Skip to main content

Lattice Quantum Chromodynamics

  • Chapter
  • First Online:

Part of the book series: Lecture Notes in Physics ((LNP,volume 936))

Abstract

Concepts and applications of lattice quantum chromodynamics (LQCD) are introduced. After discussing how to define quarks and gluons on the Euclidean hypercubic lattice, the strong coupling expansion and the weak coupling expansions are reviewed to see the vital role played by the quantum fluctuations in QCD. Fundamental techniques for numerical LQCD simulations such as the Markov Chain Monte Carlo method and the Hybrid Monte Carlo method are discussed in some details. As examples of the high precision LQCD simulations, numerical results of the heavy quark potential and the hadron masses are shown. Recent LQCD results on the baryon-baryon interactions are briefly discussed.

This is a preview of subscription content, log in via an institution.

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   99.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   129.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

Notes

  1. 1.

    Rigorous definitions are as follows. (1) The Markov chain is said to be irreducible if one can find a finite positive integer n( < ) such that (T n) ϕ ϕ > 0 for all ϕ and ϕ′. (2) The period, d(ϕ), is defined by the greatest common divisor of the set of positive integers n( ≥ 1) such that (T n) ϕ ϕ  > 0 is satisfied. If d(ϕ) = 1 for all ϕ, the Markov chain is said to be aperiodic [10].

