Jakob Schwichtenberg

Resources that helped me understand Grand Unified Theories

I recently finished my master’s thesis on dark matter in Grand Unified Theories. Here are some resources that I found particularly helpful.

Group Theoretical Preliminaries

Unification means that we embed the standard model gauge group $G_{SM} \equiv SU(3)_C \times SU(2)_L \times U(1)_Y$ in a larger gauge group $G_{GUT}$.

Thus the first important questions for me were:

  • Which groups $G_{GUT}$ can be used?
  • How can we embed $G_{SM}$ in $G_{GUT}$ and what does this actually mean?
  • How can we describe and compute the breaking of $G_{GUT}$ to $G_{SM}$? (Although there are other methods to break a symmetry, I restricted myself to the usual Higgs mechanism.)

The answer to the first question is, that we use simple groups that have non-self-conjugate representations and that are large enough such that $G_{SM}$ can be embedded. I’ve written a long post about the classification of all simple groups. A representation of a Lie group is called non-self-conjugate if it is not equivalent to the corresponding conjugated representation. This means the group elements are represented by complex matrices $R^a$ and it is impossible to get the conjugated matrices $\overline{R}^a$ from the original matrices using a similarity transformation $U R^a U^\dagger \neq \overline{R}^a$. If there is a map $U R^a U^\dagger= \overline{R}^a  $ the representation is called self-conjugate. In physics, it is conventional to call a self-conjugate representation real and a non-self-conjugate representation complex, although these notions have in mathematics a different meaning. The short version is that GUT models that put the fermions in a self-conjugate representation “will not give the observed chiral structure of weak interactions”. I was quite confused about this for several weeks because I couldn’t find a good explanation. What made it finally click for me was

The Algebra of Grand Unified Theories by John C. Baez and John Huerta

The answer to the second question is that embedding means that we identify the standard model Cartan generators among the Cartan generators of $G_{GUT}$. An embedding is best described by so-called charge axes. Unfortunately, I don’t know any good resource that explains how we can find an embedding of $G_{SM}$ in a given GUT group and I’ve planned to write about it as soon as I find some time. The standard resource on this kind of thing is

but I wasn’t able to understand his explanations. For the concept of charge axes I found

Dark Matter and Gauge Coupling Unification in Non-supersymmetric SO(10) Grand Unified Models
by Yann Mambrini, Natsumi Nagata, Keith A. Olive, Jeremie Quevillon, Jiaming Zheng

very helpful.

Unfortunately, I also don’t know any good resource that answers the third question. The breaking of $G_{GUT}$ can be described either using tensor methods or using Dynkin’s methods. In tensor language, a vev is given by a matrix $V$, which is a member of some representation $R$. Then one acts on this matrix with every generator $G$ of the group via the usual Lie algebra product $[G,V]$ (the commutator).

  • If the result $M$ of the commutator: $[G,V]=M$ is an element of $R$, i.e. of the same representation as $V$ then the generator $G$ is broken.
  • If the result $M$ of the commutator: $[G,V]=M$ is not an element of $R$, then the generator $G$ is unbroken because $G$ annihilates (by definition) the vev $V$. If acting with a generator on an element of some representation $R$ does not yield another element of $R$, the result is by definition zero. This is analogous to the ladder operators for the angular momentum in quantum mechanics. We can apply these ladder operators only until we reach the highest state. If we act with a raising operator on the highest state, we would get a state which is not part of the representation and therefore we say that the state gets annihilated.

The main task is always to write down the correct explicit matrices for the various representations. The rest of the computation, i.e. the commutator and then check if the result is in the same representation as the VEV, can be easily implemented in Mathematica. While tensor methods are easier to explain, I find Dynkin’s methods most of the time easier, especially for a non-trivial group like $SO(10)$ or $E_6$. (However, a colleague of mine recently published a Mathematica package for explicit matrix expressions of $E_6$ representations and this should simplify the computations in $E_6$ models significantly.) I don’t know any good paper or book that explains how symmetry breaking can be described in Dynkin’s framework and therefore I will write about it as soon as I find the time.

