6. Backward Error Analysis (BEA)


6.0 Overview: Why Splitting & Composition Methods?

In the previous chapters we developed geometric structure, symplecticity, Lie–algebraic flows, exponential integrators, and stability theory. We now introduce one of the most powerful and widely used classes of geometric integrators: splitting and composition methods.

These methods exploit the fact that many differential equations can be written as the sum of simpler vector fields whose individual flows are exactly integrable or preserve essential geometric properties.


6.0.1 The Basic Idea

Consider an evolution equation

\[ y' = (B + C)(y), \]

where the flows of the subfields B and C are individually computable:

\[ \Phi_B^t = e^{tB}, \qquad \Phi_C^t = e^{tC}. \]

Splitting methods approximate the combined flow by compositions of these simpler flows, e.g.

\[ e^{h(B+C)} \approx e^{hB} e^{hC}, \]

or symmetric/higher-order variants. Since each exponential preserves geometric structure (e.g. symplectic, orthogonal, or unitary), the composition inherits these invariants automatically.


6.0.2 Why Splitting? Geometry and Efficiency

Advantages of splitting approaches:


6.0.3 Compositions: From First-Order to Arbitrarily High Order

Starting from a basic (possibly low-order) splitting operator

\[ \mathcal{S}(h) = e^{hB}e^{hC}, \]

one constructs higher-order methods by nesting compositions:

\[ \mathcal{S}(a_1 h)\,\mathcal{S}(a_2 h)\cdots \mathcal{S}(a_m h). \]

Carefully chosen coefficients eliminate lower-order error terms in the Baker–Campbell–Hausdorff expansion, yielding high-order integrators with minimal additional computational cost.

Examples include:


6.0.4 Connections to Other Areas

Splitting methods reside at the intersection of:

Thus splitting and composition methods form a unifying framework for geometric time integration across physics, chemistry, applied mathematics, and computational engineering.


6.0.5 Outline of Chapter 6

This chapter develops the foundational splitting methods:

  1. Lie–Trotter Splitting: first-order, non-symmetric.
  2. Strang Splitting: symmetric second-order.
  3. Higher-Order Composition Schemes: Yoshida, Suzuki, complex-coefficient.
  4. Applications to Hamiltonian Flows: separable H = T + V, MD, celestial mechanics.
  5. Applications to PDEs: Schrödinger, heat, NLS, Klein–Gordon, Maxwell.
  6. Relation to Magnus and Exponential Integrators: BCH and high-order commutator structure.
  7. Backward Error Analysis for splitting.
  8. Modern Applications: HMC, TEBD, quantum computing gate sequences.

Splitting is thus the backbone of a huge class of geometric integrators.

6.1 Lie and Strang Splitting

In this section we study the two fundamental splitting schemes:

  1. Lie–Trotter splitting (first order),
  2. Strang splitting (second order, symmetric).

We will:

  1. Set up the abstract evolution problem and the notion of splitting,
  2. Define Lie–Trotter splitting and derive its local error using BCH,
  3. Define Strang splitting and prove it is second order and time-reversible,
  4. Discuss geometric properties (symplecticity, reversibility, invariants),
  5. Illustrate the methods on a simple Hamiltonian example with an interactive figure.

6.1.1 Problem Setting and Operator Splitting

Let \(y(t)\) solve an autonomous ODE on a manifold \(M\):

\[ \dot{y}(t) = f(y(t)), \qquad y(0) = y_0 \in M. \]

Suppose the vector field \(f\) admits a decomposition into simpler parts

\[ f = f^{[1]} + f^{[2]}, \]

such that the flows of the subsystems

\[ \dot{y} = f^{[1]}(y), \qquad \dot{y} = f^{[2]}(y) \]

can be computed (almost) exactly or very efficiently. Denote their flows by

\[ \varphi_h^{[1]} = \exp(h F^{[1]}), \qquad \varphi_h^{[2]} = \exp(h F^{[2]}), \]

where \(F^{[j]}\) are the corresponding differential operators \(F^{[j]} = f^{[j]}\!\cdot\nabla\). The full flow is

\[ \varphi_h = \exp\!\big(h(F^{[1]} + F^{[2]})\big). \]

A splitting method approximates \(\varphi_h\) by a composition of subflows, e.g.

\[ \Phi_h = \varphi_{a_1 h}^{[1]} \circ \varphi_{b_1 h}^{[2]} \circ \cdots \circ \varphi_{a_s h}^{[1]} \circ \varphi_{b_s h}^{[2]}, \]

with suitable coefficients \(a_j, b_j\). The simplest choices give the classical Lie–Trotter and Strang splittings.

Task marker: later sections will generalize this to multiple splits \(f = f^{[1]} + \cdots + f^{[m]}\) and to composition schemes of arbitrarily high order.