References

  1. K.G. Wilson, Phys. Rev. D 10, 2445 (1974)

    Article  ADS  Google Scholar 

  2. M. Creutz, Phys. Rev. D 21, 2308 (1980)

    Article  ADS  Google Scholar 

  3. N. Brambilla et al., Eur. Phys. J. C 74, 2981 (2014)

    Article  Google Scholar 

  4. K.G. Wilson, Nucl. Phys. Proc. Suppl. 140, 3 (2005)

    Article  ADS  Google Scholar 

  5. M. Creutz, Quarks, Gluons and Lattices (Cambridge University Press, Cambridge, 1985)

    Google Scholar 

  6. H.J. Rothe, World Sci. Lect. Notes Phys. 82, 1 (2012)

    Article  Google Scholar 

  7. C. Hoelbling, Acta Phys. Polon. B 45, 2143 (2014)

    Article  ADS  Google Scholar 

  8. A. Ukawa, J. Stat. Phys. 160, 1081 (2015)

    Article  ADS  Google Scholar 

  9. J.W. Negele, H. Orland, Quantum Many Particle Systems (Addison-Wesley, Redwood, 1988)

    MATH  Google Scholar 

  10. O. Häggström, Finite Markov Chains and Algorithmic Applications (Cambridge University Press, Cambridge, 2002)

    Book  MATH  Google Scholar 

  11. H. Suwa, S. Todo, Phys. Rev. Lett. 105, 120603 (2010)

    Article  ADS  Google Scholar 

  12. S. Duane, A.D. Kennedy, B.J. Pendleton, D. Roweth, Phys. Lett. B 195, 216 (1987)

    Article  ADS  Google Scholar 

  13. N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.H. Teller, E. Teller, J. Chem. Phys. 21, 1087 (1953)

    Article  ADS  Google Scholar 

  14. S. Schaefer, PoS LATTICE 2012, 001 (2012)

    Google Scholar 

  15. G.S. Bali, Phys. Rep. 343, 1 (2001)

    Article  ADS  Google Scholar 

  16. S. Durr et al., Science 322, 1224 (2008)

    Article  ADS  Google Scholar 

  17. S. Borsanyi et al., Science 347, 1452 (2015)

    Article  ADS  Google Scholar 

  18. The Review of Particle Physics (2015). http://pdg.lbl.gov/

  19. M. Asakawa, T. Hatsuda, Y. Nakahara, Prog. Part. Nucl. Phys. 46, 459 (2001)

    Article  ADS  Google Scholar 

  20. Z. Fodor, C. Hoelbling, Rev. Mod. Phys. 84, 449 (2012)

    Article  ADS  Google Scholar 

  21. R. Machleidt, Int. J. Mod. Phys. E26, 1740018 (2017)

    Article  ADS  Google Scholar 

  22. S. Aoki et al., Eur. Phys. J. C 74, 2890 (2014)

    Article  ADS  Google Scholar 

  23. M. Lüscher, Nucl. Phys. B 354, 531 (1991)

    Article  ADS  Google Scholar 

  24. N. Ishii, S. Aoki, T. Hatsuda, Phys. Rev. Lett. 99, 022001 (2007)

    Article  ADS  Google Scholar 

  25. N. Ishii et al., HAL QCD collaboration. Phys. Lett. B 712, 437 (2012)

    Article  ADS  Google Scholar 

  26. T. Iritani, HAL QCD collaboration. JHEP 1610, 101 (2016)

    Article  ADS  Google Scholar 

  27. S. Okubo, R.E. Marshak, Ann. Phys. 4, 166 (1958)

    Article  ADS  Google Scholar 

  28. M. Oka, K. Shimizu, K. Yazaki, Prog. Theor. Phys. Suppl. 137, 1 (2000)

    Article  ADS  Google Scholar 

  29. T. Inoue et al., HAL QCD collaboration. Nucl. Phys. A 881, 28 (2012)

    Article  ADS  Google Scholar 

  30. T. Inoue et al., HAL QCD collaboration. Phys. Rev. Lett. 111, 112503 (2013)

    Article  ADS  Google Scholar 

  31. A. Akmal, V.R. Pandharipande, D.G. Ravenhall, Phys. Rev. C 58, 1804 (1998)

    Article  ADS  Google Scholar 

  32. T. Doi et al., HAL QCD collaboration. PoS LATTICE2015, 086 (2016)

    Google Scholar 

Download references

Acknowledgements

The author thanks Takumi Doi and Atsushi Nakamura for useful comments and information on various aspects of LQCD simulations. He also thank the members of HAL QCD Collaboration for fruitful discussions on the hadron-hadron interactions on the lattice. This work was supported in part by MEXT SPIRE and JICFuS and also by RIKEN iTHES Project.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tetsuo Hatsuda .

Editor information

Editors and Affiliations

Appendix

Appendix

3.1.1 Four Vectors and Dirac Matrices

In the (3+1)-dimensional Minkowski spacetime, coordinates, derivatives and four vectors with μ = 0, 1, 2, 3 are

$$ \displaystyle\begin{array}{rcl} x^{\mu } = (t,\mathbf{x}),\ \ \partial ^{\mu } = (\partial _{t},-\nabla ),\ \ A^{\mu } = (A^{0},\mathbf{A}).& &{}\end{array}$$
(3.97)

In the 4-dimensional Euclidean space, we define the corresponding vectors for μ = 4, 1, 2, 3 as

$$\displaystyle{ (x_{\mu })^{\mathrm{E}} = (\tau = it,\mathbf{x}),\ \ (\partial _{\mu })^{\mathrm{E}} = (\partial _{\tau } = -i\partial _{ t},\nabla ),\ \ (A_{\mu })^{\mathrm{E}} = (A_{ 4} = iA^{0},\mathbf{A}). }$$
(3.98)

In the (3+1)-dimensional Minkowski spacetime with the metric g μ ν = diag(1, −1, −1, −1), the Dirac matrices satisfy the following relations for μ = 0, 1, 2, 3,

$$\displaystyle{ \{\gamma ^{\mu },\gamma ^{\nu }\} = 2g^{\mu \nu },\ \ \left (\gamma ^{\mu }\right )^{\dag } =\gamma ^{0}\gamma ^{\mu }\gamma ^{0},\ \ \gamma ^{5} = i\gamma ^{0}\gamma ^{1}\gamma ^{2}\gamma ^{3} =\gamma _{ 5} = \left (\gamma _{5}\right )^{\dag }. }$$
(3.99)

In the standard Dirac representation, we have

$$\displaystyle\begin{array}{rcl} \gamma ^{0} = \left (\begin{array}{cc} 1& 0 \\ 0& - 1\\ \end{array} \right ),\ \ \gamma ^{j} = \left (\begin{array}{cc} 0 & \sigma _{j} \\ -\sigma _{j}&0\\ \end{array} \right ),\ \ \gamma ^{5} = \left (\begin{array}{cc} 0&1 \\ 1&0\\ \end{array} \right ),& &{}\end{array}$$
(3.100)

where σ j are the Pauli matrices; \(\sigma _{1} = \left (\begin{array}{cc} 0&1\\ 1 &0\\ \end{array} \right ),\sigma _{2} = \left (\begin{array}{cc} 0& - i\\ i & 0\\ \end{array} \right ),\sigma _{3} = \left (\begin{array}{cc} 1& 0\\ 0 & -1\\ \end{array} \right ).\)