For the group-theoretical aspects of GUTs is found the following resources really helpful:

Grand Unified Theories in General

To get an overview I found the famous review

really helpful, although it is quite old and some things are outdated.

I know of three books about GUTs and all of them are quite old, too. Nevertheless, I think that

Grand Unification with and without Supersymmetry by Costas Kounnas, Antonio Masiero, Dimitri Nanopoulos, Keith A Olive

is still a great book and I found some chapters in

Grand Unified Theories by Graham Ross

really helpful.

The third GUT book I know of is

Unification and Supersymmetry by Rabindra N. Mohapatra 

but it didn’t help me.

Specific Topics

To understand many subtle effects in GUTs I found the review

Effective Field Theory by Howard Georgi

incredibly helpful. It helped me finally understand, why the gauge couplings should unify at all and what the GUT scale actually is.

Two other excellent papers that were immensely helpful and elucidate many important GUT problems are

An Exceptional Model for Grand Unification by Riccardo Barbieri, Dimitri V. Nanopoulos

Mass relations and neutrino oscillations in an SO(10) model by Harvey, J. A.; Reiss, D. B.; Ramond, P.

Renormalization Group Running

I’ve written three long posts about the renormalization group running and most papers that I found helpful are linked there. The most helpful paper was

Implications of LEP results for SO(10) grand unification by N.G. Deshpande, E. Keith, Palash B. Pal

and for threshold effects:

Scalar Masses

In this context, the extended survival hypothesis is extremely important. We actually don’t understand scalar masses, but we need to know them in order to compute the renormalization group equations. Therefore on usually invokes the extended survival hypothesis which states that “at every stage oft he symmetry breaking chain only those scalars are present that develop a vacuum expectation value (VEV) at the current or the subsequent levels of the spontaneous symmetry breaking.” The two standard papers on this topic are

Higgs Bosons in SO(10) and Partial Unification by F. del Aguila, Luis E. Ibanez

Higgs-boson effects in grand unified theories by Rabindra N. Mohapatra and Goran Senjanović

Breaking Chains

The scalar sector is usually too complicated in GUTs and thus one must stick to rules of thumb like the extended survival hypothesis. Another very important such rule is Michel’s conjecture. This conjecture states that states that minima of a Higgs potential correspond to vacuum expectation values that break a given algebra to a maximal subalgebra. This conjecture is explained nicely in

Group Structure of Gauge Theories by L. O’Raifeartaigh

Another important aspect explained there is the necessary condition that in order for a scalar representation to be able to break a group to the subgroups in a given breaking chain that it “must contain singlets with respect to the various subgroups G’, G”…”

Induced Vacuum Expectation Values

A third important observation in the scalar sector is that even if we assume that some scalar field that is allowed to get a VEV does not get a VEV, it is possible that is gets a small induced VEV. This is shown nicely in

Proton lifetime and fermion masses in an SO(10) model by G. Lazarides, Q. Shafi,  C. Wetterich

Aspects of the Grand Unification of Strong, Weak and Electromagnetic Interactions by A.J. Buras, John R. Ellis, M.K. Gaillard, Dimitri V. Nanopoulos

 

 

What are Quantum Numbers?

For quite some time I didn’t really understand what quantum numbers are. For example, why do we use the words “red”, “blue” and “green” for the charges of the strong interaction? Why does a gluon carry “red anti-green + green anti-red” color? From the group theoretical perspective these things actually make a lot of sense and maybe this post helps someone who is equally confused as I was a few years ago.

