GO BACK

Note on Weinberg's textbooks on QFT

Notation. The Latin indices such as , , , etc., typically span the three spatial coordinate labels, commonly denoted as 1, 2, 3. On the other hand, Greek indices like , , and so forth, usually range over the four spacetime coordinate labels, specifically 1, 2, 3, 0, where represents the time coordinate. Indices that appear twice are usually summed unless specified otherwise. The spacetime metric, denoted by , is a diagonal matrix with elements and . The d'Alembertian is represented as and defined by the equation , where is the Laplacian given by . The ‘ Levi-Civita tensor,’ symbolized by , is an entirely antisymmetric entity with . Spatial vectors in three dimensions are marked by boldface characters. A unit vector corresponding to any vector is shown with a hat, as in . A dot over a variable signifies its time derivative. The Dirac matrices adhere to , and , while . The step function yields a value of +1 when and 0 when . For a matrix or vector , the complex conjugate, transpose, and Hermitian adjoint are represented by , , and respectively. The Hermitian adjoint of an operator is marked as , except when an asterisk emphasizes that a vector or matrix of operators is not transposed. Terms like +H.c. or +c.c. appended to equations indicate the addition of the Hermitian adjoint or complex conjugate of preceding terms. A Dirac spinor with a bar over it is defined as . Apart from in Chapter ?, units are normalized such that and the speed of light are set to one. The fine structure constant is represented as , calculated as , approximately , where is the rationalized charge of the electron. Parenthetical numbers next to quoted numerical figures signify the uncertainty in the last digits. Unless otherwise stated, experimental data is sourced from ‘Review of Particle Properties,’ Phys. Rev. D50, 1173 (1994).

Chapter 1
Relativistic Quantum Mechanics

The perspective presented argues that quantum field theory exists in its current form due to its unique capability to harmonize quantum mechanics with special relativity, under some conditions. Our initial endeavor is to explore how symmetries, such as Lorentz invariance, manifest within a quantum context in the following aspects.

1.1Quantum Mechanics

Quantum field theory rests on the same foundational quantum mechanics developed by Schrödinger, Heisenberg, Pauli, Born, and other pioneers in 1925–1926.

  1. Physical states are represented by rays in complex Hilbert space (the inner product is denoted be with the first slot antilinear (conjugate-linear) and the second slot linear). Here, a ray is a set of normalized vectors (i.e. ) with and belonging to the same ray if , where is an arbitrary complex number with .

  2. Observables are represented by Hermitian operators. A state represented by a ray has a definite value for the observable represented by an operator if vectors belonging to this ray are eigenvectors of with eigenvalue :

  3. If a system is in a state represented by a ray , and an experiment is done to test whether it is in any one of the different states represented by mutually orthogonal rays (for instance, by measuring one or more observables) then the probability of finding it in the state represented by is

    where and are any vectors belongs to rays and , respectively.

1.2Symmetries

A symmetry transformation can be thought of as a shift in perspective that does not affect the outcomes of potential experiments. If an observer perceives a system in a state denoted by a ray or or ..., a corresponding observer scrutinizing the same system would view it in a different state, symbolized by a ray or or ..., respectively. However, both observers must ascertain the same probabilities:

(1.2.1)

This condition is necessary but not sufficient for a ray transformation to qualify as a symmetry; additional conditions will be elaborated upon in the following chapter. Wigner proved a significant theorem in the early 1930s, stating that for any such transformation , an operator can be defined in the Hilbert space. If is a vector in ray , then belongs to ray . The operator can either be unitary and linear:

(1.2.2)
(1.2.3)

or antiunitary and antilinear:

(1.2.4)
(1.2.5)

for all in the Hilbert space.

This finding is called the fundamental theorem of Wigner and the proof is the following:

The fundamental theorem of Wigner (1931).

Let be a Hilbert space and let

be a bijection satisying

for all rays and ; and vectors , , , and . Then there exists an operator acting on such that

for all ray and all ; and that either is unitary and linear or antiunitary and antilinear.

Proof.