6.1.2 Lie–Trotter Splitting

The Lie–Trotter splitting (or first-order splitting) uses the simple product of exponentials

\[ \Phi_h^{\mathrm{LT}} := \varphi_h^{[1]} \circ \varphi_h^{[2]} = \exp(h F^{[1]}) \exp(h F^{[2]}). \]

The corresponding one-step method is

\[ y_{n+1} = \Phi_h^{\mathrm{LT}}(y_n) = \varphi_h^{[1]}\big( \varphi_h^{[2]}(y_n)\big). \]

Local error via BCH

The Baker–Campbell–Hausdorff (BCH) formula gives

\[ \exp(h F^{[1]}) \exp(h F^{[2]}) = \exp\Big( h(F^{[1]} + F^{[2]}) + \tfrac{h^2}{2}[F^{[1]},F^{[2]}] + \mathcal{O}(h^3) \Big). \]

Thus the Lie–Trotter step is the exact flow of a modified vector field

\[ \tilde{F} = F^{[1]} + F^{[2]} + \tfrac{h}{2}[F^{[1]},F^{[2]}] + \mathcal{O}(h^2), \]

so the local truncation error is \(\mathcal{O}(h^2)\) and the method has global order 1.

Task marker: in BEA we will explicitly compute the modified Hamiltonian for Hamiltonian splittings and relate the commutator term to Poisson brackets.

Geometric properties


6.1.3 Strang Splitting (Second Order, Symmetric)

The Strang splitting (Strang 1968) is the symmetric composition

\[ \Phi_h^{\mathrm{S}} := \varphi_{h/2}^{[1]} \circ \varphi_{h}^{[2]} \circ \varphi_{h/2}^{[1]} = \exp\!\left(\tfrac{h}{2}F^{[1]}\right) \exp\!\left(h F^{[2]}\right) \exp\!\left(\tfrac{h}{2}F^{[1]}\right). \]

The corresponding scheme is:

  1. Half-step in \(f^{[1]}\): \(y^{(1)} = \varphi_{h/2}^{[1]}(y_n)\),
  2. Full step in \(f^{[2]}\): \(y^{(2)} = \varphi_{h}^{[2]}(y^{(1)})\),
  3. Half-step in \(f^{[1]}\): \(y_{n+1} = \varphi_{h/2}^{[1]}(y^{(2)})\).

Order of accuracy via BCH

Using BCH and symmetry, one can show that the odd-order error terms cancel. More concretely,

\[ \Phi_h^{\mathrm{S}} = \exp\!\Big( h(F^{[1]}+F^{[2]}) + \mathcal{O}(h^3) \Big), \]

so the local error is \(\mathcal{O}(h^3)\) and the global error is \(\mathcal{O}(h^2)\): Strang splitting has order 2.

A key general principle is:

Self-adjoint (time-symmetric) one-step methods of order ≥ 1 are always of even order.

Strang splitting is self-adjoint, hence order 2 rather than 1.

Time symmetry (reversibility)

Let \(\Phi_h^\star\) denote the adjoint method defined by

\[ \Phi_h^\star := \Phi_{-h}^{-1}. \]

A one-step method is self-adjoint (or time-symmetric) if \(\Phi_h^\star = \Phi_h\). For Strang splitting,

\[ \big(\Phi_h^{\mathrm{S}}\big)^{-1} = \varphi_{-h/2}^{[1]} \circ \varphi_{-h}^{[2]} \circ \varphi_{-h/2}^{[1]} = \Phi_{-h}^{\mathrm{S}}, \]

hence \(\Phi_h^{\mathrm{S}}\) is self-adjoint. This is crucial for good long-time behaviour, especially in reversible and Hamiltonian systems.


6.1.4 Geometric Properties of Lie and Strang Splittings

Let us summarize the main geometric features when the splitting is applied to a structured system, e.g. a Hamiltonian system with \(H = H^{[1]} + H^{[2]}\).

Task marker: in later sections we derive explicit modified Hamiltonians for separable systems \(H(q,p) = T(p) + V(q)\) under Strang splitting, and compare with standard symplectic Euler and Verlet.


6.1.5 Example: Separable Hamiltonian \(H(q,p) = T(p) + V(q)\)

A prototypical application is a separable Hamiltonian system

\[ H(q,p) = T(p) + V(q), \qquad (q,p) \in \mathbb{R}^m \times \mathbb{R}^m. \]

The Hamiltonian vector field splits as

\[ \dot{q} = \nabla_p T(p), \qquad \dot{p} = -\nabla_q V(q), \]

which we interpret as the sum of two subfields:

Their flows are explicitly available:

\[ \varphi_h^{[1]}(q,p) = \big(q + h\nabla_p T(p),\; p\big), \qquad \varphi_h^{[2]}(q,p) = \big(q,\; p - h\nabla_q V(q)\big). \]

Applying Strang splitting gives the well-known velocity Verlet / leapfrog scheme:

  1. Half kick: \( p_{n+\frac{1}{2}} = p_n - \frac{h}{2}\nabla_q V(q_n) \)
  2. Drift: \( q_{n+1} = q_n + h \nabla_p T(p_{n+\frac{1}{2}}) \)
  3. Half kick: \( p_{n+1} = p_{n+\frac{1}{2}} - \frac{h}{2}\nabla_q V(q_{n+1}) \)

This is the cornerstone integrator in molecular dynamics, celestial mechanics, and Hamiltonian Monte Carlo.

Figure 6.1 – Energy behaviour for the 1D harmonic oscillator \(H(q,p)=\tfrac{1}{2}(p^2+q^2)\): comparison of Lie–Trotter splitting and Strang splitting. The Strang (leapfrog) method shows bounded, oscillatory energy error, while Lie–Trotter shows a systematic bias.

6.1.6 Summary and References for 6.1

In this section we:

These schemes are the building blocks for higher-order composition methods and for many modern applications (Hamiltonian Monte Carlo, quantum simulation, PDE splitting, etc.). Subsequent sections generalize these ideas to multi-part splittings and higher-order factorizations.

