USING LATTICE QUANTUM CHROMODYNAMICS FOR NON-PERTURBATIVE RENORMALIZATION OF THE GLUON MOMENTUM FRACTION OF THE NUCLEON By Matthew David Zeilbeck A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Physics—Master of Science 2023 ABSTRACT Lattice quantum chromodynamics, often abbreviated as lattice QCD or LQCD, presents a non-perturbative method for calculating phenomenological quantities in quantum field theory (QFT). However, scale-dependent quantities such as the gluon momentum fraction must x ⟨ ⟩g be renormalized in order to compare lattice calculations to other theory and even other lattice calculations. The gluon momentum fraction is useful for understanding hadron structure and calculating the parton distribution function (PDF). I obtain the energy spectrum and ground state matrix elements of the gluon energy-momentum tensor (EMT) from the 2pt and 3pt correlators calculated from LQCD. Then, I calculate of the NPR renormalization constants using the regularization-independent momentum subtraction scheme (RI/MOM) for various lattice spacings with a pion mass Mπ ≈ 310MeV. This calculation required various steps in Chroma and Mathematica, and the workflow for this process is comprehensively detailed. I dedicate this thesis to anyone who has ever supported me or believed in me and my abilities and to anyone who receives any kind of benefit or positive outcome from reading it. iii ACKNOWLEDGEMENTS This work would not have been possible without the assistance and support of various people. First, I would like to thank my advisor, Dr. Huey-Wen Lin, for her support through my struggles and shortcomings. I am especially grateful for her selflessness as she allowed me to work on projects in her group despite knowing that I did not intend to complete my doctorate and that my research interests lie elsewhere. I also thank Dr. Remco Zegers, the physics graduate program director for most of my graduate education, for his support and guidance about what I could do to achieve my goals. Both of these people supported my desire to transfer in order to pursue research in quantum gravity and string theory, and even though this has yet to materialize, they nonetheless did everything in their power to set me up for success when others would have told me that this goal is too lofty and unrealistic. Next, I thank Zhouyou Fan for allowing me to take part in the NPR research that was ultimately his project. He (and Dr. Lin) gave me the opportunity to become part of this project, which allowed be to get published as a co-author. Of course, this work was also the basis for this thesis, so this thesis would not have been possible without his research acumen. Lastly, I thank my parents for their unending support and for always being there for me when I was down. Many times throughout my graduate education I needed to vent my frustrations to someone, and I could always rely on them to listen. They may not have always been able to provide the most directly useful advice, but their first and foremost concern was my happiness. Their countless sacrifices and support is why I was able to attain my educational success in the first place. I certainly wouldn’t be where I am today without them. iv TABLE OF CONTENTS CHAPTER 1 1.1 Particle Physics and Baryon Structure 1.2 The Path Integral Formulation of Quantum Mechanics INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.1 The Schrödinger Equation, Operator Formalism . . . . . . . . . . . . 1.2.2 Feynman’s Path Integral Approach . . . . . . . . . . . . . . . . . . . 1.2.3 Euclidean Time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.3.1 Quantum Mechanics in Euclidean Time 1.3 Quantum Field Theory Path Integral CHAPTER 2 QUANTUM CHROMODYNAMICS ON THE LATTICE . . . . . 2.1 Continuum QCD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.1 The QCD Lagrangian, Gauge Fields, and Color Components . . . . . 2.1.2 Local Gauge Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 QCD Actions on the Lattice . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Lattice Gauge Invariance and Link Variables . . . . . . . . . . . . . . 2.2.1.1 Fermionic Part . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1.2 Gauge Part . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.1 Calculating Lattice Quantities . . . . . . . . . . . . . . . . . . . . . . Spectral Decomposition of Lattice Quantities . . . . . . . . . . . . . . 2.3.2 2.3 Evaluating Lattice Quantities CHAPTER 3 ANALYZING LATTICE CORRELATOR DATA . . . . . . . . . . 3.1 The 2pt and 3pt Correlators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Fitting the 2pt Correlator . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Fitting the 3pt Correlator CHAPTER 4 NONPERTURBATIVE RENORMALIZATION (NPR) . . . . . . 4.1 Regularization-independent Momentum Subtraction (RI/MOM) Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Cluster-decomposition Error Reduction (CDER) . . . . . . . . . . . . . . 4.3 Renormalized Gluon Momentum Fraction . . . . . . . . . . . . . . . . . . 4.4 NPR Calculation Workflow . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.1 Chroma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Implementation of the NPR Calculation . . . . . . . . . . . 4.4.1.1 4.4.1.2 Outstanding Issues . . . . . . . . . . . . . . . . . . . . . . . 4.4.2 Mathematica . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.2.1 My Contribution to the Notebook . . . . . . . . . . . . . . . CHAPTER 5 CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v 1 1 2 2 3 4 5 7 8 8 8 9 10 10 10 12 13 13 15 17 17 18 21 26 27 28 30 32 33 38 41 41 45 46 47 CHAPTER 1 INTRODUCTION 1.1 Particle Physics and Baryon Structure All matter in the universe is comprised of particles called leptons and hadrons. Leptons have no further internal composite structure — they are fundamental — while hadrons are made of quarks and are thus composite particles. Hadrons are further classified into mesons and baryons. Mesons are lighter and are made of a quark-antiquark pair while baryons are heavier and made of three quarks. In order for quarks to bind together to make a hadron, they must interact with each other and form a bound state. The nature of these interactions is quite complicated. For example, a proton has only three valence quarks (up, up, down or uud) that are always detectable, but inside the proton, there is a complicated web of interactions involving gluons and virtual quark-antiquark pairs that affect the nature of the proton, e.g. see. Fig. 1.1. It used to be believed that the properties of hadrons like the proton were almost entirely determined by their valence quarks . However, we now know that this is not the case; it is as a result of both the properties of quarks and the nature of the interactions between them that hadrons acquire various physical properties such as spin, magnetic moment, observed mass1, effective size, etc. In addition, the gluons can carry a nontrivial fraction of the total momentum of a hadron, and calculating this fraction is part of understanding hadron structure. In addition to furthering our understanding of hadron structure, the gluon momentum fraction is also useful for understanding the gluon parton distribution function (PDF). The PDF g(x) is a probability density for finding a parton with momentum fraction x in a hadron. The gluon momentum fraction x ⟨ ⟩g is the first moment of the PDF, (cid:90) 1 x ⟨ ⟩g = 0 xg(x) dx . (1.1) The gluon momentum fraction serves another purpose besides the study of the PDF mo- 1It is primarily the energy of the interactions with gluons that we observe as the a hadron’s mass, not the sum of the bare quark masses. 