Magn. Reson. A 109, 221 (1994). doi:10.1006/jmra.1994.1159
[WU19] Wu, H., et al. "Floquet Time Crystals in Clock Models." Phys. Rev. B 100, 024310 (2019). doi:10.1103/PhysRevB.100.024310
Reference Label Convention
Citations use [FIRST_AUTHOR_SURNAME + YEAR] format:
Single author: [SHI65] = Shirley (1965)
Multiple authors: [BUK15] = Bukov et al. (2015)
Hyphenated names: Use first part (e.g., Cohen-Tannoudji → [COH92])
All in-text citations match the reference list exactly. Year format: 2-digit (65, 97, 15) for consistency.
Appendix A: Beyond-RWA Systematic Corrections [ADVANCED OPTIONAL]
[ADVANCED OPTIONAL]A.1 Magnus Expansion
The time-evolution operator for periodic can be written as:
where admits a Magnus expansion [GOL14]:
= RWA Hamiltonian (time-averaged zeroth order)
(for symmetric drives)
(Bloch-Siegert shift)
Convergence: Magnus expansion converges when [BUK15].
A.2 Van Vleck Perturbation Theory
In the Floquet picture, quasi-degenerate perturbation theory [HAE68] systematically removes off-resonant couplings order-by-order, producing effective Hamiltonians in the resonant subspace with explicit corrections.
A.3 Numerical Floquet Diagonalization
For arbitrary , direct diagonalization of the Floquet Hamiltonian (truncated at finite Fourier order ) provides exact quasienergies.
Convergence diagnostic: Monitor quasienergy shifts as increases. Typically suffices for [BUK15].
Computational cost: Scales as where is the bare Hilbert space dimension.
Recommended Use: Reference standard for graduate instruction, manuscript preparation, and peer review in AMO physics, cavity QED, trapped ions, superconducting qubits, and quantum control
Last updated