5. Exponential Integrators and Magnus Expansions


5.1 Symplectic Runge–Kutta Methods

Symplectic Runge–Kutta (SRK) methods form the second main family of geometric integrators after splitting methods. They act directly on the system of ODEs instead of decomposing the vector field. Their key virtues:

The classical example: the Gauss–Legendre collocation integrators, which are implicit, A-stable, symmetric, symplectic, and of order 2s for s stages.


5.1.1 Hamiltonian ODEs and Symplecticity Conditions

Consider an autonomous Hamiltonian system on \(\mathbb{R}^{2m}\):

\[ \dot{y} = J^{-1} \nabla H(y), \qquad J = \begin{pmatrix} 0&-I\\ I&0 \end{pmatrix}. \]

A Runge–Kutta method with Butcher tableau:

c_i a_{ij}
  b_1 b_2 b_3
Figure 5.1 – Generic Runge–Kutta tableau.

is symplectic if and only if:

\[ b_i a_{ij} + b_j a_{ji} - b_i b_j = 0 \qquad \text{for all } i,j. \tag{5.1.1} \]

This set of bilinear conditions characterises exactly those RK methods that preserve the canonical two-form \(\omega = dq \wedge dp\).


5.1.2 Gauss–Legendre Collocation Methods

The Gauss–Legendre nodes c_i in [0,1] are roots of:

\[ P_s(2c-1)=0, \]

where P_s is the Legendre polynomial of degree s.

The associated collocation method has:

The Butcher coefficients are:

\[ a_{ij} = \int_0^{c_i} \ell_j(\tau)\, d\tau,\qquad b_j = \int_0^1 \ell_j(\tau)\, d\tau, \]

where \(\ell_j\) are the Lagrange basis polynomials at the Gauss nodes.

These are the highest-order (per stage) RK methods known, and are optimal for many Hamiltonian problems.


5.1.3 Variational Derivation of Symplectic RK Methods

SRK methods arise naturally from a discrete variational principle. For a Lagrangian L(q,\dot{q}):

\[ S[q] = \int_0^h L(q,\dot q)\,dt. \]

Approximate the trajectory inside each timestep by a polynomial interpolant defined at stage points:

\[ q(t) \approx \sum_{j=1}^s q_j \ell_j(t/h). \]

Then extremise the discrete action:

\[ S_h(q_1,\dots,q_s) = h \sum_{i=1}^s b_i\, L(q_i, \dot{q}_i). \]

The resulting discrete Euler–Lagrange equations produce exactly the SRK update.

This variational origin guarantees symplecticity, momentum conservation, and excellent energy behaviour.


5.1.4 Example: Gauss–2 (Order 4) Symplectic Integrator

For s=2, the Gauss nodes are:

\[ c_{1,2} = \tfrac{1}{2} \mp \tfrac{\sqrt{3}}{6}. \]

The resulting method has:

This is one of the best all-purpose geometric ODE solvers available.


5.1.5 Phase Accuracy and Long-Time Behaviour

A hallmark of Gauss collocation methods: their superior phase accuracy for Hamiltonian oscillatory systems. For a harmonic oscillator:

\[ y' = J^{-1} \nabla \tfrac{1}{2}(q^2+p^2), \]

the Gauss-2 method yields:


5.1.6 Interactive Illustration: Gauss-2 vs Explicit RK4 for Harmonic Oscillator

Comparison of energy drift:

Figure 5.2 – Gauss-2 (blue) shows bounded energy oscillations; explicit RK4 (red) accumulates drift.

5.1.7 SRK Methods for Stiff and Semi-Stiff Hamiltonians

Gauss methods are algebraically stable and A-stable. Thus they handle:

They are often the preferred integrator when:


5.1.8 Summary


5.1.9 References for Section 5.1

  1. Hairer–Lubich–Wanner (2006). Geometric Numerical Integration.
  2. Blanes & Casas (2026). Concise Introduction to GNI.
  3. Butcher (2003). Numerical Methods for ODEs.
  4. Sanz-Serna (1988). “Runge–Kutta schemes for Hamiltonian systems”.
  5. Suris (1990–2020). Foundations of variational integrators.