1 Figure 1.1: Comparison of our understanding of the composition of the proton over time. The left image shows the previous model of the proton as consisting only of its valence quarks since gluon effects were not thought to be important. The right image shows the current understanding of the dynamics of the proton with valence quarks, virtual quarks, and gluons. The larger spheres that represent the valence quarks are made larger for emphasis and do not reflect the relative sizes of particles. The coil-like lines represent gluons that interact with the quarks to which they are attached. Image source [1]. ments. PDFs are of interest in other parts of high-energy physics, so it is worthwhile to calculate them in LQCD. One method for calculating the PDF is the “pseudo-PDF" ap- proach [2, 3, 4]. This method does not directly obtain g(x) but instead yields the quantity xg(x)/ x ⟨ ⟩g . Consequently, x ⟨ ⟩g must be known in order to eliminate it from this quantity. For these reasons, one goal of this work is to calculate the gluon momentum fraction of the nucleon. 1.2 The Path Integral Formulation of Quantum Mechanics 1.2.1 The Schrödinger Equation, Operator Formalism Lattice QCD developed from quantum field theory, which itself was developed from quan- tum mechanics. In non-relativistic quantum mechanics, one can solve physical problems by 2 obtaining the wavefunction of a system which obeys the Schrödinger equation (cid:32) N (cid:88) ℏ2 − i=1 2mi ∇ (cid:33) 2 i + V (xi, t) Ψ(xi, t) = iℏ ∂ ∂t Ψ(xi, t), (1.2) where mi is the mass of the ith particle, V (xi, t) is the potential, Ψ(xi, t) is the wavefunc- tion, and the sum is over N particles. xi is shorthand for the positions of each of the N particles xi { } . The wavefunction is overlap of the time-dependent state with the po- Ψ(t) ⟩ | sition eigenstate xi | ⟨ such that Ψ(xi, t) = xi ⟨ . Then we evaluate expectation values of Ψ(t) ⟩ | physical quantities with (t) = O ⟨ ⟩ ˆO(ˆx, ˆp) Ψ(t) Ψ(t) | | ⟨ ⟩ = (cid:90) Ψ∗(x, t)O(x, iℏ ∇ − )Ψ(x, t) d3x (1.3) where ˆO(ˆx, ˆp) is the operator associated with the observable O ⟨ , ˆp = ⟩ iℏ − ∇ in the position basis, and the hat is used to denote an operator. All operators are functions of the canonical variables (ˆx, ˆp) in operator form. 1.2.2 Feynman’s Path Integral Approach When Richard Feynman was in graduate school, he noticed a remark in Paul Dirac’s book that states that exp (cid:20) (cid:90) t2 (cid:21) i S “corresponds to" x2, t2| ⟨ x1, t1⟩ where t1 (cid:90) t2 t1 S = L(x, ˙x) ℏ , L(x, ˙x) = T ( ˙x) V (x). − In this expression, T is the kinetic energy, V is the potential energy, L is the classi- cal Lagrangian, and S is the classical action. By inserting completeness relations ˆI = (cid:90) x, t | ⟩⟨ x, t | dx and splitting up the time interval t2 − t1 into infinitesimal intervals, it can be shown that the probability amplitude of a particle located at initial position xi at time ti being found at some other final location xf at time tf is xf , tf ⟨ xi, ti | ⟩ = lim N →∞ (cid:19)N/2 (cid:90) (cid:18) m 2πi tf N − ti dx1 dx2 · · · dxN −1 eiS. (1.4) This expression is for 1 dimension, but it can easily be generalized to 3 dimensions to obtain xf , tf ⟨ xi, ti | ⟩ = lim N →∞ (cid:19)3N/2 (cid:90) (cid:18) m 2πi tf N − ti 3 d3x1 d3x2 · · · d3xN −1 eiS. (1.5) This expression becomes exact when N goes to infinity, so it may seem like the path integral is impossible and impractical to evaluate. However, this result generalizes very nicely to lattice formalism where spacetime is discrete and finite. Thus, path integrals of the form of Eq. 1.5 can be evaluated on a lattice where each integral over d3xi corresponds to a time slice ti of the Euclidean lattice. Since there is now a finite number of integrals, lattice formulation of Eq. 1.5 will have discretization errors but becomes more accurate as the lattice resolution becomes sharper and more continuum-like. 1.2.3 Euclidean Time Lattice theory is often formulated using an imaginary time t = iτ , τ > 0. This is − called Wick rotation. This means that we use a Euclidean metric gµν = δµν where δµν is the Kronecker delta, and the scalar product of two 4-vectors xµ = ( − i˜x0, x), yµ = ( − i˜y0, y) (upper vs. lower indices no longer matter with a Euclidean metric), with ˜x0, ˜y0 being the analog to τ , are now represented as xµgµνyν. When t is replaced with iτ and time derivates − are understood to be derivatives with respect to τ , the action S becomes (cid:90) (cid:90) S = (T + V ) dt = i HE dt = iSE, SE (cid:90) ≡ HE dt (1.6) where HE = 1 2m 2 (cid:12) (cid:12) (cid:12) (cid:12) dx dτ (cid:12) (cid:12) (cid:12) (cid:12) + V (x) is the Euclidean Hamiltonian. Now the path integral becomes xf , τf ⟨ xi, τi | ⟩ = lim N →∞ (cid:19)3N/2 (cid:90) (cid:18) m 2π τf N − τi d3x1 d3x2 · · · d3xN −1 e−SE . (1.7) In classical mechanics, allowed trajectories x(t) satisfy the least action principle, δS = 0. That is, they minimize the action S. In contrast, in quantum mechanics, all paths contribute to the path integral. However, eiS in Eq. 1.5 rapidly oscillates for paths with large S such that paths far from the classically allowed trajectory do not contribute much to the integral. That is, the rapidly oscillating term causes a lot of cancellation. In the Euclidean path integral Eq. 1.7, the oscillating eiS is now an exponentially decaying e−SE , so that rapid oscillations are now replaced with exponential suppression for non-classical paths in the integral where the amount of suppression increases with deviation from the classical trajectory. 4 Although I used the substitution t = iτ , in practice, t is used in LQCD equations, and − it should be understood to be the Euclidean time. 1.2.3.1 Quantum Mechanics in Euclidean Time Changing to Euclidean time affects other equations and tensorial objects in quantum mechanics in order to preserve various expressions involving 4-vectors. For instance, the Dirac equation’s gamma matrices in Minkowski spacetime γM µ satisfy (cid:8)γM µ , γM ν (cid:9) = 2ηµν (1.8) where ηµν = diag(1, 1, 1, − − 1) is the Minkowski metric. When we use Euclidean time, the − metric should be replaced with the Euclidean metric so that γµ, γν { } = 2δµν. (1.9) Thus, we must modify the γ matrices from their typical Minkowski definitions and define the Euclidean γµ. The anticommutation relation will hold if we define the Euclidean gamma matrices as − where i denotes a spatial index. Note that while 0 is used as the temporal index in Minkowski γ4 = γM 0 , γi = iγM i , i = 1, 2, 3 (1.10) spacetime, 4 is customarily used in Euclidean spacetime. Any γµ should be assumed to be defined with the Euclidean definition unless otherwise specified. This definition of the gamma matrices modifies the appearance of the Dirac equation as well. Normally the Dirac equation reads (cid:0)pµγM µ − m(cid:1) ψ(xµ) = 0 where pµ represents (p0, p) and γM µ 3 ). Summations over repeated indices (Einstein summation convention) are implied throughout. In Euclidean spacetime, represents (γM 1 , γM 0 , γM 2 , γM 4-vectors are denoted as vµ = vµ = (v1, v2, v3, v4) = (v1, v2, v3, iv0) 5 where the distinction between contravariant (upper) and covariant (lower) indices does not matter, and the components of each 4-vector are equal to their contravariant Minkowski counterparts (except that the time component is imaginary). Therefore, we can rewrite the operator in the Dirac equation as pµγM µ − m = − i (cid:0)p1γ1 + p2γ2 + p3γ3(cid:1) ip4γ4 − m = ipµγµ m. − − − Since the Euclidean metric makes no distinction between upper and lower indices, both indices in the implied summation can be written lower as they are in the above equation. However, Greek indices are still used to indicate that the sum is over 1–4 rather than