As previously stated, the adjoint of a linear operator is determined by

(1.2.6)

This definition does not apply to an antilinear operator since the right-hand side of (1.2.6) would be linear in , while the left-hand side is antilinear in . For an antilinear operator , the adjoint is instead specified as:

(1.2.7)

Given this definition, the criteria for either unitarity or antiunitarity can both be expressed as:

(1.2.8)

There exists a trivial symmetry transformation ℛ→ℛ, represented by the identity operator . This operator is naturally unitary and linear. Continuity dictates that any symmetry operation (like a rotation, translation, or Lorentz transformation) that can be reduced to a trivial transformation by continuously adjusting certain parameters (such as angles, distances, or velocities) must be characterized by a linear unitary operator , as opposed to one that is antilinear and antiunitary. (Symmetries represented by antiunitary antilinear operators are less common in physics; they all entail a reversal in the direction of time flow. See Section ? for more details.)

Specifically, a symmetry transformation that is nearly trivial on an infinitesimal scale can be depicted by a linear unitary operator that is infinitesimally close to the identity operator:

(1.2.9)

Here, is a real infinitesimal. For to be both unitary and linear, needs to be Hermitian and linear, making it a potential observable. In fact, many (if not all) physical observables, like angular momentum or momentum, are derived from symmetry transformations in this manner.

The set of symmetry transformations possesses specific characteristics that categorize it as a group. If is a transformation converting rays to , and is another transformation that maps to , then the outcome of executing both transformations consecutively is yet another symmetry transformation, denoted as , that transforms into . Additionally, any symmetry transformation that changes into has an inverse, expressed as , which reverts back to . Moreover, there exists an identity transformation, , which leaves rays unaltered.

The unitary or antiunitary operators that correspond to these symmetry transformations emulate this group structure, albeit with added complexity because operators act on vectors in Hilbert space instead of on rays. If transforms into , then applying to a vector in must result in a vector in . If then maps this ray to , must also belong to , as must . Therefore, the vectors can only differ by a phase factor , as given by:

(1.2.10)

Moreover, barring a notable exception, the linearity (or antilinearity) of specifies that these phases are state-independent. To prove this, let us consider two non-proportional vectors and and apply Equation (1.2.10) to the state:

(1.2.11)

Every unitary or antiunitary operator has an inverse (its adjoint), which is also either unitary or antiunitary. Upon left-multiplying Equation (1.2.11) by , we arrive at:

(1.2.12)

As and are linearly independent, it follows that

(1.2.13)

Consequently, the phase in Equation (1.2.10) is state-independent, leading to the operator relation:

(1.2.14)

When , this indicates that constitutes a representation of the group of symmetry transformations. For arbitrary phases , we refer to it as a ‘projective representation’ or a representation ‘up to a phase’. Whether the Lie group structure allows for state vectors to furnish an ordinary or projective representation can not be inferred from the group structure alone but will become apparent later.

The exception to the reasoning that concluded in Equation (1.2.14) lies in the possibility that the system may not be preparable in a state represented by . For example, it is generally considered unfeasible to prepare a system in a superposition of states with total angular momenta that are integers and half-integers. In such scenarios, we refer to the presence of a ‘superselection rule’ between different categories of states. As a result, the phases could be contingent on which class of states the operators and are acting upon. Further details about these phases and projective representations will be discussed in Section ?. It will be shown that any symmetry group featuring projective representations can be extended (without altering its physical meaning) to allow for all its representations to be non-projective, i.e., with . Until we reach Section ?, we will proceed with the assumption that such an extension has been applied, and will take in (1.2.14). Also, the existence of spinor is partially derived from the phase ambiguity that arises when taking absolute values and the fact that the homotopy class of the homogeneous Lorentz group.

In physics, a specific type of group known as a connected Lie group holds special significance. These are groups comprised of transformations , defined by a finite collection of real, continuous parameters, symbolized as . Each group element is linked to the identity element through a continuous path within the group itself. The multiplication rule for the group is expressed as

(1.2.15)

where is a function of both and . If denotes the coordinates of the identity, then

(1.2.16)

must hold true. In the case of such continuous groups, the transformations must be represented in the physical Hilbert space by unitary operators , rather than antiunitary ones. These unitary operators, at least in a finite vicinity of the identity, can be expressed by a power series as

(1.2.17)

Here, , and so on, are Hermitian operators independent of . Assuming that provides a standard (non-projective) representation of the transformation group, meaning

(1.2.18)

we can expand this in terms of and . In accordance with Equation (1.2.16), the second-order expansion of should be

(1.2.19)

Here, are real coefficients. Note that the presence of any or terms would be in conflict with Equation (1.2.16). Following this, Equation (1.2.18) can be articulated as:

(1.2.20)