5.2 Algebraic Order Conditions and B-Series for Symplectic RK Methods

Runge–Kutta (RK) methods admit an algebraic representation of their action on differential equations through B-series. These are formal expansions indexed by rooted trees, which encode precisely how an RK method approximates the Taylor expansion of the exact flow.

This section introduces the B-series formalism, describes order conditions, and explains why symplectic Runge–Kutta methods naturally restrict to symplectic B-series, leading to:


5.2.1 Rooted Trees and Elementary Differentials

For an autonomous ODE \(\dot{y}=f(y)\), the Taylor expansion of the exact flow can be written using rooted trees.

\[ y(h) = y(0) + h f + \frac{h^2}{2} f'f + \frac{h^3}{6} \big(f''(f,f) + f'(f'f)\big) + \cdots. \]

Each term corresponds to a rooted tree. For example:

τ₁: order 1 τ₂: order 2 τ₃: order 3
Figure 5.3 – Typical rooted trees.

For a tree \(\tau\), the associated elementary differential is denoted

\[ F(\tau)(y) \in \mathbb{R}^m. \]

The exact flow has a B-series:

\[ \phi_h(y) = y + \sum_{\tau\in\mathcal{T}} \frac{h^{|\tau|}}{\sigma(\tau)} F(\tau)(y). \tag{5.2.1} \]


5.2.2 B-Series of a Runge–Kutta Method

Any (consistent) RK method produces a B-series:

\[ \Psi_h(y) = y + \sum_{\tau\in\mathcal{T}} h^{|\tau|}\alpha(\tau)F(\tau)(y), \tag{5.2.2} \]

where \(\alpha(\tau)\) are algebraic expressions in a_{ij},b_i,c_i.

The method has order p if and only if:

\[ \alpha(\tau) = \frac{1}{\sigma(\tau)} \qquad \text{for all } |\tau|\le p. \tag{5.2.3} \]


5.2.3 The Butcher Group

B-series form a group under composition: the Butcher group \(G_B\).

Key facts:

Gauss collocation methods correspond to the exponential of the truncated pre-Lie product associated with the vector field.


5.2.4 Symplectic RK Methods and Symplectic B-Series

For Hamiltonian ODEs \(\dot{y}=J^{-1}\nabla H(y)\), the exact B-series satisfies:

\[ \phi_h^*(\omega) = \omega, \]

where \(\omega\) is the symplectic form.

An RK B-series is symplectic if and only if the Butcher coefficients satisfy:

\[ b_i a_{ij} + b_j a_{ji} - b_i b_j = 0, \tag{5.1.1 revisited} \]

These constraints force the B-series coefficients into the Hamiltonian sub-group of the Butcher group.

Consequences:


5.2.5 Why Gauss–Legendre RK Has Order 2s

A Gauss–Legendre collocation method with s stages satisfies all order conditions up to order 2s. This follows because:

  1. Collocation enforces the differential equation exactly on the space of interpolating polynomials of degree 2s-1.
  2. Quadrature is exact for polynomials up to degree 2s-1.
  3. The trees correspond to compositions of derivatives of degree up to 2s.

This gives the celebrated order-barrier optimality result:

\[ \text{For any s-stage RK method: } \quad p \le 2s. \]

Gauss methods achieve this bound with equality.


5.2.6 Interpretation via Pre-Lie Algebra of Rooted Trees

Rooted trees form a pre-Lie algebra:

\[ \tau_1 \triangleright \tau_2 = \text{sum over graftings of the root of } \tau_1 \text{ onto each node of } \tau_2. \]

The exact solution is the Lie exponential of the vector field in this algebra:

\[ \phi_h = \exp_{\triangleright}(h f). \]

Gauss methods are the truncated exponential of this pre-Lie product, preserving symplecticity by restriction to the Hamiltonian subalgebra.

