Symmetry reduction from D2 tetramer to C2 dimer: group-theoretic and thermodynamic modeling of enzyme catalytic efficiency

Research Article
Open access

Symmetry reduction from D2 tetramer to C2 dimer: group-theoretic and thermodynamic modeling of enzyme catalytic efficiency

Tianyi Qiu 1*
  • 1 Bethany Lutheran College    
  • *corresponding author vqiu@blc.edu
Published on 4 November 2025 | https://doi.org/10.54254/3029-0880/2025.29183
AORPM Vol.4 Issue 3
ISSN (Print): 3029-0880
ISSN (Online): 3029-0899

Abstract

A mathematical framework is presented to quantify the relationship between quaternary-structure symmetry, free energy, and catalytic efficiency during the transition from aD2-symmetric tetramer to aC2-symmetric dimer, exemplified with LDHA. The approach constructs explicitD2representations on subunit and interface feature spaces, derives projection operators to decompose operators and data into irreducible-representation components, and computes symmetry-resolved free-energy differences via Gaussian/statistical and harmonic/Hessian methods. Connections to kinetics are made through transition state theory with channel degeneracy. Reproducible algorithms and a workflow for mapping FoldX outputs into irrep-resolved diagnostics and efficiency predictions are provided.

Keywords:

symmetry reduction, group representation, enzyme catalytic efficiency, free energy, LDHA

Qiu,T. (2025). Symmetry reduction from D2 tetramer to C2 dimer: group-theoretic and thermodynamic modeling of enzyme catalytic efficiency. Advances in Operation Research and Production Management,4(3),1-34.
Export citation

1. Introduction

Enzyme quaternary structure often exhibits symmetry, and this structural organization can significantly affect catalytic efficiency. Understanding how symmetry reduction impacts enzymatic function is a fundamental question in structural biology and mathematical modeling. A natural mathematical framework to address this problem is group theory, which formalizes symmetry and its action on suitable data spaces.

Not all enzymes are suitable for controlled structural perturbations. To study symmetry- dependent effects concretely, tetramer-to-dimer transitions provide a classical example of symmetry reduction, where experimental and structural data are available. In this work, lactate dehydrogenase A (LDHA) is selected as the model system because it forms a  D2 -symmetric tetramer, has documented mutational studies in the literature, and structural data are available from the Protein Data Bank.

The research proceeds in the following steps:

First, to define group actions on data spaces. Structural coordinates such as collective variables, interface energies, or density fields are formalized with an inner product structure, allowing symmetry operations to act linearly and quantitatively on the space.

Second, to construct explicit representations and projectors. Irreducible representations (irreps) of  D2  (tetramer) and  C2  (dimer) are constructed, and Reynolds projection operators are derived to decompose data and operators into irrep components, isolating the contributions of each symmetry mode.

Third, to compute symmetry-resolved free-energy differences. Gaussian/statistical (covariance-based) and harmonic/Hessian (normal-mode-based) approximations quantify the free-energy change associated with symmetry reduction, with explicit contributions from each irrep.

Fourth, to bridge to catalytic efficiency. Transition state theory (TST) connects free-energy differences to predicted catalytic efficiencies, accounting for channel degeneracy and transition-state probabilities.

Fifth, to implement a computational workflow. Energy terms and structural features from FoldX outputs are mapped into the symmetry-adapted framework, and diagnostic scores identify which symmetry components dominate the observed transition.

This approach results in a set of explicit, symmetry-resolved expressions for projected covariance matrices and free-energy differences. These formulas represent a novel mathematical formalism linking group-theoretic symmetry, structural transformation, and catalytic efficiency. The framework allows prediction of which symmetry modes most strongly drive the tetramer-to-dimer transition and provides a direct quantitative bridge from structural symmetry to enzymatic function.

2. Symmetry groups, homomorphism, and action spaces

2.1. Groups and homomorphism

Consider tetramers a finite group:

D2={e,α,β,αβ}, α2=β2=e,  αβ=βα(1)

and consider dimers a finite group:

C2={e,γ}, γ2=e(2)

By construction  D2  is abelian of order  4  (isomorphic to the Klein four group  V4 ), and  C2  is the cyclic group of order  2 .

2.1.1. Group homomorphism

A map  ϕ:G→H  between groups is a homomorphism iff  ϕ(g1g2)=ϕ(g1)ϕ(g2)  for all  g1,g2∈G[1]. Define

ϕ:D2C2, ϕ(α)=γ,  ϕ(β)=e(3)

and extend multiplicatively to all of  D2 . Concretely,

ϕ(e)=e,ϕ(α)=γ,ϕ(β)=e,ϕ(αβ)=γ(4)

 ϕ  as defined above is a surjective group homomorphism.

Proof. Since  D2  is generated by  α,β  with relations  α2=β2=e  and  αβ=βα , it suffices to check that the relations are preserved:

ϕ(α2)=ϕ(α)2=γ2=e=ϕ(e), ϕ(β2)=ϕ(β)2=e(5)

and

ϕ(αβ)=ϕ(α)ϕ(β)=γe=γ=ϕ(β)ϕ(α)=ϕ(βα)(6)

Hence  ϕ  respects the defining relations and extends to a homomorphism on  D2 . Its image contains  γ  and  e , hence  Im ϕ=C2 , so  ϕ  is surjective.

2.1.2. Kernel and quotient

Recall  kerϕ={g∈D2:ϕ(g)=e}[1]. Here

kerϕ=e,β(7)

which is a subgroup of order  2 . Since  ϕ  is a homomorphism,  kerϕ  is a normal subgroup. By the First Isomorphism Theorem,

D2/kerϕIm ϕ=C2(8)

The quotient map  π:D2D2/kerϕ  followed by the isomorphism from  D2/kerϕ  to  C2  equals  ϕ .

2.1.3. Functions factoring through the quotient

If  X  is any set on which  D2  acts and    is a function space on  X , then the subspace of functions invariant under  kerϕ[1],

kerϕ:={fH: U(k)f=f, kkerϕ}(9)

is canonically identified with functions on the quotient  X/kerϕ  or, algebraically, with functions on  D2/kerϕ : every  f∈kerϕ  is constant on  kerϕ -cosets and thus descends to a function on the quotient; conversely, a function on the quotient pulls back to a  kerϕ -invariant function.

2.2. Representations, invariant projectors and factoring through quotients

2.2.1. Representations and unitarity (definitions)

Let  (H,⟨⋅,⋅⟩)  be a (complex) Hilbert space. A representation of a group  G  on    is a group homomorphism  U:G→GL(H)  satisfying  U(g1g2)=U(g1)U(g2) . The representation  U  is unitary (or orthogonal in the real case) if [2]

U(g)f,U(g)h=f,h f,hH, gG(10)

2.2.2. Averaging projector onto kerϕ -invariants

Let  K:=kerϕ  be a finite subgroup. Define the averaging operator

PK:=1|K|kKU(k)(11)

Assume  U  is unitary. Then  PK  is the orthogonal projection onto the closed subspace  K:={v∈H: U(k)v=v ∀k∈K} . Moreover  PK  commutes with  U(g)  for every  g∈G , hence the representation leaves  K  invariant.

Proof. First compute  PK2 :

PK2=1|K|2k,k'KU(k)U(k')=1|K|2hK(kK1)U(h)=1|K|hKU(h)=PK(12)

so  PK  is idempotent. Since each  U(k)  is unitary,  U(k)*=U(k)-1=U(k-1) , and taking adjoint yields  PK*=PK , hence  PK  is self-adjoint. Idempotent self-adjoint operators are orthogonal projections; their range equals  {v:PKv=v} . But  PKv=v  iff  1|K|k∈KU(k)v=v , which is equivalent to  U(k)v=v  for all  k∈K  (apply  U(k0)  and use group averaging). Thus the range is  K .

Finally, for any  g∈G ,

U(g)PKU(g)-1=1|K|kKU(g)U(k)U(g)-1=1|K|kKU(gkg-1)(13)

If  K  is normal then  gKg-1=K  and the right-hand side equals  PK . In our setting  K=kerϕ  is normal, so  PK  commutes with  U(g) . Hence  K  is  G -invariant.

2.2.3. Factoring a representation through the quotient

Let  K=kerϕ  and suppose  U:G→U(H)  is a unitary representation. Consider the restriction of  U  to  K . For  gg'∈G  with  gK=g'K , we have  g'=gk  for some  k∈K  and, for  v∈K ,

U(g')v=U(g)U(k)v=U(g)v(14)

Thus the map  U~:G/K→U(K)  given by  U~(gK):=U(g)|K  is well-defined and a representation. In other words, the representation on the  K -fixed subspace factors through the quotient group  G/K≅C2 .

