1 Introduction

When one quantizes gauge theories such as electrodynamics, Yang–Mills (YM) theory, and the Einstein–Hilbert (EH) theory of gravity, special care must be taken to cancel quantum effects that arise from unphysical gauge fields. Special “ghost” fields have been found that perform this function [1,2,3]. Even after these ghost fields have been introduced, the resulting effective action gives rise to divergent quantum effects that are removed through renormalizing the quantities that characterize the classical theory. This program works well in electrodynamics [4] and YM theory [5, 6]. With the EH action these divergences vanish on mass-shell at one loop order, but do not disappear at two loop order when the equation of motion are satisfied [7, 8].

Many different approaches, such as supergravity [9], string theory [10], loop gravity [11], and asymptotic safety [12], have been proposed to cure this problem with quantum gravity. Perhaps the simplest approach is to introduce a Lagrange multiplier field that ensures that classical equation of motion for the gravitational field is satisfied [13, 14] as then all radiative effects beyond one-loop order are absent [15] with the remaining one-loop effects being twice these one-loop effects coming from the usual quantization procedure. This approach can be applied consistently to gauge theories coupled to matter [16] or in first order form [17]. In Ref. [18], the thermal effects on gauge theory supplemented with Lagrange multiplier fields were investigated. Another example of a field theory in which a Lagrange Multiplier field is used to impose a constraint and is involved in radiative corrections is provided by Ref. [19].

A significant improvement on this approach involves the introduction of a functional determinant into the measure of the path integral that has the effect of rendering the path integral invariant under a field redefinition. It also results in the quantum effects in this approach being exactly equal to the one-loop effects normally encountered. (The factor of two mentioned above that occurs when only using a Lagrange multiplier field no longer arises when this functional determinant also is introduced.) [20].

This functional determinant can be incorporated into the effective action through the introduction of a pair of Fermionic ghosts and a single Bosonic ghost field. These ghosts are similar to Lee–Yang ghost fields [21], which appear in the context of the worldline formalism [22, 23]. In this paper, we consider the consequence of there being a gauge invariance in the classical Lagrangian. It is shown that the Lagrange multiplier field and the ghost fields all participate in this gauge transformation, with additional gauge invariances occurring in the effective Lagrangian.

We then show that if these gauge transformations are closed, then a BRST transformation [24, 25] can be defined which leaves the effective Lagrangian invariant. Such an invariance leads to the full theory being unitary [26] and renormalizable [27].

This paper is organized as follows. In Sect. 2, we review the field redefinition invariant Lagrange multiplier formalism for non-singular classical actions. The case of a classical action with a gauge symmetry is considered in Sect. 3 and we show that additional gauge invariances appear. We then quantize by using an extension of the Faddeev–Popov procedure [1]. Then, in Sect. 4, we derive the corresponding BRST transformations in the framework of the extended Lagrange multiplier formalism. In Sect. 5, we consider YM theory, a non-linear gauge theory, to demonstrate this formalism. We also find the corresponding BRST transformation. In Appendix A, we illustrate the field redefinition invariance of path integrals when using the extended Lagrange multiplier formalism. In Appendix B, we briefly review some of the material of references [13,14,15,16,17,18], to show how the Lagrange multiplier field eliminates loop effects beyond one loop order and that no one-particle irreducible contributions involving two or more Lagrange multiplier fields occur at one-loop. The propagators of the Yang–Mills theory are also derived. In Appendix C, we study the gauge algebra of YM type gauge theories in the extended Lagrange multiplier theory. A superdeterminant identity is derived in detail in the Appendix D. In Appendix E, we obtain the Zinn-Justin master equation [28] associated with the BRST transformation derived in Sect. 4. The relation with the Batalin–Vilkovisky formalism [29, 30] is briefly discussed.

2 The extended Lagrange multiplier theory

If one has a set of classical fields \( \phi _{i} (x) \) whose classical Lagrangian is \( {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) \), then a Lagrangian multiplier field \( \lambda _{i} (x) \) can be used to ensure that the classical equation of motion

$$\begin{aligned} {\mathcal {L}}_{\text {cl}, i} \equiv \frac{\partial {\mathcal {L}}_{\text {cl}}}{\partial \phi _{i}} =0 \end{aligned}$$
(2.1)

is obeyed. The path integral

$$\begin{aligned} I = \int \mathop {{\mathcal {D}} \phi _i} \mathop {{\mathcal {D}} \lambda _{i}} \exp { i \int \mathop {d^{}x}} \left( {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) + \lambda _{i} {\mathcal {L}}_{\text {cl}, i} ( \phi _{i} )\right) \end{aligned}$$
(2.2)

can in principle be evaluated using the functional analogue of the usual results

$$\begin{aligned} \int _{- \infty }^{\infty } \frac{ \mathop {d \lambda }}{2 \pi } \exp \left( i \lambda f (x)\right)&=\mathop {\delta } \left( f (x)\right) \end{aligned}$$
(2.3a)

and

$$\begin{aligned} \int _{- \infty }^{\infty } \exp ( i I[x]) \mathop {\delta } \left( f(x)\right)&= \sum _{i}^{} \exp \left( i I[x_{i}] \right) | f'(x_{i} ) |^{-1}, \end{aligned}$$
(2.3b)

where \( f( x_{i} ) = 0\). It follows that [16]

$$\begin{aligned} I = \sum _{i}^{} \exp \left( i \int \mathop {d^{}x} {\mathcal {L}}_{\text {cl}} ( {\bar{\phi }}_{i} (x))\right) \left[ \det {\mathcal {L}}_{\text {cl}, ij} ( {\bar{\phi }}_{i} )\right] ^{-1},\nonumber \\ \end{aligned}$$
(2.4)

where

$$\begin{aligned} {\mathcal {L}}_{\text {cl}, i} \left( {\bar{\phi }}_{ i} (x)\right) =0. \end{aligned}$$
(2.5)

In Eq. (2.4), the exponential is the sum of all tree-level diagrams arising from \( {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) \) alone [31]. The determinant is the square of the determinant coming from all one-loop diagrams and hence all one-loop results that normally arise acquire an extra factor of two. No contributions corresponding to higher-loop diagrams occur in the exact result of Eq. (2.4).

In Ref. [20], a further modification of the functional integral in Eq. (2.2) is introduced. The functional measure in Eq. (2.2) is supplemented by the functional determinant

$$\begin{aligned} \det \nolimits ^{1/2} [ {\mathcal {L}}_{\text {cl}, ij} ( \phi _{i} )]. \end{aligned}$$
(2.6)

This is shown to leave I invariant under a field redefinition.Footnote 1 It is also apparent from combing Eqs. (2.4) and (2.6) that all quantum effects now reduce to the one-loop effects arising from \( {\mathcal {L}}_{\text {cl}} ( \phi _{i} )\) alone.