just the spatial indices. However, Euclidean time treats time symmetrically with respect to space. As a consequence, the Dirac equation now reads ( ipµγµ − − m) ψ(xµ) = 0. (1.11) Similarly, the completeness relations for the Dirac spinors are also modified, now reading (cid:88) s=1,2 u(s)u(s) = N ( ipµγµ + m) , − (cid:88) s=1,2 v(s)v(s) = N ( − ipµγµ m) , − (1.12) where N is a normalization constant that depends on the normalization used for the bispinors. u(s) are the regular solutions to the Dirac equation ( ipµγµ m) u(s)(xµ) = 0, v(s) are the − ipµγµ + m) v(s)(xµ) = 0, and the adjoint bispinor u(s) is − negative energy solutions to ( u(s) = u(s)† γ4. − The above definition for the Euclidean γ matrices is not unique. For instance, one can define but this also requires vi = − Then, the Dirac equation is (ipµγµ − γ4 = γM 4 , γi = iγM i , vM,i such that the Euclidean 4-vector is vµ = ( vM , − m) ψ(xµ) = 0, and the completeness relations v0). − (Eq. 1.12) are similarly modified by eliminating the negative sign in ipµγµ. − 6 1.3 Quantum Field Theory Path Integral Quantum field theory (QFT) is the theory that comprehensively describes relatistic quan- tum mechanics and the fundamental interactions, excluding gravity. In classical and quantum mechanics, the primary degrees of freedom are the coordinates x, y, z. In a field theory, the degrees of freedom are the value of the field at each point in spacetime. Since spacetime is continuous, this results in infinite degrees of freedom. The Lagrangian formalism of classical mechanics can be generalized to QFT with the goal of finding a Lagrangian density that the Euler-Lagrange equations, ∂ L ∂ϕ − ∂µ ∂ L ∂(∂µϕ) = 0, such L (1.13) yield the differential equation that the field satisfies, e.g. the Klein-Gordon equation, Dirac equation, etc. The path integral can be generalized to field theory by integrating over every value of the field at every value of space in addition to every value of time: ϕ′(xµ) ϕ(yµ) ⟨ ⟩ | (cid:90) = (cid:90) (cid:90) [ϕ]eiS, D [ϕ] D ≡ (cid:90) lim N →∞ (cid:90) (cid:89) ϕ1 · · · D D ϕN −1, dϕ(x, ti) , (1.14) ϕi D (cid:90) ≡ x S = L (ϕ, ∂µϕ) d4x . 7 CHAPTER 2 QUANTUM CHROMODYNAMICS ON THE LATTICE 2.1 Continuum QCD 2.1.1 The QCD Lagrangian, Gauge Fields, and Color Components The theory of the strong interaction is quantum chromodynamics (QCD). It is mediated by gluon fields which are described by the potential Aµ(x)cd, which is a 4-vector (with Lorentz index µ) of 3 × 3 hermitian, traceless matrices (color indices c and d). These matrices are elements of the Lie algebra su(3). Together with the bispinor fields ψ(x)α c (here α is a Dirac index), the fermionic part of the QCD Lagrangian density can be written as (cid:2)ψ, ψ, A(cid:3) = F L Nf (cid:88) f =1 (f ) ψ (x)α c (cid:0)(γµ)αβDµ(x)cd + m(f )δαβδcd (cid:1) ψ(f )(x)β , d where Dµ(x)cd = δcd∂µ + igAµ(x)cd (2.1) (2.2) is the covariant derivative, g is the coupling strength between the fermionic and gauge fields, f is the flavor index, and Nf is the total number of flavors in consideration. The gauge field part of the Lagrangian density is G[A] = L 1 2 Tr[Fµν(x)Fµν(x)]. (2.3) The field strength tensor Fµν(x) is defined as Fµν(x) i[Dµ(x), Dν(x)] = ∂µAν(x) ≡ − − ∂νAµ(x) + i[Aµ(x), Aν(x)] (2.4) where the color indices have been suppressed. Since Aµ(x) su(3), we can decompose Aµ(x) ∈ into a superposition of the generators of SU(3) (which are the elements of the algebra su(3)): Aµ(x) = 8 (cid:88) i=1 A(i) µ (x)Ti, Ti = 1 2 λi, (2.5) where λi are the Gell-Mann matrices. This first summation is explicitly written despite the repeated index to emphasize that this sum runs from 1 to 8 (over color components). The 8 4-vectors A(i) µ (x) are the color components of the gauge field. Thus, the field strength tensor can be written as Fµν(x) = 8 (cid:88) i=1 F (i) µν (x)Ti µν (x) = ∂µA(i) F (i) ν (x) ∂νA(i) µ (x) − − fijkA(j) µ (x)A(k) ν (x), (2.6) where the structure constants fijk are defined by the relation [Ti, Tj] = ifijkTk. Thus, the gauge field part of the Lagrangian density can be written as G[A] = L 1 4 µν (x)F (i) F (i) µν (x), where the trace of products of Ti is evaluated with the relation Tr[TiTj] = QCD action is thus S (cid:2)ψ, ψ, A(cid:3) = (cid:90) d4x (cid:0) L F [ψ, ψ, A] + G[A](cid:1) . L (2.7) δij. The full 1 2 (2.8) 2.1.2 Local Gauge Invariance The Lagrangian is defined in the form of Eqs. 2.1 and 2.3 so that it is invariant under local SU(3) gauge transformations. These transformations modify the fields as ψ′(x) = Ω(x)ψ(x), ψ ′ (x) = ψ(x)Ω(x)†, A′ µ(x) = Ω(x)Aµ(x)Ω(x)† + i (∂µΩ(x)) Ω(x)†, (2.9) for Ω(x) ∈ SU(3). The transformation properties of Aµ(x) are defined like this so that the covariant derivative transforms as D′ µ(x) = Ω(x)Dµ(x)Ω(x)†. Because Fµν(x) is the commutator of covariant derivatives, it transforms as µν(x) = Ω(x)Fµν(x)Ω(x)†. F ′ This ensures that S (cid:104) ψ′, ψ ′ , A′(cid:105) = S (cid:2)ψ, ψ, A(cid:3). That is, the action is gauge-invariant. (2.10) (2.11) 9 2.2 QCD Actions on the Lattice The following subsection is largely based on [5], and the notation used there largely matches the notation used here. 2.2.1 Lattice Gauge Invariance and Link Variables 2.2.1.1 Fermionic Part The Lagrangian density and action introduced in the previous section apply to the con- tinuum, but the expressions in Eqs. 2.1, 2.3, and 2.8 are not suitable for the lattice in their current form. This is because of the derivative term which must be discretized on the lat- tice. Let x = an where a is the lattice spacing and n = (n, n4) is a vector of integers which denotes which lattice site is referenced, e.g. ni 1 } NT lattice. From hereon, ψ(n) is understood to mean the value of the field ψ at 0, 1, . . . , NT 0, 1, . . . N 1 } ∈ { − − , n4 ∈ { on an N 3 × position an. Thus, the gradient on the lattice is given by ∂µψ(x) = ψ(n + ˆµ) ψ(n ˆµ) − − 2a (a), + O (2.12) where ˆµ is a vector of length a in the µ-direction. Then the discretized version of the fermionic part of the action for a free fermion becomes (cid:2)ψ, ψ(cid:3) = SF (cid:90) (cid:2)ψ, ψ(cid:3) d4x F L a4 (cid:88) n∈Λ → (cid:18) ψ(n) γµ ψ(n + ˆµ) ψ(n ˆµ) − − 2a (cid:19) + mψ(n) , where Λ is the set of all spacetime points on the lattice, and the integral is replaced with a discrete sum. This presents a problem for gauge invariance because when ψ is replaced with ψ′, we are left with terms like ′ ψ (n)ψ′(n ± ˆµ) = ψ(n)Ω(n)†Ω(n ˆµ)ψ(n ± ± ˆµ) = ψ(n)ψ(n ˆµ) ± that spoil the gauge invariance that we had in the continuum theory. The solution is to introduce a new field Uµ(n) that transforms under a gauge transformation as U ′ µ(n) = Ω(n)Uµ(n)Ω(n + ˆµ)†. (2.13) 10 ̸ These are called link variables because they link site n with the neighboring site in the ˆµ- direction (site n + ˆµ) through this gauge transformation property. A negative direction may also be specified, and Uµ(n) is defined such that U−µ(n) Uµ(n ˆµ)†, − ≡ (2.14) as shown in Fig. 2.1. Figure 2.1: The link variable Uµ(n) (right) links lattice site n with site n + ˆµ and is directed in the ˆµ-direction. The link variable U−µ(n) (left) links n with n ˆµ and is directed in the − also displayed. Image source [5]. ˆµ-direction. The definition the hermitian conjuage of the link variables is − Now we note that ′ ψ (n)U ′ ±µ(n)ψ′(n ± ˆµ) = ψ(n)U±µ(n)ψ(n ˆµ) ± is a gauge-invariant construct and try as a candidate for the fermionic action (cid:2)ψ, ψ, U (cid:3) = a4 (cid:88) SF (cid:18) ψ(n) γµ Uµ(n)ψ(n + ˆµ) n∈Λ U−µ(n)ψ(n − 2a ˆµ) − (cid:19) + mψ(n) . (2.15) Although this is gauge invariant, we must still determine what U could be. It turns out that the path-ordered exponential (cid:18) (cid:90) (cid:19) P exp i Aµ dsµ , C along some contour C , has the gauge transformation property required by Eq. 2.13. Here the coupling g has been absorbed into the field Aµ. Since Uµ connects adjacent lattice sites that are a small distance apart, the integral can be approximated such that Uµ(n) = exp (iaAµ(n)) . (2.16) 11 To see that Eq. 2.15 with U instead of A is correct, we see that when Eq. 2.16 is expanded through order O (a), the action becomes (cid:2)ψ, ψ, A(cid:3) = a4 (cid:88) SF (cid:20) ψ(n) γµ (cid:18) ψ(n + ˆµ) n∈Λ ψ(n ˆµ) − − 2a + iAµ(n)ψ(n) + O (cid:19) (cid:21) (a) + mψ(n) , (2.17) which matches Eq. 2.1 if one attempts to naively discretize that expression to use on the lattice. Thus, it is clear that the correct gauge degrees of freedom on the lattice are Uµ(n) instead of Aµ(n). 