2.3. Action spaces: L2 constructions and unitarity proofs

2.3.1. Why L2 and which inner product

Let  =L2(R3)  (complex-valued functions) with the usual inner product

f,g=R3f(r)¯ g(r) dr(15)

 L2  is a Hilbert space: it is complete and admits orthogonal projections, spectral theorem for bounded self-adjoint operators, and a well-behaved theory of quadratic forms. These properties are prerequisites for using orthogonal (projector) decompositions,  logdet  formulas for Gaussian integrals, and spectral block-diagonalization of self-adjoint Hessians [3].

2.3.2. Density-field representation on L2(R3)

Definition.

Let  G  act on  R3  by orthogonal linear maps  R(g)∈O(3)  (or by rigid motions with zero translations for point groups). Define [2]

(U(g)ρ)(r):=ρ(R(g)-1r)(16)

We check that  U  is a unitary representation.

 U:G→U(L2(R3))  defined above is a unitary representation.

Proof. (Representation property.) For  g1,g2∈G ,

U(g1)U(g2)ρ(r)=U(g1)(ρ(R(g2)-1r))=ρ(R(g1)-1R(g2)-1r)=ρ((R(g2)R(g1))-1r)=U(g1g2)ρ(r)(17)

(Unitarity.) For  ρ1,ρ2L2(R3) ,

U(g)ρ1,U(g)ρ2=R3ρ1(R(g)-1r)¯ ρ2(R(g)-1r) dr(18)

Change variables  s:=R(g)-1r . Since  R(g)  is orthogonal,  detR(g)=±1  and  dr=|detR(g)| ds=ds . Thus

U(g)ρ1,U(g)ρ2=R3ρ1(s)¯ ρ2(s) ds=ρ1,ρ2(19)

Therefore each  U(g)  is unitary and  U  is a unitary representation.

2.3.3. CV (collective-variable) representation on L2(Y,μ)

Setup and unitarity criterion.

Let  Φ  be a map from full coordinates to CVs,  x↦y=Φ(x)∈Y⊂Rm , and let  μ  be a probability measure on  Y . Consider  :=L2(Y,μ)  with inner product  ⟨A,B⟩=YA(y)¯B(y) dμ(y) . Assume the group  G  acts on  Y  linearly or by permutations via  Γ:G→GL(Y) , and define

(U(g)A)(y):=A(Γ(g)-1y)(20)

 U  is unitary on  L2(Y,μ)  iff  μ  is  Γ -invariant, i.e.  μ=μ∘Γ(g)-1  for all  g∈G .

Proof. If  μ  is  Γ -invariant, then

U(g)A,U(g)B=YA(Γ(g)-1y)¯B(Γ(g)-1y) dμ(y)(21)

Change variables  z=Γ(g)-1y ; by invariance  dμ(y)=dμ(z)  and the domain is unchanged, so the integral equals  ⟨A,B⟩ . Conversely, unitarity for all  A,B  forces invariance of  μ  by testing against indicator functions.

2.4. Intertwining operators and boundedness

2.4.1. A linear sampling operator

Suppose we construct a linear sampling operator  S:L2(R3)→L2(Y,μ)  of the integral-kernel type

(Sρ)(y)=R3K(y,r) ρ(r) dr(22)

where  K:Y×R3→C  is measurable. Typical choices of  K  are localized feature kernels (e.g. Gaussian windows integrated against coordinates).

If for  μ -almost every  y  the function  r↦K(y,r)  lies in  L2(R3)  and the function  y↦∥K(y,⋅)L2(R3)2  is integrable w.r.t.  μ , then  S  is a bounded linear operator  L2(R3)→L2(Y,μ) .

Proof. By Cauchy–Schwarz,

|(Sρ)(y)|2=|K(y,r)ρ(r) dr|2K(y,)L2(R3)2 ρL2(R3)2(23)

Integrate both sides w.r.t.  y  and use the integrability assumption to obtain  ∥SρL2(Y,μ)2≤C ∥ρL2(R3)2  with  C=YK(y,⋅)L22 dμ(y)<∞ . Thus  S  is bounded.

2.4.2. Equivariance (intertwining) condition

An equivariance condition

SU(g)=Γ(g)S gG(24)

is equivalent to a symmetry constraint on the kernel  K :

K(Γ(g)-1y,r)=K(y,R(g)r)  (for a.e. y,r)(25)

Indeed,

(S(U(g)ρ))(y)=K(y,r)ρ(R(g)-1r) dr=K(y,R(g)s)ρ(s) ds(26)

while

(Γ(g)Sρ)(y)=(Sρ)(Γ(g)-1y)=K(Γ(g)-1y,s)ρ(s) ds(27)

Equality for all  ρ  forces the kernel relation above. Under that relation,  S  is an intertwiner and thus maps symmetry-adapted subspaces into symmetry-adapted subspaces.

2.5. Decomposition into irreducibles and projector formulas

2.5.1. Complete reducibility for finite group

For a finite group  G  and a unitary representation  U  on a complex Hilbert space   , Maschke’s theorem implies    decomposes as a (Hilbert) direct sum of finite-dimensional  G -invariant subspaces each of which splits into irreducible representations (isotypic decomposition) [2]:

λ∈G^λ(28)

where  G^  denotes the set of (equivalence classes of) irreducible representations and  Hλ  is the  λ -isotypic component.

2.5.2. Character projector (explicit orthogonal projector)

Let  χ  be the character of an irreducible unitary representation  Vχ  of dimension  dχ . Then the orthogonal projector onto the  χ -isotypic component of    is

Pχ=dχ|G|gGχ(g)¯ U(g)(29)

One may verify  Pχ2=Pχ ,  Pχ*=Pχ , and  PχU(h)=U(h)Pχ  for all  h∈G ; hence  Pχ  projects orthogonally onto an invariant subspace isomorphic to a direct sum of copies of  Vχ .

3. Irreducible representations and projectors

3.1. Irreps and character table of D2

3.1.1. Why all irreps are one-dimensional

The group  D2  considered here is the dihedral group of order  4 , which is isomorphic to the direct product

D2  C2×C2 = { e, α, β, αβ }, α2=β2=e,  αβ=βα(30)

Since  D2  is abelian, each element commutes with every other. A standard theorem in representation theory states:

> For a finite abelian group  G , every irreducible representation over  C  is one-dimensional.

The intuition is that the group algebra  C[G]  of an abelian group splits as a direct sum of one-dimensional eigenspaces, and thus irreps can only be  1 D characters.

Therefore, all irreps of  D2  are homomorphisms

χ:D2{+1,-1}(31)

3.1.2. Conjugacy classes and the number of irreps

For any group, the number of irreducible representations equals the number of conjugacy classes. Since  D2  is abelian, each element forms a conjugacy class by itself:

e, α, β, αβ(32)

Thus  D2  has  4  inequivalent irreps in total.

3.1.3. Explicit construction of irreps

Each one-dimensional character  χ  is determined by its values on the generators  α,β . There are  2  choices for  χ(α)  ( +1  or  -1 ), and  2  choices for  χ(β) , hence  2×2=4  distinct irreps in total.

Explicitly:

(χ(α),χ(β))(1,1), (1,-1), (-1,1), (-1,-1)(33)

The corresponding value on  αβ  is then determined multiplicatively:

χ(αβ)=χ(α)χ(β)(34)

3.1.4. The character table

It is conventional to label these four irreps as  A1,B1,B2,B3 , giving the table [2]:

Table 1. Four irrep ρ

Irrep  ρ 

χρ(e) χρ(α) χρ(β) χρ(αβ)
A1 1 1 1 1
B1 1 1 -1 -1
B2 1 -1 1 -1
B3 1 -1 -1 1

3.1.5. Orthogonality sanity check

Characters of irreps are orthogonal with respect to the inner product

χρ,χσ = 1|D2|g∈D2χρ(g)¯ χσ(g)(35)

Since all values are real ( ±1 ), orthogonality reduces to checking that each pair of rows in the table has dot product zero in  R4 , which is immediate. Thus the table is consistent.