On both sides of Equation (1.2.20), terms of order 1, , , , and correspond without issue. However, when focusing on the terms, a non-trivial condition emerges:

(1.2.21)

This reveals that if we know the group structure, specifically the function and its corresponding quadratic coefficient , we can determine the second-order terms of using the first-order generators . However, there's a requirement for consistency: the operator has to be symmetric in and , as it's the second derivative of with respect to and . Therefore, Equation (1.2.21) necessitates that

(1.2.22)

where are a set of real constants termed as structure constants, defined by

(1.2.23)

This kind of commutation relationship is termed a Lie algebra. In a later section, we will essentially demonstrate that this commutation relation (1.2.22) is the sole condition needed to perpetuate this computation. In other words, the complete power series for can be generated from an endless chain of equations like Equation (1.2.21), as long as we are aware of the first-order terms, namely the generators . While this does not mean operators are uniquely identified for all based solely on , it does signify that they are uniquely specified within a finite vicinity of the identity coordinate , such that Equation (1.2.15) holds true if and lie within this region. The discussion about extending this to all will take place in a subsequent section.

There is a particular scenario of considerable relevance that will recur frequently in our discussions. Assume the function is simply additive for some or all of the coordinates , as expressed by:

(1.2.24)

This situation is applicable, for example, in the context of spacetime translations or for rotations about a single fixed axis (but not for both simultaneously). In this special case, the coefficients from Equation (1.2.19) become zero, and likewise, the structure constants in Equation (1.2.23) also vanish. Consequently, the generators are commutative, denoted by:

(1.2.25)

Such a group is termed as Abelian. Under these conditions, computing for all becomes straightforward. According to Equations (1.2.18) and (1.2.24), for any integer , we can express:

By taking the limit as approaches infinity and retaining only the first-order term in , we obtain:

and consequently,

(1.2.26)

1.3Quantum Lorentz Transformations

Einstein's principle of relativity asserts the equivalence of specific 'inertial' frames of reference, setting it apart from the Galilean principle of relativity adhered to by Newtonian mechanics. The distinction comes from the transformation equations that link coordinate systems across different inertial frames. Given that represents the coordinates in one inertial frame—where are Cartesian spatial coordinates and is a time coordinate (assuming the speed of light equals one)—the coordinates in another inertial frame must satisfy:

(1.3.1)

or, alternatively,

(1.3.2)

In these equations, is a diagonal matrix with elements defined as:

(1.3.3)

The summation convention applies: any index like and in Equation (1.3.2) appearing twice, once as a superscript and once as a subscript, is summed over.

These transformations have the unique feature that the speed of light remains consistent—in our chosen units, equal to one—across all inertial frames. A light wave with unit speed satisfies , or in terms of the equation , which also implies and thus .

Any coordinate transformation fulfilling Eq. (1.3.2) is linear, as denoted by:

(1.3.4)

Here, are arbitrary constants, and is a constant matrix that meets the criteria:

(1.3.5)

For certain applications, it's advantageous to express the Lorentz transformation condition using an alternate formulation. The matrix possesses an inverse, designated as , which coincidentally has the same diagonal components: and .

By judiciously inserting parentheses and multiplying Eq. (1.3.5) by , we get:

Further multiplying by the inverse of the matrix yields:

(1.3.6)

These transformations constitute a group. When we initially apply a Lorentz transformation as per Eq. (1.3.4), and then follow it with another Lorentz transformation , such that

we find that the overall transformation effect is identical to performing a Lorentz transformation as described by

(1.3.7)

Here, it's worth noting that if and both meet the conditions of Eq. (1.3.5), will also be a Lorentz transformation. The bar notation is simply used to distinguish one Lorentz transformation from another. Correspondingly, the transformations on physical states obey the composition law

(1.3.8)

Calculating the determinant of Eq. (1.3.5), we arrive at

(1.3.9)

This implies that has an inverse, denoted as , which as per Eq. (1.3.5) takes the form

(1.3.10)

According to Eq. (1.3.8), the inverse of the transformation turns out to be , and naturally, the identity transformation is represented by .

Based on the dialogue in the prior section, the transformations give rise to a unitary linear transformation acting on vectors in the physical Hilbert space, represented as . These operators obey a composition law articulated as

(1.3.11)

It's worth noting that to prevent the emergence of a phase factor on the right-hand side of Eq. (1.3.11), it's generally required to extend the Lorentz group. The suitable extension for accomplishing this is discussed in Section ?.