References for Section 6.1

  1. [Str68] G. Strang, “On the construction and comparison of difference schemes,” SIAM Journal on Numerical Analysis, 5(3):506–517, 1968.
  2. [HLW06] E. Hairer, C. Lubich, G. Wanner, Geometric Numerical Integration: Structure-Preserving Algorithms for Ordinary Differential Equations, 2nd ed., Springer, 2006. (Ch. II, IV on splitting methods.)
  3. [BC26] S. Blanes, F. Casas, A Concise Introduction to Geometric Numerical Integration, 2nd ed., CRC Press, 2026. (Chapters on splitting and composition methods.)

6.2 Symmetric Composition Methods: From Strang to High Order

Strang splitting is the simplest nontrivial symmetric composition:

\[ \Phi_h^{\mathrm{S}} = \varphi^{[1]}_{h/2} \circ \varphi^{[2]}_{h} \circ \varphi^{[1]}_{h/2}, \]

which is second order because symmetry causes cancellation of all odd-order error terms in the BCH expansion. This section develops the general theory: how symmetric compositions lead to arbitrarily high-order geometric integrators.


6.2.1 General Symmetric Composition

Let \(\Psi_h\) be a basic method of order \(p\), and assume it is self-adjoint (i.e. reversible, symmetric):

\[ \Psi^{\star}_{h} = \Psi_h. \]

Consider a symmetric composition with real coefficients

\[ \Psi_h^{(m)} = \Psi_{a_1 h} \circ \Psi_{a_2 h} \circ \cdots \circ \Psi_{a_s h} \circ \Psi_{a_{s} h} \circ \cdots \Psi_{a_2 h} \circ \Psi_{a_1 h}. \tag{6.2.1} \]

Symmetry ensures the overall method is self-adjoint:

\[ \big(\Psi_h^{(m)}\big)^\star = \Psi_h^{(m)}. \]

As a consequence, all odd-order terms in the BCH expansion cancel.


6.2.2 BCH Structure and Order Conditions

Let the modified vector field of \(\Psi_h\) be

\[ \tilde{F}(h) = F + h^{p} E_{p} + h^{p+1} E_{p+1} + h^{p+2} E_{p+2} + \cdots. \tag{6.2.2} \]

Inserting scaled steps \(a_j h\) yields modified fields

\[ \tilde{F}(a_j h) = F + a_j^{p} h^{p} E_p + a_j^{p+1} h^{p+1} E_{p+1} + \cdots. \]

The BCH formula for the symmetric composition (6.2.1) has only even powers:

\[ \log\big(\Psi_h^{(m)}\big) = h F + h^{p+2} C_{p+2}(a_1,\ldots,a_s) + h^{p+4} C_{p+4}(a_1,\ldots,a_s) + \cdots. \tag{6.2.3} \]

The order conditions are the algebraic constraints on \(a_1,\ldots,a_s\) ensuring that

\[ C_{p+2} = C_{p+4} = \cdots = C_{P-1} = 0. \]

This yields an overall method of order \(P\).

Task marker: later sections will list explicit order conditions for splitting and composition up to 8th order using the Lyndon word basis for the free Lie algebra.


6.2.3 Yoshida’s Triple-Jump: Raising Order by Two

Let \(\Psi_h\) be a symmetric method of order \(2m\). Yoshida (1990) showed that the composition

\[ \Psi_h^{(m+1)} = \Psi_{\alpha h} \circ \Psi_{\beta h} \circ \Psi_{\alpha h} \tag{6.2.4} \]

is a symmetric method of order \(2m + 2\) provided \(\alpha, \beta\) satisfy

\[ 2\alpha + \beta = 1, \qquad 2\alpha^{2m+1} + \beta^{2m+1} = 0. \tag{6.2.5} \]

These equations have the explicit solution

\[ \alpha = \frac{1}{2 - 2^{1/(2m+1)}}, \qquad \beta = -\,\frac{2^{1/(2m+1)}}{2 - 2^{1/(2m+1)}}. \tag{6.2.6} \]

For example:

Base Order Resulting Order \(\alpha\) \(\beta\)
2 4 \(\displaystyle \frac{1}{2-2^{1/3}} \approx 0.675603\) \(1 - 2\alpha \approx -0.351207\)
4 6 \(\displaystyle \frac{1}{2-2^{1/5}} \approx 0.587\) \(1 - 2\alpha \approx -0.174\)

Starting from Strang (\(2^{nd}\) order), repeated Yoshida triple-jumps yield methods of orders 4, 6, 8, …


6.2.4 Suzuki’s Fractal Composition

Suzuki (1991) extended Yoshida’s idea to general “fractal” compositions. A common example is:

\[ \Psi_h^{(p+2)} = \Psi_{\gamma h}^{(p)} \circ \Psi_{\delta h}^{(p)} \circ \Psi_{\gamma h}^{(p)} \tag{6.2.7} \]

with coefficients chosen to annihilate additional BCH terms. These schemes form a recursive family generating arbitrarily high order from a base symmetric integrator.