3.2. Reynolds/character projectors: derivation and properties

3.2.1. General formula

Let  U:D2→GL(V)  be a unitary representation with character  χU(g)=Tr U(g) . For each irrep  ρ  with character  χρ , the associated projector is [2]

Pρ = dimρ|D2|gD2χρ(g)¯ U(g)(36)

3.2.2. Simplification for D2

Here  dimρ=1  and  χρ(g)∈{±1}  is real, so

PA1=1/4(I+U(α)+U(β)+U(αβ)),PB1=1/4(I+U(α)-U(β)-U(αβ)),PB2=1/4(I-U(α)+U(β)-U(αβ)),PB3=1/4(I-U(α)-U(β)+U(αβ)).(37)

Each projector is simply a linear combination of the four representation matrices, with coefficients  ±1/4 .

3.2.3. Why these are projectors

We check the key properties:

Idempotence:  Pρ2=Pρ . This follows from Schur orthogonality of characters:

1|D2|gD2χρ(g-1) χσ(hg)=δρσdimρ χρ(h)(38)

A direct substitution shows  Pρ2=Pρ .

Mutual orthogonality:  PρPσ=0  if  ρ≠σ . This again comes from the orthogonality of characters.

Resolution of identity:

ρPρ=I(39)

Thus the representation space splits as a direct sum of the four invariant subspaces.

Commutation:  U(h)Pρ=PρU(h)  for all  h∈D2 . Hence each  Pρ  is an intertwiner and respects the group symmetry.

3.2.4. Self-adjointness

If  U  is unitary, then

Pρ*=1|D2|gχρ(g) U(g)*=1|D2|gχρ(g) U(g-1)(40)

Since  χρ(g-1)=χρ(g)  for one-dimensional real characters, this equals  Pρ . Hence  Pρ  is Hermitian and therefore an orthogonal projector.

3.2.5. Multiplicity formula

The multiplicity  mρ  of  ρ  in  U  is

mρ = 1|D2|gD2χρ(g)¯ χU(g)(41)

Equivalently,  rank(Pρ)=mρ . This gives a direct computational method to decompose  U .

3.3. Concrete D2 representations in LDHA: two working examples

3.3.1. Example A: chain representation on R4

Group action. Label four chains  (A,B,C,D)  as standard basis vectors. Define:  α: A↔B C↔Dβ: A↔C B↔D 

so that  αβ  swaps  A↔D  and  B↔C . The  U(g)  are  4×4  permutation matrices.

Character computation. The trace of each permutation matrix equals the number of fixed points:  χU(e)=4, χU(α)=χU(β)=χU(αβ)=0. 

Decomposition. Using multiplicities,

mρ=1/4gχρ(g)χU(g)(42)

we obtain

mA1=mB1=mB2=mB3=1(43)

Hence

R4  A1B1B2B3(44)

Explicit basis. An orthogonal eigenbasis realizing the decomposition is

A1: (1,1,1,1), B1: (1,1,-1,-1),B2: (1,-1,1,-1), B3: (1,-1,-1,1).(45)

Applying the projectors and extracts each component.

3.3.2. Example B: interface representation on R6

Group action. List six interfaces:

(AB,AC,AD,BC,BD,CD)(46)

The action of  α,β,αβ  permutes these as:

α: ABAB,  ACBD,  ADBC,  CDCD,β: ABCD,  ACAC,  ADBC,  BDBD,αβ: ABCD,  ACBD,  ADAD,  BCBC.(47)

Characters. Counting fixed points:

χU(e)=6, χU(α)=χU(β)=χU(αβ)=2(48)

Decomposition. Multiplicities:

mA1=1/4(6+2+2+2)=3,mB1=1/4(6+2-2-2)=1,mB2=1/4(6-2+2-2)=1,mB3=1/4(6-2-2+2)=1.(49)

So

R6  3A1B1B2B3(50)

Orbit structure and basis. Interfaces split into three orbits:

AB,CD, AC,BD, AD,BC(51)

This yields:

A1 basis: s1=AB+CD,  s2=AC+BD,  s3=AD+BC,B1 basis: b1=AB-CD,B2 basis: b2=AC-BD,B3 basis: b3=AD-BC.(52)

Thus the space naturally decomposes into symmetric orbit sums ( A1 ) and antisymmetric differences ( Bi ).

Projector action. For any  w∈R6 ,

w(ρ)=Pρw(53)

where  w(A1)  averages over each orbit and  w(Bi)  extracts the signed difference. This provides an explicit computational recipe.

3.4. Applications of the projectors

3.4.1. Block diagonalization of equivariant operators

If  K  is  D2 -equivariant ( U(g)K=KU(g) ), then

K = ρPρKPρ  ρKρ,  Kρ=PρKPρ|ImPρ(54)

Thus  K  block-diagonalizes into irreducible sectors.

3.4.2. Numerical recipe

Given  U(α),U(β) , construct the four projectors. Then compute

w(ρ)=Pρw, Cρ=PρCPρ, Kρ=PρKPρ(55)

Subsequent log-determinant and free energy formulas decompose accordingly.

3.4.3. Stability

Because  Pρ  are orthogonal projectors with rational coefficients, the decomposition is numerically robust: the projections are exact up to floating-point error. For  D2 , all coefficients are  ±1/4 , ensuring excellent conditioning.

4. Explicit representations on subunit and interface spaces

4.1. Subunit (4D) representation: construction, checks, and decomposition

4.1.1. Basis and ordering

Let  Vsub=R4  with the standard basis corresponding to chains  (A,B,C,D) :

eA=(1,0,0,0), eB=(0,1,0,0), eC=(0,0,1,0), eD=(0,0,0,1)(56)

We represent a general vector as  v=(xA,xB,xC,xD) .

4.1.2. Group action by permuting chains

We define the  D2  action by permuting chain labels (orthogonal maps in the Euclidean inner product):

α: AB, CD,β: AD, BC,αβ: AC, BD.(57)

Each action is a permutation of the basis and is therefore represented by a permutation matrix  U(g)  satisfying  U(g)U(g)=I .

4.1.3. Building the matrices entry-by-entry

A permutation matrix  U(g)  is determined by where  g  sends each basis vector. For example,  α  swaps  A↔B  and  C↔D , hence

U(α)eA=eB, U(α)eB=eA, U(α)eC=eD, U(α)eD=eC(58)

so the columns of  U(α)  are the images of the basis vectors:

U(α)=[0100100000010010](59)

Proceeding identically for  β  and  αβ  gives

U(β)=[0001001001001000],  U(αβ)=[0010000110000100](60)

4.1.4. Representation sanity checks

We verify the group relations on matrices:

U(α)2=I, U(β)2=I, U(α)U(β)=U(β)U(α)=U(αβ)(61)

These hold because applying the swaps twice returns each chain to itself, and  α,β  commute on labels, so their permutation matrices commute.

4.1.5. Characters and quick multiplicity count

The character of a permutation is the number of fixed basis vectors (its trace). For the identity,  χU(e)=Tr I=4 . For each non-identity above, no chain is fixed, so  χU(α)=χU(β)=χU(αβ)=0 . Using the multiplicity formula

mρ=1|D2|gD2χρ(g) χU(g), |D2|=4(62)

and the  D2  character table, one obtains  mA1=mB1=mB2=mB3=1 . Therefore

 VsubA1B1B2B3. (63)

4.1.6. An explicit irrep eigenbasis (and why it works)

Define the four vectors

uA1=1/2[1111], uB1=1/2[11-1-1], uB2=1/2[1-11-1], uB3=1/2[1-1-11](64)

Each  uρ  is an eigenvector of every  U(g)  with eigenvalue  χρ(g)  (the one-dimensional irrep property). For instance,

U(α) uB2=1/2[0100100000010010][1-11-1]=1/2[-11-11]=- uB2(65)

while  U(β) uB2=+uB2  and  U(αβ) uB2=-uB2 , matching the character row   [1,-1,1,-1]  of  B2 . Orthogonality is immediate from the sign patterns, and the  1/2  factor normalizes the vectors to unit length.

4.1.7. Projectors (closed form matrices)

Because  D2  is abelian with real characters,

PA1=1/4(I+U(α)+U(β)+U(αβ))=1/411,PB1=1/4(I+U(α)-U(β)-U(αβ)),PB2=1/4(I-U(α)+U(β)-U(αβ)),PB3=1/4(I-U(α)-U(β)+U(αβ)),(66)

where  1=(1,1,1,1) . Explicitly,

PA1=[14141414141414141414141414141414], PB1=[1414-14-141414-14-14-14-141414-14-141414],PB2=[14-14-1414-141414-14-141414-1414-14-1414], PB3=[14-1414-14-1414-141414-1414-14-1414-1414].(67)

For any  v∈R4 , the four components  v(ρ)=Pρv  lie on the irrep lines and sum to  v .