The complete set of transformations is formally referred to as the inhomogeneous Lorentz group, also known as the Poincaré group. This group has several significant subgroups. First, transformations with naturally constitute a subgroup, described by

(1.3.12)

which is termed the homogeneous Lorentz group. Additionally, from Eq. (1.3.9), it's evident that can be either or ; transformations having inherently make up a subgroup of either the homogeneous or inhomogeneous Lorentz groups. Further scrutiny of the 00-components of Eqs. (1.3.5) and (1.3.6) yields

(1.3.13)

where ranges over 1, 2, and 3. This shows that either or . Transformations where constitute a subgroup. Observe that if and are two such matrices , then

According to Eq. (1.3.13), the three-vector has a length of , and similarly, the three-vector has a length of . Therefore, the scalar product of these two three-vectors has an upper limit given by

(1.3.14)

leading to

This subgroup, characterized by and , is identified as the proper orthochronous Lorentz group. As one cannot smoothly transition from to , or from to , any Lorentz transformation derived from the identity through a continuous variation of parameters must share the same sign for and as the identity, and thus must be a member of the proper orthochronous Lorentz group.

Every Lorentz transformation falls into one of two categories: it is either proper and orthochronous, or it can be expressed as the composition of an element from the proper orthochronous Lorentz group and one of the discrete transformations or or . Here, represents the space inversion, which has non-zero elements given by

(1.3.15)

while stands for the time-reversal matrix, with non-zero elements defined as

(1.3.16)

Therefore, a comprehensive understanding of the entire Lorentz group can be achieved by studying its proper orthochronous subgroup, along with the concepts of space inversion and time-reversal. The exploration of space inversion and time-reversal will be carried out separately in Section ?. Until that point, our focus will remain on either the homogeneous or inhomogeneous proper orthochronous Lorentz group.

1.4The Poincaré Algebra

As discussed in Section 1.2, many essential attributes of any Lie symmetry group are encapsulated in the properties of the elements in the vicinity of the identity element. In the context of the inhomogeneous Lorentz group, the identity transformation is given by and . Therefore, we aim to explore transformations that can be written as

(1.4.1)

where both and are infinitesimal. The Lorentz condition, expressed as equation (1.3.5), can be rewritten as

In this book, we adopt the convention that indices can be raised or lowered by contracting with or :

If we retain only the first-order terms in in the Lorentz condition (1.3.5), we find that this condition simplifies to the antisymmetry of :

(1.4.2)

An antisymmetric second-rank tensor in four dimensions has independent components. Coupled with the four components of , an inhomogeneous Lorentz transformation is thus characterized by parameters.

Because maps any ray onto itself, it must be proportional to the unit operator, and by a choice of phase may be made equal to it. Excluding the presence of superselection rules, we can eliminate the chance that this proportionality factor varies depending on the state acted upon by . This exclusion follows the same logic we applied in Section 1.2 to dismiss the idea that phases in projective representations of symmetry groups might depend on the states they act upon. In cases where superselection rules are relevant, it could be necessary to adjust the phase factors of depending on the sector it acts on.

For an infinitesimal Lorentz transformation as described by equation (1.4.1), must be equal to the unit operator augmented by terms that are linear in and . We express this relationship as

(1.4.3)

In this equation, and are operators that are independent of and , and the ellipsis signifies terms of higher order in and/or . For to be unitary, operators and must be Hermitian:

(1.4.4)

(Yes, the generators of boosts are observables.) Given that is antisymmetric, its coefficient can also be taken to be antisymmetric:

(1.4.5)

As we will elaborate on later, , and are the components of the momentum operators; , and are the angular momentum vector components; and is the energy operator or Hamiltonian. These identifications of angular-momentum generators are necessitated by the commutation relations of . However, the commutation relations don't prescribe a definite sign for and , making the sign choice for the term in equation (1.4.3) a matter of convention. The alignment of this choice with the standard definition of the Hamiltonian will be clarified in Section ?.