We can exponentiate the functional determinant of Eq. (2.6) through use of a Bosonic ghost field \( \chi _{i} (x) \) and a pair of Fermionic ghost fields \( \psi _{i} (x)\) and \( \theta _{i} (x)\) [32]

$$\begin{aligned}{} & {} \det \nolimits ^{1/2} {\mathcal {L}}_{\text {cl}, ij} ( \phi _{i} ) \nonumber \\{} & {} \quad = \int \mathop {{\mathcal {D}} \chi _{i}} \mathop {{\mathcal {D}} \psi _{i}} \mathop {{\mathcal {D}} \theta _{i}} \exp i \int \mathop {d^{}x} \nonumber \\{} & {} \qquad \times \left[ {\mathcal {L}}_{\text {cl}, ij} ( \phi _{i} (x)) \left( \psi _{i} (x) \theta _{j} (x) + \frac{1}{2} \chi _{i} (x) \chi _{j} (x) \right) \right] .\nonumber \\ \end{aligned}$$
(2.7)

The full classical effective Lagrangian is now

$$\begin{aligned} {\mathcal {L}}_{\text {eff}} ( \phi _{i}, \lambda _{i}, \chi _{i}, \psi _{i}, \theta _{i} )= & {} {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) + \lambda _{i} {\mathcal {L}}_{\text {cl}, i} \nonumber \\{} & {} + \left( \psi _{i} \theta _{j} + \frac{1}{2} \chi _{i} \chi _{j}\right) {\mathcal {L}}_{\text {cl},ij}. \end{aligned}$$
(2.8)

If the classical action,Footnote 2

$$\begin{aligned} S_{\text {cl}} = \int \mathop {d x} {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) \end{aligned}$$
(2.9)

is non-singular, then the quantization of the Lagrangian (2.8) can be realized through the standard path integral procedure [20]

$$\begin{aligned} Z[ {\varvec{J}} ]= & {} \int \mathop {{\mathcal {D}} {\varvec{\phi _{i}}}} \exp i \int \mathop {d x} \left( {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) + \lambda _{i} {\mathcal {L}}_{\text {cl}, i}\right. \nonumber \\{} & {} \left. + \left( \psi _{i} \theta _{j} + \frac{1}{2} \chi _{i} \chi _{j}\right) {\mathcal {L}}_{\text {cl},ij} + \bar{{\varvec{J}}}_{i} {\varvec{\phi }}_{i} \right) , \end{aligned}$$
(2.10)

where we used a compact notation in which \( {\varvec{\phi }}_{i} = ( \phi _{i}, \lambda _{i},\) \( \chi _{i}, \psi _{i}, \theta _{i} )\), \( \mathop {{\mathcal {D}} {\varvec{\phi }}_{i} } \equiv \mathop {{\mathcal {D}} \phi _{i}} \mathop {{\mathcal {D}} \lambda _{i}} \mathop {{\mathcal {D}} \chi _{i}} \mathop {{\mathcal {D}} \psi _{i}} \mathop {{\mathcal {D}} \theta _{i} }\). We also included a source \( \bar{{\varvec{J}}}_{i} \equiv ( J_{i}, J_{i} + K_{i}, L_{i}, {\bar{\eta }}_{i}, {\bar{\kappa }}_{i} )\), where \( J_{i} \), \( K_{i} \), \( L_{i} \) are ordinary sources and \( {\bar{\eta }}_{i} \), \( {\bar{\kappa }}_{i} \) are fermionic sources. Thus, the expanded form of the source term in Eq. (2.10) is given by

$$\begin{aligned} \bar{{\varvec{J}}}_{i} {\varvec{\phi }}_{i} = J_{i} (\phi _{i} + \lambda _{i} ) + K_{i} \lambda _{i} + L_{i} \chi _{i} + {\bar{\eta }}_{i} \psi _{i} + {\bar{\kappa }}_{i} \theta _{i}. \end{aligned}$$
(2.11)

3 Gauge invariance

We now consider the consequences of \( {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) \) being invariant under the local infinitesimal transformation

$$\begin{aligned} \phi _{i} \rightarrow \phi _{i} ' = \phi _{i} + H_{ij} ( \phi _{i} ) \xi _{j} \end{aligned}$$
(3.1)

so that

$$\begin{aligned} {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) = {\mathcal {L}}_{\text {cl}} ( \phi ') = {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) + {\mathcal {L}}_{\text {cl},i} H_{ij} \xi _{j}. \end{aligned}$$
(3.2)

From Eq. (3.2) it follows that

$$\begin{aligned} {\mathcal {L}}_{\text {cl}, i} H_{ij} \xi _{j} =0 \end{aligned}$$
(3.3)

and so immediately it follows that \( {\mathcal {L}}_{\text {eff}} \) in Eq. (2.8) is also invariant under the local infinitesimal gauge transformation

$$\begin{aligned} \lambda _{i} \rightarrow \lambda _{i} ' + H_{ij} ( \phi _{i} ) \zeta _{j}. \end{aligned}$$
(3.4)

From Eq. (3.1) it follows that

$$\begin{aligned} \frac{\partial {\mathcal {L}}_{\text {cl}}}{\partial \phi _{i}} = \frac{\partial \phi _{j} '}{\partial \phi _{i}} \frac{\partial {\mathcal {L}}_{\text {cl}}}{\partial \phi _{j} '} = \left( \delta _{ji} + \frac{\partial H_{jk}}{\partial \phi _{i}} \xi _{k}\right) \frac{\partial {\mathcal {L}}_{\text {cl}}}{\partial \phi _{j} '} \end{aligned}$$
(3.5)

and consequently that

$$\begin{aligned} \frac{\partial ^{2} {\mathcal {L}}_{\text {cl}}}{\partial \phi _{i} \partial \phi _{j}} = \frac{\partial \phi _{m} '}{\partial \phi _{i}} \frac{\partial }{\partial \phi _{m} '} \left[ \frac{\partial {\mathcal {L}}_{\text {cl}}}{\partial \phi _{j} '} + \frac{\partial H_{nk}}{\partial \phi _{j}} \xi _{k} \frac{\partial {\mathcal {L}}_{\text {cl}}}{\partial \phi _{n} '}\right] \end{aligned}$$
(3.6)

which to leading order in \( \xi _{i} \) reduces to

$$\begin{aligned} \frac{\partial ^{2} {\mathcal {L}}_{\text {cl}}}{\partial \phi _{i} \partial \phi _{j}}= & {} \frac{\partial ^{2} {\mathcal {L}}_{\text {cl}}}{\partial \phi _{i} ' \partial \phi _{j} '} + \left( \frac{\partial ^{2} H_{lk}}{\partial \phi _{i} \partial \phi _{j}} \xi _{k}\right) \frac{\partial {\mathcal {L}}_{\text {cl}}}{\partial \phi _{l} '} \nonumber \\{} & {} + \left( \frac{\partial H_{lk}}{\partial \phi _{i}} \frac{\partial ^{2} {\mathcal {L}}_{cl}}{\partial \phi _{j} ' \partial \phi _{l} '} + \frac{\partial H_{lk}}{\partial \phi _{j}} \frac{\partial ^{2} {\mathcal {L}}_{cl}}{\partial \phi _{i} ' \partial \phi _{l} '}\right) \xi _{k}. \end{aligned}$$
(3.7)