First, we recall that a Lie algebra representation is a map $R$ from the Lie algebra $\mathfrak{g}$ of a group $G$ to the linear operators $\mathrm{Lin}(\cdot)$ over some vector space $V$.
\begin{equation}
R: \ \mathfrak{g}  \rightarrow \mathrm{Lin}(V) \, .
\end{equation}

The easiest example is the fundamental $2$-dimensional representation of $\mathfrak{su}(2)$, which is a map

\begin{equation}
R: \ \mathfrak{su}(2)  \rightarrow \mathrm{Lin}(\mathbb{C}^2) \, .
\end{equation}

In words this means that this representation maps each element of $\mathfrak{su}(2)$ onto a $2 \times 2$ matrix that acts on $2$-dimensional vectors. A basis for this Lie algebra is given by

\begin{align}
T_1&=\frac{1}{2} \sigma_1 =  \frac{1}{2} \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix} \, , \notag \\
T_2&=\frac{1}{2} \sigma_2 =  \frac{1}{2} \begin{pmatrix} 0 & -\mathrm{i} \\ \mathrm{i} & 0 \end{pmatrix} \, , \notag  \\
T_3&=\frac{1}{2} \sigma_3 =  \frac{1}{2} \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} \, ,
\end{align}

where $\sigma_i$ denote the usual Pauli matrices.

From linear algebra we know that the eigenvectors of a linear operator always form a basis for the vector space in question. In addition, for any Lie group, one or more of the generators can be simultaneously diagonalized using similarity transformations. The set of generators that can be diagonalized simultaneously are called Cartan generators. Thus, a suggestive and particularly easy basis for the vector space of each representation is given by the eigenvectors of the Cartan generators. An easy way to label these basis vectors is to use the corresponding eigenvalues.

In particle physics the fundamental particles are (among others) labelled by their color and weak-isospin. These quantum numbers correspond to eigenvalues of the Cartan generators of the corresponding gauge groups. For $SU(2)_L$, the gauge group of weak-interactions, there is only one Cartan generator and in the fundamental two-dimensional representation it is given by

\begin{equation}
I_3 = \frac{1}{2} \sigma_3 = \frac{1}{2} \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} \, .
\end{equation}

The factor $\frac{1}{2} $ is there, because we usually normalize our generators such that the Dynkin index Tr$(T_a T_a)$ for this generator in the fundamental representation of the Lie algebra is $\frac{1}{2}$.

The corresponding eigenvalues are $\frac{1}{2}$ and $-\frac{1}{2}$. This means particles that correspond to $SU(2)_L$ eigenstates in the fundamental $2$-dimensional representation are with respect to $SU(2)_L$

\begin{equation}
\begin{pmatrix} 1 \\ 0 \end{pmatrix} \ \text{ with isospin } \frac{1}{2} \quad , \quad \begin{pmatrix} 0 \\ 1 \end{pmatrix} \ \text{ with isospin } -\frac{1}{2} \, .
\end{equation}

Less trivial is $SU(3)$, the gauge group of strong interactions, because there are two Cartan generators. In the representation that acts on the fundamental $3$-dimensional representation they can be written as

\begin{equation} \label{eq:cartansu3}
H_1=
\frac{1}{2} \left( \begin{array}{ccc}
1 & 0 & 0 \\
0 & 0 & 0 \\
0 & 0 & -1
\end{array} \right) \quad , \quad H_2 = \frac{1}{2 \sqrt{3}} \left( \begin{array}{ccc}
1 & 0 & 0 \\
0 & -2 & 0 \\
0 & 0 & 1
\end{array} \right) \, .
\end{equation}

Again, we label the trivial eigenvectors using the the eigenvalues of the Cartan generators. However, now there are two eigenvalues for each eigenvector and therefore we define objects, called weights, that collect these numbers for each eigenvector. This means the $SU(3)_C$ quantum numbers for the basis vectors of the fundamental $3$-dimensional representation are

\begin{equation}
\left(\frac{1}{2} , \frac{1}{2 \sqrt{3}} \right) \mathrm{ \ for \ }  \begin{pmatrix}
1 \\ 0 \\ 0
\end{pmatrix} \quad , \quad \left(0 , \frac{-1}{ \sqrt{3}} \right) \mathrm{ \ for \ }  \begin{pmatrix}
0 \\ 1 \\ 0
\end{pmatrix}  \quad , \quad \left(-\frac{1}{2} , \frac{1}{2 \sqrt{3}} \right) \mathrm{ \ for \ }  \begin{pmatrix}
0 \\ 0 \\ 1
\end{pmatrix}  \,  .
\end{equation}

It is conventional to replaces these weights with names: red $:=(\frac{1}{2} , \frac{1}{2 \sqrt{3}})$, blue $:=(0 , \frac{-1}{ \sqrt{3}})$, green $:=(\frac{-1}{2} , \frac{1}{2 \sqrt{3}})$. The labels for the conjugate fundamental representation $\bar 3$ are simply minus the labels of the fundamental $3$ and are called anti-red, anti-blue and anti-green.