This provides the deepest algebraic explanation for their optimal order, symmetry, and geometric behaviour.


5.2.7 Summary


5.2.8 References for Section 5.2

  1. Hairer–Lubich–Wanner (2006). Geometric Numerical Integration.
  2. Butcher (1972–2003). Tree theory and B-series.
  3. Munthe-Kaas & Wright (2008–2020). Pre-Lie algebra and Lie–Butcher series.
  4. Chartier, Hairer, Vilmart (2010–2023). Symplectic B-series integrators.
  5. Iserles, Quispel, Nørsett (1999). “Lie-group methods and B-series”.

5.3 Symmetric, Symplectic, and Energy-Preserving RK Methods

Symmetry (time reversibility), symplecticity, and energy conservation are three of the most important geometric properties for numerical integrators of Hamiltonian systems. While no Runge–Kutta method can preserve all invariants exactly, certain RK families preserve the symplectic form, quadratic invariants, and/or time-reversal symmetry, which in turn guarantees excellent long-time behaviour.

This section develops:


5.3.1 Time Symmetry (Self-Adjoint RK Methods)

A one-step method \(\Phi_h\) is symmetric (or time reversible) if:

\[ \Phi_{-h} \circ \Phi_h = \mathrm{id}. \tag{5.3.1} \]

For RK methods with tableau \((A,b,c)\), symmetry requires:

\[ c_i = 1 - c_{s+1-i}, \qquad b_i = b_{s+1-i}, \qquad a_{ij} = b_{j} - a_{s+1-i,s+1-j}. \tag{5.3.2} \]

Examples:

Key fact:

A symmetric method automatically has even order.


5.3.2 Symmetric + Symplectic → Excellent Long-Time Behaviour

When an RK method is both symplectic and symmetric:

This is why Gauss methods are the gold standard:

\[ \text{Gauss–Legendre RK} \quad \text{= symmetric + symplectic + order } 2s. \]

This combination yields qualitatively correct behaviour over astronomical time spans.


5.3.3 Backward Error Analysis: Modified Hamiltonians

A symplectic RK method applied to \(\dot{y}=J^{-1}\nabla H(y)\) exactly solves a nearby Hamiltonian system:

\[ \dot{y} = J^{-1}\nabla \tilde{H}(y), \]

where the modified Hamiltonian has a formal expansion:

\[ \tilde{H} = H + h^p H_{p+1} + h^{p+2} H_{p+3} + \cdots. \tag{5.3.3} \]

Key consequences:

H = constant ℍ̃ = constant
Figure 5.4 – Symplectic RK follows modified Hamiltonian level curves.

5.3.4 Energy-Preserving Runge–Kutta Methods

No ordinary RK method (explicit or implicit) can preserve a general Hamiltonian exactly. But several specialized families achieve exact energy conservation by modifying the discretisation:

• The Average Vector Field (AVF) Method

\[ \frac{y_{n+1}-y_n}{h} = \int_0^1 f\big(\theta y_{n+1} + (1-\theta)y_n\big)\, d\theta. \]

For Hamiltonian systems, AVF preserves energy exactly:

\[ H(y_{n+1}) = H(y_n). \]

• Discrete Gradient Methods

Choose a discrete gradient \(\bar{\nabla}H(y_{n+1},y_n)\) satisfying:

\[ H(y_{n+1}) - H(y_n) = \bar{\nabla}H(y_{n+1},y_n)^{\!\top} (y_{n+1}-y_n). \tag{5.3.4} \]

Then define an update:

\[ \frac{y_{n+1} - y_n}{h} = J^{-1} \bar{\nabla}H(y_{n+1},y_n). \]

This preserves energy exactly for any Hamiltonian.