6.2.5 Symmetric Composition for Splitting Methods

When the base integrator \(\Psi_h\) is itself a splitting, such as Strang:

\[ \Psi_h = \varphi_{h/2}^{[1]} \circ \varphi_h^{[2]} \circ \varphi_{h/2}^{[1]}, \tag{6.2.8} \]

the Yoshida and Suzuki compositions give high-order splitting methods. For example, the 4th-order symmetric splitting obtained from Strang is:

\[ \Phi_h^{(4)} = \Phi_{\alpha h}^{\mathrm{S}} \circ \Phi_{\beta h}^{\mathrm{S}} \circ \Phi_{\alpha h}^{\mathrm{S}}, \tag{6.2.9} \]

which expands (after substituting the Strang components) into a 7-factor composition of subflows \(\varphi^{[1]}\) and \(\varphi^{[2]}\).

These high-order splittings are widely used in:


6.2.6 Geometric Benefits of Symmetric Composition

Symmetric compositions retain:

High-order symplectic compositions are often competitive with (or superior to) high-order Runge–Kutta methods because the dominant error terms in Hamiltonian problems are often commutators, which compositions suppress efficiently.


6.2.7 Summary and References for 6.2

In this section we developed the general framework of symmetric composition methods. Starting from a symmetric base integrator (typically Strang), one can construct high-order geometric integrators using Yoshida triple-jumps, Suzuki fractal compositions, or more general optimized compositions.

Key takeaways:

References

  1. H. Yoshida, “Construction of higher order symplectic integrators,” Phys. Lett. A, 150(5–7):262–268, 1990.
  2. M. Suzuki, “General theory of fractal path integrals,” Phys. Lett. A, 146:319–323, 1991.
  3. E. Hairer, C. Lubich, G. Wanner, Geometric Numerical Integration, Springer, 2006.

6.3 Optimised High-Order Splitting Methods: Real & Complex Coefficients

Classical symmetric compositions (Strang, Yoshida, Suzuki) generate high-order methods with universal algebraic rules, but they are rarely optimal for a specific class of problems. In this section we present the modern theory of optimised splitting schemes, including:

These methods represent the contemporary state-of-the-art and often outperform all traditional splitting schemes for demanding applications.


6.3.1 The Algebraic Order Barrier for Real Coefficients

Consider symmetric splitting of a two-part system:

\[ \Phi_h = \prod_{j=1}^{s} \exp(a_j h F^{[1]}) \exp(b_j h F^{[2]}), \]

with the symmetry constraint

\[ a_j = a_{s+1-j}, \qquad b_j = b_{s+1-j}. \tag{6.3.1} \]

The BCH expansion of \(\log \Phi_h\) contains commutators \[ [F^{[1]},F^{[2]}], \quad [F^{[1]},[F^{[1]},F^{[2]}]], \quad \dots \] and the order conditions require vanishing of specific linear combinations of these.

Key barrier (Sheng 1989, Suzuki 1991):

No symmetric splitting of order higher than 2 exists with all real coefficients \(a_j, b_j \ge 0\).

In particular:

Thus:


6.3.2 Minimal-Stage Real-Coefficient Splittings (Optimised)

High-order real-coefficient splittings are constructed by solving the BCH order conditions while minimising either:

Fourth-order optimised real splitting (McLachlan 1995)

A minimal 4th-order symmetric scheme exists with three exponentials of each operator:

\[ \Phi_h^{(4)} = e^{a_1 h F^{[1]}} e^{b_1 h F^{[2]}} e^{a_2 h F^{[1]}} e^{b_1 h F^{[2]}} e^{a_1 h F^{[1]}}. \tag{6.3.2} \]

with coefficients

\[ a_1 = \frac{1}{2 - 2^{1/3}}, \qquad a_2 = -\frac{2^{1/3}}{2 - 2^{1/3}}, \] \[ b_1 = \frac{1}{2}. \tag{6.3.3} \]

These are Yoshida's coefficients but optimised variants exist where the error constant is minimised (e.g., McLachlan’s 1995 scheme).

Task marker: Section 6.4 derives optimal real splittings for Hamiltonian dynamics using backward error analysis and modified Hamiltonians.


6.3.3 Complex-Coefficient Splittings for High and Arbitrary Order

Allowing complex coefficients dramatically changes the landscape:

General complex symmetric splitting

\[ \Phi_h = \prod_{j=1}^s \exp(a_j h F^{[1]}) \exp(b_j h F^{[2]}), \tag{6.3.4} \]

where now \[ a_j, b_j \in \mathbb{C}, \qquad \Re(a_j), \Re(b_j) \ge 0. \]

These schemes are stable for:

Example: 4th-order complex scheme with all positive real parts

The Blanes–Casas “C4” scheme:

\[ a_1 = 0.5 + 0.28867513459 i, \qquad a_2 = 0.5 - 0.28867513459 i, \tag{6.3.5} \]

yields a 4th-order splitting with only two exponentials of each operator, outperforming all real-coefficient schemes in quantum applications.


6.3.4 Positivity Constraints for Parabolic PDEs

For diffusion or parabolic systems:

\[ \dot{u} = A u + B u, \]

the operator \( e^{a_j h A} \) must satisfy:

\[ \Re(a_j) \ge 0, \qquad \Re(b_j) \ge 0, \tag{6.3.6} \]

otherwise the method is unstable. Therefore:

SOTA: Blanes–Casas optimal CFM splittings for parabolic PDEs

They constructed 4th, 6th, and 8th-order splitting schemes with all coefficients in the right half-plane. These dominate the field in smoothing-dominated PDEs.