We can use these basis vectors and the corresponding weights to derive the basis vectors and weights for product representations like

\begin{equation} \label{eq:3x3bardecomposition}
3 \otimes \bar 3 = 1 \oplus 8 \, .
\end{equation}

The $8$ is the adjoint representation of $SU(3)$ and Eq. \ref{eq:3x3bardecomposition} tells us that we can write each element of the adjoint as $3\times 3$ matrix. For the basis vectors we use

\begin{equation}
e_{ij} = e_i \otimes e_j \, .
\end{equation}

For the quantum numbers of the product representations we use

\begin{equation}
QN \big ( (a \otimes b)_{ij} \big) = QN(a_i) + QN( b_j) \, .
\end{equation}

Formulated in terms of weights this means that we can compute the weight corresponding to the $ij$ element of the product representation $3 \otimes \bar{3}=1 \oplus 8$ by adding the weights of $3_i$ and $\bar{3}_j$

\begin{equation}
w\Big ( (3 \otimes 3)_{ij}\Big) = w(3_i) + w(3_j) \,.
\end{equation}

The $3 \times 3$ matrix that correspond to this weights is given by the Kronecker product of the $i$-th and $j$-th basis vector. Thus we have

\begin{equation}
\left(
\begin{array}{cc}
\frac{1}{2} & \frac{\sqrt{3}}{2} \\
\end{array}
\right) = \left( \begin{array}{ccc}
0 & 1 & 0 \\
0 & 0 & 0 \\
0 & 0 & 0
\end{array} \right) \, , \quad  \left(
\begin{array}{cc}
1 & 0 \\
\end{array}
\right)=  \left( \begin{array}{ccc}
0 & 0 & 1 \\
0 & 0 & 0 \\
0 & 0 & 0
\end{array} \right) \, , \quad \left(
\begin{array}{cc}
\frac{1}{2} & -\frac{\sqrt{3}}{2} \\
\end{array}
\right) = \left( \begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 1 \\
0 & 0 & 0
\end{array} \right) \, , \end{equation} \begin{equation}  \left(
\begin{array}{cc}
-\frac{1}{2} & -\frac{\sqrt{3}}{2} \\
\end{array}
\right)=
\left( \begin{array}{ccc}
0 & 0 & 0 \\
1 & 0 & 0 \\
0 & 0 & 0
\end{array} \right) \, , \quad \left(
\begin{array}{cc}
-1 & 0 \\
\end{array}
\right) = \left( \begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 0 \\
1 & 0 & 0
\end{array} \right) \, , \quad \left(
\begin{array}{cc}
-\frac{1}{2} & \frac{\sqrt{3}}{2}
\end{array}
\right) = \left( \begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 0 \\
0 & 1 & 0
\end{array} \right) \,
\end{equation}

and two zero weights $(0,0)$ that span a basis for the Cartan subalgebra

\begin{equation} \left(
\begin{array}{cc}
0 &0\\
\end{array}
\right)_1=  \frac{1}{2} \left( \begin{array}{ccc}
1 & 0 & 0 \\
0 & 0 & 0 \\
0 & 0 & -1
\end{array} \right)  \, , \quad   \left(
\begin{array}{cc}
0 & 0 \\
\end{array}
\right)_2 =  \frac{1}{2 \sqrt{3}} \left( \begin{array}{ccc}
1 & 0 & 0 \\
0 & -2 & 0 \\
0 & 0 & 1
\end{array} \right) \, .
\end{equation}

These $8$ matrices are given in a basis, known as Cartan-Weyl basis for $\mathfrak{su}(3) $. The special thing about this basis is that each matrix here is an “eigenmatrix” of the Cartan generators $H_i$, which means

\begin{equation}
H_i \circ M = [H_i,M]=\lambda_i M \, ,
\end{equation}

where $\lambda_i$ is the eigenvalue for the Cartan generator $H_i$. In physical terms, each of these $8$ matrices represents a different gluon. This is completely analogous to how the three basis vectors for $\mathbb{R}^3$ for the fundamental representation correspond to three different quarks: a red quark, a blue quark and a green quark.