In the 4-dimensional Euclidean space with the metric δ μ ν  = diag(1, 1, 1, 1), we define the Euclidean Dirac matrices as

$$\displaystyle\begin{array}{rcl} \varGamma _{\mu } \equiv \left (\gamma _{4} =\gamma ^{0},-i\boldsymbol{\gamma }\right ),\ \ \varGamma _{ -\mu }\equiv -\varGamma _{\mu },\ \ \mathrm{and}\ \ \varGamma _{5} \equiv \gamma ^{5},& &{}\end{array}$$
(3.101)

which satisfy the relations,

$$\displaystyle\begin{array}{rcl} \{\varGamma _{\mu },\varGamma _{\nu }\} = 2\delta _{\mu \nu },\ \ \varGamma _{\mu }^{\dag } =\varGamma _{\mu },\ \ \ \ (\mathrm{for}\ \mu = 1,2,3,4,5)& &{}\end{array}$$
(3.102)

3.1.2 SU(N) Algebra

Let \(\mathcal{T}^{a}\) (a = 1, ⋯ , N 2 − 1) are the Hermitian generators of the SU(N) group. They satisfy the Lie algebra

$$\displaystyle\begin{array}{rcl} \left [\mathcal{T}^{a},\mathcal{T}^{b}\right ] = if_{ abc}\mathcal{T}^{c},& &{}\end{array}$$
(3.103)

where f abc is the structure constant being totally anti-symmetric in its indices. \((\mathcal{T}^{b})^{2}\) commutes with every generator \(\mathcal{T}^{a}\) and is called the quadratic Casimir operator.

For N = 2, f abc reduces to the anti-symmetric tensor ε ijk with ε 123 = 1. For N = 3, the non-vanishing components of f abc read \(f_{123} = 1,f_{147} = -f_{156} = f_{246} = f_{257} = f_{345} = -f_{367} = 1/2,f_{458} = f_{678} = \sqrt{3}/2\).

In the fundamental representation, \(\mathcal{T}^{a}\) is written by the N × N matrices t a as

$$\displaystyle\begin{array}{rcl} t^{a} = \frac{1} {2}\lambda _{a},& &{}\end{array}$$
(3.104)

where λ a for N = 2 reduce to the Pauli matrices σ i , while those for N = 3 reduce to the Gell-Mann matrices.

Some useful relations of t a for general N are

$$\displaystyle{ \mathrm{tr}(t^{a}t^{b})\,=\,\frac{1} {2}\delta _{ab},\ \ \ t_{ij}^{a}t_{ kl}^{b}\,=\,\frac{1} {2}(\delta _{il}\delta _{jk} - \frac{1} {N}\delta _{ij}\delta _{kl}),\ \ \ (t^{a}t^{a})_{ ij} = C_{\mathrm{F}}\delta _{ij},\ \ \ \mathrm{with}\ C_{\mathrm{F}}\,=\,\frac{N^{2} - 1} {2N}. }$$
(3.105)

In the adjoint representation, \(\mathcal{T}^{a}\) is written by (N 2 − 1) × (N 2 − 1) matrices T a as

$$\displaystyle\begin{array}{rcl} (T^{a})_{ bc} = -if_{abc},& &{}\end{array}$$
(3.106)

which satisfy the relations

$$\displaystyle\begin{array}{rcl} \mathrm{tr}(T^{a}T^{b})& =& N\delta _{ ab},\ \ \ (T^{a}T^{a})_{ bc} = C_{\mathrm{A}}\delta _{bc},\ \ \ \mathrm{with}\ C_{\mathrm{A}} = N.{}\end{array}$$
(3.107)

3.1.3 Gaussian and Grassmann Integrals

Basic Gaussian and Grassmann integrals are

$$\displaystyle\begin{array}{rcl} \int _{-\infty }^{+\infty }\frac{dx} {\sqrt{ 2\pi }}\ \mathrm{e}^{-ax^{2}/2 }& =& \frac{1} {\sqrt{a}},{}\end{array}$$
(3.108)
$$\displaystyle\begin{array}{rcl} \int \frac{dz^{{\ast}}dz} {2\pi i} \ \mathrm{e}^{-b\vert z\vert ^{2} }& =& \frac{1} {b},{}\end{array}$$
(3.109)
$$\displaystyle\begin{array}{rcl} \int d\bar{\xi }d\xi \ \mathrm{e}^{-c\bar{\xi }\xi }& =& c.{}\end{array}$$
(3.110)

Here x (z) is a real (complex) number, while \(\bar{\xi }\) and ξ are anti-commuting Grassmann numbers (\(\{\xi,\bar{\xi }\}= 0\), and \(\xi ^{2} =\bar{\xi } ^{2} = 0\)). a and b are assumed to be real and positive numbers, while c is an arbitrary complex number. Equation (3.109) can be shown by rewriting the integral in terms of the real and imaginary parts of z or in terms of the polar coordinates of z. Equation (3.110) can be shown by noting that \(\mathrm{e}^{-c\bar{\xi }\xi } = 1 - c\bar{\xi }\xi\) and ∫ d ξ = ∂ ξ (integral = derivative) for Grassmann variables.