Insertion of Eqs. (3.1), (3.5) and (3.7) into Eq. (2.8) and collecting terms dependent on \( \partial {\mathcal {L}}_{\text {cl}} / \partial \phi _{i} '\) and \( \partial ^{2} {\mathcal {L}}_{\text {cl}} / \partial \phi _{i} ' \partial \phi _{j} ' \) leads to invariance of \( {\mathcal {L}}_{\text {eff}} \) under the transformations

$$\begin{aligned} \lambda _{i} \rightarrow \lambda _{i} '&= \lambda _{i} + \left[ \frac{\partial H_{ik}}{\partial \phi _{j}} \lambda _{j}\right. \nonumber \\&\left. \quad + \frac{\partial ^{2} H_{ik}}{\partial \phi _{m} \partial \phi _{n}}\left( \psi _{m} \theta _{n} + \frac{1}{2} \chi _{m} \chi _{n}\right) \right] \xi _{k} , \end{aligned}$$
(3.8a)
$$\begin{aligned} \psi _{i} \rightarrow \psi _{i} '&= \psi _{i} + \frac{\partial H_{ik}}{\partial \phi _{j}} \psi _{j} \xi _{k}, \end{aligned}$$
(3.8b)
$$\begin{aligned} \theta _{i} \rightarrow \theta _{i} '&= \theta _{i} + \frac{\partial H_{ik}}{\partial \phi _{j}} \theta _{j} \xi _{k}, \end{aligned}$$
(3.8c)
$$\begin{aligned} \chi _{i} \rightarrow \chi _{i} '&= \chi _{i} + \frac{\partial H_{ik}}{\partial \phi _{j}} \chi _{j} \xi _{k}. \end{aligned}$$
(3.8d)

In addition, there are also the gauge transformations

$$\begin{aligned} \lambda _{i} \rightarrow \lambda _{i} '&=\lambda _{i} + H_{ik, j} \left( \chi _{j} \sigma _{k} + \psi _{j} \tau _{k} - \theta _{j} \pi _{k} \right) , \end{aligned}$$
(3.9a)
$$\begin{aligned} \psi _{i} \rightarrow \psi _{i} '&= \psi _{i} + H_{ij} \pi _{j}, \end{aligned}$$
(3.9b)
$$\begin{aligned} \theta _{i} \rightarrow \theta _{i} '&=\theta _{i} + H_{ij} \tau _{j} , \end{aligned}$$
(3.9c)
$$\begin{aligned} \chi _{i} \rightarrow \chi _{i} '&= \chi _{i} + H_{ij} \sigma _{j} ; \end{aligned}$$
(3.9d)

where \( \pi _{i} \) and \( \tau _{i} \) are Fermionic gauge functions and \( \sigma _{i} \) is a Bosonic gauge function. These follow from:

$$\begin{aligned} ( {\mathcal {L}}_{\text {cl}, i} H_{ik} )_{,j} =0 \end{aligned}$$
(3.10)

which is a consequence of Eq. (3.3).

We now will introduce a gauge fixing Lagrangian

$$\begin{aligned} {\mathcal {L}}_{\text {gf}}= & {} - \frac{1}{2 \alpha } ( F_{ij} \phi _{j} )^{2} - \frac{1}{\alpha } ( F_{im} \lambda _{m} ) ( F_{i n} \phi _{n} )\nonumber \\{} & {} - \frac{1}{2 \alpha } ( F_{ij} \chi _{j} )^{2} - \frac{1}{\alpha } ( F_{im} \psi _{m} )( F_{in} \theta _{n} ). \end{aligned}$$
(3.11)

With Bosonic Nakanishi–Lautrup fields [33, 34] \( B_{i} \), \( E_{i} \) and \( G_{i} \) and similarly the Fermionic fields \( \Omega _{i} \) and \( \Xi _{i} \), this can be written as

$$\begin{aligned} {\mathcal {L}}_{\text {gf}}= & {} \frac{\alpha }{2} \left( - E_{i}^{2} + 2 B_{i} E_{i} +G_{i}^{2} - 2 \Xi _{i} \Omega _{i}\right) \nonumber \\{} & {} - E_{i} ( F_{ij} \lambda _{j} ) - B_{i} ( F_{ij} \phi _{j} ) - G_{i} ( F_{ij} \chi _{j} ) \nonumber \\{} & {} - \Xi _{i} (F_{ij} \theta _{j} ) - \Omega _{i} ( F_{ij} \psi _{j} ). \end{aligned}$$
(3.12)

Having fixed the gauge invariance of Eqs. (3.1), (3.4), (3.8) and (3.9) in this way, we now introduce the Faddeev–Popov ghost contribution to the path integral

$$\begin{aligned} I= & {} \int \mathop {{\mathcal {D}} \phi _{i}} \mathop {{\mathcal {D}} \lambda _{i}} \mathop {{\mathcal {D}} \chi _{i}} \mathop {{\mathcal {D}} \psi _{i} \mathop {{\mathcal {D}} \theta _{i}}} \exp i \int \mathop {d^{}x} \nonumber \\{} & {} \times \left[ {\mathcal {L}}_{\text {cl}} ( \phi _{i} ) + {\mathcal {L}}_{\text {cl}, i} \lambda _{i} + {\mathcal {L}}_{\text {cl}, ij} ( \phi _{i} ) \left( \psi _{ i} \theta _{j} + \frac{1}{2} \chi _{i} \chi _{j}\right) \right] .\nonumber \\ \end{aligned}$$
(3.13)

This involves first inserting the constant

(3.14)

where

$$\begin{aligned} X_{jp} = H_{jp, k} \lambda _{k} + H_{jp,mn} \left( \psi _{m} \theta _{n} + \frac{1}{2} \chi _{m} \chi _{n} \right) , \end{aligned}$$
(3.15)

is the \(5 \times 5\) matrix appearing in the argument of the \( \delta \)-function in Eq. (3.14) and is the Faddeev–Popov superdeterminant.