6.3.5 Energy-Optimised Splittings for Hamiltonian Systems

For Hamiltonian systems with \(H = T(p) + V(q)\), the goal is to minimise:

These yield:

\[ \Phi_h^{\mathrm{OMF}} = e^{\lambda h B} e^{\frac{1}{2} h A} e^{(1-2\lambda) h B} e^{\frac{1}{2} h A} e^{\lambda h B}, \tag{6.3.7} \]

with \(\lambda \approx 0.1931833275\), which dramatically reduces energy drift.


6.3.6 Summary and References for 6.3

This section presented the full modern landscape of optimised splittings:

References

  1. S. Blanes, F. Casas, J.A. Oteo, J. Ros, The Magnus Expansion and Some of its Applications, Physics Reports 470, 2009.
  2. R.I. McLachlan, “On the optimality of symplectic integrators,” SIAM J. Sci. Comput., 16, 1995.
  3. M. Suzuki, “Fractal decompositions,” J. Math. Phys., 1990–1991 series.
  4. M. Omelyan, I. Mryglod, R. Folk, Optimized Forest–Ruth- and Suzuki-type symplectic integrators, Computer Physics Communications, 2002–2003.
  5. S. Blanes, F. Casas, A Concise Introduction to GNI, 2e, CRC, 2026.

6.4 Backward Error Analysis for Splitting Methods

Backward error analysis (BEA) provides the fundamental theoretical explanation for the excellent long-time behaviour of geometric integrators. Rather than analysing the discrete numerical flow directly, BEA interprets a numerical scheme as the exact flow of a modified differential equation:

\[ \Phi_h = \exp\!\big(h \tilde{F}(h)\big), \]

where the modified vector field admits a formal series expansion \[ \tilde{F}(h) = F + h F_2 + h^2 F_3 + h^3 F_4 + \cdots. \] For geometric methods—especially symplectic mappings—the modified system inherits the geometric structure (e.g.~Hamiltonian geometry) to all orders in \(h\).


6.4.1 Modified Vector Fields from BCH Theory

Consider a splitting integrator for an autonomous system

\[ y' = (A + B)(y). \tag{6.4.1} \]

A general splitting has the form

\[ \Phi_h = \prod_{j=1}^s \exp(a_j h A)\, \exp(b_j h B). \tag{6.4.2} \]

The Baker–Campbell–Hausdorff expansion provides

\[ \log(\Phi_h) = h(A+B) + h^2 \alpha_1[A,B] + h^3\big(\alpha_2[A,[A,B]] + \alpha_3[B,[A,B]]\big) + h^4 \cdots. \tag{6.4.3} \]

Let the modified vector field be defined by

\[ h\tilde{F}(h) = \log(\Phi_h). \]

Then

\[ \tilde{F}(h) = A + B + h \alpha_1[A,B] + h^2\big(\alpha_2[A,[A,B]] + \alpha_3[B,[A,B]]\big) + h^3(\cdots). \tag{6.4.4} \]

Each coefficient \(\alpha_k\) is a known polynomial in the splitting coefficients \(a_j, b_j\). Setting these equal to zero yields the order conditions. Optimised splittings (Section 6.3) choose them to minimise the leading error terms in the modified system.


6.4.2 Geometric Structure of the Modified System

If the original system is Hamiltonian with \[ A = X_T, \qquad B = X_V, \] then all nested commutators of \(A\) and \(B\) are Hamiltonian vector fields. Thus the modified vector field is Hamiltonian:

\[ \tilde{F}(h) = X_{\tilde{H}}, \]

where the modified Hamiltonian admits the expansion

\[ \tilde{H} = H + h^2 H_2 + h^4 H_4 + h^6 H_6 + \cdots. \tag{6.4.5} \]

For symmetric (time-reversible) methods, all odd-index terms vanish:

\[ H_{2k+1} = 0. \]

Thus Strang and all symmetric compositions preserve (to very high accuracy) an even-order approximation of the true Hamiltonian, which explains their exceptional long-time energy conservation.


6.4.3 Long-Time Behavior and Near-Conservation Laws

For symplectic integrators, BEA yields the “shadow energy” theorem:

If \( \Phi_h \) is symplectic, then it exactly preserves the modified Hamiltonian \( \tilde{H} \) for exponentially long times \( T = \mathcal{O}(e^{c/h}) \), provided the solution remains in a compact set.

In practice:

This explains why symplectic integrators dominate long-time molecular dynamics, celestial mechanics, and semiclassical quantum simulations.


6.4.4 Backward Error and Optimisation of Splitting Coefficients

The leading term in the modified Hamiltonian for a symmetric splitting is

\[ H_2 = \alpha_{21}\{T,\{T,V\}\} + \alpha_{22}\{V,\{T,V\}\}. \tag{6.4.6} \]

The constants \( \alpha_{2j} \) encode the leading error of the scheme. Modern optimised splittings (Eqs.~(6.3.7) and others) choose coefficients to:

For example, the Omelyan–Mryglod–Folk (OMF) parameter \( \lambda \approx 0.1931833275 \) is chosen to minimise the modified energy error in lattice QCD simulations.


6.4.5 BEA for Complex-Coefficient Splittings

For complex splitting coefficients \(a_j, b_j \in \mathbb{C}\) with \( \Re(a_j), \Re(b_j) \ge 0 \), the modified vector field remains well-defined and admits the same BCH expansion (6.4.4).