Generalization of the above results to the case of multiple variables is straightforward. For x = (x 1, ⋯ , x n ), z = (z 1, ⋯ , z n ), ξ = (ξ 1, ⋯ , ξ n ), and \(\bar{\xi }= (\bar{\xi }_{1},\cdots \,,\bar{\xi }_{n})\) with \(\{\xi _{k},\xi _{l}\} =\{\bar{\xi } _{k},\bar{\xi }_{l}\} =\{\xi _{k},\bar{\xi }_{l}\} = 0\), we have

$$\displaystyle\begin{array}{rcl} \int \prod _{l=1}^{n}\frac{dx_{l}} {\sqrt{2\pi }}\ \mathrm{e}^{-\frac{1} {2} xAx}& =& \frac{1} {\sqrt{\mathrm{Det }\ A}},{}\end{array}$$
(3.111)
$$\displaystyle\begin{array}{rcl} \int \prod _{l=1}^{n}\frac{dz_{l}^{{\ast}}dz_{ l}} {2\pi i} \ \mathrm{e}^{-z^{{\ast}}Bz }& =& \frac{1} {\mathrm{Det}\ B},{}\end{array}$$
(3.112)
$$\displaystyle\begin{array}{rcl} \int \prod _{l=1}^{n}d\bar{\xi }_{ l}d\xi _{l}\ \mathrm{e}^{-\bar{\xi }C\xi }& =& \mathrm{Det}\ C.{}\end{array}$$
(3.113)

Here A is a non-singular and real-symmetric matrix whose eigenvalues a l satisfy a l  > 0 for all l. B is a non-singular complex matrix whose complex eigenvalues b l obtained by the biunitary transformation (UBV ) satisfy Re b l  > 0 for all l. C is an arbitrary complex matrix with no conditions. Note that B and C do not have to be Hermitian matrices. In field theories, the label “l” summarizes all possible indices including spin, flavor, color, spacetime points etc. and “Det” denotes the determinant for all these indices.

3.1.4 Method of Characteristics

We need to construct a general solution of the following partial differential equation,

$$\displaystyle\begin{array}{rcl} \left (\lambda \frac{\partial } {\partial \lambda } +\beta (g) \frac{\partial } {\partial g}\right )f(\lambda,g) = 0.& &{}\end{array}$$
(3.114)

For this purpose, we introduce the running coupling \(\bar{g}(\lambda )\) through \(\lambda d\bar{g}/d\lambda = -\beta (\bar{g})\) whose formal solution reads

$$\displaystyle\begin{array}{rcl} \lambda =\exp \left (-\int _{g}^{\bar{g}(\lambda )} \frac{dg'} {\beta (g')}\right ).& &{}\end{array}$$
(3.115)

Then the solution of Eq. (3.114) can be written as

$$\displaystyle\begin{array}{rcl} f(\lambda,g) = f(1,\bar{g}(\lambda )).& &{}\end{array}$$
(3.116)

This can be explicitly checked by applying the partial derivative on both sides,

$$\displaystyle\begin{array}{rcl} \lambda \partial _{\lambda }f(\lambda,g)& =& \lambda (\partial _{\lambda }\bar{g})(\partial _{\bar{g}}f) = -\beta (\bar{g})(\partial _{\bar{g}}f), \\ \beta \partial _{g}f(\lambda,g)& =& \beta (g)\left.(\partial \bar{g}/\partial g)\right \vert _{\lambda }(\partial _{\bar{g}}f) =\beta (\bar{g})(\partial _{\bar{g}}f).{}\end{array}$$
(3.117)

where we have used the relation \(\partial \bar{g}/\partial g =\beta (\bar{g})/\beta (g)\) obtained from Eq. (3.115).

In general, the first-order partial differential equation (PDE) can be transformed to a set of ordinary differential equations (ODEs) and can be solved by the method of characteristics. As an illustration, let us consider the following PDE,

$$\displaystyle\begin{array}{rcl} a(t,x)\partial _{t}u(t,x) + b(t,x)\partial _{x}f(t,x) = c(t,x).& &{}\end{array}$$
(3.118)

This is equivalent to the coupled ODEs,

$$\displaystyle\begin{array}{rcl} \frac{d\bar{t}} {ds} = a(\bar{t},\bar{x}),\ \ \frac{d\bar{x}} {ds} = b(\bar{t},\bar{x}),\ \ \frac{df(\bar{t},\bar{x})} {ds} = a(\bar{t},\bar{x}),& &{}\end{array}$$
(3.119)

where s parametrizes the “flow” of the coordinates. \((\bar{t}(s),\bar{x}(s))\). This is called the characteristic curve as shown in Fig. 3.15.