There is another way to denote gluons, analogous to the color notation for quarks in the fundamental $3$. We can use the fact that each gluon corresponds to a product of a basis vector of the fundamental $3$ and basis vector of the anti-fundamental $\bar{3}$, to name each gluon in terms of color, too. For example, for $i=1$ and $j=2$

\begin{equation}
8_{12} \hat= w(3_1) + w( \bar{3}_2)  = \left(\frac{1}{2} , \frac{1}{2 \sqrt{3}} \right) +\left(0 , \frac{1}{ \sqrt{3}}\right) = \text{red+ anti-blue} =\left( \frac{1}{2},  \frac{\sqrt{3}}{2} \right)  \, .
\end{equation}

Therefore just as we have red, blue and green quarks, we have $8$ gluons that we can label by color combinations of the form color-anticolor.

There is another basis for the $8$ basis matrices of the adjoint representation, more popular among physicists, called the Gell-Mann basis\footnote{The Gell-Mann basis is more popular due to the close connection to the Pauli matrices of $SU(2)$.}. In this basis the $8$ basis elements $T_a$ of $\mathfrak{su}(3) $ are given in terms of the $8$ Gell-Mann matrices $\lambda_a$ by $T_a= \frac{1}{2} \lambda_a$, where
\begin{equation}\label{lambda1-3}
\lambda_1 =
\begin{array}{ccc}
\left(
\begin{array}{ccc}
0 & 1 & 0 \\
1 & 0 & 0 \\
0 & 0 & 0
\end{array}
\right),
&
\lambda_2 =\left(
\begin{array}{ccc}
0 & -i & 0 \\
i & 0 & 0 \\
0 & 0 & 0
\end{array}
\right),
&
\lambda_3 =\left(
\begin{array}{ccc}
1 & 0 & 0 \\
0 & 0 & 0 \\
0 & 0 & -1
\end{array}
\right),
\end{array}
\end{equation}

\begin{equation}\label{lamdba4-6}
\begin{array}{ccc}
\lambda_4 =\left(
\begin{array}{ccc}
0 & 0 & 1 \\
0 & 0 & 0 \\
1 & 0 & 0
\end{array}
\right),
&
\lambda_5 =\left(
\begin{array}{ccc}
0 & 0 & i \\
0 & 0 & 0 \\
-i & 0 & 0
\end{array}
\right),
&\lambda_6 =\left(
\begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 1 \\
0 & 1 & 0
\end{array}
\right),
\end{array}
\end{equation}

\begin{equation}\label{lambda7-8}
\begin{array}{cc}
\lambda_7 =\left(
\begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & -i \\
0 & i & 0
\end{array}
\right),
&
\lambda_8=\frac{1}{\sqrt{3}}\left(
\begin{array}{ccc}
1 & 0 & 0 \\
0 & -2 & 0 \\
0 & 0 & 1
\end{array}
\right).
\end{array}
\end{equation}

Again, we can give names in the form color-anticolor to the gluon states in this basis. For example,

\begin{align}
T_1 &= \frac{1}{2}\lambda_1 =
\left(
\begin{array}{ccc}
0 & 1 & 0 \\
1 & 0 & 0 \\
0 & 0 & 0
\end{array}
\right)
= \left(
\begin{array}{ccc}
0 & 1 & 0 \\
0 & 0 & 0 \\
0 & 0 & 0
\end{array}
\right) + \left(
\begin{array}{ccc}
0 &0 & 0 \\
1 & 0 & 0 \\
0 & 0 & 0
\end{array}
\right)  \notag \\ &= \text{red anti-green} + \text{green anti-red}
\end{align}

and this is a popular way to label the gluons.