2.2.1.2 Gauge Part Since the correct gauge field to use on the lattice is U , the expression for the gauge part of the action SG[A] must also be modified. The trace of a product of link variables that trace out a closed contour is gauge-invariant, so this is the type of object that will be used to construct the action. The simplest contour of this type is the plaquette as illustrated in Fig. 2.2. It is a closed contour made of 4 link variables that make a planar square. Figure 2.2: A counter-clockwise plaquette in the ˆµ- and ˆν-directions. The links that appear to go against the orientation of the loop (Uµ(n + ˆν) at the top and Uν(n) to the left) are actually written to go in the positive ˆµ- and ˆν-directions because the diagram is meant to show the relevant link variables written in standard form such that they point in the positive directions. Thus, the link variables that appear in the plaquette are Uµ(n + ˆν)† and Uν(n)†. Image from [5]. The plaquette specifies two directions µ and ν for the directions of the links. If we wish 12 to traverse in the µ-direction first, then we write the plaquette as Uµν(n) = Uµ(n)Uν(n+ ˆµ)U−µ(n+ ˆµ+ˆν)U−ν(n+ˆν) = Uµ(n)Uν(n+ ˆµ)Uµ(n+ˆν)†Uν(n)†. (2.18) The Wilson gauge action is constructed by summing over all plaquettes in a single orientation on the lattice: SG[U ] = 2 g2 (cid:88) (cid:88) n∈Λ µ<ν Re Tr [1 − Uµν(n)] . (2.19) The second summation requires µ < ν because when µ and ν are swapped, the plaquette reverses orientation, and if µ = ν, then the plaquette has nonsensical meaning (actually it does not contribute to the summation because Uµµ = 1). If Uµν(n) is Taylor expanded and rewritten with the Baker-Campbell-Hausdorff formula up to the first term with a commuta- tor, then to lowest non-vanishing order in Eq. 2.19, we find that SG[U ] = 2 g2 (cid:88) (cid:88) n∈Λ µ<ν Re Tr [1 − Uµν(n)] = a4 2g2 (cid:88) (cid:88) (cid:0)Tr (cid:2)Fµν(n)2(cid:3) + n∈Λ µ,ν (cid:0)a2(cid:1)(cid:1) . O This matches Eq. 2.3 to lowest order, except once again, A has been rescaled to absorb a factor of g such that 1/g2 appears in the expression. The sum over µ < ν has been replaced with an unconstrained sum over µ and ν with a factor of 1/2 to account for the fact that this double-counts terms (also the sum over µ, ν is explicitly written out instead of implied like in Eq. 2.3). Note that all of these traces are over color indices only. 2.3 Evaluating Lattice Quantities 2.3.1 Calculating Lattice Quantities Now we can write the entire action as S (cid:2)ψ, ψ, U (cid:3) = SF (cid:2)ψ, ψ, U (cid:3) + SG[U ], and we can use this to evaluate the path integral on the lattice. In Eq. 1.7, there are similarities to the types of integrals that one performs in statistical physics calculations where the Boltzmann factor is e−SE . In addition, the amplitude xf , T ⟨ xi, 0 ⟩ | can be written as xf ⟨ e−T ˆH | xi | ⟩ where e−T ˆH looks like the canonical density operator ˆρ = e−β ˆH where β T . → 13 That is, the size of the time evolution/the temporal extent of the lattice acts as an inverse temperature in the Euclidean LQCD theory. This suggests defining the partition function as ZT = Tr (cid:104) e−T ˆH(cid:105) = (cid:90) ψ0D ψ0D D U0 (cid:10)ψ0, ψ0, U0 (cid:12)e−T ˆH(cid:12) (cid:12) (cid:12)ψ0, ψ0, U0 (cid:11) . (2.20) However, the matrix element in the integrand can be evaluated with the path integral for- malism if we generalize Eq. 1.7 to fields. Then the partition function on the lattice becomes (cid:90) ZT = (cid:2)ψ, ψ(cid:3) D D [U ]e−S[ψ,ψ,U], (cid:2)ψ, ψ(cid:3) = (cid:89) (cid:89) n∈Λ f,α,c D dψ(f )(n)α c dψ (f ) (n)α c , [U ] = D (cid:89) 4 (cid:89) n∈Λ µ=1 dUµ(x) . (2.21) Overall multiplicative factors that appear outside of the integral in equations like Eq. 1.7 are omitted because they cancel out in relevant calculations. This motivates defining expectation values as O2(t)O1(0) ⟩ ≡ ⟨ 1 ZT (cid:104) (cid:105) ˆρ ˆO2(t) ˆO1(0) Tr = 1 ZT (cid:104) e− ˆH(T −t) ˆO2e− ˆHt ˆO1 (cid:105) , Tr (2.22) where in the second equality, the Heisenberg representation time-evolution of operators is used, i.e. where ˆO ≡ ˆO(t) = e ˆHt ˆO(0)e− ˆHt, (2.23) ˆO(0) and the exponential is real because of Euclidean time. This particular example is a 2pt correlation function, but it can be generalized to simpler or more complex observables as well. It will be assumed that the operators ˆO are functions of the relevant fields (and not any conjugate fields, e.g. conjugate momentum), i.e. ˆO = ˆO (cid:104) ˆψ, ˆψ, ˆU (cid:105) . The same methodology used to derive Eqs. 1.5, 1.7, and 2.21 shows that Eq. 2.22 can also be evaluated in the path integral approach and becomes O2(t)O1(0) ⟩ ⟨ = (cid:90) (cid:8) 1 ZT (cid:2)ψ, ψ(cid:3) D D [U ]O2 (cid:2)ψ(n, nt), ψ(n, nt), U (n, nt)(cid:3) (cid:2)ψ(n, 0), ψ(n, 0), U (n, 0)(cid:3) e−S[ψ,ψ,U](cid:111) O1 × (2.24) 14 Because these integrals are identical in form to the integrals used to evaluate quantities in statistical physics, the same numerical techniques can be used here to calculate lattice quantities as well. However, the generation of lattice configurations and the calculation of lattice quantities from actual lattice configurations are outside the scope of this thesis. 2.3.2 Spectral Decomposition of Lattice Quantities Once an observable has been calculated from Eq. 2.24, it is analyzed through a spectral decomposition of the observable and the partition function. Now it is better to evaluate the trace using a complete basis of energy eigenstates in addition to the identity operator as the sum over projections to energy eigenstates I = becomes (cid:88) n n | .1 Thus, the partition function n | ⟩⟨ (cid:88) ZT = e−EnT 1 for sufficiently large T, (2.25) ≈ where En+1 > En. Then Eq. 2.24 becomes n O2(t)O1(0) ⟨ ⟩ ≈ (cid:88) m,n ⟨ n O2| m | ⟩ ⟨ n O1| | m ⟩ e−Em(T −t)e−Ent (cid:88) n O2| 0 | n ⟨ ⟩ ⟨ ≈ O1| n 0 ⟩ | e−Ent, (2.26) where in the last approximation it is once again assumed that T is sufficiently large. Now that the groundwork for lattice theory has been laid down, the rest of this thesis is organized as follows. Chapter 3 explains how lattice correlator data is analyzed in order to extract physical observables. I begin with the definitions of the 2pt and 3pt lattice correlators. Then I explain how the correlator data is fitted and how useful quantities such as the matrix element associated with the gluon momentum fraction are extracted from the fit (Sec. 3.1). Next, in Chapter 4, I explain how to calculate the renormalization constants to perform non-perturbative renormalization (NPR) of the gluon energy-momentum tensor (EMT) in order to renormalize the gluon momentum fraction calculated in Chapter 3. I explain what quantities are needed for the calculation and how to calculate different parts 1This definition can be modified depending on normalization conventions, e.g. I = (cid:88) Gordon fields with Lorentz-invariant normalization, i.e. p′ p | ⟨ ⟩ = 2E( | )(2π)3δ3 (p p | − n p′). | n n ⟩⟨ 2En | for Klein- 15 of the renormalization constant using the RI/MOM scheme (Secs. 4.1–4.3), including the CDER technique to improve the signal (Sec. 4.2). I also explain the workflow for the different steps required to turn the lattices into results (Sec. 4.4). Finally, I present a conclusion and summary of my work in Chapter 5. 