If the original system is Hamiltonian but the complex step destroys exact symplecticity of each subflow (e.g. for real canonical variables), the method is usually interpreted in a complexified phase space. For Schrödinger-type PDEs, complex coefficients preserve unitarity exactly, and BEA yields a modified Hamiltonian which is anti-Hermitian:

\[ \tilde{H} = H + h^2 H_2 + h^4 H_4 + \cdots, \qquad \tilde{H}^\dagger = -\tilde{H}. \tag{6.4.7} \]

This ensures exact preservation of \( \|\psi(t)\|_{L^2} \) for highly accurate quantum time propagation.


6.4.6 Resonances and Limitations of BEA

Although BEA is asymptotically valid, it breaks down when:

High-order optimized splittings typically shift the dominant resonance tongues to higher order in \(h\), improving practical stability.


6.4.7 Summary and References for 6.4

Key facts established in this section:

References

  1. E. Hairer, C. Lubich, G. Wanner, Geometric Numerical Integration, Springer, 2006.
  2. S. Blanes, F. Casas, A Concise Introduction to GNI, CRC, 2026.
  3. R.I. McLachlan, G.R.W. Quispel, “Splitting Methods,” Acta Numer., 11 (2002), 341–434.
  4. M. Omelyan, I. Mryglod, R. Folk, Computer Physics Communications, 2002–2003 series.

6.5 Splitting Methods for PDEs

Splitting schemes extend naturally from finite-dimensional ODEs to evolution PDEs. Many PDEs can be written in the semilinear form

\[ u_t = A u + B(u), \tag{6.5.1} \]

where A is a (typically linear, stiff) operator generating a strongly continuous semigroup, and B is a (possibly nonlinear) locally Lipschitz operator. If the flows of \(A\) and \(B\) are explicitly computable or efficiently approximable, splitting schemes yield structure-preserving, high-accuracy time integrators.

This section develops the rigorous theory and SOTA implementations for splitting methods applied to:

The underlying functional-analytic framework uses semigroup theory, Fourier spectral discretization, and geometric properties of flows in infinite dimensions.


6.5.1 Operator Semigroup Foundations

Let \(A: D(A)\subset X \to X\) be the generator of a \(C^0\)-semigroup \(e^{tA}\) on a Banach space \(X\). Let \(B: X \to X\) be locally Lipschitz so that the ODE

\[ u_t = B(u) \tag{6.5.2} \]

has a unique local flow \( \varphi_B^t\). Then the splitting schemes

are well-defined on \(X\).

Convergence follows from the classical theory: if \(A+B\) generates a (mild or strong) solution semigroup, then the Lie and Strang splittings converge with order 1 and 2 respectively under mild regularity assumptions.

\[ \|\Phi_h^n u_0 - e^{nh(A+B)}u_0\| \le C h^p, \qquad p = 1~\text{(Lie)},\ 2~\text{(Strang)}. \tag{6.5.3} \]


6.5.2 Fourier–Spectral Splitting

Many PDEs have linear parts diagonalizable by Fourier transform. Let \( \mathcal{F} \) denote the Fourier transform, then:

\[ A = \mathcal{F}^{-1} \Lambda(k)\, \mathcal{F}, \tag{6.5.4} \]

so that its flow is exactly computable:

\[ e^{tA} u = \mathcal{F}^{-1} \big( e^{t\Lambda(k)} \hat{u}(k) \big). \tag{6.5.5} \]

This leads to highly efficient splitting integrators:

This structure underlies spectral splittings for Schrödinger, NLS, KdV, and dispersive PDEs.


6.5.3 Splitting for Schrödinger Equations

Consider the linear Schrödinger equation

\[ i\psi_t = -\frac{1}{2}\Delta \psi + V(x)\psi. \tag{6.5.6} \]

Define

Both flows are exactly computable:

\[ e^{tA}\psi = \mathcal{F}^{-1}\left( e^{-i t |k|^2/2} \hat{\psi}(k)\right), \tag{6.5.7} \]

\[ e^{tB}\psi = e^{-i t V(x)} \psi(x). \tag{6.5.8} \]

Thus Strang splitting gives a unitary, second-order, time-reversible method. High-order complex-coefficient splittings (Section 6.3) are widely used in quantum chemistry and semiclassical dynamics.

Nonlinear Schrödinger equation (NLS)

For

\[ i \psi_t = -\frac{1}{2}\Delta \psi + g|\psi|^2\psi, \tag{6.5.9} \]

the nonlinear flow is

\[ \varphi_B^t(\psi) = e^{-i g t |\psi(x)|^2} \psi(x), \tag{6.5.10} \]

which preserves mass exactly. Strang and higher-order splittings preserve:

These are state-of-the-art and standard in Gross–Pitaevskii solvers and Bose–Einstein condensate simulations.


6.5.4 Heat and Parabolic Equations: Positivity Issues

For parabolic systems:

\[ u_t = \Delta u + B(u), \tag{6.5.11} \]

the semigroup \(e^{t\Delta}\) is contractive only for \(t\ge 0\). Therefore splitting coefficients must satisfy:

\[ \Re(a_j) \ge 0, \qquad \Re(b_j) \ge 0. \tag{6.5.12} \]

Real-coefficient high-order splittings require negative coefficients (Sheng–Suzuki barrier), which are unstable for diffusion.