4.2. Interface (6D) representation: construction, checks, and decomposition

4.2.1. Basis and ordering

Let  Vint=R6  with the basis listing the unordered interfaces in the fixed order

(AB, BC, CD, DA, AC, BD)(68)

Denote a vector as

w=(wAB,wBC,wCD,wDA,wAC,wBD)(69)

4.2.2. How to build each permutation matrix

Given a group element  g , map each unordered edge (e.g.  AB ) by acting on its endpoints (e.g.  A↦g(A) ,  B↦g(B) ), then relabel the resulting unordered edge in our fixed order. Place a  1  in row “image index” and column “original index”. Doing this for all six edges yields the  6×6  permutation matrix  U(g) .

4.2.3. Explicit edge mappings

Using your subunit actions,

α:AB, CD,  β:AD, BC,  αβ:AC, BD(70)

we obtain the following maps (written as “original edge    image edge”):

Table 2. Edge mappings

Edge

AB BC CD DA AC BD
α AB DA CD BC BD AC
β CD BC AB DA BD AC
αβ CD DA AB BC AC BD

Reading column-by-column (image of basis vectors) gives

U(α)=[100000000100001000010000000001000010], U(β)=[001000010000100000000100000001000010], U(αβ)=[001000000100100000010000000010000001](71)

4.2.4. Representation checks

Each  U(g)  is orthogonal ( U(g)U(g)=I ) since it permutes coordinates. Moreover,

U(α)2=U(β)2=I, U(α)U(β)=U(β)U(α)=U(αβ)(72)

so we indeed have a (unitary) representation of  D2 .

4.2.5. Characters (why the traces are 6,2,2,2 )

·  e  fixes all  6  edges:  χU(e)=6 .

·  α  fixes  AB  and  CD , swaps  BC↔DA  and  AC↔BD :  χU(α)=2 .

·  β  fixes  BC  and  DA , swaps  AB↔CD  and  AC↔BD :  χU(β)=2 .

·  αβ  fixes  AC  and  BD , swaps  AB↔CD  and  BC↔DA :  χU(αβ)=2 .

4.2.6. Multiplicity calculation spelled out

With the  D2  character table  {χA1,χB1,χB2,χB3}  and the above  χU ,

mρ=14(χρ(e)⋅6+χρ(α)⋅2+χρ(β)⋅2+χρ(αβ)⋅2)(73)

so

mA1=3,  mB1=1,  mB2=1,  mB3=1(74)

Hence

 Vint≅3A1 ⊕ B1 ⊕ B2 ⊕ B3. (75)

4.2.7. Orbit structure and immediate block basis

The six edges split into three  D2 -orbits of size  2 :

AB,CD,  BC,DA,  AC,BD(76)

For each orbit, form the symmetric (sum) and antisymmetric (difference) combinations:

s1=AB+CD, b1=AB-CD,s2=BC+DA, b2=BC-DA,s3=AC+BD, b3=AC-BD.(77)

Then  s1,s2,s3  span the  3A1  subspace (fixed by all  U(g) ), and  b1,b2,b3  are one-dimensional eigenlines carrying  B1,B2,B3  respectively. Indeed,

eαβαβb1: 11-1-1  (row B1)b2: 1-11-1  (row B2)b3: 1-1-11  (row B3)(78)

because  α  fixes  AB,CD  while  β  and  αβ  swap them, etc.

4.2.8. Projectors as explicit 6×6 matrices

Using

Pρ=1/4(I+χρ(α)U(α)+χρ(β)U(β)+χρ(αβ)U(αβ))(79)

we obtain (block-diagonal over the three orbits):

PA1=[120120000120120012012000012012000000121200001212], PB1=[120-12000000000-12012000000000000000000000](80)

PB2=[0000000120-12000000000-1201200000000000000], PB3=[000000000000000000000000000012-120000-1212](81)

These satisfy  Pρ2=Pρ ,  PρPσ=δρσPρ , and  ρPρ=I .

4.2.9. Worked projection formulas

For  w=(wAB,wBC,wCD,wDA,wAC,wBD) ,

PA1w=(wAB+wCD/2, wBC+wDA/2, wAB+wCD/2, wBC+wDA/2, wAC+wBD/2, wAC+wBD/2),PB1w=(wAB-wCD/2, 0, -wAB-wCD/2, 0, 0, 0),PB2w=(0, wBC-wDA/2, 0, -wBC-wDA/2, 0, 0),PB3w=(0, 0, 0, 0, wAC-wBD/2, -wAC-wBD/2).(82)

Thus  w=w(A1)+w(B1)+w(B2)+w(B3)  with each  w(ρ)  living in the claimed irrep block.

4.2.10. Why orthogonality holds

Since each  Pρ  is a polynomial in the commuting orthogonal matrices  U(g)  with real coefficients, we have  Pρ=Pρ  and  PρPσ=δρσPρ . Hence images of different  Pρ  are mutually orthogonal subspaces, providing a numerically stable decomposition for downstream energy/covariance block-diagonalization.

5. Symmetry-resolved free energy

We present two equivalent routes (Hessian / Gaussian route and statistical / covariance route), prove their equivalence under the harmonic approximation, treat practical numerical issues (zero modes, regularization, finite-sample bias), and define diagnostic irrep scores that attribute the free-energy change to symmetry sectors.

5.1. Harmonic (Hessian) route

5.1.1. Local quadratic approximation

Let  x∈Rd  denote local internal coordinates (collective coordinates, normal-mode coordinates, or small displacements) measured relative to a stable configuration  μG  that depends on the symmetry state  G∈{D2,C2} . Taylor-expand the potential energy about  μG  up to second order [4]:

EG(x) = E0(G) + 1/2(x-μG)KG(x-μG) + O(x-μG3)(83)

where  KG=2EG(μG)  is the symmetric positive-definite Hessian (we assume a local minimum so  KG≻0 ; see zero-mode handling below).  E0(G)  is the potential energy minimum (baseline).

5.1.2. Partition function in the quadratic approximation

The canonical partition function (restricting to coordinates  x ) is

Z(G)=RdeEG(x) dxeE0(G)Rdexp(-β/2(x-μG)KG(x-μG)) dx(84)

The Gaussian integral is standard:

Rdexp(-β/2xKx) dx=(2π/β)d/2 (detK)-1/2=(2πkBT)d/2 (detK)-1/2(85)

Therefore

Z(G)eE0(G) (2πkBT)d/2 (detKG)-1/2(86)

5.1.3. Free energy and logdetk term

Take  F(G)=-kBTlnZ(G)  to obtain (up to additive constants independent of  G )

F(G)  E0(G) + kBT2 lndetKG + const(87)

Here “const” contains  (d/2)kBTln(2πkBT)  and any Jacobian factors from coordinate choices; since we always compute differences between  G  states, such  G -independent constants cancel.

5.1.4. Symmetry and block-diagonalization

Suppose  U(g)  is the orthogonal/unitary representation of  G  acting on the coordinate space so that  EG  is  G -invariant in the sense  EG(U(g)x)=EG(x)  whenever  g  is a symmetry of that configuration. If the Hessian itself is  G -invariant (i.e.  U(g)KGU(g)=KG  for all  g  in the symmetry group of that structure), then  KG  commutes with the representation operators  U(g) :

[KG, U(g)]=0, gG(88)

By Schur’s lemma and the general theory of finite-group representations (Maschke’s theorem),  V=Rd  decomposes into isotypic components corresponding to irreducible representations  ρ :

V = ρ VρCmρisotypic component(89)

and any operator commuting with  U(g)  is block-diagonal with respect to this decomposition. Concretely, let  Pρ  be the orthogonal projector onto the  ρ -isotypic subspace (constructed via the character projector). Then

KG = ρKρ,G,  Kρ,G:=PρKGPρ acting on Im(Pρ)(90)

Each block  Kρ,G  is itself symmetric positive-definite on its subspace.

5.1.5. Dimension bookkeeping

Let  nρ=rank(Pρ)  denote the effective dimensionality of the  ρ -block (for abelian groups  nρ  equals the multiplicity). Then  ρnρ=d .

5.1.6. Factorization of the gaussian integral

Because  KG  is block-diagonal in this orthogonal decomposition, the determinant factorizes:

detKG=ρdetKρ,G(91)

The symmetry-resolved free energy:

F(G)E0(G)+kBT2ρlndetKρ,G+const(92)

5.1.7. Symmetry-resolved free-energy difference

Subtracting the  D2  and  C2  expressions gives

ΔFF(C2)-F(D2)=kBT2ρ[lndetKρ,C2-lndetKρ,D2] + ΔE0(93)

where  ΔE0E0(C2)-E0(D2)  is the baseline potential-energy difference (can be approximated from enthalpic terms such as FoldX).