16 CHAPTER 3 ANALYZING LATTICE CORRELATOR DATA 3.1 The 2pt and 3pt Correlators The particular lattice quantities required for the calculation of the gluon momentum frac- tion of the nucleon x ⟨ ⟩g are the 2pt and 3pt correlation functions (the correlation functions are also simply referred to as correlators). For this particular project, I analyzed spin-1/2 particles that obey the Dirac equation, and the fermion fields were normalized such that (cid:88) I = ns MN EN | ns ns , where the sum is over spin s as well as energy eigenstates n. The 2pt | ⟩⟨ correlators are defined very similarly to Eq. 2.24 as C 2pt(p, t) = χB(p, t)χB(p, 0) ⟨ ⟩ = (cid:88) n p Zn( | ) | χB(p) n 0 | | ⟨ ⟩ ⟨ n χB(p) 0 ⟩ | | e−Ent, (3.1) where χB(p, 0) is an operator that creates baryon B with momentum p at time 0 and χB(p, t) annihilates it at time t. Usually the baryon is instead created at some position on the lattice (typically the origin) and annihilated after undergoing some displacement x at time t, then Fourier transformed to the momentum of interest. The 3pt correlators are defined as C 3pt (pf , pi, tsep, tins) = = χB(pf )V (q)χB(pi) ⟨ (cid:88) pi Zm( | pf )Zn( | | ) | ⟩ m,n χB(pf ) n 0 | | ⟨ ⟩ ⟨ n V (q) m | | ⟩ ⟨ χB(pi) m 0 | ⟩ | (3.2) e−En(tsep−tins)e−Emtins. × This corresponds to a particle created at time 0 with 3-momentum pi at time 0 that interacts with operator V (q = pf − pi) at time tins (insertion time) that imparts momentum transfer q that is then annihilated at time tsep (separation time) with 3-momentum pf . tsep is also known as the source-sink separation, and tsep is also sometimes called tsnk. The V operator corresponds to some property of the particle that we want to study. As a result, the 3pt correlator is only defined for 0 tins ≤ ≤ tsep since the particle cannot be inspected after it is annihilated. The Z terms in Eqs. 3.1 and 3.2 are kinematic factors that depend on the 17 normalization, e.g. Zn( p | p ) = M/En( | | ) for spin-1/2 particles. For spin-1/2 particles, the | creation and annihilation operators have Dirac indices (not shown in Eqs. 3.1 and 3.2), so the expressions in Eqs. 3.1 and 3.2 must be contracted by a projection operator to select out the relevant spin projections. The details of the projector depend on the quantity that we are trying to calculate by using the correlators. In order to calculate the gluon momentum fraction of the nucleon, the necessary operator V is the gluon energy-momentum tensor, and in particular, we need V 0 | ⟨ , the ground 0 ⟩ | state diagonal matrix element. This can be extracted from the 3pt correlator by fitting sampled 3pt correlator data and using fitted sampled 2pt correlator data to divide out the unwanted terms and kinematic factors in Eq. 3.2. 3.2 Fitting the 2pt Correlator The following sections involve analyzing data that comes from lattice simulations of QCD, i.e. the 2pt and 3pt correlators. For my analysis, these were provided to me to analyze; I was not involved in taking the data. In addition, any values of and variables related to time (t, tins, or tsep) that appear in the following sections should be interpreted to be integer values, i.e. they correspond to lattice grid indices. This means that the exponential terms in equations like Eqs. 3.1 and 3.2 contain the unitless energy aEn and the unitless integer time t as the lattice spacing a is attached to the energy rather than the time. However, the energy will still be written as En even though it is implied to be the unitless aEn. This is reflected by the lack of units in tables containing energies. The first step to analyzing correlator data is performing 2pt correlator fits. Given Ncfg samples of the 2pt correlator as a function of time from lattice calculations, the first step is jackknife resampling of the Ncfg configurations. Jackknife resampling is important because of the nonlinearity of the quantities calculated with the correlators. By using the raw data without resampling, the sample mean of such quantities has a bias that goes as O(1), i.e. increasing the number of samples does not improve the convergence to the true mean. With jackknife resampling, the bias goes as O (1/Ncfg), so the sample mean will converge to the 18 true mean with more data [6]. For the purposes of curve fitting, Eq. 3.1 is written as C 2pt(t) = (cid:88) n | An 2e−Ent = | 2e−E0t + A0| | 2e−E1t + . . . A1| | (3.3) such that the specific matrix elements are ignored for fitting purposes and normalization or kinematic factors are absorbed into the An. The fit parameters are then An | 2 and En. | Although it is not written for brevity, the 2pt correlator and its fit parameters are functions of momentum. Thus, different sets of data for different momenta are required. The number of states/terms included in the fit depends on how many can be resolved in the data, which can be discerned by the quality of attempted multistate fits. I present an example of 2pt correlator (and subsequently 3pt correlator analysis) using the data from the lattice ensembles that are used for non-perturbative renormalization (NPR) in Chapter 4, where the ensembles are explained in more detail (see Table 4.1). The analysis in this chapter uses data from the a12m310 ensemble, and it serves to double-check the calculations in [7]. For consistency with [7], I used a 2-state fit for this data. However, not all of the data is used for fitting. The lowest values of t have too much excited state contamination to be used, and the highest values of t end up having too much noise compared to signal. Therefore, a proper subset of the range [0, Lt] is used, and selecting it is non-trivial. In order to determine a good t range to use, I start by making a so-called effective mass 1 plot. The effective mass is given by ln [C 2pt(t)/C 2pt(t + 1)]. As t increases, Eq. 3.3 becomes dominated by just the ground state term, the effective mass asymptotically approaches E0, and the plot should approach a horizontal line for this range of t 0. This is visible in ≫ Fig. 3.1 where the effective mass is plotted with the reconstructed fit band obtained by using the fitted parameters in the fitted equation. This was performed for 2pt correlators projected to pz [0, 5] × ∈ 2π/(aL). 1Effective mass is a bit of a misnomer because it only corresponds to the mass if the 2pt correlator is projected to zero momentum; otherwise it corresponds to the ground state energy, which will have a kinetic contribution. However, I’ve never heard anyone call it a ground state energy plot. 19 ∈ [0, 5] in units of 2π/(aL). Fits were performed with t Figure 3.1: Effective mass plots and reconstructed fit bands for the a12m310 ensemble for [2, 11]. No upper bound on the pz fit bands was imposed (they exceed t = 11) to demonstrate that the effective mass approaches E0. The y-axis range was chosen so that the most important features of the data can be better visualized, but this caused some of the data points for the highest t to not be visible for pz = 4, 5. ∈ The reconstructed fit bands in Fig. 3.1 are created by fitting the 2pt correlator to Eq. 3.3 and then calculating the effective mass with ln [C 2pt(t)/C 2pt(t + dt)] /dt with the fitted pa- rameters. Note that dt replaces 1 (and there is division by dt) because the fitted band uses many interpolated values at t that have smaller spacing than that of the data points. Since I used a 2 state fit, the