We turn our attention to the Lorentz transformation characteristics of and . We focus on the composite expression

where and are parameters of a new transformation, distinct from and . According to Equation (1.3.11), the operation results in , signifying that serves as the inverse of . Consequently, from (1.3.11), we obtain:

(1.4.6)

To the first order in and , this leads to:

(1.4.7)

By matching the coefficients of and on both sides of the equation and employing (1.3.10), we arrive at:

(1.4.8)
(1.4.9)

In the case of homogeneous Lorentz transformations where , these transformation laws simply indicate that behaves as a tensor and as a vector. For pure translations, where , these rules convey that remains invariant under translation, while does not. Specifically, the alteration in the spatial components of due to a spatial translation corresponds to the conventional change in angular momentum when the point of reference for measuring angular momentum is shifted.

Next, we consider the application of rules (1.4.8) and (1.4.9) to an infinitesimal transformation. Specifically, we take and , where the infinitesimals and are not related to the earlier and . Utilizing Equation (1.4.3) and retaining only first-order terms in and , Equations (1.4.8) and (1.4.9) simplify to:

(1.4.10)
(1.4.11)

By isolating the coefficients of and on both sides of these equations, we derive the commutation relations:

(1.4.12)
(1.4.13)
(1.4.14)

These equations define the Lie algebra of the Poincaré group.

In quantum mechanics, particular importance is given to those operators that are conserved, meaning they commute with the energy operator . A review of Equations (1.4.13) and (1.4.14) reveals that these conserved operators include the momentum three-vector

(1.4.15)

and the angular-momentum three-vector

(1.4.16)

as well as the energy itself. The other generators constitute what is termed the 'boost' three-vector

(1.4.17)

These are not conserved, which is why their eigenvalues are not employed to characterize physical states. Expressed in a three-dimensional notation, the commutation relations (1.4.12), (1.4.13), and (1.4.14) can be represented as:

(1.4.18)
(1.4.19)
(1.4.20)
(1.4.21)
(1.4.22)
(1.4.23)
(1.4.24)

Here, take the values 1, 2, and 3, and is the completely antisymmetric quantity where . The commutation relation (1.4.18) is identified as belonging to the angular-momentum operator.

Let us prove (1.4.22) and (1.4.24). From , (1.4.13), (1.4.15), and (1.4.17), we have

The subgroup of pure translations is a part of the inhomogeneous Lorentz group, and its group multiplication rule, as defined by (1.3.7), is

(1.4.25)

This multiplication rule is additive, similar to what is described in Equation (1.2.24). Employing Equation (1.4.3) and revisiting the arguments that led to Equation (1.2.26), we determine that finite translations in the physical Hilbert space are represented as

(1.4.26)

Likewise, a rotation through an angle around the direction specified by is represented in the physical Hilbert space as

(1.4.27)

Contrasting the Poincaré algebra with the Lie algebra of the Galilean group, the symmetry group for Newtonian mechanics, offers fascinating insights. While it is possible to derive the Galilean algebra beginning with its transformation rules and using the same methodology we used for the Poincaré algebra, a simpler path exists. Since we already possess Eqs. (1.4.18)-(1.4.24), we can more conveniently obtain the Galilean algebra as the Inönü-Wigner contraction of the Poincaré algebra in the low-velocity limit. For a set of particles with an average mass and velocity , we anticipate the momentum and angular-momentum operators to be of the order , . On the flip side, the energy operator is composed of a total mass and a non-mass energy (kinetic and potential), which are of the order , . Examining Eqs. (1.4.18)-(1.4.24) reveals that in the limit where , the commutation relations simplify to:

where scales as . It's noteworthy that in Hilbert space, the sequence of operations involving a translation and a 'boost' does not yield the expected transformation . Instead, we have:

The emergence of the phase factor indicates that we are dealing with a projective representation, which comes with a superselection rule that precludes the mixing of states with different masses. In this aspect, the mathematical framework of the Poincaré group is less complex than that of the Galilean group. Nonetheless, it is entirely feasible to extend the Galilean group formally by introducing an additional generator to its Lie algebra. This new generator would commute with all existing generators and have eigenvalues corresponding to the masses of the different states. In such a scenario, physical states would be represented through an ordinary, rather than projective, representation of the augmented symmetry group. While this might seem like a minor change in notation, it effectively eliminates the necessity for a mass superselection rule within the reinterpreted Galilean group.

1.5One-Particle States

We turn our attention to the categorization of single-particle states based on their transformation properties under the inhomogeneous Lorentz group.