We then perform the gauge transformations of Eqs. (3.1), (3.4), (3.8), (3.9) with \( (- \xi _{i}, - \zeta _{i}, - \sigma _{i}, - \pi _{i}, - \tau _{i} )\) and then insert the constant

$$\begin{aligned}{} & {} \int \mathop {{\mathcal {D}} p_{i}} \mathop {{\mathcal {D}} q_{i}} \mathop {{\mathcal {D}} r_{i}} \mathop {{\mathcal {D}} s_{i}} \mathop {{\mathcal {D}} t_{i}} \exp i \int \mathop {d^{}x} \left( - \frac{1}{2 \alpha } p_{i}^{2} \right. \nonumber \\{} & {} \left. \quad - \frac{1}{\alpha } p_{i} q_{i} - \frac{1}{2 \alpha } r_{i}^{2} - \frac{1}{\alpha } s_{i} t_{i}\right) . \end{aligned}$$
(3.16)

The integral over \( ( \xi _{i}, \zeta _{i}, \sigma _{i}, \pi _{i}, \tau _{i} ) \) are now innocuous multiplicative factors and the \( \delta \)-functions in Eq. (3.14) make it possible to integrate over \( ( p_{i}, q_{i}, r_{i}, s_{i}, t_{i} )\). This leaves us with

(3.17)

with \( {\mathcal {L}}_{\text {eff}} \) given by Eq. (2.8), \( {\mathcal {L}}_{\text {gf}}\) by Eq. (3.12), and being the Faddeev–Popov superdeterminant of Eq. (3.14). This functional determinant can be exponentiated using Fermionic ghost fields \( ( {\bar{c}}_{i}, c_{i} )\), \( ( {\bar{d}}_{i}, d_{i} )\), \( ( {\bar{e}}_{i}, e_{i} )\) and Bosonic ghost fields \( ( {\tilde{\gamma }}_{i}, \gamma _{i} )\), \( ( {\tilde{\epsilon }}_{i}, \epsilon _{i} )\) [32].Footnote 3 This leads to

(3.18)

where we used

$$\begin{aligned} \mathop {\textrm{Sdet}} \begin{pmatrix} 0 &{} A &{} 0 &{} 0 &{} 0 \\ A &{} B &{} C &{} -D &{} E \\ 0 &{} C &{} A &{} 0 &{} 0 \\ 0 &{} E &{} 0 &{} A &{} 0 \\ 0 &{} D &{} 0 &{} 0 &{} A \\ \end{pmatrix} = \mathop {\textrm{Sdet}} \begin{pmatrix} 0 &{} A &{} 0 &{} 0 &{} 0 \\ A &{} A+B &{} C &{} -D &{} E \\ 0 &{} C &{} A &{} 0 &{} 0 \\ 0 &{} E &{} 0 &{} A &{} 0 \\ 0 &{} D &{} 0 &{} 0 &{} A \\ \end{pmatrix}.\nonumber \\ \end{aligned}$$
(3.19)

The argument of the exponential in Eq. (3.18) defines the Faddeev–Popov action, \( \int \mathop {dx} {\mathcal {L}}_{\text {FP}} \).

Note that,

$$\begin{aligned} \mathop {\textrm{Sdet}} \begin{pmatrix} 0 &{} A &{} 0 &{} 0 &{} 0 \\ A &{} A &{} C &{} -D &{} E \\ 0 &{} C &{} A &{} 0 &{} 0 \\ 0 &{} E &{} 0 &{} A &{} 0 \\ 0 &{} D &{} 0 &{} 0 &{} A \\ \end{pmatrix} = \mathop {\textrm{det}} A \end{aligned}$$
(3.20)

(see the Appendix C for a derivation). Thus, the Faddeev–Popov superdeterminant in the extended Lagrange multiplier formalism is equal to \( \det F_{ij} H_{jp} \), i.e., the usual Faddeev–Popov determinant which is introduced when there is no Lagrange multiplier.

We now will consider the possibility of there being a global gauge invariance (a “BRST” invariance [24, 25]) in the total Lagrangian

$$\begin{aligned} {\mathcal {L}}_{\text {T}} = {\mathcal {L}}_{\text {eff}} + {\mathcal {L}}_{\text {gf}} + {\mathcal {L}}_{\text {FP}}. \end{aligned}$$
(3.21)

4 Global gauge invariance

In keeping with how BRST symmetry is introduced for gauge theories, we begin by replacing the gauge functions \( \xi _{i} \), \( \zeta _{i} \), \( \sigma _{i} \), \( \pi _{i} \) and \( \tau _{i} \) with the ghost fields \( c_{i} \), \( d_{i} \), \( e_{i} \), \( \gamma _{i} \) and \( \epsilon _{i} \) multiplied by \( \eta \), a constant Fermionic scalar, so that the BRST transformations of the classical fields are

$$\begin{aligned} \delta \phi _{i}&= H_{ij} c_{j} \eta , \end{aligned}$$
(4.1a)
$$\begin{aligned} \delta \lambda _{i}&= H_{ij} d_{j} \eta + \left[ H_{ij,k}\lambda _{k} \right. \nonumber \\&\left. \quad + H_{ij,mn} \left( \psi _{m} \theta _{n} + \frac{1}{2} \chi _{m} \chi _{n}\right) \right] c_j \eta \nonumber \\&\quad + H_{ij,k}( \chi _{k} e_{j} + \psi _{k} \epsilon _{j} - \theta _{k}{\gamma }_{j}) \eta ,\end{aligned}$$
(4.1b)
$$\begin{aligned} \delta \chi _{i}&= H_{ij} e_{j} \eta + H_{ij,k}\chi _{k} c_{j} \eta , \end{aligned}$$
(4.1c)
$$\begin{aligned} \delta \psi _{i}&= H_{ij} \gamma _{j} \eta + H_{ij,k} \psi _{k} c_{j} \eta , \end{aligned}$$
(4.1d)
$$\begin{aligned} \delta \theta _{i}&= H_{ij} \epsilon _{j} \eta + H_{ij,k} {\theta }_{k} c_{j} \eta . \end{aligned}$$
(4.1e)

In addition to these BRST transformations, it follows immediately that we also have these transformations:

$$\begin{aligned} \delta B_{i} = \delta E_{i} = \delta G_{i} = \delta \Xi _{i} = \delta \Omega _{i} =0 \end{aligned}$$
(4.2)

and

$$\begin{aligned} \delta {\bar{c}}_{i}= & {} - B_{i} \eta , \ \delta {\bar{d}}_{i} = - E_{i} \eta , \ \delta {\bar{e}}_{i} = - G_{i} \eta , \nonumber \\ \delta {\tilde{\gamma }}_{i}= & {} - \Omega _{i} \eta , \ \delta {\tilde{\epsilon }}_{i} = - \Xi _{i} \eta . \end{aligned}$$
(4.3)

The transformation of \( c_{i} \), \( d_{i} \), \( e_{i} \), \( \gamma _{i} \) and \( \delta _{i} \) now are determined by the requirements that

$$\begin{aligned}&\delta ( H_{ij} c_{j} ) = 0 \end{aligned}$$
(4.4a)
$$\begin{aligned}&\delta \left[ H_{ij} d_{j} + \left[ H_{ij,k}\lambda _{k} + H_{ij,mn} \left( \psi _{m} \theta _{n} + \frac{1}{2} \chi _{m} \chi _{n}\right) \right] c_j \right. \nonumber \\&\left. \quad + H_{ij,k}( \chi _{k} e_{j} + \psi _{k} \epsilon _{j} - \theta _{k}{\gamma }_{j}) \right] =0 \end{aligned}$$
(4.4b)
$$\begin{aligned}&\delta ( H_{ij} e_{j} + H_{ij,k}\chi _{k} c_{j} ) =0 , \end{aligned}$$
(4.4c)
$$\begin{aligned}&\delta (H_{ij} \gamma _{j} + H_{ij,k} \psi _{k} c_{j} ) =0, \end{aligned}$$
(4.4d)
$$\begin{aligned}&\delta ( H_{ij} \epsilon _{j} + H_{ij,k} {\theta }_{k} c_{j} ) =0 . \end{aligned}$$
(4.4e)