Thus modern high-order solvers for parabolic PDEs use:


6.5.5 KdV, Dispersive and Nonlinear Wave Equations

For KdV:

\[ u_t + 6 uu_x + u_{xxx} = 0, \tag{6.5.13} \]

split into:

The combination via Strang splitting yields a highly effective scheme used in soliton dynamics, dispersive shock wave, and Whitham modulation simulations.

Klein–Gordon and wave equations

For the Klein–Gordon equation:

\[ u_{tt} = \Delta u - m^2 u - f(u), \tag{6.5.14} \]

rewrite as a first-order Hamiltonian system in Fourier space:

\[ \frac{d}{dt} \begin{pmatrix} u \\ v \end{pmatrix} = \begin{pmatrix} v \\ \Delta u - m^2 u \end{pmatrix} + \begin{pmatrix} 0 \\ -f(u) \end{pmatrix}. \tag{6.5.15} \]

Linear part diagonalises in Fourier space; nonlinear part is pointwise. High-order splittings with FFTs are state-of-the-art for nonlinear wave propagation, spectral radiation studies, and weak turbulence.


6.5.6 Maxwell Equations and Electromagnetic Splitting

Maxwell’s equations (in vacuum):

\[ \partial_t E = \nabla\times H, \qquad \partial_t H = -\nabla\times E. \tag{6.5.16} \]

preserve:

Splitting into:

yields symplectic, structure-preserving methods suitable for:


6.5.7 Summary and References for 6.5

Key takeaways:

References

  1. Hairer, Lubich, Wanner, Geometric Numerical Integration, 2006.
  2. Blanes & Casas, A Concise Introduction to GNI, 2ed, CRC, 2026.
  3. Bao, Jin, and Markowich, “Numerical methods for the nonlinear Schrödinger equation,” SIAM Review, 2013.
  4. Holden, Karlsen, Risebro & Tao, Splitting Methods for PDEs, EMS 2011.

6.6 Relation of Splitting to Magnus Methods and Exponential Integrators

Splitting and composition methods, Magnus integrators, and exponential integrators all arise from the same operator-theoretic viewpoint: approximate the evolution operator of a differential equation by structured exponentials of simpler pieces. This section explains their common algebraic backbone and clarifies when each class is preferable.

We focus on:

  1. The shared operator framework for autonomous and non-autonomous problems,
  2. Magnus expansion as a “logarithm” of products of exponentials (splittings),
  3. Commutator-free Magnus (CFM) schemes as high-order composition methods,
  4. The link to exponential integrators for semilinear systems,
  5. Hybrid geometric schemes that combine these tools in practice.

6.6.1 Common Operator Framework

Consider a linear (possibly time-dependent) evolution equation on a Banach space \(X\):

\[ u'(t) = A(t) u(t), \qquad u(t) \in X, \]

where \(A(t)\) is a (possibly unbounded) linear operator generating an evolution family \(U(t,s)\) such that

\[ u(t) = U(t,s) u(s), \qquad U(t,s) U(s,r) = U(t,r),\quad U(s,s)=I. \]

Magnus viewpoint. For sufficiently regular \(A(t)\), we may write

\[ U(t,s) = \exp\big(\Omega(t,s)\big), \]

where \(\Omega(t,s)\) is given by the Magnus expansion, an infinite series of time integrals of nested commutators of \(A(\tau)\). Truncating \(\Omega\) yields Magnus integrators.

Splitting viewpoint. When \(A\) decomposes into simpler pieces,

\[ A(t) = A_1(t) + A_2(t) + \cdots + A_m(t), \]

whose flows (or exponentials) can be computed efficiently, we approximate \(U(t,s)\) by compositions of exponentials of the sub-operators, e.g. Lie–Trotter or Strang splittings over a time step \(h\).

Exponential integrator viewpoint. For semilinear systems

\[ u'(t) = A u(t) + g(t, u(t)), \]

exponential integrators treat \(e^{hA}\) exactly (or nearly so) and approximate the contribution of the nonlinearity \(g\) via \(\varphi\)-functions and quadrature. These methods can be interpreted as Magnus-/splitting-type approximations in the variation-of-constants formula.

In all three cases, the core object is the exponential of an operator (or sum of operators) and its algebra under composition, governed by the Baker–Campbell–Hausdorff (BCH) and Magnus series.


6.6.2 Magnus Expansion as the Logarithm of Splitting Schemes

For simplicity, consider an autonomous decomposition

\[ A = A_1 + A_2 \]

and a splitting step over a time increment \(h\):

\[ S(h) = e^{a_1 h A_1} e^{b_1 h A_2} e^{a_2 h A_1} e^{b_2 h A_2} \cdots e^{a_s h A_1} e^{b_s h A_2}. \tag{6.6.1} \]

The BCH formula ensures that there exists an operator \(Z(h)\) such that

\[ S(h) = \exp(Z(h)). \]

Expanding \(Z(h)\) as a formal power series in \(h\) gives

\[ Z(h) = h(A_1 + A_2) + h^2 \alpha_1 [A_1, A_2] + h^3 \big( \alpha_2 [A_1,[A_1,A_2]] + \alpha_3 [A_2,[A_1,A_2]] \big) + \cdots, \tag{6.6.2} \]

where the coefficients \(\alpha_j\) are polynomials in the composition coefficients \(\{a_k, b_k\}\). The order conditions for the splitting require that

Formally, \(Z(h)\) plays the same role as the Magnus generator \(\Omega(h)\) for the exact flow. For an autonomous problem, the exact Magnus generator is simply

\[ \Omega(h) = h(A_1 + A_2), \]

while for time-dependent problems \(A(t)\) the Magnus series contains time integrals of commutators. In this sense:

Splitting methods can be viewed as particular approximations of the Magnus generator by the BCH logarithm of a product of simpler exponentials.