fitting function was the first two terms of Eq. 3.4 (up to n = 1): C 2pt(t) = 2e−E0t + A0| | 2e−E1t. A1| | (3.4) In order to determine a suitable fit range, I examined the effective mass plots (without the fit bands because they are calculated from the fit which had yet to be performed) and picked t bounds such the fit range includes some of the data points before they approach a horizontal line and most of the data after this occurs. However, not all of the available data in the asymptotic region is used because eventually the data starts to become noisier and also deviates from the asymptotic value of the other nearby data points. For example, for pz = 0, the effective mass at t ∈ [12, 14] is noticeably lower than the asymptotic value approached 20 02468101214t0.50.60.70.80.91.0Meffpz=002468101214t0.50.60.70.80.91.0Meffpz=102468101214t0.70.80.91.01.11.2Meffpz=202468101214t0.80.91.01.11.21.3Meffpz=302468101214t1.01.11.21.31.41.5Meffpz=402468101214t1.21.31.41.51.61.7Meffpz=5 at t ∈ range t [6, 11]. Even though these phenomena appear at different t for different pz, the fit [2, 11] was used for all pz for consistency. All of the fits have χ2/dof < 1 except ∈ for pz = 1, 2 where they are still close to 1. This suggests that the fits are good quality, and with the lowest χ2/dof being around 0.4, the data does not appear to be overfitted. The fitted bands decrease as t increases and become asymptotically horizontal as the excited state terms decay. Eventually, the fit band approaches a horizontal line as the 2pt correlator becomes dominated by only the ground state term. The results of the fits are summarized in Table 3.1. pz 0 1 2 3 4 5 2 A0| | 7.705(97)e 2.684(28)e 8.469(120)e 2.444(57)e 6.057(205)e 1.225(56)e 10 − 9 − 9 − 8 − 8 − 7 − E0 0.681(2) 0.730(1) 0.862(2) 1.044(3) 1.245(5) 1.445(7) 2 A1| | 2.088(30)e 6.491(59)e 1.839(15)e 4.542(41)e 9.529(95)e 1.728(16)e 9 − 9 − 8 − 8 − 8 − 7 − E1 1.455(14) 1.448(10) 1.531(12) 1.696(20) 1.894(32) 2.092(44) χ2/dof 0.808(254) 1.234(322) 1.100(360) 0.692(248) 0.397(356) 0.397(587) Table 3.1: Fit parameters from fitting the 2pt correlator to Eq. 3.4 and the corresponding χ2/dof for pz t ∈ units. 2π/(aL) for the a12m310 ensemble. Fits were performed using [2, 11]. The energy parameters are actually the unitless aEn, i.e. they are in lattice [0, 5] × ∈ 3.3 Fitting the 3pt Correlator After fitting the 2pt correlator, the 3pt correlator fit is performed using parameters from the 2pt correlator fit. Just as Eq. 3.1 was rewritten as Eq. 3.3, in practice, we rewrite Eq.3.2 as C 3pt (pf , pi, tsep, tins) = (cid:88) m,n | Am(pf ) An(pi) || V m | | ⟨ e−Em(tsep−tins)e−Entins, (3.5) n ⟩ | where normalization or kinematic factors are once again absorbed into the Am and An. It is important to fit the 2pt correlator first because in practice, we do not fit the energy parame- ters in Eq. 3.2 but rather insert the values obtained from the 2pt correlator fits. This makes finding the optimal parameters much easier as the time-dependence in the 3pt correlator is more complex than that of the 2pt correlator. Thus, it is preferable to avoid having fit 21 parameters that are directly multiplied by time. For this particular work, the operator V was the gluon energy-momentum tensor (EMT) which I will label as Og, consistent with Ref. [7], given by − In particular, we used Og,tt (the time-diagonal component, alternatively written as Og,44), for Og,µν = FµρFνρ gµνFρσFρσ. (3.6) 1 4 which Og,44| 0 0 ⟩ | ⟨ is proportional to , where x ⟨ ⟩g 0 ⟩ | is the nucleon ground state and is the x ⟨ ⟩g fraction of momentum carried by the gluons. The EMT was projected to zero momentum transfer so that pi = pf . Just as with the 2pt correlator, I fitted to the first excited state diagonal term, i.e. up to m = n = 1 in Eq. 3.7. Thus, the fit function was C 3pt (tsep, tins) = + 2 A0| | A1|| | 0 ⟩ | Og 0 ⟨ | Og A0| ⟨ 0 | e−E0tsep + Og A1| ⟨ 0 | A0|| 1 ⟩ | | e−E1(tsep−tins)e−E0tins + A1| | 1 ⟩ | e−E0(tsep−tins)e−E1tins 2 Og 1 | ⟨ 1 ⟩ | e−E1tsep, (3.7) where momentum labels have been omitted for brevity. This equation simplifies from the form of Eq. 3.2 because of the identical initial and final momenta/lack of momentum transfer. Because the energies from the 2pt correlator fit are inserted into this equation, the only fitted parameters are the products || momenta are the same, the states Am | An m | | ⟨ Og m | and n | n | . Additionally, because the initial and final ⟩ are at the same momentum, so when m = n, ⟩ they describe identical states. Thus, the m = n terms do not depend on tins, and the off- ⟩ diagonal matrix elements are related as Og 1 | ⟨ matrix elements of the EMT are real, we have = 0 ⟩ | Og 1 | ⟨ Og 0 | = ⟨ 0 ⟩ | ∗. Furthermore, because the 1 ⟩ | Og 0 | . This means that there 1 ⟩ | ⟨ are only 3 parameters to fit rather than 4 (8 if the real and imaginary parts of complex parameters are considered separate) despite the presence of 4 matrix elements in Eq. 3.7. After fitting, the are divided out so that we are left with only the matrix element | (and kinematic factors that must also be removed). || Am | An In order to choose the fit range, we observe that for 0 tins ≪ ≪ tsep, Eq. 3.5 is dominated by the ground state term (m = n = 0). This corresponds to the middle-most tins between 0 (creation) and tsep (annihilation). If suitable tins and tsep are used such that both the 3pt 22 and 2pt correlators are dominated by the ground state, then the special ratio = R C 3pt(pf , pi, tsep, tins) C 2pt(pi, tsep) (cid:115) C 2pt(pi, tins)C 2pt(pi, tsep)C 2pt(pf , tsep − C 2pt(pf , tins)C 2pt(pf , tsep)C 2pt(pi, tsep − tins) tins) (3.8) approaches the ground state matrix element. Since pi = pf = p, the square root term is 1 and Eq. 3.8 becomes a simple ratio of C 3pt(p, tsep, tins)/C 2pt(p, tsep). In the limit 0 tins ≪ ≪ tsep, a plot of Eq. 3.8 as a function of tins approaches a plateau at the ground state matrix element (for pi = pf , it can look more like a stationary inflection rather than a local maximum). To determine a suitable fit range, we plot this ratio as a function of tins for mul- tiple choices of tsep and examine the plateaus. Because agreement with this limit is strongest when tins is exactly in the middle of 0 and tsep, we symmetrically exclude endpoints such that tcut points are excluded from each side, so the fit range becomes tins ∈ each tsep. Much like the effective mass plot analysis, when excited states are desired in the [tcut, tsep − tcut] for fit, we make sure to include some tins outside of the plateau region since the plateau arises when only the ground state is dominant. When 3pt correlator data with multiple tsep are fitted as part of a single fit, this is called a simultaneous fit or “sim fit," and the fit function is essentially a function of multiple variables (tins and tsep). Since we included terms in the fit function up to the first excited (two states altogether), we refer to these 3pt correlator fits as “two-sim fits." Using the same lattice configurations that were used for the 2pt correlator analysis in the previous section, I fitted the 3pt correlator for the a12m310 ensemble using [7, 11] and tcut = 1. The fitted parameters and fit functions for both the 3pt and 2pt tsep ∈ correlators are then used to make reconstructed fit bands of Eq. 3.8, and these bands are plotted with the original data points and a horizontal band to denote the location of the ground state matrix element Og,tt 0 | ⟨ 0 ⟩ | in Fig. 3.2. Only the real part of the correlators and ratio produce a signal for the gluon EMT. Because different values of tsep result in different ranges of tins such that the largest tins is tsep, the data is plotted vs. tins − centered around the middle of the plot at tins − on the same plot, they would not align very well as the data for different tsep would end at tsep/2 (rather than tins) so that the data for each tsep is tsep/2 = 0. Otherwise, if all tsep were plotted 23 ̸ ∈ [0, 5] Figure 3.2: Plots of the 3pt/2pt correlator ratio defined in Eq. 3.8 vs. tins − the data for different tsep align and can be compared with each other) for 2π/(aL) for the a12m310 ensemble. Fits were performed with tsep ∈ pz tcut = 1. Data for different tsep are offset horizontally from each other so that data points that otherwise overlap can be distinguished. The gray band denotes pz = 4, 5 increases with tsep and with momentum. 