In order to do this, we must first impose conditions on \( H_{ik} ( \phi _{i} )\). We consider the commutator of two gauge transformations of the form of Eq. (3.1),

$$\begin{aligned}{}[ \delta _{\lambda }, \delta _{\sigma } ] \phi _{i} = \left[ \left( \frac{\partial H_{ij}}{\partial \phi _{k}} \sigma _{j} \right) ( H_{kl} \lambda _{l} ) - \left( \frac{\partial H_{ij}}{\partial \phi _{k}} \lambda _{j}\right) H_{kl} \sigma _{l}\right] ,\nonumber \\ \end{aligned}$$
(4.5)

which is itself a gauge transformation, so that

$$\begin{aligned} \left( H_{ij, k} H_{kl} - H_{il, k} H_{kj}\right) \sigma _{j} \lambda _{l} = f_{jl|k} H_{ik} \lambda _{l} \sigma _{j} \end{aligned}$$
(4.6)

if the gauge transformation is “closed”. For an “open” gauge transformation, Eq. (4.6) is satisfied only if \( \phi _{i} \) satisfies the classical equation of motion (“on the mass shell”).

We will only consider gauge transformations that are closed with the additional restriction that \( H_{ij} ( \phi _{i} )\) is at most linear in \( \phi _{i} \) so that

$$\begin{aligned} H_{ij , m n}&=0, \end{aligned}$$
(4.7a)
$$\begin{aligned} f_{jl|k, i}&=0. \end{aligned}$$
(4.7b)

Gauge theories that satisfy these conditions are the so-called YM type theories [35]. Besides the YM theory, the EH action is another interesting example of a gauge theory of this type. In the next section, we will consider the quantization of the YM in the extended Lagrange multiplier formalism.

From Eq. (4.4a), we obtain that

$$\begin{aligned} \begin{aligned} \delta ( H_{ij} c_{j} )&= \frac{\partial H_{ij}}{\partial \phi _{k}} ( H_{kl} c_{l} \eta ) c_{j} + H_{ij} \delta c_{j} \\&= - \frac{1}{2} \left( \frac{\partial H_{ij}}{\partial \phi _{k}} H_{kl} - \frac{\partial H_{il}}{\partial \phi _{k}} H_{kj}\right) c_{l} c_{j} \eta + H_{ij} \delta c_{j}. \end{aligned}\nonumber \\ \end{aligned}$$
(4.8)

From Eqs. (4.6) and Eq. (4.8), we see that

$$\begin{aligned} \delta c_{j} = - \frac{1}{2} f_{mn | j} c_{m} c_{n} \eta . \end{aligned}$$
(4.9)

To obtain \( \delta e_{j} \), we now examine Eq. (4.4c),

$$\begin{aligned}{} & {} H_{ij,k} \delta \phi _{l} \chi _{k} c_{j} + H_{ij,k} ( \delta \chi _{k} c_{j} + \chi _{k} \delta c_{j} )\nonumber \\{} & {} \quad + H_{ij, k} \delta \phi _{k} e_{j} + H_{ij} \delta e_{j} =0. \end{aligned}$$
(4.10)

Upon using Eqs. (4.7a), (4.1c) and (4.9), (4.10) becomes

$$\begin{aligned}{} & {} H_{ij,k} \left[ \left( \frac{\partial H_{kl}}{\partial \phi _{m}} \chi _{m} c_{l} \right. \right. \nonumber \\{} & {} \left. \left. \quad + H_{kl} e_{l}\right) \eta c_{j} + \chi _{k} \left( - \frac{1}{2} f_{m n|j} c_{m} c_{n} \eta \right) \right] \nonumber \\{} & {} \quad + H_{ij,k} ( H_{kl} c_{l} \eta ) e_{j} + H_{ij} \delta e_{j} =0. \end{aligned}$$
(4.11)

Eq. (4.6) reduces (4.11) to simply

$$\begin{aligned} \delta e_{j}&= - f_{m n | j} c_{m} e_{n} \eta . \end{aligned}$$
(4.12a)

In a similar way, Eqs. (4.4d) and (4.4e) lead to

$$\begin{aligned} \delta \gamma _{j}&=- f_{mn |j} c_{m} \gamma _{n} \eta , \end{aligned}$$
(4.12b)
$$\begin{aligned} \delta \epsilon _{j}&=- f_{mn |j} c_{m} \epsilon _{n} \eta . \end{aligned}$$
(4.12c)

Finally, from Eq. (4.4b), we obtain

$$\begin{aligned} \delta d_{j}&=- f_{mn |j} \left( d_{m} c_{n} + \frac{1}{2} e_{m} e_{n} + \gamma _{m} \epsilon _{n}\right) \eta . \end{aligned}$$
(4.12d)

In the Appendix D, we find the corresponding Zinn-Justin master equation [28] which follows from the global gauge invariance in Eqs. (4.1), (4.2), (4.3), (4.9) and (4.12).

5 Yang–Mills theory

The classical YM Lagrangian reads

$$\begin{aligned} {\mathfrak {L}}_{\text {cl}}= - \frac{1}{4} F_{\mu \nu }^{a} F^{a \, \mu \nu }, \end{aligned}$$
(5.1)

where

$$\begin{aligned} F_{\mu \nu }^{a} = \partial _{\mu } A_{\nu }^{a} - \partial _{\nu } A_{\mu }^{a} + g f^{abc} A_{\mu }^{b} A_{\nu }^{c}. \end{aligned}$$
(5.2)

It is invariant under the gauge transformation

$$\begin{aligned} A_{\mu }^{a} \rightarrow A_{\mu }^{' a} = A_{\mu }^{a} + D_{\mu }^{ab} (A) \xi ^{b} \end{aligned}$$
(5.3)

in which \( D_{\mu }^{ab} (A) \equiv \partial _{\mu } \delta ^{ab} + g f^{apb} A_{\mu }^{p} \). Comparing Eq. (5.3) with Eq. (3.1), we can identify \( H_{ij} ( \phi _{i})\) with \( D_{\mu }^{ab} (A)\). Since \( D_{\mu }^{ab} (A)\) satisfies the conditions (4.7) and we restrict the gauge group to compact semisimple Lie groups, the quantization of the YM theory in the extended Lagrange multiplier formalism follows identically.