5.2. Statistical (Covariance) route

5.2.1. Linear-response relation between hessian and covariance

Under harmonic fluctuations at temperature  T , equipartition and Gaussian statistics give [4]

CG = (x-μG)(x-μG) = kBT KG-1(94)

provided samples are drawn from the quadratic Boltzmann weight  exKGx/2 . This is the standard fluctuation–dissipation relation.

5.2.2. Express lndetK via lndetC

Taking determinants and logarithms on each  ρ -block yields

lndetKρ,G=-lndetCρ,G+nρln(kBT)(95)

where  nρ=rank(Pρ) .

Substituting and absorbing the  nρln(kBT)  terms into the baseline gives the covariance form

 ΔF ≈ -kBT2ρ[lndetCρ,C2-lndetCρ,D2] + ΔE0', (96)

where  ΔE0'  differs from  ΔE0  by the additive constant  kBT/2ρnρln(kBT)  (which cancels in comparisons or can be included in the baseline).

Thus the Hessian and covariance routes are equivalent under the harmonic approximation and when  C  and  K  are invertible on the projected subspaces.

5.2.3. Eigenvalue (mode) representation

Let  {λρ,i(G)}i=1nρ  denote the positive eigenvalues of  Kρ,G  (Hessian modes in the  ρ  channel). Then

lndetKρ,G = i=1nρlnλρ,i(G)(97)

Equivalently, for covariance eigenvalues  {σρ,i2,(G)}  (so  σ2=kBT/λ ),

lndetCρ,G = i=1nρlnσρ,i2,(G)(98)

Therefore each mode contributes additively to  F(G)  and to  ΔF ; this allows per-mode attribution.

5.3. Diagnostic irrep scores and linear-response approximation

5.3.1. Definition of diagnostic scores

We define two complementary irrep-resolved diagnostics:

Δρ(spec):=lndet(Cρ,C2Cρ,D2-1)=lndetCρ,C2-lndetCρ,D2,Δρ(var):=tr(Cρ,C2-Cρ,D2).(99)

The contribution of irrep  ρ  to  ΔF  (covariance form) is [4]

ΔFρ = -kBT2 Δρ(spec)(100)

Large positive  Δρ(spec)  (i.e. larger  detC  in  C2 ) yields a negative contribution to  ΔF  (i.e. stabilization of the  C2  basin relative to  D2 ), and vice versa.

5.3.2. First-order (linear-response) sensitivity of lndet

If  C  is perturbed by a small symmetric  δC , then

lndet(C+δC) = lndetC + tr(C-1δC) - 1/2tr(C-1δC C-1δC)+O(δC3)(101)

Thus, when differences between  Cρ,C2  and  Cρ,D2  are small, the leading contribution is

Δρ(spec)tr(Cρ,D2-1 (Cρ,C2-Cρ,D2))(102)

This gives a practical linearized approximation to  ΔFρ  and shows the direct connection between  Δρ(var)  (trace difference) and  Δρ(spec)  via the inverse covariance weighting.

5.3.3. Interpretation of signs

If  Cρ,C2Cρ,D2  in a generalized sense (more variance in many directions), then  lndetCρ,C2>lndetCρ,D2 , so  Δρ(spec)>0  and  ΔFρ<0 : the  C2  basin gains entropic stabilization in channel  ρ .

Conversely, reduced variance in  C2  relative to  D2  (tightening) gives positive  ΔFρ  (destabilization of  C2 ).

5.4. Practical computation: projection, eigen-decomposition, and numerics

5.4.1. Algorithmic recipe

1. Choose feature / CV vector  w∈Rm  that captures interface energies, contact counts, active-site geometries, etc., and collect  N  samples  w(1),…,w(N)  from MD/ENM/resampled FoldX ensembles for each state  G .

2. Compute empirical covariance  C^G=1N-1j=1N(w(j)-w)(w(j)-w) .

3. Construct representation matrices  U(g)  on the CV space (permutation/sign or linear action) and form projectors  Pρ=dimρ|G|gχρ(g)¯U(g) .

4. Compute projected covariance blocks  C^ρ,G=PρC^GPρ .

5. Compute  slogdet(C^ρ,G)  (numerically stable log-determinant via e.g. eigenvalues or Cholesky with regularization); accumulate

ΔF^ = -kBT2ρ[slogdet(C^ρ,C2)-slogdet(C^ρ,D2)] + ΔE0^(103)

6.Estimate confidence intervals by bootstrap resampling of the sample set  {w(j)}  (resample replicates, recompute  C^ρ,G  and  ΔF^ ).

5.4.2. Zero modes and coordinate gauge

Physical systems contain trivial zero modes (overall translations and rotations) that give zero eigenvalues in  K  and divergences in  detK . Remedies:

Project out rigid-body modes from  w  (work in internal/coarse-grained coordinates) or perform computations in internal coordinates where rigid motions are absent.

Remove near-zero eigenvalues before computing log-determinant, i.e. compute the product over nonzero eigenvalues only, or add a small regularizer  εI  and track the dependence on  ε .

5.4.3. Regularization and finite-sample stability

Empirical  C^  may be rank-deficient or ill-conditioned when  N  is not much larger than the projected dimension  pρ . Use:

Ridge regularization:  C^ρ,G(ε)=C^ρ,G+εI  with  ε>0  small; compute  lndet  of this regularized matrix. Choose  ε  by cross-validation or L-curve inspection.

Shrinkage estimators (Ledoit–Wolf):  C~=(1-λ)C^+λT  with target  T  (e.g. diagonal), gives lower-variance  lndet  estimates.

Dimensionality reduction: retain only principal components that capture a large fraction of variance within each  ρ  block; compute  lndet  on the reduced block (adds a model selection step).

5.4.4. Stable evaluation of lndet

Compute the log-determinant via  slogdet  routines (Cholesky if positive-definite, or eigen-decomposition)

lndetC=i=1plnλi(104)

where  {λi}  are eigenvalues. Use numerically stable libraries (e.g. LAPACK routines) and avoid forming full dense inverses.

5.5. Baseline energy ΔE0 and mapping FoldX outputs

5.5.1. What ΔE0 represents

 ΔE0  is the difference in basin minima energies  E0(C2)-E0(D2) . In practice one often uses empirical or computed enthalpic proxies (FoldX total or interface energies) as an approximation:

ΔE0EFoldX(C2)-EFoldX(D2)(105)

with the caveat that FoldX energies are not exact free energies (lack full entropy).

5.5.2. Character-weighted baseline separation

To retain symmetry attribution in the baseline enthalpy, decompose per-group-element energies  E(g)  via character inner products:

χρ,EG = 1|G|g∈Gχρ(g)¯ E(g)(106)

A simple model for baseline difference is

ΔE0  ρdimρ|G|(χρ,EC2-⟨χρ,ED2)(107)

i.e. project the FoldX energies into irrep channels and sum the channel differences. Treat this as an enthalpic proxy to be combined with the fluctuation-derived terms [5,6].

5.6. Final expressions and per-irrep contributions

5.6.1. Final covariance-based formula

 ΔF^ = -kBT2ρ[lndetC^ρ,C2-lndetC^ρ,D2] + ΔE0^, (108)

with  C^ρ,G=PρC^GPρ  and  ΔE0^  the chosen baseline enthalpy difference.

5.6.2. Per-irrep free-energy contribution

Define

ΔFρ = -kBT2 [lndetCρ,C2-lndetCρ,D2] + ΔE0,ρ(109)

so that  ΔF=ρΔFρ . Here  ΔE0,ρ  denotes the irrep-resolved baseline enthalpy term (from FoldX projection).

5.6.3. First-order attribution using the linear approximation

If  δCρ:=Cρ,C2-Cρ,D2  is small,

ΔFρ-kBT2tr(Cρ,D2-1δCρ)+ΔE0,ρ(110)

This linear form is useful to identify dominant directions using the eigenvectors of  Cρ,D2-1  (i.e. high-sensitivity directions).

6. From free energy to efficiency: a Transition-State-Theory (TST) bridge

Notation and preliminaries. We denote by  G∈{D2,C2}  the symmetry group of the enzyme assembly (tetramer vs dimer). For a given symmetry state  G  we write:

·  QR(G)  (or  ZR(G) ) for the reactant-basin partition function (reactant ensemble),

·  Q(G)  for the transition-state (TS) partition function associated with the reactive dividing surface,

·  FR(G)=-kBTlnQR(G)  for the reactant free energy, and

·  F(G)=-kBTlnQ(G)  for the TS free energy.