The modern theory of commutator-free Magnus integrators makes this connection explicit by designing products of exponentials whose BCH logarithm reproduces a chosen truncated Magnus series.


6.6.3 Commutator-Free Magnus (CFM) Methods as High-Order Compositions

For linear non-autonomous systems

\[ u'(t) = A(t) u(t), \]

the Magnus expansion over a step \([t_n,t_{n+1}] = [t_n, t_n+h]\) is

\[ \Omega_n(h) = \int_{t_n}^{t_{n+1}} A(\tau)\, d\tau - \frac{1}{2} \int_{t_n}^{t_{n+1}}\!\!\int_{t_n}^{\tau_1} [A(\tau_1), A(\tau_2)]\, d\tau_2\, d\tau_1 + \cdots. \tag{6.6.3} \]

Classical Magnus integrators compute or approximate commutators like \([A(\tau_1), A(\tau_2)]\), which may be expensive for large matrices or operators.

Commutator-free Magnus (CFM) methods avoid explicit commutators by using products of exponentials of linear combinations of evaluations of \(A\) at quadrature nodes:

\[ U_{n+1} = \prod_{j=1}^s \exp\!\left( h \sum_{\ell=1}^r \alpha_{j\ell} A(t_n + c_\ell h) \right) U_n. \tag{6.6.4} \]

Here, \(c_\ell\) are quadrature nodes (e.g. Gauss–Legendre abscissas), and the coefficients \(\alpha_{j\ell}\) are chosen so that the BCH logarithm of this product matches the truncated Magnus series up to order \(p\).

From the perspective of Chapter 4, each factor \(\exp(h \sum_\ell \alpha_{j\ell} A(t_n+c_\ell h))\) is simply another sub-flow whose exponential can be applied efficiently (e.g. via Krylov or FFT-based methods). The overall method is a high-order composition method for the time-dependent generator.

Key properties:

Thus CFM integrators can be seen as a SOTA synthesis of Magnus ideas (logarithm of the evolution) and composition/splitting ideas (products of exponentials of simpler pieces).


6.6.4 Exponential Integrators as Splitting/Magnus Approximations

For semilinear problems

\[ u'(t) = A u(t) + g(t,u(t)), \tag{6.6.5} \]

the variation-of-constants formula gives

\[ u(t_{n+1}) = e^{hA} u(t_n) + \int_0^h e^{(h-\sigma)A} g(t_n + \sigma, u(t_n+\sigma))\, d\sigma. \tag{6.6.6} \]

Exponential Runge–Kutta and related exponential integrators approximate the integral term via quadrature and replace \(u(t_n+\sigma)\) by internal stage values. Typical schemes have the form

\[ U_{n+1} = e^{hA} U_n + h \sum_{j=1}^s \varphi_j(hA) G_j(U_n), \tag{6.6.7} \]

where the \(\varphi_j\) functions arise from integral representations involving \(e^{(h-\sigma)A}\).

Relation to splitting. A simple Lie–Trotter splitting for (6.6.5) reads

\[ U_{n+1} = \varphi_g^h\!\big(e^{hA}U_n\big), \]

where \(\varphi_g^h\) is the nonlinear flow generated by \(u' = g(t,u)\) (with appropriate time freezing or time-labelling). Higher-order Strang-type and composition splittings can be interpreted as particular quadrature approximations of (6.6.6), just as exponential integrators do, but organized as compositions of sub-flows rather than as direct quadrature of the integral.

Relation to Magnus. If \(g\) is linear in \(u\), then the full generator \(A(t)\) of (6.6.5) is linear, and Magnus/CFM schemes apply directly. For general nonlinear \(g\), Magnus ideas still influence the design of exponential integrators (e.g. time symmetry, commutator structure of linearized operators, etc.), but require more care.

In practice:


6.6.5 Hybrid Geometric Schemes and Practical Guidelines

Modern SOTA codes often blend splitting, Magnus/CFM, and exponential ideas to balance structure preservation, efficiency, and robustness.

Examples of hybrid strategies

Rule-of-thumb choice

From the geometric standpoint, all these methods are variations on a single theme: construct a discrete flow as a product of exponentials whose BCH logarithm approximates the exact generator in a way that respects invariants, symmetries, and qualitative structure.


6.6.6 References for Section 6.6

  1. E. Hairer, C. Lubich, G. Wanner, Geometric Numerical Integration, 2nd ed., Springer, 2006. (Chs. II, IV, IX: splitting, Magnus, and exponential integrators.)
  2. S. Blanes, F. Casas, A Concise Introduction to Geometric Numerical Integration, 2nd ed., Chapman & Hall/CRC, 2026. (Chs. 4–6: Magnus and commutator-free methods.)
  3. S. Blanes, F. Casas, J.A. Oteo, J. Ros, “The Magnus expansion and some of its applications,” Physics Reports 470 (2009), 151–238.
  4. M. Hochbruck, A. Ostermann, “Exponential integrators,” Acta Numerica 19 (2010), 209–286.