Replacing the YM classical Lagrangian in Eq. (5.1) into Eq. (2.8) yields the classical effective YM Lagrangian in the framework of the extended Lagrange multiplier formalism:

$$\begin{aligned} {\mathfrak {L}}_{\text {eff}}{} & {} ={\mathfrak {L}}_{\text {cl}}+ \lambda _{\mu }^{a} D^{ab }_{\nu } F^{ b \nu \mu } - \frac{1}{4} ( \partial _{\mu } \chi _{\nu }^{a} - \partial _{\nu } \chi _{\mu }^{a} )^{2}\nonumber \\{} & {} \quad - \frac{1}{2} ( \partial _{\mu } \psi _{\nu }^{a} - \partial _{\nu } \psi _{\mu }^{a} ) ( \partial ^{\mu } \theta ^{ a \, \nu } - \partial ^{\nu } \theta ^{a \, \mu } ) \nonumber \\{} & {} \quad - g f^{abc} ( \partial _{\mu } \chi _{\nu }^{a} - \partial _{\nu } \chi _{\mu }^{a} ) A^{b \, \mu } \chi ^{c \, \nu } \nonumber \\{} & {} \quad - \frac{g}{2} f^{abc} (\partial _{\mu } A_{\nu }^{a} - \partial _{\nu } A_{\mu }^{a} ) \chi ^{b \, \mu } \chi ^{c \, \nu } \nonumber \\{} & {} \quad - \frac{g^{2}}{2} f^{abc} f^{ade} A_{\mu }^{b} \chi _{\nu }^{c} A^{d \, \mu } \chi ^{e \, \nu } \nonumber \\{} & {} \quad - \frac{g^{2}}{2} f^{abc} f^{ade} A_{\mu }^{b} \chi _{\nu }^{c} \chi ^{d \, \mu } A^{e \, \nu }\nonumber \\{} & {} \quad - \frac{g^{2}}{2} f^{abc} f^{ade} A_{\mu }^{b} A_{\nu }^{c} \chi ^{d \, \mu } \chi ^{e \, \nu } \nonumber \\ {}{} & {} \quad - g f^{abc} ( \partial _{\mu } \psi _{\nu }^{a} - \partial _{\nu } \psi _{\mu }^{a} ) A^{b \, \mu } \theta ^{c \, \nu }\nonumber \\{} & {} \quad - g f^{abc} A^{b \, \mu } \psi ^{c \, \nu } ( \partial _{\mu } \theta _{\nu }^{a} \nonumber \\{} & {} \quad - \partial _{\nu } \theta _{\mu }^{a} ) - g f^{abc} (\partial _{\mu } A_{\nu }^{a} - \partial _{\nu } A_{\mu }^{a} ) \psi ^{b \, \mu } \theta ^{c \, \nu } \nonumber \\ {}{} & {} \quad - g^2 f^{abc} f^{ade} A_{\mu }^{b} \psi _{\nu }^{c} A^{d \, \mu } \theta ^{e \, \nu } \nonumber \\{} & {} \quad - g^2 f^{abc} f^{ade} A_{\mu }^{b} \psi _{\nu }^{c} \theta ^{d \, \mu } A^{e \, \nu } \nonumber \\{} & {} \quad - g^2 f^{abc} f^{ade} A_{\mu }^{b} A_{\nu }^{c} \psi ^{d \, \mu } \theta ^{e \, \nu }. \end{aligned}$$
(5.4)

By Eqs. (3.8) and (3.9), the gauge invariance (5.3) is now accompanied by

$$\begin{aligned} \lambda _{\mu }^a \rightarrow \lambda _{\mu }^{'a}&= \lambda _{\mu }^a + g f^{abc} \lambda ^{b} \xi ^c , \end{aligned}$$
(5.5a)
$$\begin{aligned} \psi _{\mu }^a \rightarrow \psi _{\mu }^{'a}&= \psi _{\mu }^a + g f^{abc} \psi ^{b} \xi ^c, \end{aligned}$$
(5.5b)
$$\begin{aligned} \theta _{\mu }^a \rightarrow \theta _{\mu }^{'a}&= \theta _{\mu }^a + g f^{ab c} \theta ^{b} \xi ^c, \end{aligned}$$
(5.5c)
$$\begin{aligned} \chi _{\mu }^a \rightarrow \chi _{\mu }^{'a}&= \chi _{\mu }^a + g f^{abc} \chi ^{b} \xi ^c, \end{aligned}$$
(5.5d)

and

$$\begin{aligned} \lambda _{\mu }^a \rightarrow \lambda _{\mu }^{'a}&=\lambda _{\mu }^a + D_{\mu }^{ab}(A) \zeta ^{b}\nonumber \\&\quad + g f^{apb} \left( \chi ^{p}_{\mu } \sigma ^{b} + \psi ^{p}_{\mu } \tau ^{b} - \theta ^{p}_{\mu } \pi ^{b} \right) , \end{aligned}$$
(5.6a)
$$\begin{aligned} \psi _{\mu }^a \rightarrow \psi _{\mu }^{'a}&= \psi _{\mu }^a + D_{\mu }^{ab}(A) \pi ^b, \end{aligned}$$
(5.6b)
$$\begin{aligned} \theta _{\mu }^a \rightarrow \theta _{\mu }^{'a}&=\theta _{\mu }^a + D_{\mu }^{ab}(A) \tau ^b , \end{aligned}$$
(5.6c)
$$\begin{aligned} \chi _{\mu }^a \rightarrow \chi _{\mu }^{'a}&= \chi _{\mu }^a + D_{\mu }^{ab}(A) \sigma ^b . \end{aligned}$$
(5.6d)

Now, in order to quantize the action in Eq. (5.4), we will employ the gauge fixing

$$\begin{aligned} F^{\mu ab} A_{\mu }^{b}\equiv & {} \partial ^{\mu } A_{\mu }^{a} = \partial ^{\mu } \lambda _{\mu }^{a} = \partial ^{\mu } \chi _{\mu }^{a} \nonumber \\= & {} \partial ^{\mu } \psi _{\mu }^{a} = \partial ^{\mu } \theta _{\mu }^{a} =0 \end{aligned}$$
(5.7)

leading to the gauge fixing Lagrangian (see Eq. (3.11))

$$\begin{aligned} {\mathfrak {L}}_{\text {gf}}= & {} - \frac{1}{2 \alpha } (\partial \cdot A^{a} )^{2} - \frac{1}{\alpha } (\partial \cdot \lambda ^{a} ) (\partial \cdot A^{a} ) \nonumber \\{} & {} - \frac{1}{2 \alpha } (\partial \cdot \chi ^{a} )^{2} - \frac{1}{\alpha } (\partial \cdot \psi ^{a} )(\partial \cdot \theta ^{a} ), \end{aligned}$$
(5.8)

where \( \partial \cdot X_{I} \equiv \partial _{\mu } X^{\mu }_{I} \) (I are internal indices).