We also define the free-energy difference already used in the manuscript:

ΔF = FR(C2)-FR(D2)(111)

6.1. TST basic formula and species ratio

Transition-state theory gives (up to the usual prefactor and a transmission coefficient) [7]

k(G) = kBTh κ(G) Q(G)QR(G) = kBTh κ(G) exp(-βΔG(G))(112)

where  κ(G)∈(0,1]  is the transmission (or recrossing) coefficient for state  G  and

ΔG(G) = F(G)-FR(G)(113)

is the activation free energy measured relative to the reactant basin.

Now compare dimer  (C2)  and tetramer  (D2) . Define also  mG  as the number of equivalent catalytic channels (active sites) per oligomer: for a homotetramer  mD2=4 , for a homodimer  mC2=2  (unless some sites are silent). The per-oligomer (or per-species) catalytic capability scales with  mG , so the ratio of specificity-like constants  (kcat/KM)  (under the rapid-equilibrium approximation for binding) may be written schematically as

 (kcat/KM)C2(kcat/KM)D2 ≈ mC2mD2κC2κD2Q(C2)/QR(C2)Q(D2)/QR(D2)=mC2mD2κC2κD2⋅exp​[(ΔGC2GD2)] . (114)

This is identical to the short boxed formula you gave; we now unpack and connect it to the  ΔF  expressions from the symmetry-resolved free-energy analysis.

6.2. Partition-function form and relation to ΔF

Using  FR(G)=-kBTlnQR(G)  and  F(G)=-kBTlnQ(G) , expands to

ΔGC2-ΔGD2=(F(C2)-F(D2))-(FR(C2)-FR(D2))(115)

Hence

Q(C2)/QR(C2)Q(D2)/QR(D2)=exp[(ΔGC2GD2)]=exp[ -β(F(C2)-F(D2))] exp[ β(FR(C2)-FR(D2))](116)

6.2.1. Interpretation

The rate ratio thus splits into two conceptually separate effects:

(TS-shift)×(reactant-baseline-shift)(117)

The reactant-baseline term contains the  ΔF  we computed in the symmetry-resolved analysis:

exp[ β(FR(C2)-FR(D2))]=exp(βΔF)(118)

Therefore the full ratio can be written as

(kcat/KM)C2(kcat/KM)D2=mC2mD2κC2κD2exp(βΔF)exp[(F(C2)-F(D2))](119)

6.2.2. Two limiting cases

1. TS is invariant under symmetry change. If the transition-state free energy is essentially the same for the two assemblies (i.e.  F(C2)≈F(D2) ), the TS-shift factor is unity and

(kcat/KM)C2(kcat/KM)D2mC2mD2κC2κD2exp(βΔF)(120)

In this scenario a higher reactant free energy  FR(C2)  (i.e.  ΔF>0 ) increases the rate of  C2  relative to  D2  because the barrier measured from the reactant basin is effectively lower.

2. Barrier shift parallels reactant shift (barrier measured in absolute energy). If  F(C2)-F(D2)≈FR(C2)-FR(D2)=ΔF  (i.e. both TS and reactant basin shift by the same absolute energy so the absolute barrier is unchanged), then the exponential terms cancel and

(kcat/KM)C2(kcat/KM)D2mC2mD2κC2κD2(121)

That is, only multiplicity and dynamical recrossing differences remain.

Important note on sign conventions: in earlier sections we defined  ΔF=FR(C2)-FR(D2) . When you see formulas of the form  exp(-βΔF)  in other parts of the manuscript, verify the context — sometimes authors report efficiency  QR  (not  ∝1/QR ). The TST result above is unambiguous once  Q  and  QR  are explicitly identified.

6.3. TS-probability surrogate: p(TS) and ΔG

A convenient experimental / data-driven surrogate for the activation free energy is the TS-region occupancy. Define a TS-like region  STS  in CV space (geometric thresholds around the dividing surface). Its Boltzmann weight relative to the reactant basin is

pTS(G) := STSe-βE(x)dxreactante-βE(x)dx  Q(G)QR(G)(122)

Hence [5]

ΔG(G) = -kBTlnQ(G)QR(G)  -kBTlnpTS(G)(123)

An immediately estimable ratio:

(kcat/KM)C2(kcat/KM)D2mC2mD2κC2κD2pTS(C2)pTS(D2)(124)

This is practical: compute or estimate  pTS  from MD or by coarse-grained CV sampling (subject to careful definition of  STS ).

6.4. Speciation / oligomeric equilibrium and observed (bulk) kinetics

6.4.1. Equilibrium between dimer and tetramer

In a preparation where both dimers (D) and tetramers (T) can exist and interconvert via

2 DT(125)

define the (dissociation-like) equilibrium constant

Kd = [D]2[T](126)

Let  Ctot  denote the total subunit concentration (monomer equivalents)

Ctot = 4 [T] + 2 [D](127)

Solving for  [D]  in terms of  Ctot  and  Kd  yields a quadratic in  d:=[D] :

4d2Kd+2d-Ctot=0(128)

Hence

d = [D] = -Kd+ Kd2+4KdCtot 4(129)

taking the physically positive root. The tetramer concentration is then

[T] = d2Kd(130)

6.4.2. Observed (bulk) catalytic rate under substrate-limited linear regime

At low substrate (initial-rate linear regime), each active site contributes approximately  (kcat/KM)[S]  to the second-order rate, so the bulk initial rate per unit volume is

v0[S](mT [T] (kcat/KM)T+mD [D] (kcat/KM)D)(131)

It is often convenient to normalize by total subunit concentration  Ctot  to obtain an observed efficiency per subunit:

(kcatKM)obs = mT [T] (kcat/KM)T+mD [D] (kcat/KM)DCtot(132)

Using the solution for  [T],[D]  above one can predict how measured bulk kinetics vary with total concentration and  Kd .

6.4.3. Alternative normalization (per oligomer molecule)

If instead one reports rate per oligomer molecule (not per subunit), define  Nmol=[T]+[D] , and the per-molecule observed efficiency is

(kcatKM)mol = mT [T] (kcat/KM)T+mD [D] (kcat/KM)D[T]+[D](133)

Choose the normalization that matches how experimental data are reported.

6.5. Assumptions, caveats, and practical recommendations

6.5.1. Assumptions made in the bridge

Rapid oligomeric equilibration: we assumed that the D  T interconversion is fast compared to catalysis (so the equilibrium distribution holds during initial-rate measurement). If not, a kinetic model including interconversion rates must be used.

Well-defined TS partition function:  Q  must be meaningfully defined (requires a reasonable dividing surface in CV space).

Separable effects: we treated multiplicity  mG , transmission  κG , and free-energy terms multiplicatively; in reality these can be coupled (e.g. interface changes may alter reaction coordinate friction and hence  κ ).

Harmonic / local approximations: when expressing  QR  via  ΔF  we typically used the Gaussian (log-det) approximation for fluctuation contributions; large anharmonic changes require more careful sampling.

7. Why the regular-character convolution fails

A tempting but incorrect formula is [8]

ΔF=?1|D2|gD2(E(g)-E(ϕ(g))) χreg(g)(134)

where  χreg  denotes the character of the regular representation of the group (here  D2 ). We unpack why this formula is mathematically and physically unsound, and we show the correct operator-based alternative.

7.1. Why the naive formula is algebraically trivial

Recall that for any finite group  G , the regular character satisfies [2]

χreg(g)={|G|,g=e,0,g≠e.(135)

Substituting this

1|D2|gD2(E(g)-E(ϕ(g))) χreg(g)=1|D2|(E(e)-E(ϕ(e))) |D2|=E(e)-E(ϕ(e))(136)

Because  ϕ  is a homomorphism with  ϕ(e)=e , the right-hand side reduces to  E(e)-E(e)=0 , so  ΔF=0  identically. Thus the formula cannot encode any nontrivial structural information — it is algebraically nullified by the properties of the regular character.

7.2. Conceptual reason: scalars vs operators

The underlying conceptual mistake is treating  E(g)  as if it were the full object controlling the free-energy change under symmetry, while the free energy (in the Gaussian/harmonic approximation) is controlled by operators (Hessians or covariance matrices) whose spectra determine  detK  or  detC  and thus the entropic part of the free energy.

Concretely:

 E(g)  is a scalar-valued function on the group elements — e.g. an interface enthalpy associated (by some choice) to the labeling induced by  g .