2π/(aL), some of the data goes outside the bounds of the plot because the error tsep/2 (so that . For 0 ⟩ | [7, 11] and Og,tt 0 | ⟨ × × 24 −4−2024tins−tsep/2−0.2−0.10.00.10.20.30.40.50.6Re[R](pz=0)a12m310tsep=7tsep=8tsep=9tsep=10tsep=11−4−2024tins−tsep/2−0.2−0.10.00.10.20.30.40.50.6Re[R](pz=1)a12m310tsep=7tsep=8tsep=9tsep=10tsep=11−4−2024tins−tsep/2−0.20.00.20.40.6Re[R](pz=2)a12m310tsep=7tsep=8tsep=9tsep=10tsep=11−4−2024tins−tsep/2−0.4−0.20.00.20.40.60.8Re[R](pz=3)a12m310tsep=7tsep=8tsep=9tsep=10tsep=11−4−2024tins−tsep/2−0.50−0.250.000.250.500.751.00Re[R](pz=4)a12m310tsep=7tsep=8tsep=9tsep=10tsep=11−4−2024tins−tsep/2−1.5−1.0−0.50.00.51.01.5Re[R](pz=5)a12m310tsep=7tsep=8tsep=9tsep=10tsep=11 different tins. With this choice of the horizontal axis, the ratio data for different tsep should overlap at tins − qualitatively if the data behaves as expected. In order to convert , so it is easier to determine 0 ⟩ | tsep/2 = 0 and approach the value of Og,tt 0 | ⟨ into , a Og,tt 0 | 0 ⟩ | ⟨ x ⟨ bare g ⟩ kinematic factor must be divided out. In Fig. 3.2, the gray band denoting Og,tt 0 | 0 ⟩ | ⟨ and plateau region (near tins − tsep/2 = 0) of the reconstructed fit bands intersect as expected. For pz 2π/(aL), the error on the 3 × ≥ data points becomes noticeably larger such that the data for the largest tsep no longer form a clear plateau. The same occurs for increasing tsep, although the increase in the amount of error is greater as pz increases. This causes the data for the largest pz and tsep to no longer form a clear plateau. As a result, the vertical axis range for pz = 4, 5 2π/(aL) is reduced × so that the gray band and data with smaller error can be seen more clearly. This results in error bars for the largest tsep extending past the plots’ vertical range. All of the fits have χ2/dof < 1. The fitted parameters and matrix element are displayed in Table 3.2. 2 2 pz 0 1 2 3 4 5 A0| M00 | 3.488(460)e 10 − 1.320(129)e 9 − 4.146(494)e 9 − 1.369(218)e 8 − 3.767(917)e 8 − 9.538(3569)e 8 − A1|M01 A0|| | 5.950(933)e 10 − 2.127(246)e 9 − 6.639(781)e 9 − 2.014(275)e 8 − 4.928(963)e 8 − 1.203(307)e 7 − Table 3.2: Fit parameters for the 3pt correlator two-sim fits to Eq. 3.7 and the corresponding χ2/dof for pz A1| M11 | 1.158(741)e 8 − 2.629(1409)e 8 − 1.255(395)e 7 − 3.196(1627)e 7 − 5.387(7204)e 7 − 3.281(28207)e 7 − Og,tt 0 0 ⟨ ⟩ | | 0.453(59) 0.492(48) 0.490(58) 0.560(89) 0.622(151) 0.778(289) 2π/(aL) for the a12m310 ensemble. − − − − − − χ2/dof 0.541(507) 0.307(386) 0.454(401) 0.259(252) 0.354(360) 0.221(194) ij = i ⟨ Og,tt | j | [0, 5] . Fits were performed using tsep ∈ × ∈ M ⟩ pz), the error on some of the parameters is on the same order as the parameters themselves. For pz = 5, the error on 2 A1| | than the parameter. must be divided by a kinetic factor to convert it to M11 is actually an order of magnitude larger bare x g ⟩ ⟨ Og,tt 0 | 0 ⟩ | ⟨ . [7, 11]. For the noisiest data (largest 25 CHAPTER 4 NONPERTURBATIVE RENORMALIZATION (NPR) Some quantities calculated via LQCD must be renormalized in order to be compared to the same quantities calculated by other means, including other lattice calculations that use different definitions of quantities. For example, the gluon momentum fraction of the nucleon is independent of the definition of the gauge energy-momentum tensor (EMT) (e.g. clover vs. plaquette definitions) after nonperturbative renormalization (NPR) is applied to the EMT [8]. This chapter references a paper of which I am a co-author [7], and the use of first person singulars like “I" indicates that I am referring to the work that I specifically completed whereas plurals like “we" refer to the work in that paper completed mostly by the primary author (though my contributions involved double-checking some of the work). We used lattice configurations of Nf = 2 + 1 + 1 highly-improved staggered quarks (HISQ) [9] generated by the MILC collaboration [10]. The a12m310 ensemble analyzed in Chapter 3 comes from these configurations, and their parameters are summarized in Table 4.1. ensemble a (fm) L3 Lt × (MeV) M val π (MeV) M val ηs Ncfg a09m310 0.0888(8) 323 96 × 313.1(13) 698.0(7) 1009 a12m310 0.1207(11) 243 64 × 309.0(11) 684.1(6) 1013 a15m310 0.1510(20) 163 48 × 319.1(31) 687.3(13) 900 Table 4.1: Lattice spacing a in fm, lattice size L3 × and M val in MeV, and number of configurations Ncfg for the lattice ensembles analyzed in ηs Chapters 3 and 4. Ncfg applies only to the nucleon correlators analyzed in Chapter 3 and not to the gluon functions analyzed in Chapter 4. Data taken from Ref. [7]. Lt, valence pion and ηs masses M val π Before any calculations are performed, the lattice gauge links undergo 5 steps of hyper- cubic smearing (5-HYP) [11] because this is the smearing that was applied to the hadronic matrix element to reduce statistical uncertainties [12]. The gluon fields are then gauge-fixed 26 in the Landau gauge by imposing ∂µAµ(x + ˆµ/2) = 0 with (cid:21) (cid:18) (cid:19) (cid:20)Uµ(x) Uµ(x)† Aµ x + ˆµ = 1 2 − 2ig0a , traceless (4.1) where the term in brackets has been made traceless (over color). The gauge-fixed gluon fields are eventually used to calculate Aµ(p) with the Fourier transform (cid:19) (cid:18) Aµ(p) = a4 (cid:88) eip·(x+ 1 2 ˆµ)Aµ x + ˆµ . 1 2 x∈Λ (4.2) 4.1 Regularization-independent Momentum Subtraction (RI/MOM) Scheme In my particular work, I calculated the necessary renormalization factors for converting the bare gluon momentum fraction of the nucleon, denoted as bare g x ⟨ ⟩ (where x is the momen- tum fraction), calculated from LQCD to the the modified minimal substraction scheme MS (MS-bar) using the regularization-independent momentum-subtraction (RI/MOM) scheme [8]. The gluon momentum fraction is equal to the overlap with the ground state of the temporal component of the energy-momentum tensor (EMT), Og,tt, after kinematic factors that arise from the normalization of the fields and the choice of projection are divided out. That is, x ⟨ ⟩g ∝ ⟨ Og,tt 0 | , where 0 ⟩ | Og,µν = FµρFνρ 1 4 − gµνFρσFρσ, where Fµν is the gluon field strength tensor and gµν is the metric. Thus, renormalizing is equivalent to renormalizing the operator Og. (4.3) x ⟨ ⟩g The renormalization procedure outlined here is derived in Refs. [8, 13] and also utilized in our work [7]. The renormalized gluon momentum fraction in the MS scheme can be written as x ⟨ MS g = Z MS Og ⟩ (cid:0)µ2, µ2 R (cid:1) g = RMS (cid:0)µ2, µ2 bare ⟩ R x ⟨ (cid:1) Z RI Og (cid:1) (cid:0)µ2 R x ⟨ bare g ⟩ , (4.4) where the Z’s are renormalization constants, µ is the matching scale, µR is the renormaliza- tion scale, and RMS(µ, µR) is the one-loop perturbative matching ratio given by (cid:19) (cid:19) (cid:21) g2Nf 16π2 (cid:20) 2 3 ln (cid:18) µ2 µ2 R − + 10 9 g2Nc 16π2 − (cid:18) 4 2ξ + 3 − ξ2 4 RMS (cid:0)µ2, µ2 (cid:1) = 1 R . (4.5) 27 In this equation, derived in Ref. [14], Nf is the number of flavors, Nc is the number of colors, g2 = 4πα(µ) is the coupling strength (with α(µ) as the coupling constant), and ξ is the parameter from Riemann’s zeta function. In this work, Nf = 4, Nc = 3, ξ = 0 in the Landau gauge, and µ = 2 GeV. The RI/MOM renormalization constant Z RI g (µ2 R) can be obtained from the normalization condition Zg (cid:0)p2(cid:1) Z RI Og (cid:0)p2(cid:1) Λbare Og (cid:0)p2(cid:1) (cid:16) Λtree Og (cid:0)p2(cid:1)(cid:17)−1(cid:12) (cid:12) (cid:12) (cid:12)p2=µ2 R = 1, (4.6) where Zg (p2) is the gluon-field renormalization constant, and ΛOg (p2)’s are amputated Green’s functions for the operator Og for gluons in the Landau gauge. In Refs. [8, 13], from this condition, it is shown that (cid:16) (cid:17)−1 (cid:0)µ2 R (cid:1) = Z RI Og p2 (Og,µµ ⟨ 2(p2 µ − Og,νν) Tr[Aτ (p)Aτ ( − − p2 p)] Tr[Aτ (p)Aτ ( ν) ⟩ ⟨ − p)] ⟩ (cid:12) (cid:12) (cid:12) (cid:12)p2=µ2 R,τ ̸=µ̸=ν,pτ =0 , (4.7) where there is no summation over repeated indices. The quantity Tr[Aµ(p)Aν( − gluon propagator or gluon 2pt function and is also written as Dg,µν(p) = Tr[Aµ(p)Aν( − p)], such that the quantity used in Eq. 4.7 is Dg,τ τ . 