In the YM theory, we identify \( H_{ij} ( \phi ) \mapsto D_{\mu }^{ab} (A)\) and \( F_{ij} \mapsto F^{\mu ab} \). Using these relation in Eq. (3.14), we find that the Faddeev–Popov superdeterminant is

(5.9)

where we used that \( H_{ij,k} = g f^{ikj} \). We obtain the ghost Lagrangian by replacing Eq. (5.9) in Eq. (3.18) which reads

$$\begin{aligned} {\mathfrak {L}}_{\text {gh}}= & {} {\bar{c}}^{a} \partial \cdot D^{ab} (A + \lambda ) c^{b} + {\bar{d}}^{a} \partial \cdot D^{ab} (A) c^{b} + {\bar{c}}^{a} \partial \cdot D^{ab} (A) d^{b} \nonumber \\{} & {} + {\bar{e}}^{a} \partial \cdot D^{ab} (A)e^{b} + {\tilde{\gamma }}^{a} \partial \cdot D^{ab} (A)\gamma ^{b} + {\tilde{\epsilon }}^{a} \partial \cdot D^{ab} (A)\epsilon ^{b} \nonumber \\{} & {} \quad {+} {\bar{c}}^{a} g f^{apb} \partial \cdot \chi ^p e^{b} {+} {\bar{e}}^{a} g f^{apb} \partial \cdot \chi ^p c^{b} {-}{\bar{c}}^{a} g f^{apb} \partial \cdot \theta ^p \gamma ^{b}\nonumber \\{} & {} {+} {\tilde{\gamma }}^{a} g f^{apb} \partial \cdot {\psi }^p c^{b} {+}{\bar{c}}^{a} g f^{apb} \partial \cdot {\psi }^p \epsilon ^{b} {+} {\tilde{\epsilon }}^{a} g f^{apb} \partial \cdot \theta ^p c^{b}.\nonumber \\ \end{aligned}$$
(5.10)

Thus, the generating functional of the YM theory in this framework is given by

$$\begin{aligned} Z [ {\varvec{J}}, {\varvec{\eta }}, \bar{{\varvec{\eta }}} ]= & {} \int \mathop {{\mathcal {D}} {\varvec{A}}_{\mu }^{a}} \mathop {{\mathcal {D}} {\varvec{c}}^{a} } \mathop {{\mathcal {D}} \bar{{\varvec{c}}}^{a} } \exp i \int \mathop {d x} \left( {\mathfrak {L}}_{ \text {eff} } {+} {\mathfrak {L}}_{\text {gf} } {+} {\mathfrak {L}}_{\text {gh}}\right. \nonumber \\{} & {} \left. + \bar{{\varvec{J}}}^{a \, \mu } {\varvec{A}}_{\mu }^{a} + \bar{{\varvec{\eta }}}^{a} {\varvec{c}}^{a} + \bar{{\varvec{c}}}^{a} {\varvec{\eta }}^{a} \right) , \end{aligned}$$
(5.11)

where we have used a compact notation in which \( {\varvec{A}}_{\mu }^{a} = ( A_{\mu }^{a}, \lambda _{\mu }^{a}, \chi _{\mu }^{a}, \psi _{\mu }^{a}, \theta _{\mu }^{a} )\), \( {\varvec{c}}^{a} = ( c^{a}, d^{a}, e^{a}, \gamma ^{a}, \epsilon ^{a} )\) and

$$\begin{aligned} \mathop {{\mathcal {D}} {\varvec{A}}_{\mu }^{a} }\equiv & {} \mathop {{\mathcal {D}} A_{\mu }^{a} } \mathop {{\mathcal {D}} \lambda _{\mu }^{a}} \mathop {{\mathcal {D}} \chi _{\mu }^{a}} \mathop {{\mathcal {D}} \psi _{\mu }^{a}} \mathop {{\mathcal {D}} \theta _{\mu }^{a}}, \quad \nonumber \\ \mathop {{\mathcal {D}} {\varvec{c}}^{a} }\equiv & {} \mathop {{\mathcal {D}} c^{a}} \mathop {{\mathcal {D}} d^{a}} \mathop {{\mathcal {D}} e^{a}} \mathop {{\mathcal {D}} \gamma ^{a}} \mathop {{\mathcal {D}} \epsilon ^{a}}, \quad \nonumber \\ \mathop {{\mathcal {D}} \bar{{\varvec{c}}}^{a} }\equiv & {} \mathop {{\mathcal {D}} {\bar{c}}^{a}} \mathop {{\mathcal {D}} {\bar{d}}^{a}} \mathop {{\mathcal {D}} {\bar{e}}^{a}} \mathop {{\mathcal {D}} {\tilde{\gamma }}^{a}} \mathop {{\mathcal {D}} {\tilde{\epsilon }}^{a}}. \end{aligned}$$
(5.12)

The sources terms are defined as

$$\begin{aligned} \bar{{\varvec{J}}}^{a \, \mu } {\varvec{A}}_{\mu }^{a} \equiv {}&J^{a \, \mu } (A_{\mu }^{a} + \lambda _{\mu }^{a} )+ K^{a \, \mu } \lambda _{\mu }^{a} \nonumber \\&+ L^{a \, \mu } \chi _{\mu }^{a} + {\bar{\eta }}^{a \, \mu } \psi _{\mu }^{a} + {\bar{\kappa }}^{a \, \mu } \theta _{\mu }^{a}, \end{aligned}$$
(5.13a)
$$\begin{aligned} \bar{{\varvec{\eta }}}^{a} {\varvec{c}}^{a} \equiv {}&{\bar{\eta }}^{a} (c^{a} + d^{a} )+ {\bar{\kappa }}^{a} d^{a} + {\bar{\upsilon }}^{a} e^{a} + {\tilde{J}}^{a} \gamma ^{a} + {\tilde{K}}^{a} \epsilon ^{a}, \end{aligned}$$
(5.13b)
$$\begin{aligned} \bar{{\varvec{c}}}^{a} {\varvec{\eta }}^{a} \equiv {}&( {\bar{c}}^{a} + {\bar{d}}^{a} ) \eta ^{a} + {\bar{d}}^{a} \kappa ^{a} + {\bar{e}}^{a} \upsilon ^{a} + {\tilde{\gamma }}^{a} J^{a} + {\tilde{\epsilon }}^{a} K^{a}, \end{aligned}$$
(5.13c)

where (\( J^{a \, \mu } \), \( K^{a \, \mu } \), \( L^{a \, \mu } \)) are real ordinary sources, (\( {J}^{a} \), \( {K}^{a} \)) are complex ordinary sources, and (\( \eta ^{a \, \mu } \), \( \kappa ^{a \, \mu } \), \( \eta ^{a} \), \( \kappa ^{a} \), \( \upsilon ^{a} \)) are Fermionic sources. The Feynman rules derived from Eq. (5.11) are presented in Appendix B.