Fig. 3.15
figure 15

Schematic illustration of the characteristic curve

The function f can be obtained by integrating the last equation of Eq. (3.119) on the characteristic curve from the initial point (t i, x i) to the final point (t F, x F) ≡ (t, x),

$$\displaystyle\begin{array}{rcl} f(t,x) = f(t_{\mathrm{i}},x_{\mathrm{i}}) + h(t,x,t_{\mathrm{i}},x_{\mathrm{i}})& &{}\end{array}$$
(3.120)

where h stands for an integration of the known function \(c(\bar{t},\bar{x})\) on the characteristic curve. Equation (3.120) implies that the desired function at (t, x) is obtained essential by a “pullback” of the point to (t i, x i) along the characteristic curve. It is a straightforward exercise to generalize the above derivation to the system with more coordinates, (t, x).

3.1.5 Leapfrog Integrator in Molecular Dynamics

Let us start with a Tayler expansion of the field ϕ:

$$\displaystyle\begin{array}{rcl} \phi (s+\varepsilon )& =& \phi (s) +\varepsilon \dot{\phi } (s) + \frac{\varepsilon ^{2}} {2}\ddot{\phi }(s) + O(\varepsilon ^{3}), \\ & =& \phi (s) +\varepsilon \pi (s) + \frac{\varepsilon ^{2}} {2}\dot{\pi }(s) + O(\varepsilon ^{3}), \\ & =& \phi (s) +\varepsilon \pi (s +\varepsilon /2) + O(\varepsilon ^{3}),{}\end{array}$$
(3.121)

where we have used the equation of motion, \(\dot{\phi }(s) \equiv d\phi (s)/ds =\pi (s)\). To evaluate π(s +ɛ∕2), we take the midpoint prescription which does not have O(ɛ 2) error,

$$\displaystyle\begin{array}{rcl} \pi (s +\varepsilon /2)& =& \pi (s -\varepsilon /2) +\varepsilon \dot{\pi } (s) + O(\varepsilon ^{3}) \\ & =& \pi (s -\varepsilon /2) -\varepsilon \frac{\delta S(\phi )} {\delta \phi (s)} + O(\varepsilon ^{3}).{}\end{array}$$
(3.122)

Equations (3.121) and (3.122) give a procedure to move the molecular dynamics one-step forward, (ϕ(s), π(sɛ∕2)) → (ϕ(s +ɛ), π(s +ɛ∕2)). The initial and final steps need to receive special care,

$$\displaystyle{ \pi (\varepsilon /2) =\pi (0) -\frac{1} {2}\varepsilon \frac{\delta S(\phi )} {\delta \phi (s)} + O(\varepsilon ^{2}),\ \ \ \pi (s_{\mathrm{ F}}) =\pi (s_{\mathrm{F}} -\varepsilon /2) -\frac{1} {2}\varepsilon \frac{\delta S(\phi )} {\delta \phi (s_{\mathrm{F}})} + O(\varepsilon ^{2}), }$$
(3.123)

which have only O(ε 2) accuracy. An illustration of this leapfrog integrator is shown in Fig. 3.16. Since the initial and final steps introduce O(ɛ 2) error irrespective of the length of the MD trajectory, and the intermediate steps introduce O(ɛ 3) ×ɛ −1 = O(ɛ 2) error as a whole, one finds Δ H = O(ɛ 2) after one MD trajectory before the Metropolis test.

The leapfrog integrator satisfies the reversibility and symplectic property, which can be checked explicitly by using the above definitions (Exercise 3.13).

Fig. 3.16
figure 16

The leapfrog integrator

Rights and permissions

Reprints and permissions

Copyright information

© 2017 Springer International Publishing AG

About this chapter

Cite this chapter

Hatsuda, T. (2017). Lattice Quantum Chromodynamics. In: Hjorth-Jensen, M., Lombardo, M., van Kolck, U. (eds) An Advanced Course in Computational Nuclear Physics. Lecture Notes in Physics, vol 936. Springer, Cham. https://doi.org/10.1007/978-3-319-53336-0_3

Download citation

Publish with us

Policies and ethics