Solving the Renormalization Group Equations for the Gauge Couplings

We have already discussed why the gauge couplings depend on the energy scale and how we can compute the renormalization group equations (RGEs) that describe how the couplings change with energy. In this post we talk about how we can solve the RGEs.

The Standard Model RGEs

To solve the RGEs, we need boundary conditions. One such condition is always given by the measured values of the coupling constants. Usually we use the couplings strength at the energy scale that is given by the mass of the $Z$-boson:

\begin{align}
\omega_{1Y}(M_Z)& = 59.0116 \, ,\notag \\
\omega_{2L}(M_Z)& = 29.5874 \, , \notag\\
\omega_{3C}(M_Z)& = 8.4388  \, , \notag \\
M_Z &= 91.1876 \text{ GeV} \, ,
\end{align}

These values are taken from the Review of Particle Physics.

Then, using the general formula described in this post, we can derive the RGEs, for example, for the standard model gauge couplings. The Standard Model RGE coefficients are

\begin{align}
a_{SM}=\left(
\begin{array}{c}
\frac{41}{10} \\ -\frac{19}{6} \\ -7
\end{array}
\right) \qquad , \qquad
b_{SM}= \left(
\begin{array}{ccc}
\frac{199}{50} & \frac{27}{10} & \frac{44}{5} \\
\frac{9}{10} & \frac{35}{6} & 12 \\
\frac{11}{10} & \frac{9}{2} & -26 \\
\end{array}
\right)
\end{align}

and putting them into Equation 2 of this post yields differential equations that can be solved, for example, using Mathematica.

 

My Mathematica notebook that solves the 1-loop and 2-loop Standard Model RGEs numerically and plots the solutions can be downloaded here.

 

The solutions of the $2$-loop RGEs for the Standard Model gauge couplings are shown in the figure below.

sm2loop

 

We can see that the couplings do not meet exactly at one point. This is famous result that in non-supersymmetric GUTs one needs at least one intermediate scale or additional particles to achieve the unification of the gauge couplings.

Before we can solve the RGEs with an intermediate scale, we need to discuss the matching conditions at the various scales.

Matching Conditions

As a first approximation, we can compute the unification scale by determining where the gauge couplings meet at one point. This would mean, for example, that we use the boundary condition

\begin{equation} \omega_{1Y}(M_{GUT}) = \omega_{2L}(M_{GUT})= \omega_{3C}(M_{GUT})  \end{equation}

In this first approximation $M_{GUT}$ is the scale where the heavy vector bosons and scalars get integrated out. At scales far above this mass scale, the spontaneous symmetry breaking has a negligible effect. This procedure is sufficient when we use the $1$-loop RGEs.

Unfortunately, if we want to determine the GUT scale with higher accuracy and use the $2$-loop RGEs, this picture is vastly oversimplified. It is unlikely that all vector bosons and scalars have exactly the same mass. If their masses are not degenerate the correct matching conditions for a breaking $G \rightarrow \prod_i G_i$ are
\begin{equation}
\label{eq:thresholddef}
\omega_{G_i}=\omega_{G}-\frac{\lambda_i(\mu)}{12 \pi} ,
\end{equation}
where
\begin{eqnarray}
\label{eq:lambdasthresholds}
\lambda_i(\mu)= \underbrace{\left( C_2(A_G)-C_2(A_i) \right)}_{\lambda_i^G} -21 \underbrace{ \; T(V)\ln \frac{M_V}{\mu}}_{\lambda_i^V}  + \underbrace{T(S) \ln \frac{M_S}{\mu}}_{\lambda_i^S} + 8 \underbrace{T(F) \ln \frac{M_F}{\mu} }_{\lambda_i^F} \, .
\end{eqnarray}

Here $V$, $S$ and $F$ denote the vector, scalar and fermion subgroup representations that get integrated out at the matching scale $\mu$ and $M_V$, $M_S$, $M_F$ their masses. Once more $C_2(r)$ and $T(r)$ denote the quadratic Casimir invariant and the Dynkin index of the representation $r$. Further, $A_G$ and $A_i$ denote the adjoint representation of the group $G$ and subgroup $i$. In words this means the GUT scale is no longer where the gauge coupling meet at one point, but can lie above or below this point.