5.1 BRST transformation

We also can obtain the BRST transformation that leaves the action in Eq. (5.11) invariant. Replacing \(H_{ij} ( \phi _{i}) \mapsto D_{\mu }^{ab} (A) \) in Eqs. (4.1), (4.6), (4.9) and (4.12), we find that the total Lagrangian in Eq. (5.11) must be invariant under the following transformationFootnote 4

$$\begin{aligned} \delta A_{\mu }^{a}&= D_{\mu }^{ab}(A) c^{b} \eta , \end{aligned}$$
(5.14a)
$$\begin{aligned} \delta \lambda _{\mu }^{a}&= D_{\mu }^{ab} (A)d^{b} \eta + g f^{abc} \lambda ^{b} c^{c} \eta \nonumber \\&\quad + g f^{abc} ( \chi ^b e^c + \psi ^b \epsilon ^c - \theta ^b{\gamma }^c) \eta , \end{aligned}$$
(5.14b)
$$\begin{aligned} \delta \chi _{\mu }^a&= D_{\mu }^{ab} (A)e^b \eta + g f^{abc}\chi ^b c^c \eta , \end{aligned}$$
(5.14c)
$$\begin{aligned} \delta \psi _{\mu }^a&= D_{\mu }^{ab}(A) \gamma ^b \eta + g f^{abc} \psi ^b c^c \eta , \end{aligned}$$
(5.14d)
$$\begin{aligned} \delta \theta _{\mu }^a&=D_{\mu }^{ab} (A)\epsilon ^b \eta + g f^{abc} {\theta }^b c^c \eta , \end{aligned}$$
(5.14e)
$$\begin{aligned} \delta c_{\mu }^a&= -\frac{g}{2} f^{abc} c^b c^c \eta \end{aligned}$$
(5.14f)
$$\begin{aligned} \delta e_{\mu }^a&= - g f^{abc} c^b e^c \eta \end{aligned}$$
(5.14g)
$$\begin{aligned} \delta \gamma _{\mu }^a&=- g f^{abc} c^b \gamma ^c \eta , \end{aligned}$$
(5.14h)
$$\begin{aligned} \delta \epsilon _{\mu }^a&=- gf^{abc} c^b \epsilon ^c \eta \end{aligned}$$
(5.14i)
$$\begin{aligned} \delta d_{\mu }^a&=- gf^{abc} \left( d^b c^c + \frac{1}{2} e^b e^c + \gamma ^b \epsilon ^c\right) \eta , \end{aligned}$$
(5.14j)

and we also have

$$\begin{aligned} \delta {\bar{c}}^a= & {} - \frac{1}{\alpha } \partial \cdot {(A + \lambda )}^{a} \eta , \nonumber \\ \delta {\bar{d}}^a= & {} -\frac{1}{\alpha }\partial \cdot A^{a} \eta , \ \delta {\bar{e}}^a = -\frac{1}{\alpha }\partial \cdot \chi ^{a} \eta , \nonumber \\ \delta {\tilde{\gamma }}^a= & {} -\frac{1}{\alpha } \partial \cdot \theta ^{a} \eta , \ \delta {\tilde{\epsilon }}^a = \frac{1}{\alpha } \partial \cdot \psi ^a \eta . \end{aligned}$$
(5.15)

This transformation can be used to show that unitarity is retained in the YM theory in the framework of the extended Lagrange multiplier theory [16].

6 Discussion

It has been shown [13,14,15] that by using a Lagrange multiplier field to ensure that the equations of motion are satisfied, radiative effects beyond one-loop order are suppressed. Furthermore, it has been demonstrated that the path integral associated with this action becomes form invariant under a change of variable if the measure of the path integral is suitably altered. It follows that the resulting effective action coincides exactly with the usual one-loop effective action [20]. In this paper, we have extended this result to theories in which the classical action possesses a gauge invariance. The gauge transformations in the tree level action (including all Lagrange multiplier and ghost fields) as well as the BRST transformation have been derived.

Having the quantum effects beyond one-loop order suppressed is of particular relevance when discussing quantum gravity. In quantum electrodynamics and YM theory, the divergences that arise due to quantum effects are proportional to terms appearing in the initial action, and hence they can be absorbed into these terms through the process of “renormalization” [4,5,6]. When the EH action is quantized using the Faddeev–Popov procedure, then at one-loop order divergences have a different character; they vanish when external fields satisfy the classical equation of motion (are “on mass-shell”) [36]. This no longer holds at two-loop order and beyond [7, 8]. However, if the EH action is supplemented by a Lagrange multiplier term, then this one-loop divergence can be absorbed by the Lagrange multiplier field [13]. No difficulty arises at higher loop order as no effects beyond one-loop order arise. The resulting effective action is free of divergences, and as it possesses a BRST invariance, it is also unitary. This approach can also be employed when matter fields are present in addition to the gravitational field [16].

We note that when a Lagrange multiplier field is used in a similar way with a YM theory, the divergences which now arise exclusively at one-loop order, lead to closed-form expressions for the renormalization group \( \beta \)-function associated with the running coupling. Since at one-loop order, this running coupling develops unphysical “Landau poles”, one cannot restrict YM theory to one-loop order using a Lagrange multiplier field in a physically consistent manner. This difficulty does not arise when one uses a Lagrange multiplier field in conjunction with EH action.

However, YM and gravity (based on the EH action) are both non-linear gauge theories, which have several other well-known similarities [37,38,39]. With this in mind, we have considered the YM theory in the extended Lagrange multiplier theory to illustrate the general procedure introduced in Sect. 3. With the usual covariant gauge fixing condition (5.7), which is used to fix all the gauge invariances present in the classical effective action in Eq. (5.4), we have obtained the generating functional of the YM theory in the framework of the extended Lagrange multiplier theory (5.11). We also found the BRST transformation in Eqs. (5.14) and (5.15) that leaves the resulting quantum effective action \( {\mathfrak {L}}_{\text {eff}} + {\mathfrak {L}}_\text {gf} + {\mathfrak {L}}_{ \text {gh} } \), which appears in Eq. (5.11), invariant.

The principle result of this paper is that when a Lagrange multiplier field is used to eliminate quantum effects beyond one-loop order in a gauge theory, and a suitable modification of the measure of the path-integral is made, then the path integral is form-invariant under a change of integration variables, that the effective action possesses a BRST invariance if the gauge algebra is closed, and that the quantum effects coincide with the usual one-loop effects arising from the classical action alone. This is of special significance when using the Faddeev–Popov procedure to quantize the EH action (when using the extended Lagrange multiplier theory) as it results in a theory that is unitary and renormalizable with the classical EH action intact. There is no need to postulate the existence of extra dimensions or degrees of freedom, or to invoke special non-perturbative effects (see, for example, [40]).

We note that the approach outlined in Sect. 3 can be straightforwardly applied to the case of the EH action. The EH action is invariant under diffeomorphisms in which the metric transforms as

(6.1)

which enables us to identify \( g^{\mu \nu } \) and in Eq. (6.1) with \(\phi _{i} \) and \( H_{ij} ( \phi _{i} )\) in Eq. (3.1).

However, in the case of the Einstein–Cartan action [41, 42], which describes spacetimes with torsion, applying this approach is non-trivial. This action, which is described in terms of the tetrad field \( e_{\mu }^{a} \) and the spin connection \( \omega _{\mu ab} \), is invariant under two distinct gauge transformations, namely, the diffeomorphisms and local Lorentz transformations [43, 44]. Hence, the quantization of this theory, in the extended Lagrange multiplier formalism, requires an extension of the Faddeev–Popov procedure outlined in Sect. 3 to accommodate two gauge symmetries. This extension would be analogous to what was done in Section III of [45] for the first order form of the Einstein–Cartan theory.Footnote 5