Energy-preserving ≠ Symplectic

There is a fundamental incompatibility: an RK method cannot be both symplectic and energy-preserving unless the Hamiltonian is quadratic.

Thus one chooses:


5.3.5 Comparison Chart: Symmetry, Symplecticity, Energy Preservation

Method Symplectic? Symmetric? Exact Energy? Order
Gauss–s Yes Yes No 2s
Radau IIA No No No 2s-1
Implicit midpoint Yes Yes No 2
AVF / DG Methods No Depends Yes 2
Table 5.1 – Comparison of geometric properties of key RK-type methods.

5.3.6 Summary


5.3.7 References for Section 5.3

  1. Hairer–Lubich–Wanner (2006). Geometric Numerical Integration.
  2. Blanes & Casas (2026). Concise Introduction to GNI.
  3. Sanz-Serna & Calvo (1994). Numerical Hamiltonian Problems.
  4. Quispel, Turner (2010–2021). Discrete gradient methods.
  5. Gonzalez (1996–2000). Average Vector Field method and variants.

5.4 Partitioned and Symplectic Partitioned Runge–Kutta Methods

Many important ODEs have a natural partitioned structure. Examples include:

For such systems, Partitioned Runge–Kutta (PRK) methods assign different RK tableaux to each component, while maintaining a coupled update.

When the system is Hamiltonian, special algebraic conditions give rise to Symplectic PRK (SPRK) methods.


5.4.1 Partitioned ODEs and PRK Schemes

Consider a partitioned system:

\[ \begin{aligned} q' &= f(q,p), \\ p' &= g(q,p). \end{aligned} \tag{5.4.1} \]

A partitioned Runge–Kutta method is defined by two tableaux:

A
c
b
Figure 5.5 – A general Partitioned Runge–Kutta pair.

The stage values satisfy:

\[ \begin{aligned} Q_i &= q_n + h\sum_{j=1}^s a_{ij}\,f(Q_j,P_j), \\ P_i &= p_n + h\sum_{j=1}^s \tilde{a}_{ij}\,g(Q_j,P_j), \end{aligned} \tag{5.4.2} \]

and the step update is:

\[ \begin{aligned} q_{n+1} &= q_n + h\sum_{i=1}^s b_i\,f(Q_i,P_i), \\ p_{n+1} &= p_n + h\sum_{i=1}^s \tilde{b}_i\,g(Q_i,P_i). \end{aligned} \tag{5.4.3} \]


5.4.2 Hamiltonian Structure and SPRK Conditions

For a Hamiltonian system in canonical form:

\[ q' = +\nabla_p H(p,q), \qquad p' = -\nabla_q H(p,q), \tag{5.4.4} \]

the exact flow is symplectic:

\[ \Phi_h^*(dq \wedge dp) = dq \wedge dp. \]

A partitioned RK method is symplectic if and only if the coefficients satisfy:

\[ b_i \tilde{a}_{ij} + \tilde{b}_j a_{ji} = b_i \tilde{b}_j, \qquad 1\le i,j\le s. \tag{5.4.5} \]

This is the symplectic PRK condition.

Special cases:


5.4.3 Separable Hamiltonians and the SPRK Midpoint Family

For separable systems:

\[ H(p,q) = T(p) + V(q), \tag{5.4.6} \]

the ODE is:

\[ q' = +T_p(p), \qquad p' = -V_q(q). \tag{5.4.7} \]

Since the two components depend only on one variable each, SPRK methods simplify dramatically.

• The Symplectic Euler A/B pair

Two 1-stage PRK methods:

Their compositions give:

Verlet is the most widely used SPRK of order 2.