Given that the components of the energy-momentum four-vector commute among themselves, it is logical to represent physical state-vectors using eigenvectors of the four-momentum. To do this, we introduce a label to account for any additional degrees of freedom, leading us to consider state-vectors such that

(1.5.1)

For more complex states, like those comprising multiple free particles, the label would need to accommodate both continuous and discrete values. In this discussion, we are focusing solely on one-particle states, whose definition includes that the label is purely discrete. It is worth noting that specific bound states of two or more particles, like the ground state of a hydrogen atom, are also considered one-particle states in this context. While such states are not elementary particles, the distinction between composite and elementary particles is irrelevant for our current purposes.

Equations (1.5.1) and (1.4.26) inform us about the transformation behavior of these states under homogeneous Lorentz transformations.

Applying equation (1.4.9), we find that when a quantum homogeneous Lorentz transformation or equivalently acts on , it yields a four-momentum eigenvector with eigenvalue :

(1.5.2)

Therefore, must be expressible as a linear combination of state-vectors :

(1.5.3)

Generally, one might be able to construct suitable linear combinations of such that the matrix becomes block-diagonal. In other words, with values within a single block could constitute a representation of the inhomogeneous Lorentz group on their own. It makes sense to associate the states of a particular particle type with components of an irreducible representation of the inhomogeneous Lorentz group, meaning it can't be further broken down in this manner.

It should be noted that different types of particles may be related to isomorphic representations, which means their matrices could be identical or transformed into one another by a similarity transformation. In certain scenarios, particle types might be defined as irreducible representations of larger groups, which include the inhomogeneous proper orthochronous Lorentz group as a subgroup. For example, for massless particles whose interactions are invariant under space inversion, it's common to treat all components of an irreducible representation of the inhomogeneous Lorentz group as a single particle type.

The next step in our investigation is to elucidate the structure of the coefficients in irreducible representations of the inhomogeneous Lorentz group.

For our objectives, it's crucial to recognize that the only functions of left invariant by all proper orthochronous Lorentz transformations are the invariant square , and for , also the sign of . Therefore, for each specific value of , and when , each sign of , we can select a 'standard' four-momentum denoted as . Any within this category can then be represented as

(1.5.4)

where is a particular standard Lorentz transformation depending on and, implicitly, on our chosen standard . Consequently, the states having momentum can be defined as

(1.5.5)

where is a numerical normalization factor, the specifics of which will be determined later. Up to this juncture, no details have been provided about how the labels are connected across varying momenta; Equation (1.5.5) now addresses this absence.

When applying an arbitrary homogeneous Lorentz transformation to equation (1.5.5), we obtain:

(1.5.6)

The purpose of this last step is to show that the Lorentz transformation first maps to , then to , and finally back to . This transformation belongs to a subgroup within the homogeneous Lorentz group, characterized by Lorentz transformations that keep invariant:

(1.5.7)

This subgroup is termed the little group. For any that satisfies Equation (1.5.7), we find that:

(1.5.8)

The coefficients serve as a representation of the little group. Specifically, for any elements , the relationship

is satisfied, and hence

(1.5.9)

Particularly, we can apply Equation (1.5.8) to the little-group transformation

(1.5.10)

resulting in:

or, recalling definition (1.5.5):

(1.5.11)

Aside from normalization issues, the task of identifying the coefficients in transformation rule (1.5.3) has now been distilled down to finding the representations of the little group. This technique, which involves deriving representations of a larger group like the inhomogeneous Lorentz group from the representations of its little group, is known as the method of induced representations.

Standard Little Group
(a)
(b)
(c) ,
(d) ,
(e)
(f)

Table 1.5.1. Standard four-momenta and their associated little groups for different categories of four-momenta are discussed. In this context, represents an arbitrary positive energy, for instance, 1 eV. The little groups are generally straightforward to understand: is the regular three-dimensional rotation group that comprises Lorentz transformations which keep a zero-momentum particle stationary. Meanwhile, and are Lorentz groups in (2+1)-dimensions and (3+1)-dimensions, respectively. The group is the set of transformations in Euclidean geometry, which includes both rotations and translations in two dimensions. Its role as the little group for cases where will be elaborated on later.

Table 1.5.1 provides a suitable selection for the standard four-momentum along with the associated little group for different categories of four-momenta.

Out of the six categories of four-momenta, only types (a), (c), and (f) have any recognized implications for physical states. For class (f) — where —it pertains to the vacuum state, which is essentially unchanged by . Our subsequent discussion will be confined to cases (a) and (c), which correspond to particles with mass and massless particles, respectively.