The free energy in the harmonic regime is

F(G)E0(G)+kBT2ρlndetKρ,G(137)

so it depends on determinants of matrices  Kρ,G  (or equivalently the spectra of covariance blocks  Cρ,G ). Spectral information cannot be recovered from a single scalar per group element.

A simple counterexample (illustrative). Consider two different Hessians  K(1)  and  K(2)  that, under some ad-hoc mapping, yield identical scalar lists  {E(g)}  but have different eigenvalue spectra. Their  lndetK  will differ, hence their  F  differ, while the scalar convolution returns zero (or the same trivial value) and misses the actual free-energy difference.

7.3. Correct object: projectors acting on operators

The correct symmetry-aware decomposition acts on operators, not scalars. Construct the character projectors

Pρ = dimρ|G|gGχρ(g)¯ U(g)(138)

and apply them to the operator of interest (Hessian  K  or covariance  C ):

Kρ,G = Pρ KG Pρ, Cρ,G = Pρ CG Pρ(139)

Then each  ρ -block contains the full spectral information for that symmetry channel, and the free-energy difference is recovered from the  logdet  of those blocks:

ΔF=-kBT2ρ[lndetCρ,C2-lndetCρ,D2]+ΔE0(140)

8. Free energy difference as a symbolic expression

8.1. Setup and notation

Let  G∈{D2,C2}  denote the oligomeric symmetry state. The six inter-subunit interface variables are collected into

wR6, μ(G)=E[wG], Σ(G)=Cov(wG)(141)

We will denote by  n  the dimension of the fluctuating subspace under consideration (here  n≤6  after removing any constrained/rigid modes). For each irreducible representation  ρ  of  D2  define the projection operator

Pρ=dimρ|D2|gD2χρ(g)¯ U(g)(142)

where  U(g)  are the permutation (orthogonal) matrices on interface space and  χρ  the characters. These satisfy [2]

PρPσ=δρσPρ,  Pρ=Pρ,  ρPρ=I(143)

We define the symmetry-resolved covariances and means by

Σρ(G)=Pρ Σ(G) Pρ,  μρ(G)=Pρ μ(G)(144)

A convenient parameterization of the mean vector is to assign one variable to each orbit of interfaces under  D2 :

μ(G)=[s1(G)s2(G)s1(G)s2(G)s3(G)s3(G)](145)

corresponding to the interfaces (AB, BC, CD, DA, AC, BD).

8.2. Gaussian (harmonic) partition function — full derivation

Let the microscopic conformational coordinate be  x∈Rd . The energy function  EG(x)  respects the symmetry action  U(g)  of group  G . The constrained partition function (or Z(G)) is expressed as a Reynolds average to ensure only  G -equivalent configurations contribute:

Z(G) = 1|G|gG Rdexp(-β EG(U(g)x)) dx,  β=(kBT)-1(146)

Assume the reactant-basin energy is quadratic about its minimum  w0  (or  μG ):

E(w)E0(G)+1/2(w-w0)KG(w-w0)(150)

with symmetric positive-semidefinite Hessian  KG  of size  n×n  (restricted to the fluctuation subspace). The reactant partition function in the harmonic approximation is

QR(G)=Z(G)=Rne-βE(w) dweE0(G)Rne-β2yKGy dy=eE0(G) (2π)n/2 (det(βKG))-1/2,(151)

where we used the Gaussian integral identity  e-1/2yAydy=(2π)n/2(detA)-1/2  for  A≻0 . Taking the free energy  F(G)=-kBTlnQR(G)  yields

F(G)=E0(G)+kBT2lndetKG+kBTn2lnβ-kBTn2ln(2π)(152)

The last two terms are  G -independent constants (they depend only on  n  and  T ) and may be absorbed into “const” in what follows.

It is often convenient to express  F  in terms of the covariance matrix

Σ(G)=yy=(βKG)-1(153)

whence  detKG=β-ndet(Σ(G))-1  and

F(G)=E0(G)-kBT2lndetΣ(G)+const(154)

This form makes the entropic role of the covariance explicit: larger covariance    larger  detΣ⇒  lower  F  (more entropy).

8.3. Symmetry-resolved blocks and factorization

If the basin/hessian  KG  is invariant under the group action (i.e.  U(g)KGU(g)=KG  for all  g ), then  KG  commutes with every  U(g) . By standard representation theory one may choose an orthonormal basis that simultaneously block-diagonalizes all  U(g)  and  KG  so that

KGρKρ,G,  Σ(G)ρΣρ(G)(155)

with

Kρ,G=PρKGPρ,  Σρ(G)=PρΣ(G)Pρ(156)

Determinants factorize over blocks, we obtain the symmetry-resolved free energy

F(G)=E0(G)-kBT2ρlndetΣρ(G)+const(157)

Subtracting the two symmetry states yields the central symbolic expression:

ΔF = F(C2)-F(D2) = ΔE0 - kBT2ρ(lndetΣρ(C2)-lndetΣρ(D2)),(158)

where  ΔE0=E0(C2)-E0(D2) . (Note: For positive-definite matrices,  lndetΣ=ln|Σ| .)

Remark on zero modes / pseudo-determinants. If  KG  has zero eigenvalues (rigid translations/rotations or constrained directions), the Gaussian integral is formally divergent. Practically one removes those rigid modes (restrict to fluctuation subspace) and uses the pseudodeterminant

pdet A=λi>0λi(159)

replacing  det . Equivalently, fix gauges or integrate out rigid coordinates—this affects only the absorbed “const” and not the difference between symmetry states if the zero-mode count is the same.

8.4. Orbit-based parameterization and explicit closed form

Label the six unordered edges as

(AB, BC, CD, DA, AC, BD)(160)

which form three  D2 -orbits of size two:

O1={AB,CD},  O2={BC,DA},  O3={AC,BD}(161)

Assume the covariance is block-diagonal by orbit (no cross-orbit covariances):

Σ(G)=diag(Σ1(G),Σ2(G),Σ3(G)),  Σi(G)=[vi(G)ci(G)ci(G)vi(G)], i=1,2,3(162)

Σ(G)=[v1(G)c1(G)0000c1(G)v1(G)000000v2(G)c2(G)0000c2(G)v2(G)000000v3(G)c3(G)0000c3(G)v3(G)](163)

Introduce normalized symmetric / antisymmetric orbit coordinates for orbit  i :

us,i=ei,1+ei,22,  ub,i=ei,1-ei,22(164)

where  ei,1,ei,2  are the standard basis vectors for the two edges in orbit  i  (e.g. for  O1 ,  e1,1=eAB,e1,2=eCD ). Compute the variances in these coordinates:

Var(us,i)=vi(G)+ci(G),  Var(ub,i)=vi(G)-ci(G)(165)

Under the orbit-decoupling assumption the projection onto irreps yields

ΣA1(G)=diag(v1(G)+c1(G), v2(G)+c2(G), v3(G)+c3(G))(166)

and the nontrivial one-dimensional irreps give

ΣB1(G)=(v1(G)-c1(G)), ΣB2(G)=(v2(G)-c2(G)), ΣB3(G)=(v3(G)-c3(G))(167)

Hence the determinants are products of these scalar entries, and substituting into explicit orbit form

ΔF=ΔE0-kBT2i=13ln(vi(C2)+ci(C2))(vi(C2)-ci(C2))(vi(D2)+ci(D2))(vi(D2)-ci(D2)).(168)

This is the closed-form symbolic expression in terms of the orbit-parameters  {si(G),vi(G),ci(G)}  and the enthalpic baseline difference  ΔE0 , enabling sensitivity analysis without specific numerical values.

Positivity constraint. For physical positive-definiteness one requires

vi(G)>|ci(G)|,  i=1,2,3,(169)

so that each orbit-block is SPD and the logarithms are well-defined.

8.5. Sensitivity (partial derivatives) — how each parameter affects ΔF

Differentiating gives closed-form sensitivities. For a fixed orbit  i  and varying the  C2  parameters,

∂ΔFvi(C2)=-kBT2(1vi(C2)+ci(C2)+1vi(C2)-ci(C2))=-kBT vi(C2)(vi(C2))2-(ci(C2))2,(170)

and

∂ΔFci(C2)=-kBT2(1vi(C2)+ci(C2)-1vi(C2)-ci(C2))=-kBT ci(C2)(vi(C2))2-(ci(C2))2.(171)

Analogous formulas (with opposite sign inside the big parentheses) hold for derivatives with respect to  vi(D2)  and  ci(D2) .

8.5.1. Interpretation

·Increasing  vi(C2)  (more variance on orbit  i  in the dimer) decreases  ΔF  if  vi(C2)>0  (entropy stabilizes  C2  relative to  D2 .