For the moment, we neglect the logarithmic terms and discuss them in a later post. Initially it was assumed that the logarithmic terms are small and therefore negligibly. However, in GUTs there are ususually lots of scalar fields and altough the contribution from each individual scalar field is small, many small contributions can add up to a large term. The corrections coming from these logarithmic terms are usually called threshold corrections.

Without the logarithmic terms, we can derive, for example, for the breaking chain ${SO(10) \rightarrow SU(4) \times SU(2)_L \times SU(2)_R \rightarrow \text{SM}}$ the following matching conditions at the $SO(10)$ scale

\begin{align} \label{eq:so10matching}
\omega_{SU(4)_C}(\mu_u) – \frac{4}{12\pi}&=  \omega_{SO(10)}(\mu_u) – \frac{8}{12\pi} \, ,  \notag \\
\omega_{SU(2)_L}(\mu_u) – \frac{2}{12\pi}&=   \omega_{SO(10)}(\mu_u) – \frac{8}{12\pi} \, , \notag \\
\omega_{SU(2)_R}(\mu_u) – \frac{2}{12\pi}  &=  \omega_{SO(10)}(\mu_u) – \frac{8}{12\pi}
\end{align}

and at the Pati-Salam scale

\begin{align} \label{eq:patiso10matching}
\omega_{SU(3)_C}(\mu_i) – \frac{3}{12\pi}&=  \omega_{SU(4)_C}(\mu_i) – \frac{4}{12\pi} \, , \notag \\
\omega_{SU(2)_L}(\mu_i) – \frac{2}{12\pi}  &= \omega_{SU(2)_L}(\mu_i) – \frac{2}{12\pi} \, , \notag \\
\omega_{U(1)_Y}(\mu_i)  &=  \frac{3}{5}\left( \omega_{SU(2)_R}(\mu_i) – \frac{2}{12\pi}  \right) + \frac{2}{5}\left( \omega_{SU(4)_C}(\mu_i) – \frac{4}{12\pi}  \right) \, .
\end{align}

To derive these, we have simply used Eq. \ref{eq:lambdasthresholds} and the numerical values for the quadratic Casimirs, which are listed, for example, in the last section of this post.

The RGEs in Models with enlarged Gauge Symmetry

Now equipped with these matching conditions, we can finally solve the RGEs in an $SO(10)$ model with a Pati-Salam intermediate symmetry. We use the coefficients computed in this post for the running between the Pati-Salam and the $SO(10)$ scale:

\begin{align} \label{eq:paticoefficients}
a_{PS} = \left(
\begin{array}{c}
\frac{26}{3} \\ \frac{26}{3} \\ \frac{2}{3}
\end{array}
\right)  \, , \qquad
b_{PS}=  \left(
\begin{array}{ccc}
\frac{779}{3} & 48 & \frac{1277}{2} \\
48 & \frac{779}{3} & \frac{1277}{2} \\
\frac{249}{2} & \frac{249}{2} & \frac{3541}{6} \\
\end{array}
\right).
\end{align}

The Standard Model RGEs are unchanged and therefore as described in the first section. Using the matching conditions from above, we can solve the RGEs, for example, using Mathematica. The result is shown in the following figure.

patirgewithoutthresholdssmall

We can see that through the intermediate symmetry we can achieve unifcation of the gauge couplings. This is possible, because we now have one additional fit paramter: the intermediate scale $M_{PS}$.

This breaking chain was thought to be ruled out, because the $SO(10)$ scale is quite low and therefore the proton lifetime is below the present bound from the Super Kamiokande experiment. However as already mentioned above, there can be large threshold corrections, when not all scalar particles that get integrated out at a given scale habe exactly the same mass. Therefore this breaking chain was recently reanalysed in this paper. The authors found that the proton lifetime can be well above the present bound from Super Kamiokande through the threshold corrections.