5.4.4 The Canonical Example: Stoermer–Verlet as a SPRK Method

Taking the tableaux:

\[ A = \begin{pmatrix}0\end{pmatrix},\quad b=\begin{pmatrix}1\end{pmatrix}, \qquad \tilde{A} = \begin{pmatrix}\frac12\end{pmatrix},\quad \tilde{b}=\begin{pmatrix}1\end{pmatrix}, \tag{5.4.8} \]

yields the scheme:

\[ \begin{aligned} p_{n+\frac12} &= p_n - \tfrac{h}{2} \nabla V(q_n),\\ q_{n+1} &= q_n + h\, \nabla_p T(p_{n+\frac12}),\\ p_{n+1} &= p_{n+\frac12} - \tfrac{h}{2} \nabla V(q_{n+1}). \end{aligned} \tag{5.4.9} \]

This is exactly the Stoermer–Verlet integrator, which is:


5.4.5 Block Structure, Trees, and Order Conditions

PRK order conditions are described by two-coloured rooted trees:

Two-coloured rooted tree
Figure 5.6 – A typical order-3 PRK rooted tree.

Exactly as in the classical B-series case:

\[ \Psi_h = y + \sum_{\tau\in \mathcal{T}_{2}} h^{|\tau|}\,\alpha(\tau)\,F(\tau), \tag{5.4.10} \]

where \(\mathcal{T}_{2}\) is the set of two-coloured rooted trees.

Using coloured trees splits the order conditions into mixed and diagonal classes, allowing high-order symmetric & symplectic PRK methods to be constructed systematically.


5.4.6 Higher-Order SPRK Methods

Examples include:

A typical high-order symmetric composition:

\[ \Psi_h^{(4)} = \Phi_{\gamma_1 h}\circ \Phi_{\gamma_2 h}\circ \Phi_{\gamma_1 h}, \tag{5.4.11} \]

with \(\gamma_1 = 1/(2-2^{1/3})\), \(\gamma_2 = -2^{1/3}\gamma_1\).

Such compositions retain symplecticity and symmetry by construction.


5.4.7 Summary


5.4.8 References for Section 5.4

  1. Hairer–Lubich–Wanner (2006). Geometric Numerical Integration.
  2. Blanes & Casas (2026). Concise Introduction to GNI.
  3. Sanz-Serna & Calvo (1994). Numerical Hamiltonian Problems.
  4. McLachlan & Quispel (2002–2023). Survey of symplectic splitting & SPRK methods.
  5. Laskar & Robutel (2001). SABA/SBAB integrators.

5.5 Exponential Runge–Kutta (ERK), Lawson, and Integrating Factor Methods

Many evolution equations have the semilinear structure:

\[ y' = Ly + N(y), \qquad y(0)=y_0 \tag{5.5.1} \]

where:

Typical examples include:

In such cases, classical RK methods suffer order reduction and stability problems. Exponential Runge–Kutta (ERK) and integrating factor / Lawson methods resolve this by treating the stiff linear part exactly.


5.5.1 The Integrating Factor Transformation

The exact flow of the linear part is:

\[ \phi_L^h(y) = e^{hL}y. \]

Introduce the change of variables:

\[ v(t) = e^{-tL} y(t). \tag{5.5.2} \]

Then:

\[ v' = e^{-tL}N\!\big(e^{tL}v\big), \tag{5.5.3} \]

which is nonstiff if N is nonstiff. Ordinary RK applied to v'(t) gives the classical Lawson method.

This is the foundation of all exponential integrators.


5.5.2 Lawson Methods (Explicit Exponential RK)

Apply an s-stage RK method to (5.5.3):

\[ v_{n+1} = v_n + h\sum_{i=1}^s b_i\,e^{-c_i h L} N\!\big(e^{c_i h L} V_i\big), \tag{5.5.4} \]

with stages:

\[ V_i = v_n + h\sum_{j=1}^s a_{ij}\,e^{-c_j h L} N\!\big(e^{c_j hL} V_j\big). \tag{5.5.5} \]

Return to y via:

\[ y_{n+1} = e^{hL} v_{n+1}. \tag{5.5.6} \]

Advantages:


5.5.3 Exponential Runge–Kutta (ERK): ϕ-Functions

ERK methods express the solution using the variation-of-constants formula:

\[ y(t+h)=e^{hL}y(t) +\int_0^h e^{(h-\tau)L} N(y(t+\tau))\,d\tau. \tag{5.5.7} \]

This integral naturally leads to the phi functions:

\[ \varphi_0(z)=e^z,\qquad \varphi_1(z)=\frac{e^z-1}{z},\qquad \varphi_2(z)=\frac{e^z - z - 1}{z^2}, \dots \tag{5.5.8} \]

Higher-order ERK methods are linear combinations of \(\varphi_k(hL)\) acting on nonlinear terms.

• General s-stage ERK Method

\[ Y_i = e^{c_i h L} y_n + h\sum_{j=1}^s a_{ij}(hL)\,N(Y_j), \tag{5.5.9} \]

and

\[ y_{n+1} = e^{hL}y_n + h\sum_{i=1}^s b_i(hL)\,N(Y_i), \tag{5.5.10} \]

where the coefficients are matrix functions:

\[ a_{ij}(hL) = \int_0^{c_i}e^{(c_i-\tau)hL}\ell_j(\tau)\,d\tau, \tag{5.5.11} \]

and similarly for b_i(hL).


5.5.4 Geometric Structure of Exponential Integrators

If the linear part L generates a symmetry group of the system, exponential integrators inherit invariances automatically:

• L skew-symmetric → Norm/energy preservation

\[ L^T = -L \;\Longrightarrow\; \|e^{hL}y\|=\|y\|. \]

• L Hamiltonian → Symplecticity of ehL

If L is a Hamiltonian matrix:

\[ e^{hL} \in \mathrm{Sp}(2n). \tag{5.5.12} \]

Then exponential integrators preserve linear symplectic structure automatically.

• L representing a Lie algebra action → ERK respects group symmetry

Example: rotation, Schrödinger flow, Maxwell–Bloch dynamics.


5.5.5 Stability: Why ERK Beats Classical RK

Classical RK stability ERK stability (nearly optimal)
Figure 5.7 – ERK inherits stability from exact linear flow.

Because ERK treats the stiff part exactly, its stability region includes the entire left half-plane for linear test problems y' = \lambda y with stiff \(\lambda\). This eliminates order reduction for parabolic PDEs.


5.5.6 Relation to Magnus and Splitting Methods

ERK is closely linked with:

For stiff linear operators, ERK methods are effectively splitting methods with exact L-flow and approximated nonlinear flow.


5.5.7 Summary


5.5.8 References for Section 5.5

  1. Hochbruck & Ostermann (2010), Exponential Integrators.
  2. Blanes & Casas (2026), Concise Intro to GNI.
  3. Lawson (1967), original integrating factor RK formulation.
  4. Cox & Matthews (2002), ETD–RK methods.
  5. Iserles (2008–2015), Lawson–Magnus & geometric exponential methods.

5.6 Magnus–Runge–Kutta, Commutator-Free Magnus Methods, and Lie–Algebraic Exponential Integrators

For non-autonomous linear systems

\[ y'(t) = A(t)\, y(t), \tag{5.6.1} \]

the exact solution is given by a time-ordered exponential:

\[ y(t+h) = \mathcal{T} \exp\!\left(\int_t^{t+h}A(\tau)\,d\tau\right)y(t), \tag{5.6.2} \]

which lies in the Lie group generated by the Lie algebra containing A(t). This structure must be preserved for:

This section develops three families of integrators that preserve the Lie-group geometry:

  1. Magnus integrators — exact exponential of a series of commutators.
  2. Commutator-free Magnus (CFM) methods — high-order but exponential-only.
  3. Magnus–Runge–Kutta hybrids — nonlinear / semilinear geometric integrators.