·Increasing positive covariance  ci(C2)>0  increases  ΔF .

·The sensitivity scales as  1/(v2-c2)  and therefore grows if the block becomes nearly singular; this indicates directions where small structural changes produce large free-energy effects.

8.5.2. Numerical stability and regularization

When  (vi2-ci2)  is very small, add a small Tikhonov regularizer  ϵ>0  to  Σ  (i.e. replace  Σ↦Σ+ϵI ) to stabilize logarithms and derivatives; carry this through the algebra if needed for numerical work.

The formulas above provide a fully symbolic, algebraically explicit map

{ ΔE0, si(G), vi(G), ci(G) }  ΔF(172)

suitable for sensitivity analysis, uncertainty propagation, and for guiding mutational design that targets particular orbit variances or correlations.

9. Conclusion

9.1. Summary of symbolic results

Starting from a harmonic basin approximation and symmetry-resolved block diagonalization we derived

F(G)=E0(G)-kBT2ρlndetΣρ(G)+const(173)

and therefore

ΔF=ΔE0-kBT2ρ(lndetΣρ(C2)-lndetΣρ(D2)).(174)

Under orbit-decoupling this reduces to the explicit orbit expression in terms of  {si(G),vi(G),ci(G)} .

9.2. Bridge to transition-state theory and efficiency

In the TST approximation the catalytic-efficiency ratio can be expressed schematically as

(kcat/KM)C2(kcat/KM)D2mC2mD2κC2κD2exp[-β(ΔGC2GD2)](175)

Using the identity

ΔGC2-ΔGD2=(F(C2)-F(D2))-(FR(C2)-FR(D2))(176)

and inserting the symmetry-resolved expression for  FR(G)  (i.e.  ΔF ), we isolate the reactant-baseline contribution:

(kcat/KM)C2(kcat/KM)D2=mC2mD2κC2κD2exp(βΔF)exp[(F(C2)-F(D2))](177)

Two useful limiting scenarios are:

1. TS invariant: if  F(C2)≈F(D2)  then

(kcat/KM)C2(kcat/KM)D2mC2mD2κC2κD2exp(βΔF)(178)

2. Absolute-barrier preserved: if  F(C2)-F(D2)≈ΔF  (i.e. both TS and reactant shift in parallel) then exponential terms cancel and only multiplicity and dynamical factors remain:

(kcat/KM)C2(kcat/KM)D2mC2mD2κC2κD2(179)

Absorbing the main symmetry effect into  ΔF , a simplified statistical–kinetic bridge for the efficiency ratio  η=kcat/KM  is

 ηC2ηD2 ≈ (|C2||D2|)1/2exp​(-ΔFkBT) . (180)

The prefactor  |C2|/|D2|  reflects reduced symmetry volume, and the exponential encodes the thermodynamic penalty.

9.3. Connection to FoldX data

For each group element  g , let  E(g)  denote FoldX energy components. Define the character inner product:

χρ,EG=1|G|g∈Gχρ(g)¯ E(g)(181)

ΔE0ρdimρ|G|( ⟨χρ,EC2-⟨χρ,ED2)(182)

Similarly,  Σρ,G=PρΣ(G)Pρ  with  w  the feature vector ensemble.

9.4. Final synthesis

(i) Pρ=dimρ|G|gχρ(g)¯U(g), Σρ,G=Pρ Σ(G) Pρ;(ii) ΔF=-kBT2ρ[lndetΣρ,C2-lndetΣρ,D2]+ΔE0;(iii) ηC2ηD2(|C2||D2|)1/2exp(-ΔF/kBT).(183)

Acknowledgements

First and foremost, I would like to thank my parents for their unconditional support at all times and for inspiring me in mathematics and making me decide to take this path. Standing on their shoulders, I was able to glimpse higher and farther into the future, and embrace more possibilities. It was my parents who, with their unconditional love and tolerance, embraced every moment of my confusion, exhaustion, and anxiety, mended and healed the cracks in my life, and transformed the hardships and sufferings of my growth into a warm spiritual journey. It was they who made me understand that I never walk alone. No matter where I go, no matter whether I succeed or fail, there is always a light waiting for me to return home, giving me the courage and confidence to accept my imperfections and move on to the next intersection in life.

I would like to thank the professors who have guided me over the past four years, Professor Espinoza and Professor Kjeer. They have led me into the vast and profound world of mathematics, allowing me to experience the warmth of the wisdom left by our predecessors. It was their trust and recognition that soothed my anxieties and allowed me to see my own strengths. They have always accompanied me on my academic journey, leaving each tender commentary on my immature and clumsy steps. Even amidst my own setbacks, they have taught me that "there is no need to fear the infinite truth; every step forward brings its own joy." That ever-burning light has illuminated my small corner of refuge and given me the resilience to pursue my studies.

I would also like to thank my high school friend Yang Xilin for his help in biology during my research and thesis.


References

[1]. Gallian, J. A. (1990). Contemporary abstract algebra. New Delhi: Cengage Learning.

[2]. Fulton, W., & Harris, J. (2013). Representation theory: a first course (Vol. 129).Springer Science & Business Media.

[3]. Conway, J. B. (2019). A course in functional analysis (Vol. 96).Springer.

[4]. Horn, R. A., & Johnson, C. R. (2012). Matrix analysis. Cambridge University Press.

[5]. Berman, H. M., Westbrook, J., Feng, Z., Gilliland, G., Bhat, T. N., Weissig, H., ... & Bourne, P. E. (2000). The protein data bank.Nucleic Acids Research, 28(1), 235–242.

[6]. Berman, H. M., Henrick, K., Kleywegt, G., Nakamura, H., & Markley, J. (2012). RCSB PDB.Nucleic Acids Research, 40(D1), D445–D452.

[7]. Truhlar, D. G., Garrett, B. C., & Klippenstein, S. J. (1996). Current status of transition-state theory.The Journal of Physical Chemistry, 100(31), 12771–12800.

[8]. Tinkham, M. (2003). Group theory and quantum mechanics. Courier Corporation.


Cite this article

Qiu,T. (2025). Symmetry reduction from D2 tetramer to C2 dimer: group-theoretic and thermodynamic modeling of enzyme catalytic efficiency. Advances in Operation Research and Production Management,4(3),1-34.

Data availability

The datasets used and/or analyzed during the current study will be available from the authors upon reasonable request.

Disclaimer/Publisher's Note

The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of EWA Publishing and/or the editor(s). EWA Publishing and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

About volume

Journal:Advances in Operation Research and Production Management

Volume number: Vol.4
Issue number: Issue 3
ISSN:3029-0880(Print) / 3029-0899(Online)

© 2024 by the author(s). Licensee EWA Publishing, Oxford, UK. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license. Authors who publish this series agree to the following terms:
1. Authors retain copyright and grant the series right of first publication with the work simultaneously licensed under a Creative Commons Attribution License that allows others to share the work with an acknowledgment of the work's authorship and initial publication in this series.
2. Authors are able to enter into separate, additional contractual arrangements for the non-exclusive distribution of the series's published version of the work (e.g., post it to an institutional repository or publish it in a book), with an acknowledgment of its initial publication in this series.
3. Authors are permitted and encouraged to post their work online (e.g., in institutional repositories or on their website) prior to and during the submission process, as it can lead to productive exchanges, as well as earlier and greater citation of published work (See Open access policy for details).

References

[1]. Gallian, J. A. (1990). Contemporary abstract algebra. New Delhi: Cengage Learning.

[2]. Fulton, W., & Harris, J. (2013). Representation theory: a first course (Vol. 129).Springer Science & Business Media.

[3]. Conway, J. B. (2019). A course in functional analysis (Vol. 96).Springer.

[4]. Horn, R. A., & Johnson, C. R. (2012). Matrix analysis. Cambridge University Press.

[5]. Berman, H. M., Westbrook, J., Feng, Z., Gilliland, G., Bhat, T. N., Weissig, H., ... & Bourne, P. E. (2000). The protein data bank.Nucleic Acids Research, 28(1), 235–242.

[6]. Berman, H. M., Henrick, K., Kleywegt, G., Nakamura, H., & Markley, J. (2012). RCSB PDB.Nucleic Acids Research, 40(D1), D445–D452.

[7]. Truhlar, D. G., Garrett, B. C., & Klippenstein, S. J. (1996). Current status of transition-state theory.The Journal of Physical Chemistry, 100(31), 12771–12800.

[8]. Tinkham, M. (2003). Group theory and quantum mechanics. Courier Corporation.