4.2 Cluster-decomposition Error Reduction (CDER) p)] is the (4.8) Naïve attempts to calculate Z MS Og proved difficult as a signal-to-noise ratio under 100% was obtained [8, 7]. In order to improve the signal, a technique was developed in Ref. [8] called cluster-decomposition error reduction (CDER). The motivation for this method is that the correlators decay exponentially in the distance between operator insertions (e.g. gluon fields and the EMT), so there is no point in integrating beyond the correlation length as it will only pick up noise. This is implemented by imposing cutoffs in the spatial integrals used 28 to calculate the correlators. In this specific work, they become [7] Dg,τ τ = Tr[Aτ (p)Aτ ( ⟨ C CDER 3 (p) = Og,µν Tr[Aτ (p)Aτ ( ⟨ p)] − ⟩ ≈ p)] − ⟩ ≈ (cid:28)(cid:90) |r′| p2fitmin defined in the “Parameters" section. This section also makes a plot of the data points and the reconstructed fit band on separate plots. To combine them, use the Show function. This can be done by copying each image and pasting them as the arguments of Show or by saving each plot as a variable and using those variables as the arguments. Separate plots are also produced for the alternate definition of momentum in Eq. 4.15. 44 4.4.2.1 My Contribution to the Notebook The plots in Fig. 4.1 were produced using Mathematica notebooks that were written as described in this subsection. I did not originally write them, but I made significant modifica- tions and improvmenets to them. My work involved improving the readability and efficiency of the code, adding additional comments, double-checking calculations and formulas, and reducing the required runtime. I improved efficiency and runtime in various ways. I utilized more efficient data structures for the required information that we needed to store (namely maps, implemented as the Association data type in Mathematica). This was sensible com- pared to the list, which functions as a map but with keys restricted to sequential integers, because the p2 in integer units are not necessarily sequential and not necessarily in ascending order. Thus, the map also kept the data ordered in terms of ascending momentum. I also implemented the methods for parsing text files by using the fact that the needed line num- bers are the same for each type of object (FµνFρσ or Dg,µν(p)). This improved the runtime of the notebooks from the order of days to the order of hours. In addition, the modularity of the code allows for one to make additions or modifications, particularly in the last section where the plots are produced. 45 CHAPTER 5 CONCLUSION Lattice QCD provides a method for theoretical calculations of phenomenological parameters and observables. I outlined the basics of LQCD to demonstrate that it reduces to continuum QCD in the limit of zero spacing, and thus, it is a viable method for calculations of physical and phenomenological quantities. Then I defined the 2pt and 3pt correlators and fitted them in order to calculate the matrix element Og,tt 0 | ⟨ , which is used to calculate 0 ⟩ | x ⟨ bare g ⟩ . The effective mass and correlator ratio plots combined with the reasonable χ2/dof indicate that the fits are of good quality. Afterward, I calculated the necessary renormalization constants using the RI/MOM method to convert the results to the MS scheme. Using truncated lattices inspired by the CDER technique, I was able to obtain satisfactory data with a suitable signal- noise ratio such that the calculated quantities have reasonable error. Lastly, I documented the many steps involved in calculating the NPR constants and improved the efficiency of the Mathematica notebook that was previously written for this purpose. This allows for our results to be replicated and for our method to be applied to other future work, such as the gluon momentum fraction of mesons. 46 REFERENCES [1] Karen McNulty Walsh. Physicists zoom in on gluons’ contribution to proton spin, Feb 2016. [2] Zhouyou Fan, Rui Zhang, and Huey-Wen Lin. Nucleon gluon distribution function from 2 + 1 + 1-flavor lattice qcd. International Journal of Modern Physics A, 36(13):2150080, 2021. [3] Zhouyou Fan and Huey-Wen Lin. Gluon parton distribution of the pion from lattice qcd. Physics Letters B, 823:136778, 2021. [4] Tanjib Khan, Raza Sabbir Sufian, Joseph Karpie, Christopher J. Monahan, Colin Egerer, Bálint Joó, Wayne Morris, Kostas Orginos, Anatoly Radyushkin, David G. Richards, Eloy Romero, and Savvas Zafeiropoulos. Unpolarized gluon distribution in the nucleon from lattice quantum chromodynamics. Phys. Rev. D, 104:094516, Nov 2021. [5] Christof Gattringer and Christian B. Lang. Quantum Chromodynamics on the Lattice: An Introductory Presentation. Lecture Notes in Physics. Springer, 2010. [6] B.A. Berg. Markov Chain Monte Carlo Simulations and Their Statistical Analysis: With Web-based Fortran Code. World Scientific, 2004. [7] Zhouyou Fan, Huey-Wen Lin, and Matthew Zeilbeck. Nonperturbatively renormalized nucleon gluon momentum fraction in the continuum limit of Nf = 2 + 1 + 1 lattice qcd. Phys. Rev. D, 107:034505, Feb 2023. [8] Yi-Bo Yang, Ming Gong, Jian Liang, Huey-Wen Lin, Keh-Fei Liu, Dimitra Pefkou, and Phiala Shanahan. Nonperturbatively renormalized glue momentum fraction at the physical pion mass from lattice qcd. Phys. Rev. D, 98:074506, Oct 2018. [9] E. Follana, Q. Mason, C. Davies, K. Hornbostel, G. P. Lepage, J. Shigemitsu, H. Trot- tier, and K. Wong. Highly improved staggered quarks on the lattice with applications to charm physics. Phys. Rev. D, 75:054502, Mar 2007. [10] A. Bazavov, C. Bernard, J. Komijani, C. DeTar, L. Levkova, W. Freeman, Steven Gottlieb, Ran Zhou, U. M. Heller, J. E. Hetrick, J. Laiho, J. Osborn, R. L. Sugar, D. Toussaint, and R. S. Van de Water. Lattice qcd ensembles with four flavors of highly improved staggered quarks. Phys. Rev. D, 87:054505, Mar 2013. [11] Anna Hasenfratz and Francesco Knechtli. Flavor symmetry and the static potential with hypercubic blocking. Phys. Rev. D, 64:034504, Jul 2001. [12] Zhou-You Fan, Yi-Bo Yang, Adam Anthony, Huey-Wen Lin, and Keh-Fei Liu. Gluon quasi-parton-distribution functions from lattice qcd. Phys. Rev. Lett., 121:242001, Dec 2018. [13] P. E. Shanahan and W. Detmold. Gluon gravitational form factors of the nucleon and the pion from lattice qcd. Phys. Rev. D, 99:014511, Jan 2019. 47 [14] Yi-Bo Yang, Michael Glatzmaier, Keh-Fei Liu, and Yong Zhao. The 1-loop correction of the qcd energy momentum tensor with the overlap fermion and hyp smeared iwasaki gluon, 2016. [15] Santanu Mondal, Rajan Gupta, Sungwoo Park, Boram Yoon, Tanmoy Bhattacharya, and Huey-Wen Lin. Moments of nucleon isovector structure functions in 2 + 1 + 1-flavor qcd. Phys. Rev. D, 102:054512, Sep 2020. 48