5.6.1 The Magnus Expansion

Magnus (1954) showed that the solution of (5.6.1) can be written exactly as:

\[ y(t+h) = \exp\!\left(\Omega(t,h)\right)y(t), \tag{5.6.3} \]

where

\[ \Omega(t,h)=\int_t^{t+h}A(\tau)d\tau - \frac12 \int_t^{t+h}\!\!\int_t^{\tau_1} [A(\tau_1),A(\tau_2)]\,d\tau_2 d\tau_1 + \cdots. \tag{5.6.4} \]

Properties:

Practical truncation:

The 2nd-order Magnus method:

\[ \Omega^{[2]} = h\, A(t+\tfrac{h}{2}). \tag{5.6.5} \]

The 4th-order Magnus method:

\[ \Omega^{[4]} = h\Big(\tfrac12(A_1+A_2) - \tfrac{\sqrt{3}}{12}h [A_2, A_1]\Big), \tag{5.6.6} \]

where A_1, A_2 are samples at Gauss–Legendre nodes.


5.6.2 Commutator-Free Magnus (CFM) Methods

Standard Magnus integrators require evaluating matrix commutators [A_i, A_j] and nested commutators, which can be expensive.

CFM integrators avoid commutators entirely: instead of \(\Omega\) being a sum of Lie brackets, we use sums of exponentials:

\[ y_{n+1} = \exp\!\left(h\sum_k \alpha_k A(t_n + c_k h)\right) y_n, \tag{5.6.7} \]

with coefficients chosen to match the Magnus expansion up to order p.

Advantages:

Example: 4th-order CFM

\[ y_{n+1} = \exp\!\big(h(\tfrac12 A_1 + \tfrac12 A_2)\big)\; y_n, \tag{5.6.8} \]

with A_1, A_2 chosen at Gauss nodes. No commutators appear, yet the method is 4th order.


5.6.3 Magnus–Runge–Kutta Methods for Nonlinear Systems

For semilinear systems:

\[ y' = A(t) y + N(y,t), \tag{5.6.9} \]

one can combine a Magnus integrator for the linear part with a Runge–Kutta method for the nonlinear part:

\[ y_{n+1} = \exp\!\left(\Omega(t_n,h)\right) \left( y_n + h\sum_{i=1}^s b_i\, \tilde{N}_i \right), \tag{5.6.10} \]

where \(\tilde{N}_i\) are nonlinear stages evaluated in a transformed frame:

\[ \tilde{N}_i = N\Big( \exp(\Omega_i)\big(y_n + \cdots\big),\, t_n + c_i h \Big). \tag{5.6.11} \]

These hybrid methods:


5.6.4 High-Order Commutator-Free Magnus–RK Methods

Blanes, Casas, Iserles, and others developed high-order methods combining:

A typical 6th-order CFM–RK method uses:

\[ y_{n+1} = \exp\!\left( h(\alpha_1 A_1 + \alpha_2 A_2 + \alpha_3 A_3) \right) \Big( y_n + h\sum b_i N(Y_i) \Big). \tag{5.6.12} \]

All coefficients arise from matching the Magnus series through order 6, but no commutators are used.


5.6.5 Preservation of Lie-Group Structure

If the exact flow satisfies:

\[ y(t) \in G, \qquad G = \exp(\mathfrak{g}), \]

then Magnus and CFM–RK methods ensure:

\[ y_{n+1} = \exp(\xi_n) y_n \in G, \tag{5.6.13} \]

because \(\xi_n \in \mathfrak{g}\) for all steps.

Thus:


5.6.6 Summary


5.6.7 References for Section 5.6

  1. Blanes, Casas, Oteo, & Ros (2009), The Magnus Expansion.
  2. Iserles (2008–2022), Lie-group Methods and Magnus Integrators.
  3. Blanes & Casas (2026), Concise Intro to GNI.
  4. Hochbruck–Ostermann (2010), Exponential Integrators: Review.
  5. Alvermann & Fehske (2011), Commutator-free Magnus methods for large quantum systems.