Now is an appropriate time to discuss the normalization of these states. Employing the standard orthonormalization procedure from quantum mechanics, we can select states with standard momentum to be orthonormal as denoted by the equation:

(1.5.12)

(Let me remark that is the standard momentum and runs over all possibilities such that , so, for example, we can not use (1.5.12) to calculate . Also and are normalized such that (1.5.12) holds) The presence of the delta function arises because and are eigenstates of a Hermitian operator with eigenvalues and , respectively. As a direct outcome, the representation of the little group in Eqs. (1.5.8) and (1.5.11) must be unitary.

(1.5.13)

For and , the little groups and do not possess any non-trivial finite-dimensional unitary representations. Hence, if there were states with a specific momentum having or that non-trivially transform under the little group, an infinite number of such states would be required.

Regarding the scalar products for generic momenta, the unitarity of the operator as expressed in Eqs. (1.5.5) and (1.5.11) provides the following formula for the scalar product:

Here, (Hence, ). (Let me remark that here is just the one in

although here we use

This is correct as

which gives

thereby getting

Since as well, the delta function is proportional to . The presence of implies that only the coefficient when matters, as otherwise the inner product vanishes. Hence, with , we have

(1.5.14)

The next step involves determining the proportionality factor that links to .

When integrating an arbitrary scalar function over four-momenta subject to and (corresponding to cases (a) or (c)), the Lorentz-invariant integral takes the form:

Here, is the step function: for and for .

When integrating over the 'mass shell' , the invariant volume element becomes:

(1.5.15)

By the definition of the delta function,

we find that the invariant delta function is

(1.5.16)

Given that and are connected to and through a Lorentz transformation , we arrive at the following equation:

Consequently, the scalar product becomes:

(1.5.17)

The normalization constant is occasionally set to . However, in doing so, one would need to account for the term in scalar products. In this context, we will use the more common convention where:

(1.5.18)

With this choice, the scalar product simplifies to:

(1.5.19)

Next, we turn our attention to the two physically relevant cases: particles with mass and particles with zero mass.

1.5.1Mass Positive-Definite

In this context, the little group is represented by the three-dimensional rotation group. Its unitary representations can be decomposed into a direct sum of irreducible unitary representations, denoted by , having dimensions of , where takes values 0, , 1, etc. These representations can be constructed from the standard matrices for infinitesimal rotations , where is infinitesimal. The representation is given by:

(1.5.20)
(1.5.21)
(1.5.22)

where varies over the set . gives the component of angular momentum in the three-axis. For a particle having mass and spin , Equation (1.5.11) is transformed to:

(1.5.23)

Here, the little-group element — often referred to as the Wigner rotation — is given by Equation (1.5.10) as:

Let

be the Lorentz factor (w.r.t the particle with 4-momentum ). Note that the relativistic mass with 4-momentum (w.r.t the particle with 4-momentum ) is

Hence, together with

we can rewrite the Lorentz factor to be

which gives

Let

Then a choice of that take to could be

Then from this we can determine the Wigner rotation and hence the representation with spin , .

1.5.2Mass Zero

Note that an infinitesimal rotation around the two-axis followed by an infinitesimal boost along the one-axis leaves unchange as

Also an infinitesimal rotation around the two-axis followed by an infinitesimal boost along the one-axis leaves unchange. And clearly, an infinitesimal rotation around the three axis leaves . Hence, an infinitesimal small group transformation can be rewritten as

where

We see that the commutators for these generators are

Hence, we simultaneously diagonalized and by their eigenstates such that

However, if one of and is not zero, then we can find a continuum of spectrum of and , i.e.

where

which contradicts to our assumption that is of discrete (experiment does not find a continuum of for one-particle states). Hence, for physical states, we must have

(For the case when or , see arXiv:1302.1198.) Hence, for a physical state , we must have

Here is assumed to be the eigenvalue of (now that , is a common eigenstate for both , , and , although neither and commute nor and ), such that

Note that is in the three-axis, gives the component of angular momentum in the direction of motion. is called the helicity.

We are now ready to find the representation of the little group.

Hence,

Therefore,

where is determined by

Instead of unitary operator acting on the Hilbert space, we prefer using the following Lorentz transformation identity.

where

where

On the other hand,

where .

Therefore,

Hence,

where we choose

to take to , where

is a pure boost along the three-direction and

with

is a pure rotation that takes the three axis into the direction of .